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

Corresponding Abelian Extensions of Integrally Equivalent Number Fields

Shaver Phagan Department of Mathematics
Purdue University, West Lafayette, IN
[email protected]

Abstract

Extensive work has been done to determine necessary and sufficient conditions for a bijective correspondence of abelian extensions of number fields to force an isomorphism of the base fields. However, explicit examples of correspondences over non-isomorphic fields are rare. Integrally equivalent number fields admit an induced correspondence of abelian extensions. Studying this correspondence using idelic class field theory and linear algebra, we show that the corresponding extensions share features similar to those of arithmetically equivalent fields, and yet they are not generally weakly Kronecker equivalent. We also extend a group cohomological result of Arapura et. al. and present geometric and arithmetic applications.

1. Introduction and Background

1.1. Summary

D. Prasad noted an isomorphism of the idele class groups of some non-isomorphic number fields [Pra17], which induces a correspondence of their abelian extensions. This article is principally concerned with determining the arithmetic similarity of extensions corresponding under this bijection. In doing so, one is led to the notions of arithmetic, Kronecker, and weak Kronecker equivalence as means of comparing the splitting behavior of rational primes. In Section 2, we record the fact that a generic correspondence of abelian extensions of number fields cannot generally preserve arithmetic equivalence classes. Then, in Section 4, we prove that our corresponding extensions are not even weakly Kronecker equivalent in general, yet they share features similar to those of arithmetically equivalent fields. Some necessary technical facts on group schemes are recorded in Section 3. In Section 5, we conclude by extending a group cohomological result of Arapura et. al. This is motivated by a geometric application to manifolds with contractible universal cover, but the techniques shed further light on the arithmetic correspondence. We describe below some of the essential ingredients in this article, reviewing relevant facts and establishing context.

1.2. Corresponding Abelian Extensions

A triple of groups (G,G1,G2)(G,G_{1},G_{2}) is called an integral Gassmann triple if there is an isomorphism [G/G1][G/G2]\mathbb{Z}[G/G_{1}]\simeq\mathbb{Z}[G/G_{2}] of G\mathbb{Z}G-modules, where the GG-action is permutation of cosets induced by left multiplication. Let K/K/\mathbb{Q} be a Galois extension of number fields with group GG. For a subfield FF of KK, let GFG_{F} be the subgroup of GG fixing FF. Subfields K1K_{1} and K2K_{2} of KK are called integrally equivalent if (G,GK1,GK2)(G,G_{K_{1}},G_{K_{2}}) is an integral Gassmann triple. Integral equivalence of number fields K1K_{1} and K2K_{2} induces an isomorphism CK1CK2C_{K_{1}}\simeq C_{K_{2}} of idele class groups, and this gives a correspondence of their abelian extensions, by class field theory (c.f. Section 4). Letting ΔF\Delta_{F} be the discriminant of a number field FF, and Kn(F)K_{n}(F) its nthn^{th} KK-group (c.f. Section 1.4), corresponding abelian extensions of integrally equivalent number fields enjoy the following relations.

Theorem 1.1.

If Li/KiL_{i}/K_{i}, i=1,2i=1,2 are corresponding abelian extensions of integrally equivalent number fields, then, for m=[Li:Ki]m=[L_{i}:K_{i}],

  1. (1)

    L1,L2L_{1},L_{2} have the same degree, Galois closure, and maximal abelian subextension over \mathbb{Q}

  2. (2)

    L1L_{1} and L2L_{2} have the same signature for mm odd

  3. (3)

    Kn(L1)Kn(L2)K_{n}(L_{1})\simeq K_{n}(L_{2}) for odd n3n\geq 3 when mm is odd

  4. (4)

    ΔL1\Delta_{L_{1}} and ΔL2\Delta_{L_{2}} have the same prime divisors

  5. (5)

    There is a set 𝒫\mathcal{P} of rational primes of positive density, so that if ζLi(s)=+ai(n)n\zeta_{L_{i}}(s)=\sum_{\mathbb{N}_{+}}\frac{a_{i}(n)}{n}, and every prime divisor of nn is in 𝒫\mathcal{P}, then a1(n)=a2(n)a_{1}(n)=a_{2}(n) (c.f. Corollary 4.4). Furthermore, the LiL_{i} are ultra-coarsely arithmetically equivalent (c.f. Section 1.3).

Readers acquainted with the topic of arithmetic similarity (c.f. Section 1.3, and [Per77, Kli78, Jeh77, Loc95, Loc94b, Loc94a, Kli98]) might have already noted the similarity of Theorem 1.1 to the following.

Theorem 1.2.

[Per77] If F1,F2F_{1},F_{2} are arithmetically equivalent number fields, then

  1. (1)

    F1,F2F_{1},F_{2} have the same degree, Galois closure, and maximal Galois subextension over \mathbb{Q}

  2. (2)

    F1F_{1} and F2F_{2} have the same signature

  3. (3)

    Kn(F1)Kn(F2)K_{n}(F_{1})\simeq K_{n}(F_{2}) for odd n3n\geq 3

  4. (4)

    ΔF1=ΔF2\Delta_{F_{1}}=\Delta_{F_{2}}

  5. (5)

    ζF1=ζF2\zeta_{F_{1}}=\zeta_{F_{2}}.

This resemblance between arithmetically equivalent number fields and corresponding abelian extensions of integrally equivalent number fields is especially striking, since we show in Theorem 4.16 that the corresponding extensions are not even weakly Kronecker equivalent in general. On the other hand, integral equivalence appears to be a significantly stronger condition than arithmetic equivalence, and, since it induces a correspondence of abelian extensions, one might guess that certain qualities of arithmetic similarity are inherited.

A triple of groups (G,G1,G2)(G,G_{1},G_{2}) is called a Gassmann triple if there is an isomorphism of G\mathbb{Q}G-modules [G/G1][G/G2]\mathbb{Q}[G/G_{1}]\simeq\mathbb{Q}[G/G_{2}]. Number fields F1F_{1} and F2F_{2} are arithmetically equivalent if and only if (G,GF1,GF2)(G,G_{F_{1}},G_{F_{2}}) is a Gassmann triple, where GG is the group of the Galois closure of the FiF_{i}, i=1,2i=1,2, over the rational numbers. From the perspective of group (co)homology, the constituents G1,G2G_{1},G_{2} of an integral Gassmann triple (G,G1,G2)(G,G_{1},G_{2}) enjoy equality of (co)homology groups with trivial action on torsion coefficients, but this is not always so for an ordinary Gassmann triple (G,G1,G2)(G,G_{1},G_{2}) (c.f. Corollary 5.2 in this article, and [BP16]). Furthermore, there are several constructions which produce Gassmann triples (c.f. [Per77, Sun85]), but all known examples of integral Gassmann triples fundamentally arise from a nuclear triple, originally due to L. Scott [Sco93], and whose existence appears to the author a miracle. Scott noticed that G=PSL(2,𝔽29)G=\operatorname{PSL}(2,\mathbb{F}_{29}) has non-conjugate subgroups G1G_{1} and G2G_{2}, both isomorphic to A5A_{5}, such that (G,G1,G2)(G,G_{1},G_{2}) is an integral Gassmann triple. It was D. Zywina who then showed that GG is the Galois group of an extension of number fields [Zyw15], and D. Prasad who first noted that this implies the isomorphism of idele class groups of what we are calling integrally equivalent number fields [Pra17]. We recall some history motivating our study of the induced correspondence of abelian extensions.

For number fields K1,K2K_{1},K_{2} with absolute Galois groups Γ1,Γ2,\Gamma_{1},\Gamma_{2}, respectively, the celebrated Neukirch-Uchida Theorem [Uch76] says that Γ1Γ2\Gamma_{1}\simeq\Gamma_{2} if and only if K1K2K_{1}\simeq K_{2}, but it is known that an isomorphism of abelianized Galois groups Γ1abΓ2ab\Gamma_{1}^{ab}\simeq\Gamma_{2}^{ab} does not imply an isomorphism K1K2K_{1}\simeq K_{2}. Indeed, Γ1abΓ2ab\Gamma_{1}^{ab}\simeq\Gamma_{2}^{ab} for non-isomorphic integrally equivalent K1,K2K_{1},K_{2}, since ΓiabCKi^\Gamma_{i}^{ab}\simeq\widehat{C_{K_{i}}}, where CKiC_{K_{i}} is the idele class group, and the hat indicates profinite completion. Now, Cornelissen et. al. have devoted considerable effort to reconstructing global fields from their abelianized Galois groups and related data (c.f. [Cor13, CLMS17, CM14]), culminating in [CdSL+18], whose main theorem says the existence of an isomorphism Γ1abΓ2ab\Gamma_{1}^{ab}\simeq\Gamma_{2}^{ab} inducing an LL-function preserving correspondence of Dirichlet characters is equivalent to an isomorphism K1K2K_{1}\simeq K_{2}, for global fields K1K_{1} and K2K_{2}. Furthermore, they prove the following theorem in the number field case.

Theorem 1.3.

[CdSL+18] Given a number field KK and an integer m3m\geq 3, there is a character χ\chi of ΓKab\Gamma_{K}^{ab} of order mm such that if FF is a number field admitting a character χ\chi^{\prime} with LK(χ)=LF(χ)L_{K}(\chi)=L_{F}(\chi^{\prime}), then FKF\simeq K.

As a corollary, one can deduce that a number field is uniquely determined up to isomorphism by the set of Dedekind zeta functions of its finite abelian extensions, as observed by Solomatin in [Sol19]. We recall the argument below for the convenience of the reader (c.f. Section 2). This result suggests asking how arithmetically similar corresponding abelian extensions can be when the base fields are non-isomorphic. Despite the cited work of Cornelissen et. al., examples of correspondences of abelian extensions do not abound in the literature, so this converse question has not received much attention. We explore it here and provide an answer with Theorems 1.1 and 4.16.

Some final remarks on notation and terminology. Throughout the sequel, the symbol \prod is used to indicate a direct product, except in the proofs of Theorem 3.3 and Corollary 4.4. Furthermore, an integral Gassmann triple (Γ,Γ1,Γ2)(\Gamma,\Gamma_{1},\Gamma_{2}) is assumed to have [Γ:Γi]<[\Gamma:\Gamma_{i}]<\infty, and number fields K1,K2K_{1},K_{2} will be called integrally equivalent over FF if (G,GK1,GK2)(G,G_{K_{1}},G_{K_{2}}) is an integral Gassmann triple, where GG is the group of the Galois closure of Ki/FK_{i}/F, i=1,2i=1,2. Lastly, 00\in\mathbb{N}, and +={0}\mathbb{N}_{+}=\mathbb{N}-\{0\}.

1.3. Comparing Splitting Types

Let p𝒪K=𝔭1e1𝔭nenp\mathcal{O}_{K}=\mathfrak{p}_{1}^{e_{1}}\cdots\mathfrak{p}_{n}^{e_{n}} be the decomposition of the rational prime pp into prime ideals in the ring of integers of a number field KK. Define the splitting type SK(p)S_{K}(p) as the multiset of residues of KK over pp, that is SK(p)={{fi}}i=1nS_{K}(p)=\{\{f_{i}\}\}_{i=1}^{n}, with fi=[𝒪K/𝔭i:/p]f_{i}=[\mathcal{O}_{K}/\mathfrak{p}_{i}:\mathbb{Z}/p\mathbb{Z}]. Arithmetic equivalence of number fields F1,F2F_{1},F_{2} can be characterized as an equality SF1(p)=SF2(p)S_{F_{1}}(p)=S_{F_{2}}(p) for all but finitely many rational primes pp. Call F1,F2F_{1},F_{2} Kronecker equivalent if 1SF1(p) if and only if 1SF2(p)1\in S_{F_{1}}(p)\text{ if and only if }1\in S_{F_{2}}(p) for all but finitely many pp, and weakly Kronecker equivalent if gcd(SF1(p))=gcd(SF2(p))\text{gcd}(S_{F_{1}}(p))=\text{gcd}(S_{F_{2}}(p)) for all but finitely many pp. We record a useful theorem for reference later.

Theorem 1.4 ([Loc94a] Theorem 33^{\prime} ).

If F1F_{1} and F2F_{2} are weakly Kronecker equivalent number fields, then for any prime pp, gcd(SF1(p))=1\gcd(S_{F_{1}}(p))=1 if and only if gcd(SF2(p))=1\gcd(S_{F_{2}}(p))=1.

In fact, F1F_{1} and F2F_{2} are Kronecker equivalent if and only if 𝒩F1,p=𝒩F2,p\mathcal{N}_{F_{1},p}=\mathcal{N}_{F_{2},p} for all pp, where 𝒩Fi,p={jnjfj | nj,fjSFi(p)}\mathcal{N}_{F_{i},p}=\{\sum_{j}n_{j}f_{j}\text{ }|\text{ }n_{j}\in\mathbb{N},f_{j}\in S_{F_{i}}(p)\} [Loc95]. In particular, Kronecker equivalence implies weak Kronecker equivalence, and, clearly, arithmetic equivalence implies Kronecker equivalence. Furthermore, for Kronecker equivalent F1,F2F_{1},F_{2}, if ζFi(s)=n+ai(n)ns\zeta_{F_{i}}(s)=\sum_{n\in\mathbb{N}_{+}}\frac{a_{i}(n)}{n^{s}}, then a1(n)0a_{1}(n)\neq 0 if and only if a2(n)0a_{2}(n)\neq 0, so that Kronecker equivalence is a sort of approximate arithmetic equivalence. Weak Kronecker equivalence does not support a similar interpretation in a nice way, but a related notion does. We call number fields F1F_{1} and F2F_{2} such that lcm(SF1(p))=lcm(SF2(p))\operatorname{lcm}(S_{F_{1}}(p))=\operatorname{lcm}(S_{F_{2}}(p)) for all but finitely many pp ultra-coarsely arithmetically equivalent. This terminology is motivated by Proposition 1.6. As it turns out, number fields are ultra-coarsely arithmetically equivalent if and only if they have the same Galois closure.

Proposition 1.5.

Number fields K1,K2K_{1},K_{2} are ultra-coarsely arithmetically equivalent if and only if they have the same Galois closure over \mathbb{Q}.

