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

On the Mordell–Weil ranks of supersingular abelian varieties over p2{\mathbb{Z}}_{p}^{2}-extensions

Cédric Dion Département de Mathématiques et de Statistique
Université Laval, Pavillion Alexandre-Vachon
1045 Avenue de la Médecine
Québec, QC
Canada G1V 0A6
cedric.dion.1@ulaval.ca
 and  Jishnu Ray Harish Chandra Research Institute, A CI of Homi Bhabha National Institute, Chhatnag Road, Jhunsi, Prayagraj (Allahabad) 211 019 India jishnuray@hri.res.in; jishnuray1992@gmail.com
Abstract.

Let pp be a fixed odd prime and let KK be an imaginary quadratic field in which pp splits. Let AA be an abelian variety defined over KK with supersingular reduction at both primes above pp in KK. Under certain assumptions, we give a growth estimate for the Mordell–Weil rank of AA over finite extensions inside the p2{\mathbb{Z}}_{p}^{2}-extension of KK. In the last section, written by Chris Williams, he includes some speculative remarks on the pp-adic LL-functions for GSp(4)\mathrm{GSp}(4) corresponding to the multi-signed Selmer groups constructed in this paper.

Key words and phrases:
supersingular abelian varieties, Mordell–Weil ranks, Coleman maps, Selmer groups.
2010 Mathematics Subject Classification:
Primary: 11R23; Secondary: 11G10, 11R20

1. Introduction

1.1. Work of Lei–Ponsinet

Let pp be an odd prime, LL be a number field and LcycL_{\mathrm{cyc}} be the cyclotomic p{\mathbb{Z}}_{p}-extension of LL. Let Lcyc=nLnL_{\mathrm{cyc}}=\cup_{n}L_{n} where Gal(Ln/L)/pn{\mathrm{Gal}}(L_{n}/L)\cong{\mathbb{Z}}/p^{n}{\mathbb{Z}}. Let AA be an abelian variety over LL. Suppose AA has good ordinary reduction at the primes of LL above pp. Then it is conjectured by Mazur [1] that the classical pp-primary Selmer group Selp(A/Lcyc){\mathrm{Sel}}_{p}(A/L_{\mathrm{cyc}}) is cotorsion as a p[[Gal(Lcyc/L)]]{\mathbb{Z}}_{p}[[{\mathrm{Gal}}(L_{\mathrm{cyc}}/L)]]-module (follows by the work of Kato [2] and Rohrlich [3] for elliptic curves defined over {\mathbb{Q}} and L/L/{\mathbb{Q}} abelian). Under this assumption Mazur's control theorem [1] on these classical pp-primary Selmer groups implies that the Mordell–Weil rank A(Ln)A(L_{n}) is bounded independently of nn. However, this technique only works under the ordinary assumption. If the abelian variety has supersingular reduction at one of the primes above pp, then the Selp(A/Lcyc){\mathrm{Sel}}_{p}(A/L_{\mathrm{cyc}}) is not cotorsion and Mazur's control theorem will not hold. Now suppose that pp is unramified in LL and AA has supersingular reduction at all primes of LL above pp. Then Büyükboduk and Lei [4] constructed signed Selmer groups for AA; these generalize plus and minus Selmer groups of Kobayashi [5] for ap=0a_{p}=0 and Sprung's /\sharp/\flat-Selmer groups for pap0p\mid a_{p}\neq 0 [6]. The former are proved to be cotorsion for elliptic curves over {\mathbb{Q}} and the latter are conjectured to be cotorsion as well. In the case when pap0p\mid a_{p}\neq 0, at least one of the /\sharp/\flat-Selmer groups is known to be cotorsion [6]. In the case of ap=0a_{p}=0, Kobayashi also proved an analogue of the control theorem for ±\pm-Selmer groups. Combined, those results lead to a bound independent of nn for the growth of the rank for supersingular elliptic curves [5, Corollary 10.2]. Under the assumption that the signed Selmer groups of Büyükboduk and Lei are cotorsion, in [7], Lei and Ponsinet give an explicit sufficient condition in terms of Coleman maps attached to AA which ensures that the rank of the Mordell–Weil group of the dual abelian variety A(Ln)A^{\vee}(L_{n}) stays bounded as nn varies. Further they show that when the Frobenius operator on the Dieudonné module at pp can be expressed as certain block anti-diagonal matrix (cf. [7, Section 3.2.1]) then their explicit condition is satisfied; hence their result applies to abelian varieties of GL2GL_{2}-type.

1.2. Work of Lei–Sprung

In [8] (see also [9] for a less technical exposition), Lei and Sprung consider the special case of elliptic curves defined over {\mathbb{Q}} and the base field F=KF=K, an imaginary quadratic field in which pp splits. The Mordell–Weil rank of EE at the nn-th layer of a p{\mathbb{Z}}_{p}-extension of KK is know to be of the form apn+O(1)ap^{n}+O(1) where aa is an integer called the growth number. When EE has ordinary reduction at pp, this growth number aa is equal to 0 for the cyclotomic p{\mathbb{Z}}_{p}-extension of KK by the works of Kato and Rohrlich. But works of [10], [11], [12], [13] and [14] exhibit cases where a=1a=1 along the anticyclotomic extension of KK when the root number of E/KE/K is 1-1 as it is predicted by Mazur's growth number conjecture [15, §18, p.201]. In the supersingular p5p\geq 5 case, [16] also finds a=1a=1 under some assumptions. So along the anticyclotomic tower, the Mordell–Weil rank may be unbounded. Let KK_{\infty} be the p2{\mathbb{Z}}_{p}^{2}-extension of KK obtained as a compositum of the cyclotomic and the anticyclotomic extension of KK. Let KnK_{n} be a subfield of KK_{\infty} such that Gal(Kn/K)(/pn)2{\mathrm{Gal}}(K_{n}/K)\cong({\mathbb{Z}}/p^{n}{\mathbb{Z}})^{2}. Suppose the elliptic curve EE has good supersingular reduction at pp, Lei and Sprung constructed four /\sharp/\flat-Selmer groups for elliptic curves over the p2\mathbb{Z}_{p}^{2}-extension KK_{\infty}, generalizing works of Kim for the case ap=0a_{p}=0 [17]. Assume that EE is an elliptic curve over {\mathbb{Q}} with conductor prime to pp and the class number of KK is coprime to pp. When p|app|a_{p}, but ap0a_{p}\neq 0, suppose that one of the four ±±\pm\pm-Selmer groups defined by Kim [17] is cotorsion. Then, Lei and Sprung show that rankE(Kn)=O(pn)\mathrm{rank}E(K_{n})=O(p^{n}). In the case ap=0a_{p}=0, they show that rankE(Kn)=O(pn)\mathrm{rank}E(K_{n})=O(p^{n}) under the assumption that all of the four //\sharp/\flat\sharp/\flat-Selmer groups they defined are cotorsion.

1.3. Our work in this article

Kazim Büyükboduk and Antonio Lei jointly started a program to generalize the plus/minus theory of Kobayashi and Pollack in the case of motives crystalline at pp. This led to their series of joint papers [4], [18] and [19]. Note that [4] and [18] (including the work of Lei–Ponsinet mentioned above) are over the cyclotomic p{\mathbb{Z}}_{p}-extension and [19] is for representations coming from modular forms. In this paper our results add to that program giving us results for abelian varieties over a maximal abelian pro-pp extension of a number field that is unramified outside pp.

We restrict ourselves to the case of imaginary quadratic field KK where pp splits. But instead of working with elliptic curves we work with abelian varieties with supersingular reduction at both primes above pp. In this case, we have at our disposal the signed Selmer groups of Büyükboduk and Lei [4] which are defined only over the cyclotomic extension of KK. The objectives in our paper are threefold. We

  • define multi-signed Selmer groups for AA over the p2{\mathbb{Z}}_{p}^{2}-extension KK_{\infty}. In order to achieve this we construct multi-signed Coleman maps attached to the abelian variety AA. This gives us the local condition at primes above pp.