Proof.

If K1,K2K_{1},K_{2} are ultra-coarsely arithmetically equivalent number fields, a rational prime splits completely in K1K_{1} if and only if it does so in K2K_{2}, with at most finitely many exceptions, so the Galois closure of K1/K_{1}/\mathbb{Q} is equal to that of K2/K_{2}/\mathbb{Q}. Now, suppose K1,K2K_{1},K_{2} are number fields with the same Galois closure K/K/\mathbb{Q}, and let G=Gal(K/)G=\operatorname{Gal}(K/\mathbb{Q}). Letting λi,p=lcm(SKi(p))\lambda_{i,p}=\operatorname{lcm}(S_{K_{i}}(p)), if pp is a rational prime unramified in KK with Frobenius class FpGF_{p}\subset G, we have λi,p|n\lambda_{i,p}|n if and only if FpnGKiF_{p}^{n}\subset G_{K_{i}}, which is equivalent to Fpn=1F_{p}^{n}=1, since KK is the Galois closure of KiK_{i}, and FpnF_{p}^{n} is a GG-conjugacy class. In particular, λ1,p=o(Fp)=λ2,p\lambda_{1,p}=o(F_{p})=\lambda_{2,p}. ∎

Proposition 1.6.

Suppose F1,F2F_{1},F_{2} are ultra-coarsely arithmetically equivalent number fields, and let SS be the finite set of rational primes pp such that lcm(SF1(p))lcm(SF2(p))\operatorname{lcm}(S_{F_{1}}(p))\neq\operatorname{lcm}(S_{F_{2}}(p)). If ζFi(s)=n+ai(n)ns\zeta_{F_{i}}(s)=\sum_{n\in\mathbb{N}_{+}}\frac{a_{i}(n)}{n^{s}}, and nn is a positive integer with no prime divisor in SS, we may write λp=lcm(SFi(p))\lambda_{p}=\operatorname{lcm}(S_{F_{i}}(p)) unambiguously for pp dividing nn, and, furthermore, if a1(n)0a_{1}(n)\neq 0, there is m+m\in\mathbb{N}_{+} with a2(m)0a_{2}(m)\neq 0 and nΛ1<m<nΛn\Lambda^{-1}<m<n\Lambda, where Λ=p1λp1pkλpk\Lambda=p_{1}^{\lambda_{p_{1}}}\cdots p_{k}^{\lambda_{p_{k}}} and p1,,pkp_{1},...,p_{k} are the prime divisors of nn.

Proof.

Given a prime pp not in SS, we know λp+𝒩Fi,p\lambda_{p}\mathbb{N}_{+}\subset\mathcal{N}_{F_{i},p}, i=1,2i=1,2, so if cp𝒩F1,pc_{p}\in\mathcal{N}_{F_{1},p}, there is dp𝒩F2,pd_{p}\in\mathcal{N}_{F_{2},p} such that |cpdp|<λp|c_{p}-d_{p}|<\lambda_{p}. If n=p1c1pkckn=p_{1}^{c_{1}}\cdots p_{k}^{c_{k}}, we know that cj𝒩F1,pjc_{j}\in\mathcal{N}_{F_{1},p_{j}}, so there is dj𝒩F2,pjd_{j}\in\mathcal{N}_{F_{2},p_{j}} such that |djcj|<λpj|d_{j}-c_{j}|<\lambda_{p_{j}}, for 1jk1\leq j\leq k. Setting m=p1d1pkdkm=p_{1}^{d_{1}}\cdots p_{k}^{d_{k}}, we know that mm is the absolute norm of some ideal in 𝒪F2\mathcal{O}_{F_{2}}, so that a2(m)0a_{2}(m)\neq 0, and furthermore, mn=p1d1c1pkdkck\frac{m}{n}=p_{1}^{d_{1}-c_{1}}\cdots p_{k}^{d_{k}-c_{k}}, so that Λ1<mn<Λ\Lambda^{-1}<\frac{m}{n}<\Lambda. ∎

1.4. KK-Groups of Odd Index

The KK-theory of a number field FF is a more contemporary invariant. We will restrict our attention here to groups Kn(F)K_{n}(F) with nn odd. We show that arithmetically equivalent fields have the same KK-groups with odd n3n\geq 3 and record some useful facts about the groups Kn(F)K_{n}(F) for the convenience of the reader. More details can be found in Section 5.3 of [Wei] or Chapter VI of [Wei13].

The KK-groups are abelian, and K1(F)=F×K_{1}(F)=F^{\times}. For odd n3n\geq 3, Kn(F)K_{n}(F) is a finitely generated group, given by the following rule.

Kn(F){r1+r2/wi,n1 mod 8r2(/2)r11/2wi,n3 mod 8r1+r2/12wi,n5 mod 8r2/wi,n7 mod 8K_{n}(F)\simeq\begin{cases}\mathbb{Z}^{r_{1}+r_{2}}\bigoplus\mathbb{Z}/w_{i}\mathbb{Z},&n\equiv 1\text{ mod }8\\ \mathbb{Z}^{r_{2}}\bigoplus(\mathbb{Z}/2\mathbb{Z})^{r_{1}-1}\bigoplus\mathbb{Z}/2w_{i}\mathbb{Z},&n\equiv 3\text{ mod }8\\ \mathbb{Z}^{r_{1}+r_{2}}\bigoplus\mathbb{Z}/\frac{1}{2}w_{i}\mathbb{Z},&n\equiv 5\text{ mod }8\\ \mathbb{Z}^{r_{2}}\bigoplus\mathbb{Z}/w_{i}\mathbb{Z},&n\equiv 7\text{ mod }8\end{cases}

where r1r_{1} (r2)(r_{2}) is the number of real (complex) places of FF, i=(n+1)/2i=(n+1)/2, and

(1) vp(wi)=max{ν | Gal(F(ζpν)/F) has exponent dividing i},v_{p}(w_{i})=\text{max}\{\nu\text{ }|\text{ Gal}(F(\zeta_{p^{\nu}})/F)\text{ has exponent dividing }i\},

ζr\zeta_{r} is a primitive rthr^{th} root of unity, and vpv_{p} is the pp-adic valuation associated to the rational prime pp.

Proposition 1.7.

If F1,F2F_{1},F_{2} are arithmetically equivalent number fields, then for odd n3n\geq 3, we have Kn(F1)Kn(F2)K_{n}(F_{1})\simeq K_{n}(F_{2}).

Proof.

Since the FiF_{i} are arithmetically equivalent, we know they have the same signature. Furthermore, given a primitive root of unity μ\mu, we know that F1(μ)=F2(μ)F_{1}\cap\mathbb{Q}(\mu)=F_{2}\cap\mathbb{Q}(\mu), so Gal(F1(μ)/F1)Gal(F2(μ)/F2)\text{Gal}(F_{1}(\mu)/F_{1})\simeq\text{Gal}(F_{2}(\mu)/F_{2}), since Gal(Fi(μ)/Fi)Gal((μ)/Fi(μ))\text{Gal}(F_{i}(\mu)/F_{i})\simeq\text{Gal}(\mathbb{Q}(\mu)/F_{i}\cap\mathbb{Q}(\mu)). ∎

1.5. Diagrams in (Co)Homology and Corresponding Abelian Covers

We say that manifolds M1,M2M_{1},M_{2} are integrally equivalent if there are coverings M1,M2MM_{1},M_{2}\rightarrow M with normal closure NMN\rightarrow M a GG-covering, where (G,G1,G2)(G,G_{1},G_{2}) is an integral Gassmann triple, and MiM_{i} is the subcover corresponding to GiG_{i}. Arapura et. al. construct non-isometric integrally equivalent closed hyperbolic manifolds [AKMS19]. Their results imply that for integrally equivalent manifolds M1,M2M_{1},M_{2} with contractible universal cover, there are isomorphisms in cohomology H(M1)H(M2)H^{*}(M_{1})\simeq H^{*}(M_{2}) natural with respect to restriction (corestriction) from (to) H(M)H^{*}(M). These already suggest a correspondence of the abelian covers when the manifolds are closed and orientable, by Poincare duality. We show that there are similar isomorphisms in homology and define a correspondence of the abelian covers of integrally equivalent manifolds M1M_{1} and M2M_{2}, as follows. Covers MiMiM_{i}^{\prime}\rightarrow M_{i} correspond if UiH1(Mi)U_{i}\subset H_{1}(M_{i}), i=1,2i=1,2 are finite index subgroups such that the isomorphism H1(M1)H1(M2)H_{1}(M_{1})\simeq H_{1}(M_{2}) restricts to an isomorphism U1U2U_{1}\simeq U_{2}, and π1(Mi)\pi_{1}(M_{i}^{\prime}) is the preimage of UiU_{i} under the projection π1(Mi)H1(Mi)\pi_{1}(M_{i})\rightarrow H_{1}(M_{i}). The diagrams in homology and cohomology are obtained using group (co)homology arguments 111We pursue group (co)homological relations afforded by integral Gassmann triples insofar as they elucidate our geometric correspondence, but there is much more to be said. Homological relations in the spirit of Stallings’ Theorem and applications thereof are explored in [GM23]. 222A conceivable approach to our study of corresponding abelian extensions would be to lead with the group (co)homological results from Section 5 and then define our correspondence using the isomorphism in first homology from Corollary 5.2, with (Γ,Γ1,Γ2)(\Gamma,\Gamma_{1},\Gamma_{2}) an integral Gassmann triple of absolute Galois groups of number fields. As we will see, the adelic approach taken in Section 4 allows us to easily circumvent the coprimality assumptions needed in Theorem 5.4 (c.f. Theorem 4.2). in Section 5, and the more geometrically flavored Theorem 1.8 is obtained as a quick corollary.

Theorem 1.8.

If MM is a manifold with contractible universal cover, and M1,M2MM_{1},M_{2}\rightarrow M are integrally equivalent degree dd covers, then corresponding abelian covers MiMiM_{i}^{\prime}\rightarrow M_{i}, i=1,2i=1,2 of degree dd^{\prime} have the same normal closure over MM when (d,d)=1(d,d^{\prime})=1.

Acknowledgements

I would like to thank my advisor Ben McReynolds for all his help, and I would like to thank Daniel Le for his helpful comments and a clarifying conversation about algebraic tori. I would also like to thank Milana Golich and Justin Katz for conversation on this work. Thanks also to Zachary Selk and Dustin Lee Enyeart for helpful comments that improved the quality of exposition.

2. Solomatin’s Theorem

This section is devoted to proving the following theorem. We take liberties with certain details of the exposition and proof but mostly follow [Sol19]. Given a group CC, we use Cˇ\check{C} to denote the character group of homomorphisms C×C\rightarrow\mathbb{C^{\times}}.

Theorem 2.1.

[Sol19] If ZFZ_{F} is the set of Dedekind zeta functions of finite abelian extensions of a number field FF, then ZK1=ZK2Z_{K_{1}}=Z_{K_{2}} if and only if K1K2K_{1}\simeq K_{2}.

Note that ZK1=ZK2Z_{K_{1}}=Z_{K_{2}} implies ζK1=ζK2\zeta_{K_{1}}=\zeta_{K_{2}}. Indeed, there is an abelian extension L2/K2L_{2}/K_{2} such that ζL2=ζK1\zeta_{L_{2}}=\zeta_{K_{1}}, but then ζK2=ζL1\zeta_{K_{2}}=\zeta_{L_{1}} for some abelian extension L1/K1L_{1}/K_{1}, so that

[L2:]=[K1:][L1:]=[K2:],[L_{2}:\mathbb{Q}]=[K_{1}:\mathbb{Q}]\leq[L_{1}:\mathbb{Q}]=[K_{2}:\mathbb{Q}],

and therefore L2=K2L_{2}=K_{2}. Before proving Theorem 2.1, we review preliminary results and set terminology.

Given a finite cyclic group CC and a Galois extension K/K/\mathbb{Q} of number fields with subfield FKF\subset K, letting G=Gal(K/)G=\operatorname{Gal}(K/\mathbb{Q}), we call KK^{\prime} a wreathing extension of K/F/K/F/\mathbb{Q} by CC if the following four conditions are satisfied.

  1. (1)

    KKK\subset K^{\prime}

  2. (2)

    K/K^{\prime}/\mathbb{Q} is Galois with group Gal(K/)=C[G/GF]G\operatorname{Gal}(K^{\prime}/\mathbb{Q})=C[G/G_{F}]\rtimes G

  3. (3)

    Gal(K/F)=C[G/GF]GF\operatorname{Gal}(K^{\prime}/F)=C[G/G_{F}]\rtimes G_{F}

  4. (4)

    Gal(K/K)=C[G:GF]\operatorname{Gal}(K^{\prime}/K)=C^{[G:G_{F}]}

where the action of the semi-direct product is permutation of cosets according to left multiplication. When knowledge of KK is either implicit or unnecessary, we refer to KK^{\prime} as a wreathing extension of FF by CC.

Theorem 2.2.

[CdSL+18] Given a finite Galois extension K/K/\mathbb{Q} of number fields, a subfield FKF\subset K, and a finite cyclic group CC, there is a wreathing extension of K/F/K/F/\mathbb{Q} by CC.

Given a wreathing extension KK^{\prime} of K/F/K/F/\mathbb{Q} by CC, set G=Gal(K/)G^{\prime}=\operatorname{Gal}(K^{\prime}/\mathbb{Q}), so that GF=C[G/GF]GFG^{\prime}_{F}=C[G/G_{F}]\rtimes G_{F}, and arrange so that the first coordinate of C[G/GF]C[G/G_{F}] is fixed by GFG_{F} (i.e. the first coordinate corresponds to the coset GFG/GFG_{F}\in G/G_{F}). Letting n=[G:GF]n=[G:G_{F}], to KK^{\prime} we associate the morphism

χK:GFC:(c1,,cn,g)c1.\chi_{K^{\prime}}:G^{\prime}_{F}\rightarrow C:(c_{1},...,c_{n},g)\mapsto c_{1}.

If C=/mC=\mathbb{Z}/m\mathbb{Z}, a homomorphism φ:C{e2πik/m}k=0m1×\varphi:C\rightarrow\{e^{2\pi ik/m}\}_{k=0}^{m-1}\subset\mathbb{C}^{\times} determines a one-dimensional representation φχK\varphi\circ\chi_{K^{\prime}} of GFG^{\prime}_{F}, which we denote χφ\chi_{\varphi}. Note that each χφ\chi_{\varphi} gives a Dirichlet character of ΓF\Gamma_{F} via the composition

ΓFGF/ker(χφ)χφ¯×,\Gamma_{F}\rightarrow G_{F}^{\prime}/\ker(\chi_{\varphi})\xrightarrow{\overline{\chi_{\varphi}}}\mathbb{C}^{\times},

which we also refer to as χφ\chi_{\varphi}. We will further abuse notation and identify χK\chi_{K^{\prime}} with χφ\chi_{\varphi^{\prime}}, where φ(k)=e2πik/m\varphi^{\prime}(k)=e^{2\pi ik/m}. Two lemmas are needed for the proof of the theorem.

Lemma 2.3.

If KK^{\prime} is a wreathing extension of a number field FF by a finite cyclic group of order greater than 22, and FF^{\prime} is a number field admitting a Dirichlet character χ\chi^{\prime} satisfying LF(χ)=LF(χK)L_{F^{\prime}}(\chi^{\prime})=L_{F}(\chi_{K^{\prime}}), then FFF\simeq F^{\prime}.

Proof.

Follows from the proof of Theorem 10.1 in [CdSL+18]. ∎

Lemma 2.4.

If KK^{\prime} is a wreathing extension of a number field FF by a finite cyclic group CC, and φCˇ\varphi\in\check{C} is non-trivial, then the induced representation V=IndGFG(χφ)V=\operatorname{Ind}_{G^{\prime}_{F}}^{G^{\prime}}(\chi_{\varphi}) is irreducible.

Proof.

We use ee to denote the identity element of GG. Let g1,,gnGg_{1},...,g_{n}\in G be a complete set of representatives for G/GFG/G_{F}, and assume without loss of generality that g1GFg_{1}\in G_{F}. Then gi=(0,gi)g_{i}^{\prime}=(0,g_{i}), i=1,,ni=1,...,n, is a complete set of coset representatives for G/GFG^{\prime}/G^{\prime}_{F}, and VV decomposes as

V=i=1ngi.V=\bigoplus\limits_{i=1}^{n}g^{\prime}_{i}\mathbb{C}.

We argue that the action of G\mathbb{C}G^{\prime} on VV is transitive by showing g1G.vg^{\prime}_{1}\in\mathbb{C}G^{\prime}.v for 0vV0\neq v\in V. Recall that we arrange so that the first coordinate of C[G/GF]C[G/G_{F}] is fixed by GFG_{F}, and let λ=χφ(1)\lambda=\chi_{\varphi}(1). Note that if v0v\neq 0, there is gGg\in G^{\prime} such that the coefficient of g1g^{\prime}_{1} in g.vg.v is non-zero, so it suffices to prove the claim under the assumption v=igiciv=\sum_{i}g^{\prime}_{i}c_{i} with c10c_{1}\neq 0. Under this assumption, let v=(1,0,,0,e).vv^{\prime}=(1,0,...,0,e).v. Compute v=λg1c1+i=2ngiciv^{\prime}=\lambda g^{\prime}_{1}c_{1}+\sum_{i=2}^{n}g^{\prime}_{i}c_{i}, so that (vv)=(1λ)g1c1(v-v^{\prime})=(1-\lambda)g^{\prime}_{1}c_{1}. Since φ\varphi is not trivial, we know λ1\lambda\neq 1, and hence g1G.vg^{\prime}_{1}\in\mathbb{C}G^{\prime}.v. ∎

We may now prove Solomatin’s Theorem. Any properties of Artin LL-functions used below can be found in Lemma 10.2 in [CdSL+18].

Proof of Theorem 2.1.

We need only show that ZK1=ZK2Z_{K_{1}}=Z_{K_{2}} implies K1K2K_{1}\simeq K_{2}, as the converse is obvious. Let K1K_{1} be a number field and KK^{\prime} a wreathing extension of K1K_{1} by a cyclic group C1C_{1} of order m3m\geq 3, and take L1/K1L_{1}/K_{1} to be the abelian extension given by GL1=ker(χK)G^{\prime}_{L_{1}}=\ker(\chi_{K^{\prime}}), where G=Gal(K/)G^{\prime}=\operatorname{Gal}(K^{\prime}/\mathbb{Q}). Observe that Gal(L1/K1)C1\operatorname{Gal}(L_{1}/K_{1})\simeq C_{1} via χK\chi_{K^{\prime}}. Now suppose ZK1=ZK2Z_{K_{1}}=Z_{K_{2}} for a number field K2K_{2}, so that ζK1=ζK2\zeta_{K_{1}}=\zeta_{K_{2}} and ζL1=ζL2\zeta_{L_{1}}=\zeta_{L_{2}} for some abelian extension L2/K2L_{2}/K_{2}. Letting 11 denote the trivial representation and ρF=IndGFG\rho_{F}=\operatorname{Ind}_{G^{\prime}_{F}}^{G^{\prime}}, we know that

ζKi=LKi(1)=L(ρKi(1)),\zeta_{K_{i}}=L_{K_{i}}(1)=L_{\mathbb{Q}}(\rho_{K_{i}}(1)),

so that L(ρK1(1))=L(ρK2(1))L_{\mathbb{Q}}\left(\rho_{K_{1}}(1)\right)=L_{\mathbb{Q}}\left(\rho_{K_{2}}(1)\right) and hence

(2) ρK1(1)ρK2(1).\rho_{K_{1}}(1)\simeq\rho_{K_{2}}(1).

Similarly, ζL1=ζL2\zeta_{L_{1}}=\zeta_{L_{2}} implies

(3) ρL1(1)ρL2(1).\rho_{L_{1}}(1)\simeq\rho_{L_{2}}(1).

Now, ρLi(1)=ρKi(ΛLi/Ki)\rho_{L_{i}}(1)=\rho_{K_{i}}(\Lambda_{L_{i}/K_{i}}), where ΛLi/Ki\Lambda_{L_{i}/K_{i}} is the permutation representation given by the left multiplication action of GKiG_{K_{i}}^{\prime} on GKi/GLiG_{K_{i}}^{\prime}/G_{L_{i}}^{\prime}, which factors through the left regular representation of Gal(Li/Ki)\operatorname{Gal}(L_{i}/K_{i}). Letting C2=Gal(L2/K2)C_{2}=\operatorname{Gal}(L_{2}/K_{2}), equation 3 can therefore be rewritten

φCˇ1ρK1(χφ)ψCˇ2ρK2(ψ),\bigoplus_{\varphi\in\check{C}_{1}}\rho_{K_{1}}(\chi_{\varphi})\simeq\bigoplus_{\psi\in\check{C}_{2}}\rho_{K_{2}}(\psi^{\prime}),

which, along with equation 2 implies

(4) 1φCˇ1ρK1(χφ)1ψCˇ2ρK2(ψ),\bigoplus_{1\neq\varphi\in\check{C}_{1}}\rho_{K_{1}}(\chi_{\varphi})\simeq\bigoplus_{1\neq\psi\in\check{C}_{2}}\rho_{K_{2}}(\psi^{\prime}),

where ψ\psi^{\prime} is the character GK2×G^{\prime}_{K_{2}}\rightarrow\mathbb{C}^{\times} induced by ψC2ˇ\psi\in\check{C_{2}}. Now, one has |C2|=|C1|=m|C_{2}|=|C_{1}|=m, so that either side of equation 4 has m1m-1 direct summands, and by Lemma 2.4, each summand on the left hand side is an irreducible representation, so that the right hand side of equation 4 also consists of irreducible representations. Since ρK1(χK)\rho_{K_{1}}(\chi_{K^{\prime}}) is a summand of the left hand side of equation 4, we conclude there is ψCˇ2\psi\in\check{C}_{2} such that ρK1(χK)ρK2(ψ)\rho_{K_{1}}(\chi_{K^{\prime}})\simeq\rho_{K_{2}}(\psi^{\prime}). But then

LK1(χK)=L(ρK1(χK))=L(ρK2(ψ))=LK2(ψ),L_{K_{1}}(\chi_{K^{\prime}})=L_{\mathbb{Q}}(\rho_{K_{1}}(\chi_{K^{\prime}}))=L_{\mathbb{Q}}(\rho_{K_{2}}(\psi^{\prime}))=L_{K_{2}}(\psi^{\prime}),

so that K1K2K_{1}\simeq K_{2}, by Lemma 2.3. ∎

3. Quasi-split Tori and Idele Norm

In this section, we prove that taking adelic points of a certain group scheme recovers the idele norm. This result is likely known to experts, but for the sake of completeness and for lack of a suitable reference, we include a proof. We write KsK^{s} for the separable closure of a field KK, and 𝔾m\mathbb{G}_{m} will denote the multiplicative group scheme. We are interested in relating Weyl restriction of 𝔾m\mathbb{G}_{m} to the idele norm. For further discussion of the lemmas, see Chapter 2 in [PR94]. Given a subset SS of an abelian group, we will use the notation S\sum S to denote the sum sSs\sum_{s\in S}s. We will also write γg\gamma^{g} in place of g1γg,g^{-1}\gamma g, given elements γ,g\gamma,g of a group GG.

Let KK be a local or global field. It is well-known that finite-dimensional KK-tori correspond to finitely generated free abelian groups, equipped with an action of 𝒢=Gal(Ks/K)\mathcal{G}=\text{Gal}(K^{s}/K), via a contravariant equivalence of categories. In particular, the group-module corresponding to a KK-torus TT is its character group Hom(TKs,𝔾m)\text{Hom}(T_{K^{s}},\mathbb{G}_{m}). Since TT is finite-dimensional over KK, it splits over a finite Galois extension E/KE/K. Furthermore, if TT is a so-called quasi-split torus and G=Gal(E/K)G=\text{Gal}(E/K), there is a finite GG-set AA and a GG-equivariant isomorphism TKs𝔾mAT_{K^{s}}\simeq\mathbb{G}_{m}^{A}, where the right hand side is a direct product of copies of 𝔾m\mathbb{G}_{m}, indexed by AA with GG-action given by permutation of coordinates. This induces an action of 𝒢\mathcal{G} on the character group. The set of quasi-split tori is precisely the set of products of Weyl restrictions of 𝔾m\mathbb{G}_{m} (c.f. Chapter 2 in [PR94]). A straightforward adaptation of the proof of Theorem 7.5 in [Wat79] yields the following.

Lemma 3.1.

Given E/KE/K Galois with group GG and subextension K/KK^{\prime}/K, letting, Ω=G/GK\Omega=G/G_{K^{\prime}} and T=ResK/K(𝔾m)T^{\prime}=\text{Res}_{K^{\prime}/K}(\mathbb{G}_{m}), the character group of TT^{\prime} is Ω\mathbb{Z}\Omega.

We will need an additional observation before proving the main theorem. Let α=Ω\alpha=\sum\Omega, and let N:T𝔾mN:T^{\prime}\rightarrow\mathbb{G}_{m} be the KK-scheme corresponding to the morphism ι:Ω:1α\iota:\mathbb{Z}\rightarrow\mathbb{Z}\Omega:1\mapsto\alpha. The KsK^{s}-form of NN has associated Hopf algebra morphism

NKs:Ks[X,X1]Ks[Xi,Xi1]:XX1Xn,N^{*}_{K^{s}}:K^{s}[X,X^{-1}]\rightarrow K^{s}[X_{i},X_{i}^{-1}]:X\mapsto X_{1}\cdots X_{n},

where XiX_{i} is the coordinate given by g¯i\overline{g}_{i}, and g1,,gng_{1},...,g_{n} is a full set of coset representatives for G/GKG/G_{K^{\prime}}, so that taking KK-points of NN recovers the usual field norm. In summary:

Lemma 3.2.

The KK-scheme NN corresponding to the inclusion of modules ι\iota has KK-points given by the field norm N(K)=NK/K:(K)×K×N(K)=N_{K^{\prime}/K}:(K^{\prime})^{\times}\rightarrow K^{\times}.

With the lemmas recorded, we are ready to state and prove the theorem.

Theorem 3.3.

Let E/KE/K be a Galois extension of number fields with group GG, and suppose GKGG_{K^{\prime}}\subset G is the stabilizer of the subfield KEK^{\prime}\subset E, Ω=G/GK\Omega=G/G_{K^{\prime}}, and α=Ω\alpha=\sum\Omega. If ι\iota is the morphism Ω\mathbb{Z}\rightarrow\mathbb{Z}\Omega of GG-modules given by 1α1\mapsto\alpha, and NN is the scheme corresponding to ι\iota, then N(𝔸K):𝕀K𝕀KN(\mathbb{A}_{K}):\mathbb{I}_{K^{\prime}}\rightarrow\mathbb{I}_{K} is the idele norm.

Proof.

It suffices to check componentwise, so we need to compute N(Kν)N(K_{\nu}) for a place ν\nu of KK. To do so, we view Ω\mathbb{Z}\Omega as a 𝒟ω/ν\mathcal{D}_{\omega/\nu}-module, where ω\omega is a place of EE over ν\nu, and 𝒟ω/νG\mathcal{D}_{\omega/\nu}\subset G is the associated decomposition group. Let g1,,gmg_{1},...,g_{m} be a complete set of representatives in GG for the double cosets 𝒟ω/ν\G/GK\mathcal{D}_{\omega/\nu}\backslash G/G_{K^{\prime}}. If Ωi\Omega_{i} is the 𝒟ω/ν\mathcal{D}_{\omega/\nu}-orbit of g¯i\overline{g}_{i} in Ω\Omega, then Ω=iΩi\mathbb{Z}\Omega=\bigoplus_{i}\mathbb{Z}\Omega_{i}. Furthermore, if αi=Ωi\alpha_{i}=\sum\Omega_{i}, then α=iαi\alpha=\sum_{i}\alpha_{i}, so that, if ιi\iota_{i} is the morphism Ωi:1αi\mathbb{Z}\rightarrow\mathbb{Z}\Omega_{i}:1\mapsto\alpha_{i}, we have the decomposition

ι=i=1mιi.\iota=\sum\limits_{i=1}^{m}\iota_{i}.