  • give an explicit sufficient condition in terms of Coleman maps (hypothesis (H-large)) mentioned in the text) attached to AA which ensures that the Mordell–Weil rank is bounded by a function which is O(pn)O(p^{n}) along the tower KnK_{n} (see theorem 5.5).

  • show that this explicit condition (H-large) is satisfied when the Frobenius on the Dieudonné module at primes above pp can be expressed in a certain block diagonal form. This special case can be thought of as an analogue for the case ap=0a_{p}=0 for supersingular elliptic curves and happens for abelian varieties of the GL2GL_{2}-type (see section 4.3). When our signed Selmer groups for AA are cotorsion, we hence deduce rankA(Kn)=O(pn)\mathrm{rank}A^{\vee}(K_{n})=O(p^{n}) (see theorem 5.6).

In order to state the main results more precisely, let us introduce some more notation first. Let AA^{\vee} be the dual abelian variety and let SelJ¯(A/K){\mathrm{Sel}}_{\underline{J}}(A^{\vee}/K_{\infty}) be the multi-signed Selmer group attached to AA constructed in section 3.3. Let 𝒳J¯(A/K)\mathcal{X}_{\underline{J}}(A^{\vee}/K_{\infty}) denote the Pontryagin dual

Homcts(SelJ¯(A/K),p/p).\mathrm{Hom}_{\mathrm{cts}}({\mathrm{Sel}}_{\underline{J}}(A^{\vee}/K_{\infty}),\mathbb{Q}_{p}/\mathbb{Z}_{p}).

Let Λ\Lambda be the completed group ring pGal(K/K)\mathbb{Z}_{p}\llbracket{\mathrm{Gal}}(K_{\infty}/K)\rrbracket. Then, the first of our main theorems is the following:

Theorem A (Theorem 5.5).

Suppose that 𝒳J¯(A/K)\mathcal{X}_{\underline{J}}(A^{\vee}/K_{\infty}) is a torsion Λ\Lambda-module for some J¯\underline{J} and that hypothesis (H-large) is satisfied. We have rankA(Kn)=O(pn).\mathrm{rank}A^{\vee}(K_{n})=O(p^{n}).

To state our next theorem, we need to introduce some further notation. Let Tp(A)T_{p}(A) be the Tate module of AA. Suppose that p=𝔭𝔭cp={\mathfrak{p}}{\mathfrak{p}}^{c} where 𝔭{\mathfrak{p}} and 𝔭c{\mathfrak{p}}^{c} are prime ideals of KK. Let 𝔻cris,𝔭(Tp(A))\mathbb{D}_{{\mathrm{cris}},{\mathfrak{p}}}(T_{p}(A)) (reps. 𝔻cris,𝔭c(Tp(A))\mathbb{D}_{{\mathrm{cris}},{\mathfrak{p}}^{c}}(T_{p}(A))) be the Dieudonné module of Tp(A)T_{p}(A) seen as a representation of Gal(K𝔭¯/K𝔭){\mathrm{Gal}}(\overline{K_{\mathfrak{p}}}/K_{\mathfrak{p}}) (resp. Gal(K𝔭c¯/K𝔭c){\mathrm{Gal}}(\overline{K_{{\mathfrak{p}}^{c}}}/K_{{\mathfrak{p}}^{c}})). These modules are naturally equipped with an action of a Frobenius operator. Let C𝔭C_{\mathfrak{p}} and C𝔭cC_{{\mathfrak{p}}^{c}} be the matrices defined in section 2.2 arising from the Frobenius action on 𝔻cris,𝔭(Tp(A))\mathbb{D}_{{\mathrm{cris}},{\mathfrak{p}}}(T_{p}(A)) and 𝔻cris,𝔭c(Tp(A))\mathbb{D}_{{\mathrm{cris}},{\mathfrak{p}}^{c}}(T_{p}(A)) respectively.

Theorem B (Theorem 5.6).

Suppose C𝔭C_{\mathfrak{p}} and C𝔭cC_{{\mathfrak{p}}^{c}} are block anti-diagonal matrices and the Pontryagin dual of the Selmer groups SelJ¯(A/K){\mathrm{Sel}}_{\underline{J}}(A^{{\vee}}/K_{\infty}) for J¯{I¯0,I¯1,I¯mix0,1,I¯mix1,0}\underline{J}\in\{\underline{I}_{0},\underline{I}_{1},\underline{I}_{\mathrm{mix}_{0,1}},\underline{I}_{\mathrm{mix}_{1,0}}\} are all Λ\Lambda-torsion. Then rankA(Kn)=O(pn).\mathrm{rank}A^{\vee}(K_{n})=O(p^{n}).

The Selmer groups appearing in theorem B are all multi-signed Selmer groups for particular choices of the indexing set J¯\underline{J} (see the discussion before proposition 4.14).

Let the abelian variety AA be of dimension gg and such that C𝔭C_{\mathfrak{p}} and C𝔭cC_{{\mathfrak{p}}^{c}} are block anti-diagonal. The indexing set J¯:=(J𝔭,J𝔭c)\underline{J}:=(J_{\mathfrak{p}},J_{{\mathfrak{p}}^{c}}) where J𝔭J_{\mathfrak{p}} and J𝔭cJ_{{\mathfrak{p}}^{c}} are subsets of {1,,2g}\{1,...,2g\} such that |J𝔭|+|J𝔭c|=2g|J_{\mathfrak{p}}|+|J_{{\mathfrak{p}}^{c}}|=2g. Note that the Selmer groups SelJ¯(A/K){\mathrm{Sel}}_{\underline{J}}(A^{\vee}/K_{\infty}) depend not only on the indexing set J¯\underline{J} but also on the choice of Hodge-compatible bases of the Dieudonné modules 𝔻cris,𝔭(Tp(A))\mathbb{D}_{\mathrm{cris},{\mathfrak{p}}}(T_{p}(A)) and 𝔻cris,𝔭c(Tp(A))\mathbb{D}_{\mathrm{cris},{\mathfrak{p}}^{c}}(T_{p}(A)) (see section 3.3 and remark 3.4). This is also the case for the multi-signed Selmer groups defined in [4] and [7] for supersingular abelian varieties in the cyclotomic case. When 2g=42g=4, upon fixing a Hodge-compatible bases, the indexing set J¯\underline{J} has cardinality 7070. But all of these 7070 multi-signed Selmer groups are not canonically defined. Defining multi-signed Selmer groups canonically is crucial in framing corresponding Iwasawa main conjectures relating our multi-signed Selmer groups to canonically defined pp-adic LL-functions in the analytic side. A new key observation made in this paper is that the specific four multi-signed Selmer groups mentioned in theorem B are independent of the choice of the Hodge-compatible bases and hence canonically defined (see proposition 4.18). 111We thank David Loeffler for suggesting this to us. Therefore, this gives a strong incentive for regarding these particular multi-signed choices as ``more fundamental and arithmetically interesting" than the other multi-signed Selmer groups. (In the case of elliptic curves over the p2{\mathbb{Z}}_{p}^{2}-extension KK_{\infty}, they correspond to the usual four /\sharp/\flat signed Selmer groups of Lei and Sprung, generalizing four +/+/- signed Selmer groups of Kim when ap=0a_{p}=0.) Apart from these four multi-signed Selmer groups, we also show that the two other multi-signed Selmer groups corresponding to the choices J¯=({1,,2g},)\underline{J}=(\{1,...,2g\},\emptyset) and J¯=(,{1,,2g})\underline{J}=(\emptyset,\{1,...,2g\}) are also canonically defined and independent of the choice of Hodge-compatible bases (see proposition 4.18). These two Selmer groups should correspond to BDP type pp-adic LL-functions (see section 6).

We now give an outline of the paper. In section 2, we introduce our notations and recall the basic preliminaries on pp-adic Hodge theory that we will need throughout our article. The definition of multi-signed Selmer groups and Coleman maps are given in section 3. In section 4, under the hypothesis (H-large), we use multi-signed Coleman maps and logarithmic matrices to study the growth of

𝒴n\colonequalsCoker(HIw1(K,T)Γnv|pH1(Kn,v,T)Hf1(Kn,v,T))\mathcal{Y}_{n}^{\prime}\colonequals{\mathrm{Coker}}\left(H_{\mathrm{Iw}}^{1}(K_{\infty},T)_{\Gamma_{n}}\to\prod_{v|p}\frac{H^{1}(K_{n,v},T)}{H^{1}_{f}(K_{n,v},T)}\right)

where Hf1(Kn,v,T)H^{1}_{f}(K_{n,v},T) is the Bloch-Kato local condition defined using the kernel of the dual exponential map. To achieve this, we generalized the work of [7] to the setting of p2\mathbb{Z}_{p}^{2}-extensions by using inputs from [8]. We find that rankp(𝒴n)=O(pn)\mathrm{rank}_{\mathbb{Z}_{p}}(\mathcal{Y}_{n}^{\prime})=O(p^{n}). In section 5, we consider the short exact sequence

0𝒴n𝒳n𝒳n000\to\mathcal{Y}_{n}\to\mathcal{X}_{n}\to\mathcal{X}_{n}^{0}\to 0

where 𝒳n\mathcal{X}_{n} is the Pontryagin dual of the classical pp-Selmer group of AA^{\vee} over KnK_{n}, 𝒳n0\mathcal{X}_{n}^{0} is the Pontryagin dual of the fine Selmer group of AA^{\vee} and

𝒴n\colonequalsCoker(HΣ1(Kn,T)v|pH1(Kn,v,T)Hf1(Kn,v,T)).\mathcal{Y}_{n}\colonequals{\mathrm{Coker}}\left(H^{1}_{\Sigma}(K_{n},T)\to\prod_{v|p}\frac{H^{1}(K_{n,v},T)}{H^{1}_{f}(K_{n,v},T)}\right).

Under the assumption that at least one of the Selmer groups SelJ¯(A/K){\mathrm{Sel}}_{\underline{J}}(A^{\vee}/K_{\infty}) is Λ\Lambda-cotorsion, we show that rankp(𝒳n0)\mathrm{rank}_{\mathbb{Z}_{p}}(\mathcal{X}_{n}^{0}) is also O(pn)O(p^{n}). Using the natural surjection 𝒴n𝒴n\mathcal{Y}_{n}^{\prime}\to\mathcal{Y}_{n}, we deduce that the growth of 𝒴n\mathcal{Y}_{n} is at most O(pn)O(p^{n}). We are then able to conclude that the Mordell-Weil rank of AA^{\vee} is O(pn)O(p^{n}) along KK_{\infty}.

One expects that there should be analytic analogues of all of the above, corresponding to our (algebraic) results under appropriate Iwasawa main conjectures. In section 6, written by Chris Williams, we give some speculative remarks on the shape of the analytic theory.

The above concludes the outline. Now, before going to the next section, we would like to make further remarks. In this paper we stick to the case of p2{\mathbb{Z}}_{p}^{2}-extension of an imaginary quadratic field instead of working more generally over pd{\mathbb{Z}}_{p}^{d}-extension of a number field. Our guess is that the growth condition will be O(p(d1)n)O(p^{(d-1)n}), but we will need to make the hypotheses that the multi-signed Selmer groups are cotorsion over the corresponding Iwasawa algebra. Such hypotheses are currently only known for elliptic curves under non-vanishing of signed pp-adic LL-functions over p2{\mathbb{Z}}_{p}^{2}-extension of an imaginary quadratic field (see remark 4.3). The analytic side for general abelian varieties are even more mysterious for d>2d>2 (see section 6), and in this case (the Euler system machinery is also unavailable) as per our knowledge, there are currently no results in the literature towards cotorsion-ness of signed Selmer groups. Even if we assume that the signed Selmer groups are cotorsion, the present techniques in this paper will not generalize verbatim for d>2d>2. We also faced technical difficulty generalizing (4.5) for general pd{\mathbb{Z}}_{p}^{d}-extensions.

Acknowledgements

We are grateful to Antonio Lei for answering many of our questions. We thank David Loeffler and Chris Williams for several useful conversations. We also thank Kâzım Büyükboduk, Eknath Ghate, Mahesh Kakde, Chan-Ho Kim, Debanjana Kundu, Meng Fai Lim, Filippo Nuccio, Gautier Ponsinet, Dipendra Prasad, Anwesh Ray and Romyar Sharifi for comments and corrections that helped improving the quality of the paper. Finally, we are grateful for the referee's many constructive comments. The first named author's research is supported by the Canada Graduate Scholarships – Doctoral program from the Natural Sciences and Engineering Research Council of Canada. The second author gratefully acknowledges support from the Inspire Research Grant, DST, Govt. of India.

2. Preliminaries

2.1. Local and Global setup

Let p3p\geq 3 be a fixed prime number for the rest of the paper. Let KK be an imaginary quadratic field in which pp splits into the primes 𝔭{\mathfrak{p}} and 𝔭c{\mathfrak{p}}^{c} of KK. Here, cc denotes the complex conjugation. We will always use 𝔮{\mathfrak{q}} to mean an element of the set {𝔭,𝔭c}\{{\mathfrak{p}},{\mathfrak{p}}^{c}\}. Let KK_{\infty} be the unique p2\mathbb{Z}_{p}^{2}-extension of KK with Galois group Γ\colonequalsGal(K/K)\Gamma\colonequals{\mathrm{Gal}}(K_{\infty}/K). If 𝔞\mathfrak{a} is an ideal of 𝒪K\mathcal{O}_{K}, K(𝔞)K(\mathfrak{a}) will denote the ray class field of KK of conductor 𝔞\mathfrak{a}. If n0n\geq 0 is an integer, write Γn\colonequalsΓpn\Gamma_{n}\colonequals\Gamma^{p^{n}} and Kn\colonequalsKΓn=K(pn+1)Gal(K(1)/K)K_{n}\colonequals K_{\infty}^{\Gamma_{n}}=K(p^{n+1})^{{\mathrm{Gal}}(K(1)/K)}. We make the hypothesis that K(1)K=KK(1)\bigcap K_{\infty}=K. It follows that the Galois group Γ\Gamma is isomorphic to G𝔭×G𝔭cG_{{\mathfrak{p}}}\times G_{{\mathfrak{p}}^{c}} where G𝔮G_{\mathfrak{q}} is the Galois group of the extension K(𝔮)K/KK({\mathfrak{q}}^{\infty})\bigcap K_{\infty}/K. We fix topological generators γ𝔭\gamma_{\mathfrak{p}} and γ𝔭c\gamma_{{\mathfrak{p}}^{c}} respectively for these groups. Let Λ\colonequalspΓpγ𝔭1,γ𝔭c1\Lambda\colonequals\mathbb{Z}_{p}\llbracket\Gamma\rrbracket\cong\mathbb{Z}_{p}\llbracket\gamma_{\mathfrak{p}}-1,\gamma_{{\mathfrak{p}}^{c}}-1\rrbracket. Sometimes, we shall also use the notation Λ(Γ)\Lambda(\Gamma) for Λ\Lambda when we want to emphasize that Λ\Lambda is the Iwasawa algebra of the group Γ\Gamma. More generally, if GG is any profinite group, we denote by Λ(G)\Lambda(G) the completed group ring pG\mathbb{Z}_{p}\llbracket G\rrbracket. Write Γ𝔮\Gamma_{\mathfrak{q}} for the decomposition group of 𝔮{\mathfrak{q}} in Γ\Gamma. Let μpn\mu_{p^{n}} denote the set of pnp^{n}th roots of unity and let μp\colonequalsn1μpn\mu_{p^{\infty}}\colonequals\bigcup_{n\geq 1}\mu_{p^{n}}.