Therefore, letting NiN_{i} be the KνK_{\nu}-scheme corresponding to ιi\iota_{i}, we have

N(Kν)=i=1mNi(Kν).N(K_{\nu})=\prod\limits_{i=1}^{m}N_{i}(K_{\nu}).

It thus remains to determine Ni(Kν)N_{i}(K_{\nu}). Now, it is well known that each gig_{i} corresponds to a place ηi\eta_{i} of KK^{\prime} over ν\nu. More explicitly, we can let ωi=gi1ω\omega_{i}=g_{i}^{-1}\omega and take ηi\eta_{i} to be the place of KK^{\prime} divided by ωi\omega_{i}, so that gi1𝒟ω/νgi=𝒟ωi/νg_{i}^{-1}\mathcal{D}_{\omega/\nu}g_{i}=\mathcal{D}_{\omega_{i}/\nu} and 𝒟ωi/ηi=𝒟ωi/νGK\mathcal{D}_{\omega_{i}/\eta_{i}}=\mathcal{D}_{\omega_{i}/\nu}\cap G_{K^{\prime}}. Letting Ωi=𝒟ωi/ν/𝒟ωi/ηi\Omega_{i}^{\prime}=\mathcal{D}_{\omega_{i}/\nu}/\mathcal{D}_{\omega_{i}/\eta_{i}}, observe that the maps

𝒟ω/ν𝒟ωi/ν:γγgi\mathcal{D}_{\omega/\nu}\rightarrow\mathcal{D}_{\omega_{i}/\nu}:\gamma\mapsto\gamma^{g_{i}}
ΩiΩi:γg¯iγgi¯\Omega_{i}\rightarrow\Omega_{i}^{\prime}:\gamma\overline{g}_{i}\mapsto\overline{\gamma^{g_{i}}}

determine an isomorphism of Galois representations (𝒟ω/ν,Ωi)(𝒟ωi/ν,Ωi)(\mathcal{D}_{\omega/\nu},\Omega_{i})\rightarrow(\mathcal{D}_{\omega_{i}/\nu},\Omega_{i}^{\prime}) allowing an identification Ni(Kν)=Ni(Kν)N_{i}(K_{\nu})=N_{i}^{\prime}(K_{\nu}), where NiN_{i}^{\prime} is the KνK_{\nu}-scheme corresponding to the morphism Ωi:1Ωi\mathbb{Z}\rightarrow\mathbb{Z}\Omega_{i}^{\prime}:1\mapsto\sum\Omega_{i}^{\prime}. By Lemma 3.2, we know that Ni(Kν)N_{i}^{\prime}(K_{\nu}) is the field norm (Kηi)×Kν×(K_{\eta_{i}}^{\prime})^{\times}\rightarrow K_{\nu}^{\times}. It follows that N(Kν)N(K_{\nu}) is the component over ν\nu of the idele norm 𝕀K𝕀K\mathbb{I}_{K^{\prime}}\rightarrow\mathbb{I}_{K}. ∎

4. Corresponding Abelian Extensions

Let K1K_{1} and K2K_{2} be number fields integrally equivalent over FF. Writing Ti=T_{i}=Res(𝔾m)Ki/F{}_{K_{i}/F}(\mathbb{G}_{m}), the isomorphism [G/GK1][G/GK2]\mathbb{Z}[G/G_{K_{1}}]\simeq\mathbb{Z}[G/G_{K_{2}}] gives an isomorphism T1T2T_{1}\simeq T_{2}, by Lemma 3.1. The FF-points of the TiT_{i} are therefore isomorphic K1×K2×K_{1}^{\times}\simeq K_{2}^{\times}, as are the 𝔸F\mathbb{A}_{F}-points 𝕀K1𝕀K2\mathbb{I}_{K_{1}}\simeq\mathbb{I}_{K_{2}}, where 𝔸F\mathbb{A}_{F} denotes the ring of FF-adeles. Furthermore, the isomorphism of ideles respects the diagonal embeddings Ki×𝕀KiK_{i}^{\times}\rightarrow\mathbb{I}_{K_{i}}, so we can quotient to get an isomorphism of idele class groups φ:CK1CK2\varphi:C_{K_{1}}\rightarrow C_{K_{2}} [Pra17]. By class field theory, a finite abelian extension Li/KiL_{i}/K_{i} is uniquely determined by the finite-index open subgroup of CKiC_{K_{i}} given by the image of the idele class norm NLi/KiN_{L_{i}/K_{i}}. Moreover, every finite-index open subgroup of CKiC_{K_{i}} is such an image:

Li/KiNLi/Ki(CLi)CKi.L_{i}/K_{i}\leftrightarrow N_{L_{i}/K_{i}}(C_{L_{i}})\subset C_{K_{i}}.

We say that L1/K1L_{1}/K_{1} and L2/K2L_{2}/K_{2} correspond if NL2/K2(CL2)=φ(NL1/K1(CL1))N_{L_{2}/K_{2}}(C_{L_{2}})=\varphi(N_{L_{1}/K_{1}}(C_{L_{1}})). Observe that φ\varphi induces an isomorphism of Galois groups Gal(L1/K1)Gal(L2/K2)\text{Gal}(L_{1}/K_{1})\rightarrow\text{Gal}(L_{2}/K_{2}), so that, in particular, [L1:K1]=[L2:K2][L_{1}:K_{1}]=[L_{2}:K_{2}], which implies [L1:F]=[L2:F][L_{1}:F]=[L_{2}:F]. We state the following proposition without proof, as it is an immediate consequence of our definition of corresponding abelian extensions.

Proposition 4.1.

Suppose Li/KiL_{i}/K_{i}, i=1,2i=1,2 correspond, and Li/Ki,i=1,2L_{i}^{\prime}/K_{i},i=1,2 correspond. Then LiLi/KiL_{i}L_{i}^{\prime}/K_{i}, i=1,2i=1,2 correspond and LiLi/KiL_{i}\cap L_{i}^{\prime}/K_{i}, i=1,2i=1,2 correspond.

The remainder of this section is devoted to fleshing out the relations between corresponding L1L_{1} and L2L_{2}. Specifically, in Section 4.1, we prove Theorem 1.1, and then in Section 4.2, we demonstrate that the LiL_{i} are not even weakly Kronecker equivalent in general. Throughout the sequel, Li/KiL_{i}/K_{i}, i=1,2i=1,2 will denote corresponding abelian extensions of integrally equivalent number fields. We will also freely identify Ki,ν×\prod K_{i,\nu}^{\times} (product over some finite set of places) with its image under the projection 𝕀KiCKi\mathbb{I}_{K_{i}}\rightarrow C_{K_{i}}. Whether we are working in the idele group or the idele class group should be clear from context. Lastly, given an abelian extension Li/KiL_{i}/K_{i} and a place ν\nu of KiK_{i}, we will refer to the completion of LiL_{i} at a place over ν\nu simply as Li,νL_{i,\nu}, since Li,η1Li,η2L_{i,\eta_{1}}\simeq L_{i,\eta_{2}} for places η1,η2\eta_{1},\eta_{2} of LiL_{i} over ν\nu.

4.1. Arithmetic Similarity

Given a place ω\omega of FF, all direct products from here on are over the places ν\nu (or η\eta) of KiK_{i} dividing ω\omega, unless otherwise stated. We know that Ti(Fω)=Ki,ν×T_{i}(F_{\omega})=\prod K_{i,\nu}^{\times}, so applying T1T2T_{1}\simeq T_{2} to FωF_{\omega} gives an isomorphism

(5) φω:ν|ωK1,ν×η|ωK2,η×.\varphi_{\omega}:\prod\limits_{\nu|\omega}K_{1,\nu}^{\times}\rightarrow\prod\limits_{\eta|\omega}K_{2,\eta}^{\times}.

Define Ni,ωCKiN_{i,\omega}\subset C_{K_{i}} by Ni,ω=NLi/Ki(CLi)Ki,ν×N_{i,\omega}=N_{L_{i}/K_{i}}(C_{L_{i}})\cap\prod K_{i,\nu}^{\times}. Recall that

(6) Ki,ν×/(Ni,ωKi,ν×)Gal(Li,ν/Ki,ν)K_{i,\nu}^{\times}/(N_{i,\omega}\cap K_{i,\nu}^{\times})\simeq\text{Gal}(L_{i,\nu}/K_{i,\nu})

and

(7) |𝒪Ki,ν×/(Ni,ω𝒪Ki,ν×)|=e(Li,ν/Ki,ν).|\mathcal{O}_{K_{i,\nu}}^{\times}/(N_{i,\omega}\cap\mathcal{O}_{K_{i,\nu}}^{\times})|=e(L_{i,\nu}/K_{i,\nu}).

See Chapter X in [Tat08] and Chapter V in [Ser79] for details. The functions φω\varphi_{\omega} provide a means of probing the LiL_{i} locally. This proves fruitful, allowing us to compare, respectively, ζLi\zeta_{L_{i}}, ΔLi\Delta_{L_{i}}, and the Galois closure of Li/L_{i}/\mathbb{Q}, i=1,2i=1,2. Then, in Theorem 4.9 we will make use of the Norm Limitation Theorem from global class field theory to conclude that corresponding extensions have the same maximal abelian sub-extension over FF, which allows us to relate their KK-groups. We begin with an observation shedding light on both the Galois closures and the ζLi\zeta_{L_{i}}.

Proposition 4.2.

If K1,K2K_{1},K_{2} are integrally equivalent number fields over FF, with corresponding abelian extensions Li/KiL_{i}/K_{i}, i=1,2i=1,2, then L1/FL_{1}/F and L2/FL_{2}/F have the same Galois closure.

Proof.

Recall that the Galois closure of an extension of number fields E/FE/F is uniquely determined by the set of primes in FF which split completely in EE. Thus, it suffices to show that a prime ω\omega of FF splits completely in L1L_{1} if and only if it does so in L2L_{2}. Now, ω\omega splits completely in LiL_{i} if and only if it splits completely in KiK_{i} and Gal(Li,ν/Ki,ν)(L_{i,\nu}/K_{i,\nu}) is trivial for each place ν\nu of KiK_{i} over ω\omega. But then, since K1K_{1} and K2K_{2} are integrally equivalent over FF, we know that ω\omega splits completely in K1K_{1} if and only if it does so in K2K_{2}. Furthermore, by equation 6, Gal(Li,ν/Ki,ν)=1(L_{i,\nu}/K_{i,\nu})=1 for every ν\nu over ω\omega if and only if Ni,ω=Ki,ν×N_{i,\omega}=\prod K_{i,\nu}^{\times}. Since φω\varphi_{\omega} in equation 5 restricts to an isomorphism N1,ωN2,ωN_{1,\omega}\rightarrow N_{2,\omega}, we know that N1,ω=K1,ν×N_{1,\omega}=\prod K_{1,\nu}^{\times} if and only if N2,ω=K2,η×N_{2,\omega}=\prod K_{2,\eta}^{\times}. Thus ω\omega splits completely in L1L_{1} if and only if it does so in L2L_{2}, so the LiL_{i} indeed have the same Galois closure over FF. ∎

Corollary 4.3.

Corresponding abelian extensions are ultra-coarsely arithmetically equivalent in the sense of Proposition 1.6.

Before stating and proving the next corollary, we fix some terminology. Say that a rational prime pp splits relatively completely in Li/KiL_{i}/K_{i} if every prime of KiK_{i} over pp splits completely in LiL_{i}, and observe that arguments in the proof of Proposition 4.2 guarantee that a rational prime splits relatively completely in L1/K1L_{1}/K_{1} if and only if it does so in L2/K2L_{2}/K_{2}.

Corollary 4.4.

If ζLi(s)=n+ai(n)ns\zeta_{L_{i}}(s)=\sum_{n\in\mathbb{N}_{+}}\frac{a_{i}(n)}{n^{s}}, and nn is an integer whose prime divisors all split relatively completely in the Li/KiL_{i}/K_{i}, i=1,2i=1,2, then a1(n)=a2(n)a_{1}(n)=a_{2}(n).

Proof.

Let 𝒫\mathcal{P} consist of those rational primes which split relatively completely in the Li/KiL_{i}/K_{i}, i=1,2i=1,2. For Re(s)>1\text{Re}(s)>1, let

ζLi,𝒫(s)=𝔭|p𝒫(1N(𝔭)s)1,\zeta_{L_{i},\mathcal{P}}(s)=\prod\limits_{\mathfrak{p}|p\in\mathcal{P}}(1-N(\mathfrak{p})^{-s})^{-1},

where the primes 𝔭\mathfrak{p} are in 𝒪Li\mathcal{O}_{L_{i}}. A positive integer ff is in SLi(p)S_{L_{i}}(p) with multiplicity equal to the number of prime ideals in 𝒪Li\mathcal{O}_{L_{i}} with absolute norm pfp^{f}, and ζLi,𝒫\zeta_{L_{i},\mathcal{P}} counts the number of ideals in 𝒪Li\mathcal{O}_{L_{i}} whose absolute norm is only divisible by primes in 𝒫\mathcal{P}. For p𝒫p\in\mathcal{P}, SL1(p)=SL2(p)S_{L_{1}}(p)=S_{L_{2}}(p), since SK1(p)=SK2(p)S_{K_{1}}(p)=S_{K_{2}}(p) and pp splits relatively completely in Li/KiL_{i}/K_{i}, i=1,2i=1,2. Therefore, we have ζL1,𝒫=ζL2,𝒫\zeta_{L_{1},\mathcal{P}}=\zeta_{L_{2},\mathcal{P}}. Now, if every prime divisor of nn is in 𝒫\mathcal{P}, then ai(n)a_{i}(n) is the coefficient of nsn^{-s} in ζLi,𝒫\zeta_{L_{i},\mathcal{P}}, and we conclude a1(n)=a2(n)a_{1}(n)=a_{2}(n). ∎

Remark 4.5.