We let FF_{\infty} and kcyck_{\text{cyc}} be the unramified p\mathbb{Z}_{p}-extension and the cyclotomic p\mathbb{Z}_{p}-extension of p\mathbb{Q}_{p} respectively. Let kk_{\infty} be the compositum of FF_{\infty} and kcyck_{\text{cyc}}. For n0n\geq 0, knk_{n} and FnF_{n} denote the subextension of kcyck_{\text{cyc}} and FF_{\infty} such that [kn:p]=pn[k_{n}:\mathbb{Q}_{p}]=p^{n} and [Fn:p]=pn[F_{n}:\mathbb{Q}_{p}]=p^{n}. Write Γp\colonequalsGal(k/p)p2\Gamma_{p}\colonequals{\mathrm{Gal}}(k_{\infty}/\mathbb{Q}_{p})\cong\mathbb{Z}_{p}^{2}, Γur\colonequalsGal(F/p)=σ¯p\Gamma_{\text{ur}}\colonequals{\mathrm{Gal}}(F_{\infty}/\mathbb{Q}_{p})=\overline{\langle\sigma\rangle}\cong\mathbb{Z}_{p} and Γcyc\colonequalsGal(kcyc/p)=γ¯p\Gamma_{\text{cyc}}\colonequals{\mathrm{Gal}}(k_{\text{cyc}}/\mathbb{Q}_{p})=\overline{\langle\gamma\rangle}\cong\mathbb{Z}_{p}. We identify K𝔮K_{\mathfrak{q}} with p\mathbb{Q}_{p}, Γ𝔮\Gamma_{\mathfrak{q}} with Γp\Gamma_{p} and G𝔮G_{\mathfrak{q}} with Γcyc\Gamma_{\text{cyc}} via γ𝔮γ\gamma_{\mathfrak{q}}\mapsto\gamma. Let (Γcyc)\mathcal{H}(\Gamma_{\text{cyc}}) be the set of power series

n0cn(γ1)n\sum_{n\geq 0}c_{n}\cdot(\gamma-1)^{n}

with coefficients in 𝐐p\mathbf{Q}_{p} such that n0cnXn\sum_{n\geq 0}c_{n}X^{n} converges on the open unit disk.

2.2. pp-adic Hodge Theory

For this subsection only, let KK be any number field where all primes above pp are unramified. Note that for our purpose, pp will be totally split in KK and so we can take Kv=pK_{v}=\mathbb{Q}_{p} in the following discussion where vv is a prime of KK above pp. Denote by 𝒪Kv\mathcal{O}_{K_{v}} the ring of integers of KvK_{v} and let χ:Gal(Kv(μp)/Kv)p×\chi:{\mathrm{Gal}}(K_{v}(\mu_{p^{\infty}})/K_{v})\to\mathbb{Z}_{p}^{\times} be the cyclotomic character. Let /K\mathcal{M}_{/K} be a motive defined over KK in the sense of [20] and p\mathcal{M}_{p} its pp-adic realization. Let TT be a GKG_{K}-stable p\mathbb{Z}_{p}-lattice inside p\mathcal{M}_{p}. We shall denote by T\colonequalsHom(T,μp)T^{\dagger}\colonequals\mathrm{Hom}(T,\mu_{p^{\infty}}) the cartier dual of TT and by T(1)\colonequalsHom(T,p(1))T^{\ast}(1)\colonequals\mathrm{Hom}(T,\mathbb{Z}_{p}(1)) the Tate dual of TT. Suppose that

(H.crys) p\mathcal{M}_{p} is crystalline at all the primes vv above pp in KK.

For simplicity, suppose that the dimension of IndK/p\mathrm{Ind}_{K/\mathbb{Q}}\mathcal{M}_{p} over KvK_{v} is even. This will be the case for example when the motive \mathcal{M} is the motive associated to an abelian variety. Let 2g\colonequalsdimKv(IndK/p)2g\colonequals\mathrm{dim}_{K_{v}}(\mathrm{Ind}_{K/\mathbb{Q}}\mathcal{M}_{p}) and let g±:=dimKv(IndK/p)c=±1g_{\pm}:=\mathrm{dim}_{K_{v}}(\mathrm{Ind}_{K/\mathbb{Q}}\mathcal{M}_{p})^{c=\pm 1}. Let vv be a prime above pp in KK and let 𝔻cris,v(p)\mathbb{D}_{{\mathrm{cris}},v}(\mathcal{M}_{p}) be (𝔹crispp)GKv(\mathbb{B}_{\mathrm{cris}}\otimes_{\mathbb{Q}_{p}}\mathcal{M}_{p})^{G_{K_{v}}} where 𝔹cris\mathbb{B}_{\mathrm{cris}} is the crystalline period ring defined by Fontaine. It admits the structure of a filtered φ\varphi-module. Let 𝔸Kv+\colonequals𝒪Kvπ\mathbb{A}_{K_{v}}^{+}\colonequals\mathcal{O}_{K_{v}}\llbracket\pi\rrbracket where π\pi is seen as a formal variable equipped with a semilinear action by a Frobenius φ\varphi which acts as the absolute Frobenius on 𝒪Kv\mathcal{O}_{K_{v}} and on π\pi by φ(π)=(π+1)p1\varphi(\pi)=(\pi+1)^{p}-1, and with an action of gGal(Kv(μp)/Kv)g\in{\mathrm{Gal}}(K_{v}(\mu_{p^{\infty}})/K_{v}) given by g(π)=(π+1)χ(g)1g(\pi)=(\pi+1)^{\chi(g)}-1. Let v(T)\mathbb{N}_{v}(T) be the Wach module of TT whose existence and properties are shown in [21, Proposition 2.1.1]. It is a free 𝔸Kv+\mathbb{A}_{K_{v}}^{+}-module of rank 2g2g. Furthermore, the quotient v(T)/πv(T)\mathbb{N}_{v}(T)/\pi\mathbb{N}_{v}(T) is identified with a 𝒪Kv\mathcal{O}_{K_{v}}-lattice of 𝔻cris,v(p)\mathbb{D}_{{\mathrm{cris}},v}(\mathcal{M}_{p}). We denote by 𝔻cris,v(T)\mathbb{D}_{{\mathrm{cris}},v}(T) this 𝒪Kv\mathcal{O}_{K_{v}}-lattice. It is equipped with a filtration of 𝒪Kv\mathcal{O}_{K_{v}}-modules {Fili𝔻cris,v(T)}i\{\mathrm{Fil}^{i}\mathbb{D}_{{\mathrm{cris}},v}(T)\}_{i\in\mathbb{Z}} and a Frobenius operator φ\varphi. If we suppose that

(H.HT) the Hodge–Tate weights of p\mathcal{M}_{p} are either 0 or 11,

the filtration takes the form

Fili𝔻cris,v(T)={0if i1,𝔻cris,v(T)if i1.\mathrm{Fil}^{i}\mathbb{D}_{{\mathrm{cris}},v}(T)=\begin{cases}0&\text{if $i\geq 1$,}\\ \mathbb{D}_{{\mathrm{cris}},v}(T)&\text{if $i\leq-1$.}\end{cases}

Note that 𝔻cris,v(V)=𝔻cris,v(T)𝒪KvKv\mathbb{D}_{{\mathrm{cris}},v}(V)=\mathbb{D}_{{\mathrm{cris}},v}(T)\otimes_{\mathcal{O}_{K_{v}}}K_{v}. We also make the following assumptions:

(H.Frob) The slopes of the Frobenius on the Dieudonné module 𝔻cris,v(p)\mathbb{D}_{{\mathrm{cris}},v}(\mathcal{M}_{p}) lie inside [1,0)[-1,0) and that 11 is not an eigenvalue;

(H.P) g+=gg_{+}=g_{-} (=g=g) and dimKvFil0𝔻cris,v(p)=g\mathrm{dim}_{K_{v}}\mathrm{Fil}^{0}\mathbb{D}_{{\mathrm{cris}},v}(\mathcal{M}_{p})=g.

Definition 2.1.

Choose an 𝒪Kv\mathcal{O}_{K_{v}}-basis {v1,,v2g}\{v_{1},\ldots,v_{2g}\} of 𝔻cris,v(T)\mathbb{D}_{{\mathrm{cris}},v}(T) such that {v1,,vg}\{v_{1},\ldots,v_{g}\} is a 𝒪Kv\mathcal{O}_{K_{v}}-basis of Fil0𝔻cris,v(T)\mathrm{Fil}^{0}\mathbb{D}_{{\mathrm{cris}},v}(T). Such a basis is called Hodge-compatible.

The matrix of φ\varphi with respect to this basis is of the form

Cφ,v=Cv[Ig001pIg]C_{\varphi,v}=C_{v}\left[\begin{array}[]{c|c}I_{g}&0\\ \hline\cr 0&\frac{1}{p}I_{g}\end{array}\right]

for some CvGL2g(𝒪Kv)C_{v}\in\mathrm{GL}_{2g}(\mathcal{O}_{K_{v}}) and where IgI_{g} is the identity g×gg\times g matrix.

3. Multi-Signed Selmer groups over p2{\mathbb{Z}}_{p}^{2}-extension

In this section, we define multi-signed Selmer groups for motives /K\mathcal{M}_{/K} satisfying (H.crys), (H.HT), (H.Frob) and (H.P) over p2{\mathbb{Z}}_{p}^{2}-extension of an imaginary quadratic field generalizing an earlier work of Büyükboduk and Lei [19].

Let KK be an imaginary quadratic field where (p)=𝔭𝔭c(p)={\mathfrak{p}}{\mathfrak{p}}^{c} splits. We write Kcyc/KK^{\text{cyc}}/K and Kac/KK^{\text{ac}}/K for the cyclotomic and anticyclotomic p\mathbb{Z}_{p}-extensions contained in KK_{\infty} respectively.

3.1. Yager module

Let L/pL/\mathbb{Q}_{p} be a finite unramified extension. For x𝒪Lx\in\mathcal{O}_{L}, define

yL/p(x)\colonequalsτGal(L/p)τ(x)τ1𝒪L[Gal(L/p)].y_{L/\mathbb{Q}_{p}}(x)\colonequals\sum_{\tau\in{\mathrm{Gal}}(L/\mathbb{Q}_{p})}\tau(x)\cdot\tau^{-1}\in\mathcal{O}_{L}[{\mathrm{Gal}}(L/\mathbb{Q}_{p})].

Let SL/pS_{L/\mathbb{Q}_{p}} be the image of yL/py_{L/\mathbb{Q}_{p}} in 𝒪L[Gal(L/p)]\mathcal{O}_{L}[{\mathrm{Gal}}(L/\mathbb{Q}_{p})]. In [22, Section 3.2], it is shown that there is an isomorphism yF/py_{F_{\infty}/\mathbb{Q}_{p}} of Λ(Γur)\Lambda(\Gamma_{\text{ur}})-modules

limpLF𝒪LSF/p\colonequalslimpLFSL/p\varprojlim_{\mathbb{Q}_{p}\subseteq L\subseteq F_{\infty}}\mathcal{O}_{L}\cong S_{F_{\infty}/\mathbb{Q}_{p}}\colonequals\varprojlim_{\mathbb{Q}_{p}\subseteq L\subseteq F_{\infty}}S_{L/\mathbb{Q}_{p}}

where the inverse limit is taken with respect to the trace maps on the left and the projection maps Gal(L/p)Gal(L/p){\mathrm{Gal}}(L^{\prime}/\mathbb{Q}_{p})\to{\mathrm{Gal}}(L/\mathbb{Q}_{p}) for LLL\subseteq L^{\prime} on the right. By [22, Proposition 3.2], limpLF𝒪L\varprojlim_{\mathbb{Q}_{p}\subseteq L\subseteq F_{\infty}}\mathcal{O}_{L} is a free Λ(Γur)\Lambda(\Gamma_{\text{ur}})-module of rank one, thus the Yager module SF/pS_{F_{\infty}/\mathbb{Q}_{p}} is also free of rank one over Λ(Γur)\Lambda(\Gamma_{\text{ur}}).

3.2. Interpolation property of Loeffler–Zerbes' big logarithm map

Let {Ωp}\{\Omega_{\mathbb{Q}_{p}}\} be a basis of the Yager module SF/pS_{F_{\infty}/\mathbb{Q}_{p}}. Recall that TT is a GKG_{K}-stable p\mathbb{Z}_{p}-lattice inside p\mathcal{M}_{p}. Let T\mathcal{L}_{T}^{\infty} be the two-variable big logarithm map of Loeffler–Zerbes [22]

T:HIw1(k,T)Ωp((Γcyc)^Λ(Γur))p𝔻cris(T).\mathcal{L}_{T}^{\infty}:H^{1}_{\mathrm{Iw}}(k_{\infty},T)\to\Omega_{\mathbb{Q}_{p}}\cdot\left(\mathcal{H}(\Gamma_{\text{cyc}})\widehat{\otimes}\Lambda(\Gamma_{\text{ur}})\right)\otimes_{\mathbb{Z}_{p}}\mathbb{D}_{\mathrm{cris}}(T).

It is a morphism of Λ(Γp)\Lambda(\Gamma_{p})-modules. The completed tensor product (Γcyc)^Λ(Γur)\mathcal{H}(\Gamma_{\text{cyc}})\widehat{\otimes}\Lambda(\Gamma_{\text{ur}}) is isomorphic to (Γp)\mathcal{H}(\Gamma_{p}), the set of power series in γ1\gamma-1 and σ1\sigma-1 with coefficients in p\mathbb{Q}_{p} converging on the open unit disk. Thus, one may see the big logarithm map as an application sending elements of the Iwasawa cohomology to two-variables power series tensored with 𝔻cris(T)\mathbb{D}_{\mathrm{cris}}(T). For knk_{n} a finite subextension of kk_{\infty}, let expT:H1(kn,T)kn𝔻cris(T)\exp_{T}^{\ast}:H^{1}(k_{n},T)\to k_{n}\otimes\mathbb{D}_{\text{cris}}(T) be the Bloch–Kato dual exponential map. We say that the cyclotomic part of a finite order character η:Γp¯p×\eta:\Gamma_{p}\to\overline{{\mathbb{Q}}}_{p}^{\times} is of conductor pn+1p^{n+1} if η\eta sends the topological generator γ\gamma to a primitive pnp^{n}-th root of unity. Then, the big logarithm map enjoys the following interpolation property (see [22, Theorem 4.15]):

Proposition 3.1.

Let xHIw1(k,T)x\in H^{1}_{\mathrm{Iw}}(k_{\infty},T). Let η\eta be a character on Γp\Gamma_{p} whose cyclotomic part is of conductor pn+1>1p^{n+1}>1. Then we have

Φn1T(x)(η)=ε(η1)expT(eηx)\Phi^{-n-1}\mathcal{L}_{T}^{\infty}(x)(\eta)=\varepsilon(\eta^{-1})\exp_{T}^{\ast}(e_{\eta}x)

where ε(η1)\varepsilon(\eta^{-1}) is the ε\varepsilon-factor of η1\eta^{-1}, eηe_{\eta} is the idempotent corresponding to η\eta and Φ\Phi is the operator which act as the arithmetic Frobenius σ\sigma on FmF_{m} and φ\varphi on 𝔻cris(T)\mathbb{D}_{\mathrm{cris}}(T).