By Grunwald-Wang, given a finite set SS of rational primes, the Li/KiL_{i}/K_{i} can be chosen to split relatively completely over each pSp\in S. In particular, given M+M\in\mathbb{N}_{+}, there are corresponding abelian extensions L1,L2L_{1},L_{2} such that a1(n)=a2(n)a_{1}(n)=a_{2}(n) for any n+n\in\mathbb{N}_{+} with nMn\leq M.

Arithmetically equivalent K1,K2K_{1},K_{2} have the same signature and contain the same roots of unity, so there is an abstract isomorphism 𝒪K1×𝒪K2×\mathcal{O}_{K_{1}}^{\times}\simeq\mathcal{O}_{K_{2}}^{\times}. For integrally equivalent K1,K2K_{1},K_{2}, it turns out that the isomorphism φ:K1×K2×\varphi:K_{1}^{\times}\rightarrow K_{2}^{\times} in fact restricts to an isomorphism 𝒪K1×𝒪K2×\mathcal{O}_{K_{1}}^{\times}\simeq\mathcal{O}_{K_{2}}^{\times}. The keystone is Proposition 4.6, which may also be leveraged to relate the discriminants of the LiL_{i}.

Proposition 4.6.

φω\varphi_{\omega} restricts to an isomorphism 𝒪1,ν×𝒪2,η×\prod\limits\mathcal{O}_{1,\nu}^{\times}\simeq\prod\limits\mathcal{O}_{2,\eta}^{\times}.

Proof.

Let nn be the number of distinct prime divisors of ω\omega in KiK_{i} (note this is independent of whether i=1i=1 or i=2i=2), Hi=Ki,ν×H_{i}=\prod\limits K_{i,\nu}^{\times}, Ui=𝒪i,ν×U_{i}=\prod\limits\mathcal{O}_{i,\nu}^{\times}. We know HinUiH_{i}\simeq\mathbb{Z}^{n}\bigoplus U_{i}, so H1/U1nH_{1}/U_{1}\simeq\mathbb{Z}^{n}, and therefore H2/φω(U1)nH_{2}/\varphi_{\omega}(U_{1})\simeq\mathbb{Z}^{n}. But then, letting π\pi denote the projection H2H2/φω(U1)H_{2}\rightarrow H_{2}/\varphi_{\omega}(U_{1}), we know that π(H2)π(n)π(U2)\pi(H_{2})\simeq\pi(\mathbb{Z}^{n})\pi(U_{2}) is free abelian of rank nn and π(U2)U2/U2φω(U1)\pi(U_{2})\simeq U_{2}/U_{2}\cap\varphi_{\omega}(U_{1}). Now, the UiU_{i} are virtually pro-pp, so they have no nontrivial free abelian quotients. Therefore, π(U2)=0\pi(U_{2})=0, so U2φω(U1)U_{2}\subset\varphi_{\omega}(U_{1}). The inclusion φω(U1)U2\varphi_{\omega}(U_{1})\subset U_{2} is a consequence of the decomposition of H2H_{2} above, because otherwise U1U_{1} would surject a nontrivial free abelian group. ∎

Corollary 4.7.

The isomorphism K1×K2×K_{1}^{\times}\simeq K_{2}^{\times} restricts to an isomorphism 𝒪K1×𝒪K2×\mathcal{O}_{K_{1}}^{\times}\simeq\mathcal{O}_{K_{2}}^{\times}.

Proof.

Follows from Proposition 4.6, along with the facts that 𝒪Ki×\mathcal{O}_{K_{i}}^{\times} is precisely the set of elements of Ki×K_{i}^{\times} with ν\nu-adic valuation equal to 0 for every place ν\nu of KiK_{i}, and Ki×𝒪Ki,ν×K_{i}^{\times}\cap\mathcal{O}_{K_{i,\nu}}^{\times} is precisely the set of elements of Ki×K_{i}^{\times} with ν\nu-adic valuation equal to 0. ∎

Corollary 4.8.

A rational prime divides ΔL1\Delta_{L_{1}} if and only if it divides ΔL2\Delta_{L_{2}}.

Proof.

Note by equation 7 that a rational prime pp is unramified in LiL_{i} if and only if it is unramified in KiK_{i} and Ui,pNi,pU_{i,p}\subset N_{i,p}, where Ui,p=ν|p𝒪Ki,ν×U_{i,p}=\prod_{\nu|p}\mathcal{O}^{\times}_{K_{i,\nu}}, and Ni,p=NLi/Ki(CLi)ν|pKi,ν×N_{i,p}=N_{L_{i}/K_{i}}(C_{L_{i}})\cap\prod_{\nu|p}K_{i,\nu}^{\times}. But then the KiK_{i} are arithmetically equivalent, so they are unramified over the same rational primes, and Proposition 4.6 says that U1,pN1,pU_{1,p}\subset N_{1,p} if and only if U2,pN2,pU_{2,p}\subset N_{2,p}. Hence, a rational prime is unramified in L1L_{1} if and only if it is unramified in L2L_{2}. Since the prime divisors of ΔLi\Delta_{L_{i}} are exactly those rational primes which ramify in LiL_{i}, the claim holds. ∎

We have now come to the main theorem of this section, whose proof requires Theorem 3.3 and class field theory. Theorem 4.9 is used to relate the odd KK-groups of corresponding abelian extensions.

Theorem 4.9.

If NL/KN_{L/K} denotes the idele class norm CLCKC_{L}\rightarrow C_{K}, then NL1/F(CL1)=NL2/F(CL2)N_{L_{1}/F}(C_{L_{1}})=N_{L_{2}/F}(C_{L_{2}}). In particular, L1/FL_{1}/F and L2/FL_{2}/F contain the same maximal abelian subextension.

Proof.

Let Ωi=G/GKi\Omega_{i}=G/G_{K_{i}}, AA be the GG-equivariant linear isomorphism Ω1Ω2\mathbb{Z}\Omega_{1}\rightarrow\mathbb{Z}\Omega_{2}, and define

αi=Ωi.\alpha_{i}=\sum\Omega_{i}.

Since the action of GG on Ωi\Omega_{i} is transitive, αiΩi\mathbb{Z}\alpha_{i}\subset\mathbb{Z}\Omega_{i} is the only rank 1 submodule fixed pointwise by GG. But then gAα1=Agα1=Aα1gA\alpha_{1}=Ag\alpha_{1}=A\alpha_{1} for each gGg\in G, so Aα1=nα2A\alpha_{1}=n\alpha_{2} for some nn\in\mathbb{Z}. Notice that αi\alpha_{i} corresponds to the vector (1,1,,1)(1,1,...,1) when we use the cosets as a basis for Ωi\mathbb{Z}\Omega_{i}, so in fact nn is an eigenvalue of AA. Since AA and A1A^{-1} are each represented in this basis by an invertible integer matrix, n=±1n=\pm 1. Letting Ni:Ti𝔾mN_{i}:T_{i}\rightarrow\mathbb{G}_{m} be the scheme corresponding to the GG-module morphism Ωi:1αi\mathbb{Z}\rightarrow\mathbb{Z}\Omega_{i}:1\mapsto\alpha_{i}, we therefore have the commutative diagram of schemes

T1T2N1N2𝔾m𝔾m\begin{CD}T_{1}@>{}>{}>T_{2}\\ @V{N_{1}}V{}V@V{}V{N_{2}}V\\ \mathbb{G}_{m}@>{}>{}>\mathbb{G}_{m}\end{CD}

where the bottom isomorphism is either the identity or inversion. Taking 𝔸F\mathbb{A}_{F}-points gives

𝕀K1𝕀K2𝕀F𝕀F\begin{CD}\mathbb{I}_{K_{1}}@>{}>{}>\mathbb{I}_{K_{2}}\\ @V{}V{}V@V{}V{}V\\ \mathbb{I}_{F}@>{}>{}>\mathbb{I}_{F}\end{CD}

where the vertical arrows are the idele norms, by Theorem 3.3. The diagonal embeddings are respected, so we can pass to idele class groups and obtain the following commutative diagram.

(8) CK1φCK2NK1/FNK2/FCFCF\begin{CD}C_{K_{1}}@>{\varphi}>{}>C_{K_{2}}\\ @V{N_{K_{1}/F}}V{}V@V{}V{N_{K_{2}/F}}V\\ C_{F}@>{}>{}>C_{F}\end{CD}

From the fact that NLi/F=NKi/FNLi/KiN_{L_{i}/F}=N_{K_{i}/F}\circ N_{L_{i}/K_{i}} and the commutativity of diagram (8), we find that NL1/F(CL1)=NL2/F(CL2)N_{L_{1}/F}(C_{L_{1}})=N_{L_{2}/F}(C_{L_{2}}), so L1/FL_{1}/F and L2/FL_{2}/F have the same maximal abelian subextension, by the Norm Limitation Theorem (c.f. Theorem 7.3.10 in [Ked], Theorem 7 in Chapter XIV of [Tat08]). ∎

Corollary 4.10.

Fixing a separable closure FsF^{s} of FF, if αFs\alpha\in F^{s}, and F(α)F(\alpha) is abelian, then K1(α)K_{1}(\alpha) and K2(α)K_{2}(\alpha) correspond.

Proof.

Suppose L1=K1(α)L_{1}=K_{1}(\alpha). By Theorem 4.9, L1/FL_{1}/F and L2/FL_{2}/F have the same maximal abelian subextension, so αL2\alpha\in L_{2}. Hence [K2(α):K2][L2:K2][K_{2}(\alpha):K_{2}]\leq[L_{2}:K_{2}]. But then, [L2:K2]=[K1(α):K1][L_{2}:K_{2}]=[K_{1}(\alpha):K_{1}], so [K2(α):K2][K1(α):K1][K_{2}(\alpha):K_{2}]\leq[K_{1}(\alpha):K_{1}]. We can just as easily argue that [K1(α):K1][K2(α):K2][K_{1}(\alpha):K_{1}]\leq[K_{2}(\alpha):K_{2}], so in fact [K1(α):K1]=[K2(α):K2][K_{1}(\alpha):K_{1}]=[K_{2}(\alpha):K_{2}], and thus [K2(α):K2]=[L2:K2][K_{2}(\alpha):K_{2}]=[L_{2}:K_{2}], so L2=K2(α)L_{2}=K_{2}(\alpha). By symmetry we conclude that L1=K1(α)L_{1}=K_{1}(\alpha) if and only if L2=K2(α)L_{2}=K_{2}(\alpha). In other words, K1(α)K_{1}(\alpha) and K2(α)K_{2}(\alpha) are corresponding abelian extensions. ∎

Proposition 4.11.

If [Li:Ki][L_{i}:K_{i}] is odd, then L1L_{1} and L2L_{2} have the same signature.

Proof.

Let F=ν|Fν𝔸FF_{\infty}=\prod_{\nu|\infty}F_{\nu}\subset\mathbb{A}_{F} and set Ti,=Ti(F)T_{i,\infty}=T_{i}(F_{\infty}) and Ni,=NLi/Ki(CLi)Ti,N_{i,\infty}=N_{L_{i}/K_{i}}(C_{L_{i}})\cap T_{i,\infty}. If [Li:Ki][L_{i}:K_{i}] is odd, then the image of Ti,T_{i,\infty} under the map CKiGal(Li/Ki)C_{K_{i}}\rightarrow\operatorname{Gal}(L_{i}/K_{i}) is trivial, as Ti,/Ni,T_{i,\infty}/N_{i,\infty} is a 22-group. Therefore, Ni,=Ti,N_{i,\infty}=T_{i,\infty}. In particular, no archimedean prime of KiK_{i} ramifies in LiL_{i}. Since K1K_{1} and K2K_{2} have the same signature, so do L1L_{1} and L2L_{2}.

Alternatively, let GG be the group of the Galois closure of the LiL_{i} over \mathbb{Q}, and let 𝒟G\mathcal{D}\subset G be a subgroup of order 11 or 22. Because K1K_{1} and K2K_{2} are integrally equivalent over FF, they are arithmetically equivalent, so by Proposition 2.6 in [Sut],

|{g𝒟g1GK1 | gG}|=|{g𝒟g1GK2 | gG}|.|\{g\mathcal{D}g^{-1}\subset G_{K_{1}}\text{ }|\text{ }g\in G\}|=|\{g\mathcal{D}g^{-1}\subset G_{K_{2}}\text{ }|\text{ }g\in G\}|.

But then the image of 𝒟GKi\mathcal{D}\cap G_{K_{i}} is trivial under the projection GKiGKi/GLiG_{K_{i}}\rightarrow G_{K_{i}}/G_{L_{i}}, so 𝒟GKi\mathcal{D}\subset G_{K_{i}} if and only if 𝒟GLi\mathcal{D}\subset G_{L_{i}}, and therefore

|{g𝒟g1GL1 | gG}|=|{g𝒟g1GL2 | gG}|.|\{g\mathcal{D}g^{-1}\subset G_{L_{1}}\text{ }|\text{ }g\in G\}|=|\{g\mathcal{D}g^{-1}\subset G_{L_{2}}\text{ }|\text{ }g\in G\}|.

By Proposition 2.2 in [Sut], it follows that G/GL1G/GL2G/G_{L_{1}}\simeq G/G_{L_{2}} as 𝒟\mathcal{D}-sets. This is then true, in particular, for 𝒟\mathcal{D} a decomposition group of the archimedean prime. ∎

Proposition 4.12.

If Kn,t(F)K_{n,t}(F) denotes the torsion part of Kn(F)K_{n}(F) for a number field FF, then for odd n3n\not\equiv 3 mod 88, Kn,t(L1)Kn,t(L2)K_{n,t}(L_{1})\simeq K_{n,t}(L_{2}). Furthermore, if [Li:Ki][L_{i}:K_{i}] is odd, then Kn(L1)Kn(L2)K_{n}(L_{1})\simeq K_{n}(L_{2}) for odd n3n\geq 3.

Proof.