For simplicity, let us suppose that the dimension of 𝔻cris(T)\mathbb{D}_{\mathrm{cris}}(T) as a p\mathbb{Z}_{p}-module is even as it will be the case for the applications we have in mind. Write 2g2g for the dimension of 𝔻cris(T)\mathbb{D}_{\mathrm{cris}}(T) as a p\mathbb{Z}_{p}-module where gg is some strictly positive integer. Choose {vi}i=12g\{v_{i}\}_{i=1}^{2g} a Hodge-compatible p\mathbb{Z}_{p}-basis of 𝔻cris(T)\mathbb{D}_{\mathrm{cris}}(T) as in section 2.2. Let LL be a finite unramified extension of p\mathbb{Q}_{p} and let LcycL_{\mathrm{cyc}} denote the cyclotomic p\mathbb{Z}_{p}-extension of LL. Then, in [4], the authors show the existence of one-variable Coleman maps

ColT,L,i:HIw1(Lcyc,T)𝒪LΛ(Γcyc)\mathrm{Col}_{T,L,i}:H^{1}_{\mathrm{Iw}}(L_{\mathrm{cyc}},T)\to\mathcal{O}_{L}\otimes\Lambda(\Gamma_{\mathrm{cyc}})

for 1i2g1\leq i\leq 2g. Those maps are compatible with the corestriction maps HIw1(Fm,cyc,T)HIw1(Fm1,cyc,T)H^{1}_{\mathrm{Iw}}(F_{m,\mathrm{cyc}},T)\to H^{1}_{\mathrm{Iw}}(F_{m-1,\mathrm{cyc}},T) and the trace maps 𝒪FmΛ(Γcyc)𝒪Fm1Λ(Γcyc)\mathcal{O}_{F_{m}}\otimes\Lambda(\Gamma_{\mathrm{cyc}})\to\mathcal{O}_{F_{m-1}}\otimes\Lambda(\Gamma_{\mathrm{cyc}}) (see [23, Appendix A.1] where the Coleman maps ColT,Fm,i\mathrm{Col}_{T,F_{m},i} are denoted by Col/,Fm\mathrm{Col}_{\sharp/\flat,F_{m}} instead). We define two-variable Coleman maps by taking the inverse limit of the ColT,Fm,i\mathrm{Col}_{T,F_{m},i} as FmF_{m} runs through the finite extensions between FF_{\infty} and p\mathbb{Q}_{p}. In order to get a family of maps landing in Λ(Γp)\Lambda(\Gamma_{p}), we further compose it with yF/py_{F_{\infty}/\mathbb{Q}_{p}}. More precisely, the two-variable Coleman maps are defined by

ColT,i:HIw1(k,T)\displaystyle\mathrm{Col}_{T,i}^{\infty}:H^{1}_{\mathrm{Iw}}(k_{\infty},T) ΩpΛ(Γp)\displaystyle\to\Omega_{\mathbb{Q}_{p}}\cdot\Lambda(\Gamma_{p})
(zm)\displaystyle(z_{m}) (yF/p1)limFmColT,Fm,i(zm).\displaystyle\mapsto(y_{F_{\infty}/\mathbb{Q}_{p}}\otimes 1)\circ\varprojlim_{F_{m}}\mathrm{Col}_{T,F_{m},i}(z_{m}).

where (zm)limFmHIw1(Fm,cyc,T)(z_{m})\in\varprojlim_{F_{m}}H^{1}_{\mathrm{Iw}}(F_{m,\mathrm{cyc}},T). By identifying ΩpΛ(Γp)\Omega_{\mathbb{Q}_{p}}\cdot\Lambda(\Gamma_{p}) with Λ(Γp)\Lambda(\Gamma_{p}), we omit Ωp\Omega_{\mathbb{Q}_{p}} from the notation and see ColT,i\mathrm{Col}_{T,i}^{\infty} as taking value in Λ(Γp)\Lambda(\Gamma_{p}). By combining [4, Theorem 2.13] and [22, Theorem 4.7 (1)], we see that these Coleman maps decompose the big logarithm map

T=(v1,,v2g)MT[ColT,1ColT,2g].\mathcal{L}_{T}^{\infty}=(v_{1},\ldots,v_{2g})\cdot M_{T}\cdot\begin{bmatrix}\mathrm{Col}_{T,1}^{\infty}\\ \vdots\\ \mathrm{Col}_{T,2g}^{\infty}\end{bmatrix}.

The matrix MTM_{T} is called a logarithmic matrix and is defined in the following way: Let Φpn(1+X)\Phi_{p^{n}}(1+X) denote the pnp^{n}th cyclotomic polynomial. For n1n\geq 1, first define the matrices

Cn\colonequals[Ig00Φpn(1+X)Ig]C1C_{n}\colonequals\left[\begin{array}[]{c|c}I_{g}&0\\ \hline\cr 0&\Phi_{p^{n}}(1+X)I_{g}\end{array}\right]C^{-1}

and Mn\colonequals(Cφ)n+1CnC1M_{n}\colonequals(C_{\varphi})^{n+1}C_{n}\cdots C_{1} where CC and CφC_{\varphi} are defined as in section 2.2 with v=pv=p. In [4], it is shown that the sequence {Mn}\{M_{n}\} converges to a 2g×2g2g\times 2g matrix with entries in (Γcyc)\mathcal{H}(\Gamma_{\text{cyc}}) which we call MTM_{T}. Let 𝔮{𝔭,𝔭c}{\mathfrak{q}}\in\{{\mathfrak{p}},{\mathfrak{p}}^{c}\}. If we include the primes 𝔭{\mathfrak{p}} and 𝔭c{\mathfrak{p}}^{c} in our notation, then we have the Coleman maps

(3.1) ColT,𝔮,i:H1(K𝔮,TΛ(Γ𝔮)ι)p[[Γp]]\mathrm{Col}_{T,{\mathfrak{q}},i}^{\infty}:H^{1}(K_{\mathfrak{q}},T\otimes\Lambda(\Gamma_{\mathfrak{q}})^{\iota})\rightarrow{\mathbb{Z}}_{p}[[\Gamma_{p}]]

for i{1,,2g}i\in\{1,...,2g\}. They satisfy

T,𝔮=(v1,,v2g)MT,𝔮[ColT,𝔮,1ColT,𝔮,2g].\mathcal{L}_{T,{\mathfrak{q}}}^{\infty}=(v_{1},\ldots,v_{2g})\cdot M_{T,{\mathfrak{q}}}\cdot\begin{bmatrix}\mathrm{Col}_{T,{\mathfrak{q}},1}^{\infty}\\ \vdots\\ \mathrm{Col}_{T,{\mathfrak{q}},2g}^{\infty}\end{bmatrix}.

For 𝔮=𝔭{\mathfrak{q}}={\mathfrak{p}}, MT,𝔭=limrM𝔭,rM_{T,{\mathfrak{p}}}=\lim_{r\rightarrow\infty}M_{{\mathfrak{p}},r}. Here M𝔭,r\colonequals(Cφ,𝔭)r+1C𝔭,rC𝔭,1M_{{\mathfrak{p}},r}\colonequals(C_{\varphi,{\mathfrak{p}}})^{r+1}C_{{\mathfrak{p}},r}\cdots C_{{\mathfrak{p}},1},

C𝔭,r\colonequals[Ig00Φpr(γ𝔭)Ig]C𝔭1.C_{{\mathfrak{p}},r}\colonequals\left[\begin{array}[]{c|c}I_{g}&0\\ \hline\cr 0&\Phi_{p^{r}}(\gamma_{\mathfrak{p}})I_{g}\end{array}\right]C_{\mathfrak{p}}^{-1}.

For 𝔮=𝔭c{\mathfrak{q}}={\mathfrak{p}}^{c} we have expressions MT,𝔭c=limsM𝔭c,sM_{T,{\mathfrak{p}}^{c}}=\lim_{s\rightarrow\infty}M_{{\mathfrak{p}}^{c},s}, where M𝔭c,s\colonequals(Cφ,𝔭c)s+1C𝔭c,sC𝔭c,1M_{{{\mathfrak{p}}^{c}},s}\colonequals(C_{\varphi,{\mathfrak{p}}^{c}})^{s+1}C_{{\mathfrak{p}}^{c},s}\cdots C_{{{\mathfrak{p}}^{c}},1},

C𝔭c,s\colonequals[Ig00Φps(γ𝔭c)Ig]C𝔭c1.C_{{\mathfrak{p}}^{c},s}\colonequals\left[\begin{array}[]{c|c}I_{g}&0\\ \hline\cr 0&\Phi_{p^{s}}(\gamma_{{\mathfrak{p}}^{c}})I_{g}\end{array}\right]C_{{\mathfrak{p}}^{c}}^{-1}.

For any r1r\geq 1, rr an integer, define H𝔭,r=C𝔭,rC𝔭,1H_{{\mathfrak{p}},r}=C_{{\mathfrak{p}},r}\cdots C_{{\mathfrak{p}},1}. Similarly, for any integer s1s\geq 1, we define H𝔭c,s=C𝔭c,sC𝔭c,1H_{{\mathfrak{p}}^{c},s}=C_{{\mathfrak{p}}^{c},s}\cdots C_{{\mathfrak{p}}^{c},1}.
Let Hr,sH_{r,s} be the block diagonal matrix where the blocks on the diagonal are given by H𝔭,rH_{{\mathfrak{p}},r} and H𝔭c,sH_{{\mathfrak{p}}^{c},s}. Consider the tuples I¯=(I𝔭,I𝔭c)\underline{I}=(I_{\mathfrak{p}},I_{{\mathfrak{p}}^{c}}) and J¯=(J𝔭,J𝔭c)\underline{J}=(J_{\mathfrak{p}},J_{{\mathfrak{p}}^{c}}) where I𝔮{1,,2g}I_{\mathfrak{q}}\subseteq\{1,...,2g\} and J𝔮{1,,2g}J_{\mathfrak{q}}\subseteq\{1,...,2g\} for 𝔮{𝔭,𝔭c}{\mathfrak{q}}\in\{{\mathfrak{p}},{\mathfrak{p}}^{c}\}. Let us suppose that |I𝔭|+|I𝔭c|=2g|I_{\mathfrak{p}}|+|I_{{\mathfrak{p}}^{c}}|=2g and |J𝔭|+|J𝔭c|=2g|J_{\mathfrak{p}}|+|J_{{\mathfrak{p}}^{c}}|=2g.
We define HI¯,J¯,r,sH_{\underline{I},\underline{J},r,s} to be the (I¯,J¯)th(\underline{I},\underline{J})^{th}-minor of Hr,sH_{r,s}. We also define I¯0\colonequals(I𝔭,0,I𝔭c,0)\underline{I}_{0}\colonequals(I_{{\mathfrak{p}},0},I_{{\mathfrak{p}}^{c},0}) where I𝔮,0={1,,g}I_{{\mathfrak{q}},0}=\{1,\ldots,g\}.

3.3. Definitions and properties

The multi-signed Selmer group can be defined without the assumption that K(1)K=KK(1)\cap K_{\infty}=K. For this subsection only, we work in more generality and drop this assumption. Suppose that there are ptp^{t} primes above 𝔭{\mathfrak{p}} and 𝔭c{\mathfrak{p}}^{c} in KK_{\infty}. Fix a choice of coset representatives γ1,,γpt\gamma_{1},\ldots,\gamma_{p^{t}} and δ1,,δpt\delta_{1},\ldots,\delta_{p^{t}} for Γ/Γ𝔭\Gamma/\Gamma_{\mathfrak{p}} and Γ/Γ𝔭c\Gamma/\Gamma_{{\mathfrak{p}}^{c}} respectively. Since pp splits in KK, we can identify K𝔮K_{\mathfrak{q}} with p\mathbb{Q}_{p} and Γ𝔮\Gamma_{\mathfrak{q}} with Γp\Gamma_{p}. Consider the ``semi-local" decomposition coming from Shapiro's lemma

H1(K𝔭,TΛι)=j=1ptH1(K𝔭,TΛ(Γ𝔭)ι)γjv|𝔭HIw1(K,v,T)H^{1}(K_{\mathfrak{p}},T\otimes\Lambda^{\iota})=\bigoplus_{j=1}^{p^{t}}H^{1}(K_{\mathfrak{p}},T\otimes\Lambda(\Gamma_{\mathfrak{p}})^{\iota})\cdot\gamma_{j}\cong\bigoplus_{v|{\mathfrak{p}}}H^{1}_{\text{Iw}}(K_{\infty,v},T)

where vv runs through the primes above 𝔭{\mathfrak{p}} in KK_{\infty}. By choosing a Hodge-compatible basis of 𝔻cris,𝔭(T)\mathbb{D}_{\mathrm{cris},{\mathfrak{p}}}(T), define the Coleman map for TT at 𝔭{\mathfrak{p}} by

ColT,𝔭,ik:H1(K𝔭,TΛι)\displaystyle\mathrm{Col}_{T,{\mathfrak{p}},i}^{k_{\infty}}:H^{1}(K_{\mathfrak{p}},T\otimes\Lambda^{\iota}) Λ(Γ)\displaystyle\to\Lambda(\Gamma)
x=j=1ptxjγj\displaystyle x=\sum_{j=1}^{p^{t}}x_{j}\cdot\gamma_{j} j=1ptColT,𝔭,i(xj)γj\displaystyle\mapsto\sum_{j=1}^{p^{t}}\mathrm{Col}_{T,{\mathfrak{p}},i}^{\infty}(x_{j})\cdot\gamma_{j}

for all 1i2g1\leq i\leq 2g.

Let T,𝔭k=j=1ptT,𝔭γj\mathcal{L}_{T,{\mathfrak{p}}}^{k_{\infty}}=\oplus_{j=1}^{p^{t}}\mathcal{L}_{T,{\mathfrak{p}}}^{\infty}\cdot\gamma_{j}. Define ColT,𝔭c,ik\mathrm{Col}_{T,{\mathfrak{p}}^{c},i}^{k_{\infty}} and T,𝔭ck\mathcal{L}_{T,{\mathfrak{p}}^{c}}^{k_{\infty}} in an analogous way. Let J𝔮J_{\mathfrak{q}} denote a subset of {1,,2g}\{1,\ldots,2g\} and let

ColT,J𝔮:H1(K𝔮,TΛι)\displaystyle\mathrm{Col}_{T,J_{\mathfrak{q}}}:H^{1}(K_{\mathfrak{q}},T\otimes\Lambda^{\iota}) i=1|J𝔮|Λ(Γ)\displaystyle\to\bigoplus_{i=1}^{|J_{\mathfrak{q}}|}\Lambda(\Gamma)
z\displaystyle z (ColT,𝔮,ik(z))iJ𝔮.\displaystyle\mapsto(\mathrm{Col}^{k_{\infty}}_{T,{\mathfrak{q}},i}(z))_{i\in J_{\mathfrak{q}}}.