Let Li/KiL_{i}/K_{i}, i=1,2i=1,2, be corresponding abelian extensions and nn an odd integer. If n=1n=1, then we know the first claim holds, since the LiL_{i} have the same roots of unity, by Theorem 4.9, so assume additionally that n3n\geq 3. Letting 𝒞i,p={(Gal(Li(ζpk)/Li),k)}k\mathcal{C}_{i,p}=\{(\text{Gal}(L_{i}(\zeta_{p^{k}})/L_{i}),k)\}_{k\in\mathbb{N}}, where ζpk\zeta_{p^{k}} is a primitive pkp^{k}-th root of unity, we know by equation 1 that the torsion part of Kn(Li)K_{n}(L_{i}) is determined by 𝒞i,p\mathcal{C}_{i,p} for odd nn not congruent to 33 mod 88. From Proposition 4.1 and Corollary 4.10, we know that L1(ζ)/K1L_{1}(\zeta)/K_{1} and L2(ζ)/K2L_{2}(\zeta)/K_{2} correspond for any primitive root of unity ζ\zeta, since Li(ζ)=LiKi(ζ)L_{i}(\zeta)=L_{i}K_{i}(\zeta). Therefore, Gal(L1(ζ)/K1)Gal(L2(ζ)/K2)\text{Gal}(L_{1}(\zeta)/K_{1})\simeq\text{Gal}(L_{2}(\zeta)/K_{2}), and this isomorphism restricts to an isomorphism Gal(L1(ζ)/L1)Gal(L2(ζ)/L2)\text{Gal}(L_{1}(\zeta)/L_{1})\simeq\text{Gal}(L_{2}(\zeta)/L_{2}), by the definition of corresponding abelian extensions. Therefore, 𝒞1,p=𝒞2,p\mathcal{C}_{1,p}=\mathcal{C}_{2,p} for each prime pp. This verifies that Kn,t(L1)Kn,t(L2)K_{n,t}(L_{1})\simeq K_{n,t}(L_{2}) for n3n\not\equiv 3 mod 88. Supposing now that [Li:Ki][L_{i}:K_{i}] is odd, then by Proposition 4.11, the signatures of the LiL_{i} are the same, so Kn(L1)Kn(L2)K_{n}(L_{1})\simeq K_{n}(L_{2}) for any odd n3n\geq 3. ∎

4.2. Corresponding Abelian Extensions That Are Not Weakly Kronecker Equivalent

We prove that non-isomorphic integrally equivalent number fields always possess corresponding abelian extensions that are not weakly Kronecker equivalent. Before proving the main result of this subsection, we need to set the stage with some lemmas. We use eie_{i} to denote the ithi^{th} standard basis vector.

Lemma 4.13.

If (G,G1,G2)(G,G_{1},G_{2}) is a non-trivial integral Gassmann triple333i.e. the GiG_{i} are not conjugate in GG of finite groups, and AGLn()A\in\text{GL}_{n}(\mathbb{Z}) is the matrix expressing the [G]\mathbb{Z}[G]-module isomorphism [G/G1][G/G2]\mathbb{Z}[G/G_{1}]\rightarrow\mathbb{Z}[G/G_{2}] using cosets as bases, then each row of AA contains multiple non-zero components.

Proof.

Since the groups are finite and G1,G2G_{1},G_{2} are not conjugate in GG, we know that for every gGg\in G, there is hgG2g1h\in gG_{2}g^{-1} such that hG1h\notin G_{1}. In particular, letting e1ie_{1}^{i} be the basis element corresponding to the coset GiG/GiG_{i}\in G/G_{i} and g1,,gng_{1},...,g_{n} be a complete set of coset representatives for G/G2G/G_{2}, then if Ae11=iaigie12Ae_{1}^{1}=\sum_{i}a_{i}g_{i}e_{1}^{2} and aj0a_{j}\neq 0, there is ggjG2gj1g\in g_{j}G_{2}g_{j}^{-1} such that gG1g\notin G_{1}, so that ge11e11ge_{1}^{1}\neq e_{1}^{1} and A(ge11)=gA(e11)=ajgje12+ijaiggie12A(ge_{1}^{1})=gA(e_{1}^{1})=a_{j}g_{j}e_{1}^{2}+\sum_{i\neq j}a_{i}gg_{i}e_{1}^{2}. In particular, the jthj^{th} row of AA has at least two non-zero components. By GG-equivariance, we conclude that any row with a non-zero component must have multiple nonzero components. But then AA is invertible, so every row has at least one non-zero component.

Lemma 4.14.

Call a sublattice of the standard lattice in Euclidean space normal if it has a basis consisting of integral multiples of the standard basis. The maximal normal sublattice \mathcal{L} of the lattice given by the \mathbb{Z}-span of the columns of MGL(r,)M(r,)M\in\text{GL}(r,\mathbb{Q})\cap M(r,\mathbb{Z}) is imi\bigoplus_{i}m_{i}\mathbb{Z}, where mim_{i} it the least positive integer mm such that mM1eirmM^{-1}e_{i}\in\mathbb{Z}^{r}.

Proof.

The ithi^{th} component of any element of \mathcal{L} is an integer multiple of the least positive mm such that meime_{i}\in\mathcal{L}, since \mathcal{L} is a normal sublattice. However, meime_{i}\in\mathcal{L} if and only if mM1eirmM^{-1}e_{i}\in\mathbb{Z}^{r}. ∎

Lemma 4.15.

Let pp be a rational prime that splits completely in integrally equivalent number fields KiK_{i}, i=1,2i=1,2, and set i,p=Ti(p)/Ui,p\mathcal{L}_{i,p}=T_{i}(\mathbb{Q}_{p})/U_{i,p}. Identifying i,p\mathcal{L}_{i,p} with n\mathbb{Z}^{n}, the automorphism nn\mathbb{Z}^{n}\rightarrow\mathbb{Z}^{n} determined by the linear isomorphism 1,p2,p\mathcal{L}_{1,p}\rightarrow\mathcal{L}_{2,p} induced by φp\varphi_{p} in equation 5 is precisely the transpose of the matrix expressing the GG-equivariant linear isomorphism A:[G/GK2][G/GK1]A:\mathbb{Z}[G/G_{K_{2}}]\rightarrow\mathbb{Z}[G/G_{K_{1}}] with bases given by the cosets G/GKiG/G_{K_{i}}.

Proof.

A map of sets s:S1S2s:S_{1}\rightarrow S_{2} determines an extension to a linear map p[S1]p[S2]\mathbb{Q}_{p}[S_{1}]\rightarrow\mathbb{Q}_{p}[S_{2}], which we also refer to as ss. Let XiX_{i} be the character group of TiT_{i}, and suppose gi1,,ginGg_{i1},...,g_{in}\in G is a full set of representatives for G/GiG/G_{i}, with xijXix_{ij}\in X_{i} the preimage of gijGig_{ij}G_{i} under the isomorphism φi:Xi[G/Gi]\varphi_{i}:X_{i}\rightarrow\mathbb{Z}[G/G_{i}]. Since pp splits completely in KiK_{i}, we know Ti(p)=Hom(p[Xi],p)T_{i}(\mathbb{Q}_{p})=\text{Hom}(\mathbb{Q}_{p}[X_{i}],\mathbb{Q}_{p}), and a morphism f:p[Xi]pf:\mathbb{Q}_{p}[X_{i}]\rightarrow\mathbb{Q}_{p} is uniquely determined by the nn-tuple (f(xi1),,f(xin))(p×)n(f(x_{i1}),...,f(x_{in}))\in(\mathbb{Q}_{p}^{\times})^{n}. Denote by φ\varphi the isomorphism Hom(p[X1],p)Hom(p[X2],p)\text{Hom}(\mathbb{Q}_{p}[X_{1}],\mathbb{Q}_{p})\rightarrow\text{Hom}(\mathbb{Q}_{p}[X_{2}],\mathbb{Q}_{p}), and let A:X2X1A^{\prime}:X_{2}\rightarrow X_{1} be given by A=φ11Aφ2A^{\prime}=\varphi_{1}^{-1}\circ A\circ\varphi_{2}. Observe that φ(f)=fA\varphi(f)=f\circ A^{\prime}, so that, if fHom(p[X1],p)f\in\text{Hom}(\mathbb{Q}_{p}[X_{1}],\mathbb{Q}_{p}) is given by f(x1j)=αjf(x_{1j})=\alpha_{j}, then φ(f)(x2j)=(α1,,αn)colj(A)\varphi(f)(x_{2j})=(\alpha_{1},...,\alpha_{n})^{\text{col}_{j}(A)}, where for vectors a=(a1,,an),b=(b1,,bn)a=(a_{1},...,a_{n}),b=(b_{1},...,b_{n}), we are writing ab=a1b1anbna^{b}=a_{1}^{b_{1}}\cdots a_{n}^{b_{n}}. Now, write αi=pmiαi\alpha_{i}=p^{m_{i}}\alpha_{i}^{\prime}, where αi\alpha_{i}^{\prime} is a unit. Passing to the quotient (p×)n/(p×)nn(\mathbb{Q}_{p}^{\times})^{n}/(\mathbb{Z}_{p}^{\times})^{n}\simeq\mathbb{Z}^{n}, we see that the image of α=(α1,,αn)\alpha=(\alpha_{1},...,\alpha_{n}) is identified with m=(m1,,mn)m=(m_{1},...,m_{n}), while the image of (αcol1(A),,αcoln(A))(\alpha^{\text{col}_{1}(A)},...,\alpha^{\text{col}_{n}(A)}) is identified with ATmA^{T}m. ∎

Recall from Corollary 4.8 that a rational prime pp is unramified in L1L_{1} if and only if it is unramified in L2L_{2}. With Ni,pN_{i,p} and Ui,pU_{i,p} as in Corollary 4.8, let i,p=Ni,p/Ui,p\mathcal{L}_{i,p}^{\prime}=N_{i,p}/U_{i,p} and i,p=Ti(p)/Ui,p\mathcal{L}_{i,p}=T_{i}(\mathbb{Q}_{p})/U_{i,p}. Identifying i,p\mathcal{L}_{i,p} with n\mathbb{Z}^{n}, view i,p\mathcal{L}_{i,p}^{\prime} as a sublattice. Note that the orders of the relative decomposition groups for Li/KiL_{i}/K_{i} over a rational prime pp unramified in LiL_{i} are exactly the positive multiples of the standard basis vectors forming a basis of the largest normal sublattice in i,p\mathcal{L}_{i,p}^{\prime}, by equation 6. Though i,p\mathcal{L}_{i,p}^{\prime} certainly can be a normal sublattice in i,p\mathcal{L}_{i,p}, this is by no means the case in general (c.f. [Tat08] Chapters VII and X).

Theorem 4.16.

Given non-isomorphic integrally equivalent number fields, there exist corresponding abelian extensions which are not weakly Kronecker equivalent.

Proof.

Suppose K1,K2K_{1},K_{2} are non-isomorphic, integrally equivalent number fields whose Galois closure over \mathbb{Q} has group GG, and let A:[G/GK2][G/GK1]A:\mathbb{Z}[G/G_{K_{2}}]\rightarrow\mathbb{Z}[G/G_{K_{1}}] be a GG-equivariant isomorphism inducing a correspondence of the abelian extensions of the KiK_{i}. If pp is an odd rational prime that splits completely in the KiK_{i}, by a Grunwald-Wang argument (for instance, intersecting appropriate subgroups of CK1C_{K_{1}} guaranteed to exist by Theorem 6 in Chapter X of [Tat08]), given a prime qq coprime to every cofactor of AA, there is an abelian extension L1L_{1} of K1K_{1} in which pp is unramified with 1,p=qq\mathcal{L}_{1,p}^{\prime}=\mathbb{Z}\bigoplus q\mathbb{Z}\bigoplus\cdots\bigoplus q\mathbb{Z}. In particular, 1,p\mathcal{L}_{1,p}^{\prime} is equal to its maximal normal sublattice, and SL1(p)={{1,q,,q}}S_{L_{1}}(p)=\{\{1,q,...,q\}\}. Now, let M=ATdiag(1,q,,q)M=A^{T}\text{diag}(1,q,...,q). By Lemma 4.15, we know that 2,p\mathcal{L}_{2,p}^{\prime} is given by the \mathbb{Z}-span of the columns of MM, so the splitting type of pp in L2L_{2} is given by the set of least positive integers mim_{i} such that miM1eim_{i}M^{-1}e_{i} is integral, by Lemma 4.14. Since AA is GG-equivariant, so is A1A^{-1}. Furthermore, because K1≄K2K_{1}\not\simeq K_{2}, every row of A1A^{-1} has multiple non-zero entries, by Lemma 4.13. If CC denotes the cofactor matrix of AA, we know that A1=±CTA^{-1}=\pm C^{T}, so every column of CC has multiple non-zero entries. Now, M1=±diag(1,q1,,q1)CM^{-1}=\pm\text{diag}(1,q^{-1},...,q^{-1})C, and every column of CC has multiple non-zero entries, so miM1eim_{i}M^{-1}e_{i} is integral if and only if q|miq|m_{i}, by choice of qq, so the splitting type of pp in L2L_{2} is {{q,,q}}\{\{q,...,q\}\}. Therefore, gcd(SL1(p))=1\gcd(S_{L_{1}}(p))=1, while gcd(SL2(p))=q\gcd(S_{L_{2}}(p))=q, so the LiL_{i} are not weakly Kronecker equivalent, by Theorem 1.4. ∎

Remark 4.17.

We know that a non-trivial correspondence of abelian extensions cannot preserve arithmetic equivalence classes (c.f. Theorem 2.1). We have also just seen that a correspondence of abelian extensions induced by a non-trivial integral equivalence of number fields cannot preserve weak Kronecker equivalence classes. The question remains whether or not this is a general phenomenon. In other words, letting 𝒦(F)\mathcal{K}(F) denote the weak Kronecker class of a number field FF, if K1,K2K_{1},K_{2} are number fields admitting a bijective correspondence of abelian extensions φ:L1/K1φ(L1/K1)\varphi:L_{1}/K_{1}\mapsto\varphi(L_{1}/K_{1}) such that 𝒦(φ(L1))=𝒦(L1)\mathcal{K}(\varphi(L_{1}))=\mathcal{K}(L_{1}) for every abelian extension L1L_{1} of K1K_{1}, does it follow that K1K2K_{1}\simeq K_{2}?

5. Diagrams in Homology and Cohomology

Given an integral Gassmann triple (Γ,Γ1,Γ2)(\Gamma,\Gamma_{1},\Gamma_{2}) and a Γ\Gamma-module AA, Arapura et. al proved that there are isomorphisms in cohomology H(Γ1,A)H(Γ2,A)H^{*}(\Gamma_{1},A)\simeq H^{*}(\Gamma_{2},A) compatible with corestriction to and restriction from H(Γ,A)H^{*}(\Gamma,A), where the Γi\Gamma_{i}-module structure on AA is ResΓiΓ\text{Res}_{\Gamma_{i}}^{\Gamma}(A) [AKMS19]. We show that there are in fact compatible isomorphisms φ:H(Γ1,A)H(Γ2,A)\varphi:H(\Gamma_{1},A)\rightarrow H(\Gamma_{2},A), where HH denotes either HH_{*} or HH^{*}. Furthermore, for ΓN\Gamma_{N} a finite index normal subgroup of Γ\Gamma contained in Γ1Γ2\Gamma_{1}\cap\Gamma_{2}, and AA a finite abelian group of order coprime to [Γ:Γi][\Gamma:\Gamma_{i}], we demonstrate that φ\varphi is also compatible with restriction and corestriction to and from, respectively, H(ΓN,A)H(\Gamma_{N},A). If UiH1(Γi)U_{i}\subset H_{1}(\Gamma_{i}), i=1,2i=1,2 are index t<t<\infty subgroups such that φ:H1(Γ1)H1(Γ2)\varphi:H_{1}(\Gamma_{1})\rightarrow H_{1}(\Gamma_{2}) restricts to an isomorphism U1U2U_{1}\rightarrow U_{2}, we say that U1U_{1} and U2U_{2} correspond. Furthermore, if Γi\Gamma_{i}^{\prime} is the preimage of UiU_{i} in Γi\Gamma_{i} under the natural projection ΓiH1(Γi)\Gamma_{i}\rightarrow H_{1}(\Gamma_{i}), then we say Γ1\Gamma_{1}^{\prime} and Γ2\Gamma_{2}^{\prime} correspond if U1U_{1} and U2U_{2} do. We show in Theorem 5.4 that corresponding Γ1\Gamma_{1}^{\prime} and Γ2\Gamma_{2}^{\prime} have the same normal core in Γ\Gamma, assuming tt is coprime to [Γ:Γi][\Gamma:\Gamma_{i}]. Theorem 1.8 is then proved as a corollary. In Proposition 5.6, we conclude with an arithmetic application of the techniques in this section. Throughout, [Γ:Γi]<[\Gamma:\Gamma_{i}]<\infty, =\bigotimes=\bigotimes_{\mathbb{Z}}, and if coefficients are not specified for a (co)homology group, the coefficients are \mathbb{Z} with trivial action.

Lemma 5.1.

If Ω1,Ω2\Omega_{1},\Omega_{2} are finite and transitive GG-sets, then an isomorphism φ:Ω1Ω2\varphi:\mathbb{Z}\Omega_{1}\rightarrow\mathbb{Z}\Omega_{2} of G\mathbb{Z}G-modules, or its negative, is natural with respect to both the diagonal embeddings Δi:Ωi:nnΩi\Delta_{i}:\mathbb{Z}\rightarrow\mathbb{Z}\Omega_{i}:n\mapsto n\sum\Omega_{i} and the augmentations εi:Ωi:ωΩinωωωΩinω\varepsilon_{i}:\mathbb{Z}\Omega_{i}\rightarrow\mathbb{Z}:\sum_{\omega\in\Omega_{i}}n_{\omega}\omega\mapsto\sum_{\omega\in\Omega_{i}}n_{\omega}. Precisely, either φΔ1=Δ2\varphi\circ\Delta_{1}=\Delta_{2} and ε2φ=ε1\varepsilon_{2}\circ\varphi=\varepsilon_{1}, or (φ)Δ1=Δ2(-\varphi)\circ\Delta_{1}=\Delta_{2} and ε2(φ)=ε1\varepsilon_{2}\circ(-\varphi)=\varepsilon_{1}.

Proof.

Let AA be the matrix expressing the isomorphism φ\varphi with respect to the bases Ωi\Omega_{i}. Observe as in the proof of Proposition 4.9 that, by GG-equivariance, (1,,1)(1,...,1) is an eigenvector of AA, and therefore A(1,,1)=±(1,,1)A(1,...,1)=\pm(1,...,1). Furthermore, letting Ωi\mathbb{Z}\Omega_{i}^{*} denote the dual GG-module to Ωi\mathbb{Z}\Omega_{i} consisting of all \mathbb{Z}-linear maps Ωi\mathbb{Z}\Omega_{i}\rightarrow\mathbb{Z}, with GG-action given by gf(ω)=f(g1ω)gf(\omega)=f(g^{-1}\omega), for fΩi,ωΩif\in\mathbb{Z}\Omega_{i}^{*},\omega\in\Omega_{i}, gGg\in G, then the induced G\mathbb{Z}G-module isomorphism φ:Ω2Ω1\varphi^{*}:\mathbb{Z}\Omega_{2}^{*}\rightarrow\mathbb{Z}\Omega_{1}^{*} is expressed by ATA^{T} with respect to the dual bases Ωi\Omega_{i}^{*}. We conclude that (1,,1)(1,...,1) is an eigenvector of ATA^{T}. Observe that (1,,1)Ωi(1,...,1)\in\mathbb{Z}\Omega_{i}^{*} is precisely εi\varepsilon_{i}, so ε2φ=±ε1\varepsilon_{2}\circ\varphi=\pm\varepsilon_{1}. Finally, note that if φΔ1=cΔ2\varphi\circ\Delta_{1}=c\Delta_{2}, and ε2φ=cε1\varepsilon_{2}\circ\varphi=c^{\prime}\varepsilon_{1}, then

c|Ω2|=(ε2φΔ1)(1)=c|Ω1|.c|\Omega_{2}|=(\varepsilon_{2}\circ\varphi\circ\Delta_{1})(1)=c^{\prime}|\Omega_{1}|.

Since |Ω1|=|Ω2||\Omega_{1}|=|\Omega_{2}|, we conclude c=cc=c^{\prime}. ∎

Corollary 5.2.

If (Γ,Γ1,Γ2)(\Gamma,\Gamma_{1},\Gamma_{2}) is an integral Gassmann triple, then, given a Γ\Gamma-module AA, there is a commutative diagram

H(Γ,A){H(\Gamma,A)}H(Γ1,A){H(\Gamma_{1},A)}H(Γ2,A){H(\Gamma_{2},A)}ResΓ2Γ\scriptstyle{Res_{\Gamma_{2}}^{\Gamma}}ResΓ1Γ\scriptstyle{Res_{\Gamma_{1}}^{\Gamma}}CorΓ1Γ\scriptstyle{Cor^{\Gamma}_{\Gamma_{1}}}\scriptstyle{\simeq}CorΓ2Γ\scriptstyle{Cor^{\Gamma}_{\Gamma_{2}}}

where HH denotes either HH^{*} or HH_{*}, and the Γi\Gamma_{i}-module structure on AA is ResΓiΓ\text{Res}_{\Gamma_{i}}^{\Gamma}(A).

Proof.

Letting Ωi=Γ/Γi\Omega_{i}=\Gamma/\Gamma_{i}, recall that restriction ResΓiΓ\text{Res}^{\Gamma}_{\Gamma_{i}} (resp. corestriction CorΓiΓ\text{Cor}^{\Gamma}_{\Gamma_{i}}) in (co)homology is obtained by tensoring the diagonal embedding Δi\Delta_{i} (resp. augmentation morphism εi\varepsilon_{i}) with AA, then applying H(Γ,)H(\Gamma,-) and Shapiro’s Lemma (c.f. [Bro82] III.5.6 and III.9.A). By Lemma 5.1, there is a commutative diagram of Γ\mathbb{Z}\Gamma-module morphisms

{\mathbb{Z}}Ω1{\mathbb{Z}\Omega_{1}}Ω2{\mathbb{Z}\Omega_{2}}Δ2\scriptstyle{\Delta_{2}}Δ1\scriptstyle{\Delta_{1}}ε1\scriptstyle{\varepsilon_{1}}\scriptstyle{\simeq}ε2\scriptstyle{\varepsilon_{2}}

Tensoring with AA then applying H(Γ,)H(\Gamma,-) and Shapiro’s Lemma yields the desired result. ∎

From here on, (Γ,Γ1,Γ2)(\Gamma,\Gamma_{1},\Gamma_{2}) is an integral Gassmann triple, m=[Γ:Γi]m=[\Gamma:\Gamma_{i}], and AA is a Γ\Gamma-module, unless stated otherwise. We use φ\varphi to denote the isomorphism H(Γ1,A)H(Γ2,A)H(\Gamma_{1},A)\rightarrow H(\Gamma_{2},A) from Corollary 5.2.

Lemma 5.3.

Let AA be a finite abelian group of order coprime to mm, and let ΓN\Gamma_{N} be a finite index normal subgroup of Γ\Gamma contained in Γ1Γ2\Gamma_{1}\cap\Gamma_{2}. There is a commutative diagram of the following form.

H(Γ1,A){H(\Gamma_{1},A)}H(Γ2,A){H(\Gamma_{2},A)}H(ΓN,A){H(\Gamma_{N},A)}φ\scriptstyle{\varphi}ResΓNΓ1\scriptstyle{Res_{\Gamma_{N}}^{\Gamma_{1}}}ResΓNΓ2\scriptstyle{Res_{\Gamma_{N}}^{\Gamma_{2}}}CorΓNΓ1\scriptstyle{Cor^{\Gamma_{1}}_{\Gamma_{N}}}CorΓNΓ2\scriptstyle{Cor^{\Gamma_{2}}_{\Gamma_{N}}}
Proof.

Because ΓNΓ\Gamma_{N}\trianglelefteq\Gamma, and ΓNΓi\Gamma_{N}\subset\Gamma_{i}, i=1,2i=1,2, the action of ΓN\Gamma_{N} on Γi\Γ\Gamma_{i}\backslash\Gamma by right multiplication is trivial, so that |Γi\Γ/ΓN|=m|\Gamma_{i}\backslash\Gamma/\Gamma_{N}|=m. Now, we know that

ResΓiΓCorΓNΓ(z)=CorΓNΓiResΓNΓN(gz),\text{Res}^{\Gamma}_{\Gamma_{i}}\text{Cor}^{\Gamma}_{\Gamma_{N}}(z)=\sum\text{Cor}^{\Gamma_{i}}_{\Gamma_{N}}\text{Res}^{\Gamma_{N}}_{\Gamma_{N}}(gz),

where the sum is over representatives gΓg\in\Gamma for the set of double cosets Γi\Γ/ΓN\Gamma_{i}\backslash\Gamma/\Gamma_{N} (c.f. Proposition 9.5 in Chapter III of [Bro82]). Clearly, ResΓNΓN(gz)=gz\text{Res}^{\Gamma_{N}}_{\Gamma_{N}}(gz)=gz, and, because ΓNΓ\Gamma_{N}\trianglelefteq\Gamma, we know that CorΓNΓi(gz)=CorΓNΓi(z)\text{Cor}^{\Gamma_{i}}_{\Gamma_{N}}(gz)=\text{Cor}^{\Gamma_{i}}_{\Gamma_{N}}(z), so ResΓiΓCorΓNΓ(gz)=mCorΓNΓi(z)\text{Res}^{\Gamma}_{\Gamma_{i}}\text{Cor}^{\Gamma}_{\Gamma_{N}}(gz)=m\text{Cor}^{\Gamma_{i}}_{\Gamma_{N}}(z). Since φResΓ1Γ=ResΓ2Γ\varphi\circ\text{Res}^{\Gamma}_{\Gamma_{1}}=\text{Res}^{\Gamma}_{\Gamma_{2}}, we conclude mφ(CorΓNΓ1(z))=mCorΓNΓ2(z)m\varphi(\text{Cor}_{\Gamma_{N}}^{\Gamma_{1}}(z))=m\text{Cor}_{\Gamma_{N}}^{\Gamma_{2}}(z). But then mm is coprime to |A||A|, so φ(CorΓNΓ1(z))=CorΓNΓ2(z)\varphi(\text{Cor}_{\Gamma_{N}}^{\Gamma_{1}}(z))=\text{Cor}_{\Gamma_{N}}^{\Gamma_{2}}(z), as desired.

Let zH(Γ1,C)z\in H(\Gamma_{1},C), and recall that CorΓ1Γ(z)=CorΓ2Γ(φ(z))\text{Cor}_{\Gamma_{1}}^{\Gamma}(z)=\text{Cor}_{\Gamma_{2}}^{\Gamma}(\varphi(z)), so that

(9) ResΓNΓCorΓ1Γ(z)=ResΓNΓCorΓ2Γ(φ(z)).\text{Res}^{\Gamma}_{\Gamma_{N}}\text{Cor}_{\Gamma_{1}}^{\Gamma}(z)=\text{Res}^{\Gamma}_{\Gamma_{N}}\text{Cor}_{\Gamma_{2}}^{\Gamma}(\varphi(z)).

Similarly as before, we know that

ResΓNΓCorΓiΓ(z)=CorΓNΓNResΓNgΓig1(gz).\text{Res}^{\Gamma}_{\Gamma_{N}}\text{Cor}^{\Gamma}_{\Gamma_{i}}(z)=\sum\text{Cor}^{\Gamma_{N}}_{\Gamma_{N}}\text{Res}^{g\Gamma_{i}g^{-1}}_{\Gamma_{N}}(gz).

Now, ResΓNgΓig1(gz)=ResΓNΓi(z),\text{Res}^{g\Gamma_{i}g^{-1}}_{\Gamma_{N}}(gz)=\text{Res}^{\Gamma_{i}}_{\Gamma_{N}}(z), and CorΓNΓN\text{Cor}^{\Gamma_{N}}_{\Gamma_{N}} is the identity map, so we conclude ResΓNΓCorΓiΓ(z)=mResΓNΓi(z)\text{Res}^{\Gamma}_{\Gamma_{N}}\text{Cor}^{\Gamma}_{\Gamma_{i}}(z)=m\text{Res}^{\Gamma_{i}}_{\Gamma_{N}}(z). Therefore, by equation 9, we have mResΓNΓ1(z)=mResΓNΓ2(φ(z))m\text{Res}^{\Gamma_{1}}_{\Gamma_{N}}(z)=m\text{Res}^{\Gamma_{2}}_{\Gamma_{N}}(\varphi(z)), and we see that ResΓNΓ1(z)=ResΓNΓ2(φ(z))\text{Res}^{\Gamma_{1}}_{\Gamma_{N}}(z)=\text{Res}^{\Gamma_{2}}_{\Gamma_{N}}(\varphi(z)). ∎

Recall that restriction and corestriction are induced by chain maps. With the hypotheses of Lemma 5.3 and AA an abelian group with trivial Γ\Gamma-action, viewing H(Γi)AH_{*}(\Gamma_{i})\bigotimes A as a subgroup of H(Γi,A)H_{*}(\Gamma_{i},A) as in the the universal coefficient theorem, the diagram

H(Γ1)A{H_{*}(\Gamma_{1})\bigotimes A}H(Γ2)A{H_{*}(\Gamma_{2})\bigotimes A}H(ΓN)A{H_{*}(\Gamma_{N})\bigotimes A}φidA\scriptstyle{\varphi\otimes id_{A}}

commutes, by Lemma 5.3 and naturality of the inclusion H(Γi)AH(Γi,A)H_{*}(\Gamma_{i})\bigotimes A\rightarrow H_{*}(\Gamma_{i},A). The upward maps are CorΓNΓiidA\text{Cor}^{\Gamma_{i}}_{\Gamma_{N}}\bigotimes\text{id}_{A}, and the downward maps are ResΓNΓiidA\text{Res}^{\Gamma_{i}}_{\Gamma_{N}}\bigotimes\text{id}_{A}. Recall that a finite abelian group is naturally a commutative ring with identity, and let AA and BB be abelian groups, with AA finite. For us, the map BBAB\rightarrow B\bigotimes A is always the abelian group homomorphism bb1b\mapsto b\otimes 1. We may now prove the main theorem of this section.

Theorem 5.4.

If ΓiΓi\Gamma_{i}^{\prime}\subset\Gamma_{i}, i=1,2i=1,2 are corresponding index tt subgroups, then Γ1\Gamma_{1}^{\prime} and Γ2\Gamma_{2}^{\prime} have the same normal core in Γ\Gamma if (m,t)=1(m,t)=1.

Proof.

Suppose Γi/ΓiA\Gamma_{i}/\Gamma_{i}^{\prime}\simeq A for AA a finite abelian group of order coprime to mm, and fix projections πi:H1(Γi)A\pi_{i}:H_{1}(\Gamma_{i})\rightarrow A such that π1=π2φ\pi_{1}=\pi_{2}\circ\varphi and Γi\Gamma_{i}^{\prime} is the preimage of kerπi\ker\pi_{i} under the natural map ΓiH1(Γi)\Gamma_{i}\rightarrow H_{1}(\Gamma_{i}). Apply the functor A-\bigotimes A, viewing AA as a trivial Γ\Gamma-module, to the corresponding commutative diagram, and extend the result to the commutative diagram

Γ1{\Gamma_{1}}H1(Γ1)A{H_{1}(\Gamma_{1})\bigotimes A}N{N}H1(N)A{H_{1}(N)\bigotimes A}AA{A\bigotimes A}Γ2{\Gamma_{2}}H1(Γ2)A{H_{1}(\Gamma_{2})\bigotimes A}φidA\scriptstyle{\varphi\otimes id_{A}}ι1\scriptstyle{\iota_{1}}ι2\scriptstyle{\iota_{2}}

where ιi\iota_{i} is inclusion, NN is the normal core of Γi\Gamma_{i} in Γ\Gamma, and the horizontal maps are of the form GH1(G)H1(G)AG\rightarrow H_{1}(G)\rightarrow H_{1}(G)\bigotimes A. The proof of Theorem 1 in [Per77] guarantees that NN is independent of choice of i=1,2i=1,2. Since the map AAA:aa1A\rightarrow A\bigotimes A:a\mapsto a\otimes 1 is injective, and the diagram

H1(Γi){H_{1}(\Gamma_{i})}H1(Γi)A{H_{1}(\Gamma_{i})\bigotimes A}A{A}AA{A\bigotimes A}πi\scriptstyle{\pi_{i}}

commutes for i=1,2i=1,2, the kernel of the map ΓiAA\Gamma_{i}\rightarrow A\bigotimes A is Γi\Gamma_{i}^{\prime}, and we conclude NΓ1=NΓ2N\cap\Gamma_{1}^{\prime}=N\cap\Gamma_{2}^{\prime}. Now, if NiN_{i} is the normal core of Γi\Gamma_{i}^{\prime} in Γ\Gamma, we know that NiNN_{i}\subset N, so that NiN_{i} is equal to the normal core of NΓiN\cap\Gamma_{i}^{\prime} in Γ\Gamma, and therefore N1=N2N_{1}=N_{2}. ∎

Remark 5.5.

Letting πi\pi^{\prime}_{i} be the projection ΓiA\Gamma_{i}\rightarrow A, note the argument above shows that in fact π1(n)=π2(n)\pi^{\prime}_{1}(n)=\pi^{\prime}_{2}(n) for nNn\in N when |A||A| is coprime to [Γ:Γi][\Gamma:\Gamma_{i}]. It is not clear to the author how one could draw a similar conclusion without the coprimality assumption.

We can now easily prove Theorem 1.8, which is a sort of geometric analog of Theorem 1.1.1.

Proof of Theorem 1.8.

Recall the usual identification H(π1(M))=H(M)H(\pi_{1}(M))=H(M) for a manifold with contractible universal cover, and observe that the map H1(M)H1(M)H_{1}(M^{\prime})\rightarrow H_{1}(M) induced by a covering MMM^{\prime}\rightarrow M is corestriction H1(π1(M))H1(π1(M))H_{1}(\pi_{1}(M^{\prime}))\rightarrow H_{1}(\pi_{1}(M)). The subgroup of π1(M)\pi_{1}(M) corresponding to the normal closure of MiMM_{i}^{\prime}\rightarrow M under the covering space Galois correspondence is precisely the normal core of π1(Mi)\pi_{1}(M_{i}^{\prime}) in π1(M)\pi_{1}(M). The rest follows from Theorem 5.4 and the fact that (π1(M),π1(M1),π1(M2))(\pi_{1}(M),\pi_{1}(M_{1}),\pi_{1}(M_{2})) is an integral Gassmann triple. ∎

We conclude with an application to the arithmetic correspondence. The proof is only sketched, as all the requisite ideas have been discussed in greater detail in this article.

Proposition 5.6.

If Li/KiL_{i}/K_{i}, i=1,2i=1,2 are corresponding abelian extensions of number fields K1,K2K_{1},K_{2} integrally equivalent over FF, letting KK be the Galois closure of the KiK_{i} over FF, if [Li:Ki][L_{i}:K_{i}] is coprime to [Ki:F][K_{i}:F], then L1K=L2KL_{1}K=L_{2}K.

Proof.

Let K/FK/F be a Galois extension with group GG and subextension K/FK^{\prime}/F. If Ω=G/GK\Omega=G/G_{K^{\prime}}, and jj is the scheme corresponding to the augmentation map Ω\mathbb{Z}\Omega\rightarrow\mathbb{Z}, one can show that j(𝔸F)j(\mathbb{A}_{F}), modulo principle ideles, gives the map CFCKC_{F}\rightarrow C_{K^{\prime}} induced by inclusion FKF\rightarrow K^{\prime}. This observation, along with Theorem 3.3 and Lemma 5.1, reveals that for integrally equivalent number fields K1K_{1} and K2K_{2}, there is an isomorphism CK1CK2C_{K_{1}}\rightarrow C_{K_{2}} such that following diagram commutes.

CF{C_{F}}CK1{C_{K_{1}}}CK2{C_{K_{2}}}\scriptstyle{\simeq}

where the upward arrows are idele class norms and the downward arrows are the maps induced by inclusion. Functoriality of the global Artin map (c.f. [RV99] Section 6.4) then guarantees commutativity of the following diagram.

ΓFab{\Gamma_{F}^{ab}}ΓK1ab{\Gamma_{K_{1}}^{ab}}ΓK2ab{\Gamma_{K_{2}}^{ab}}\scriptstyle{\simeq}

where, identifying abelianization with the functor H1()H_{1}(-), the downward arrows are restriction and the upward arrows are corestriction in first homology. From here, reason similarly as we have in this section to conclude that ΓL1ΓK=ΓL2ΓK\Gamma_{L_{1}}\cap\Gamma_{K}=\Gamma_{L_{2}}\cap\Gamma_{K}. ∎

References

  • [AKMS19] D. Arapura, J. Katz, D. B. McReynolds, and P. Solapurkar. Integral Gassman equivalence of algebraic and hyperbolic manifolds. Mathematische Zeitschrift, 291(1):179–194, February 2019.
  • [BP16] Alex Bartel and Aurel Page. Torsion homology and regulators of isospectral manifolds. Journal of Topology, 9(4):1237–1256, 2016. _eprint: https://onlinelibrary.wiley.com/doi/pdf/10.1112/jtopol/jtw023.
  • [Bro82] Kenneth S. Brown. Cohomology of Groups, volume 87 of Graduate Texts in Mathematics. Springer, New York, NY, 1982.
  • [CdSL+18] Gunther Cornelissen, Bart de Smit, Xin Li, Matilde Marcolli, and Harry Smit. Characterization of global fields by Dirichlet L-series. Research in Number Theory, 5(1):7, November 2018.
  • [CLMS17] Gunther Cornelissen, Xin Li, Matilde Marcolli, and Harry Smit. Reconstructing global fields from dynamics in the abelianized Galois group, June 2017. arXiv:1706.04517 [math].
  • [CM14] Gunther Cornelissen and Matilde Marcolli. Quantum Statistical Mechanics, L-Series and Anabelian Geometry I: Partition Functions. In Vincenzo Ancona and Elisabetta Strickland, editors, Trends in Contemporary Mathematics, volume 8, pages 47–57. Springer International Publishing, Cham, 2014. Series Title: Springer INdAM Series.
  • [Cor13] Gunther Cornelissen. Curves, dynamical systems, and weighted point counting. Proceedings of the National Academy of Sciences, 110(24):9669–9673, June 2013. Publisher: Proceedings of the National Academy of Sciences.
  • [GM23] Milana Golich and D. B. McReynolds. Diamonds: Homology and the Central Series of Groups, October 2023. arXiv:2310.08283 [math].
  • [Jeh77] Wolfram Jehne. Kronecker classes of algebraic number fields. Journal of Number Theory, 9(3):279–320, August 1977.
  • [Ked] Kiran S. Kedlaya. Notes on class field theory.
  • [Kli78] Norbert Klingen. Zahlkörper mit gleicher Primzerlegung. Journal für die reine und angewandte Mathematik, 0299_0300:342–384, 1978.
  • [Kli98] Norbert Klingen. Arithmetical Similarities: Prime Decomposition and Finite Group Theory. Oxford Mathematical Monographs. Oxford University Press, Oxford, New York, April 1998.
  • [Loc94a] Manfred Lochter. Weakly Kronecker equivalent number fields. Acta Arithmetica, 67(4):295–312, 1994.
  • [Loc94b] Manfred Lochter. Weakly Kronecker equivalent number fields and global norms. Acta Arithmetica, 67(2):105–121, 1994.
  • [Loc95] M. Lochter. New Characterizations of Kronecker Equivalence. Journal of Number Theory, 53(1):115–136, July 1995.
  • [Per77] Robert Perlis. On the equation ζk(s) = ζk’(s). Journal of Number Theory, 9(3):342–360, August 1977.
  • [PR94] V. P. Platonov and A. S. Rapinchuk. Algebraic groups and number theory. Number v. 139 in Pure and applied mathematics. Academic Press, Boston, 1994.
  • [Pra17] Dipendra Prasad. A refined notion of arithmetically equivalent number fields, and curves with isomorphic Jacobians. Advances in Mathematics, 312:198–208, May 2017.
  • [RV99] Dinakar Ramakrishnan and Robert J. Valenza. Fourier Analysis on Number Fields, volume 186 of Graduate Texts in Mathematics. Springer, New York, NY, 1999.
  • [Sco93] Leonard Scott. Integral equivalence of permutation representations. Group Theory, January 1993.
  • [Ser79] Jean-Pierre Serre. Local Fields, volume 67 of Graduate Texts in Mathematics. Springer New York, New York, NY, 1979.
  • [Sol19] Pavel Solomatin. A Note on Number Fields Sharing the List of Dedekind Zeta-Functions of Abelian Extensions with some Applications towards the Neukirch-Uchida Theorem, January 2019. arXiv:1901.09243 [math].
  • [Sun85] Toshikazu Sunada. Riemannian Coverings and Isospectral Manifolds. Annals of Mathematics, 121(1):169–186, 1985. Publisher: Annals of Mathematics.
  • [Sut] Andrew V. Sutherland. Stronger arithmetic equivalence. discrete Analysis, 2021. arXiv:2104.01956 [math].
  • [Tat08] Emil Artin and John Tate. Class Field Theory. American Mathematical Society, Providence, RI, 2nd revised edition edition, December 2008.
  • [Uch76] Koji Uchida. Isomorphisms of Galois groups. Journal of the Mathematical Society of Japan, 28(4), October 1976.
  • [Wat79] William C. Waterhouse. Introduction to Affine Group Schemes, volume 66 of Graduate Texts in Mathematics. Springer, New York, NY, 1979.
  • [Wei] Charles Weibel. Algebraic K-Theory of Rings of Integers in Local and Global Fields.
  • [Wei13] Charles A. Weibel. The K-book: an introduction to algebraic K-theory. Graduate studies in mathematics ; v. 145. American Mathematical Society, Providence, Rhode Island, 2013.
  • [Zyw15] David Zywina. The inverse Galois problem for PSL2(Fp). Duke Mathematical Journal, 164(12):2253–2292, September 2015. Publisher: Duke University Press.