Tate's local pairing induces a pairing

v|𝔮HIw1(K,v,T)×v|𝔮H1(K,v,T)p/p\bigoplus_{v|{\mathfrak{q}}}H^{1}_{\mathrm{Iw}}(K_{\infty,v},T)\times\bigoplus_{v|{\mathfrak{q}}}H^{1}(K_{\infty,v},T^{\dagger})\to\mathbb{Q}_{p}/\mathbb{Z}_{p}

for all places vv of KK_{\infty} above 𝔮{\mathfrak{q}}. We define HJ𝔮1(K,𝔮,T)v|𝔮H1(K,v,T)H^{1}_{J_{\mathfrak{q}}}(K_{\infty,{\mathfrak{q}}},T^{\dagger})\subseteq\bigoplus_{v|{\mathfrak{q}}}H^{1}(K_{\infty,v},T^{\dagger}) as the orthogonal complement of kerColT,J𝔮\ker\mathrm{Col}_{T,J_{\mathfrak{q}}} under the previous pairing. If LL is a finite extension of K𝔮K_{\mathfrak{q}}, define

Hur1(L,p(1))\colonequalsker(H1(L,p(1))H1(Lur,p(1)))H^{1}_{\mathrm{ur}}(L,\mathcal{M}_{p}^{\ast}(1))\colonequals\ker\left(H^{1}(L,\mathcal{M}_{p}^{\ast}(1))\to H^{1}(L_{\mathrm{ur}},\mathcal{M}_{p}^{\ast}(1))\right)

where LurL_{\mathrm{ur}} is the maximal unramified extension of LL. Let Hur1(L,T)H^{1}_{\mathrm{ur}}(L,T^{\dagger}) be the image of Hur1(L,p(1))H^{1}_{\mathrm{ur}}(L,\mathcal{M}_{p}^{\ast}(1)) under the natural map H1(L,p(1))H1(L,T)H^{1}(L,\mathcal{M}_{p}^{\ast}(1))\to H^{1}(L,T^{\dagger}). Let

Hur1(K,w,T)\colonequalslimpLK,wHur1(L,T)H^{1}_{\mathrm{ur}}(K_{\infty,w},T^{\dagger})\colonequals\varinjlim_{\mathbb{Q}_{p}\subseteq L\subseteq K_{\infty,w}}H^{1}_{\mathrm{ur}}(L,T^{\dagger})

where ww is a place in KK_{\infty} and LL runs through finite extensions of p\mathbb{Q}_{p} contained in K,wK_{\infty,w}. Let J¯\colonequals(J𝔭,J𝔭c)\underline{J}\colonequals(J_{\mathfrak{p}},J_{{\mathfrak{p}}^{c}}) where as before J𝔭J_{\mathfrak{p}} and J𝔭cJ_{{\mathfrak{p}}^{c}} are subsets of {1,,2g}\{1,\ldots,2g\} such that |J𝔭|+|J𝔭c|=2g|J_{\mathfrak{p}}|+|J_{{\mathfrak{p}}^{c}}|=2g. Fix Σ\Sigma a finite set of prime of KK containing the primes above pp, the archimedean primes and the primes of ramification of TT^{\dagger}. Let KΣK_{\Sigma} be the maximal extension of KK unramified outside Σ\Sigma. Let Σ\Sigma^{\prime} be the set of primes of KK_{\infty} lying above the primes in Σ\Sigma. If MM is any Gal(KΣ/K){\mathrm{Gal}}(K_{\Sigma}/K_{\infty})-module, we denote H1(KΣ/K,M)H^{1}(K_{\Sigma}/K_{\infty},M) by HΣ1(K,M)H^{1}_{\Sigma}(K_{\infty},M).

Definition 3.2.

Let

𝒫Σ,J¯(T/K)\colonequalswΣ,wpH1(K,w,T)Hur1(K,w,T)×𝔮|pv|𝔮H1(K,v,T)HJ𝔮1(K,𝔮,T).\mathcal{P}_{\Sigma,\underline{J}}(T^{\dagger}/K_{\infty})\colonequals\prod_{w\in\Sigma^{\prime},w\nmid p}\frac{H^{1}(K_{\infty,w},T^{\dagger})}{H^{1}_{\mathrm{ur}}(K_{\infty,w},T^{\dagger})}\times\prod_{{\mathfrak{q}}|p}\frac{\bigoplus_{v|{\mathfrak{q}}}H^{1}(K_{\infty,v},T^{\dagger})}{H^{1}_{J_{\mathfrak{q}}}(K_{\infty,{\mathfrak{q}}},T^{\dagger})}.

The J¯\underline{J}-Selmer group of TT^{\dagger} over KK_{\infty} is

SelJ¯(T/K)\colonequalsker(HΣ1(K,T)𝒫Σ,J¯(T/K)).{\mathrm{Sel}}_{\underline{J}}(T^{\dagger}/K_{\infty})\colonequals\ker\left(H^{1}_{\Sigma}(K_{\infty},T^{\dagger})\to\mathcal{P}_{\Sigma,\underline{J}}(T^{\dagger}/K_{\infty})\right).

We define the fine Selmer group over KK_{\infty} as

Selp0(T/K):=Ker(HΣ1(K,T)vH1(K,v,T)){\mathrm{Sel}}_{p}^{0}(T^{\dagger}/K_{\infty}):={\mathrm{Ker}}\Big{(}H^{1}_{\Sigma}(K_{\infty},T^{\dagger})\rightarrow\prod_{v}H^{1}(K_{\infty,v},T^{\dagger})\Big{)}

where vv runs through all the places of KK_{\infty}.

Remark 3.3.

In the literature, Selp0(T/K){\mathrm{Sel}}_{p}^{0}(T^{\dagger}/K_{\infty}) is sometimes referred as the strict Selmer group instead.

Let 𝒳J¯(T/K)\mathcal{X}_{\underline{J}}(T^{\dagger}/K_{\infty}) and 𝒳0(T/K)\mathcal{X}_{0}(T^{\dagger}/K_{\infty}) denote the Pontryagin dual of the J¯\underline{J}-Selmer group and fine Selmer group respectively. Recall that SelJ¯(T/K){\mathrm{Sel}}_{\underline{J}}(T^{\dagger}/K_{\infty}) (resp. Selp0(T/K){\mathrm{Sel}}_{p}^{0}(T^{\dagger}/K_{\infty})) is said to be Λ\Lambda-cotorsion if 𝒳J¯(T/K)\mathcal{X}_{\underline{J}}(T^{\dagger}/K_{\infty}) (resp. 𝒳0(T/K)\mathcal{X}_{0}(T^{\dagger}/K_{\infty})) is a torsion module over Λ\Lambda.

Remark 3.4.

Change of basis. The Selmer groups SelJ¯(T/K){\mathrm{Sel}}_{\underline{J}}(T^{\dagger}/K_{\infty}) depend both on the indexing set J¯\underline{J} and the choice of the Hodge-compatible basis of the Dieudonné module 𝔻cris,𝔮(T)\mathbb{D}_{\mathrm{cris},{\mathfrak{q}}}(T) (defined after (H.P)). A change of Hodge-compatible basis will affect the Coleman maps ColT,𝔮,i\mathrm{Col}_{T,{\mathfrak{q}},i}^{\infty} in the same way as described in [4, Section 2.4].

Remark 3.5.

Under the hypothesis K(1)K=KK(1)\cap K_{\infty}=K, there is a unique prime above 𝔮{\mathfrak{q}} in KK_{\infty}. In this case,

H1(K𝔮,TΛι)HIw1(K,v,T)H^{1}(K_{\mathfrak{q}},T\otimes\Lambda^{\iota})\cong H^{1}_{\text{Iw}}(K_{\infty,v},T)

where vv is the unique prime above 𝔮{\mathfrak{q}}. Furthermore, Γ=Γ𝔮\Gamma=\Gamma_{\mathfrak{q}} and we get Coleman maps

ColT,𝔮,ik:HIw1(K,v,T)Λ.\mathrm{Col}_{T,{\mathfrak{q}},i}^{k_{\infty}}:H^{1}_{\text{Iw}}(K_{\infty,v},T)\to\Lambda.

3.4. Poitou–Tate exact sequence

By [24, Proposition A.3.2], we have the following Poitou–Tate exact sequence

(3.2) HIw1(K,T)H1(K𝔭,TΛι)kerColT,J𝔭H1(K𝔭c,TΛι)kerColT,J𝔭c𝒳J¯(T/K)𝒳0(T/K)0,H^{1}_{\text{Iw}}(K_{\infty},T)\to\frac{H^{1}(K_{\mathfrak{p}},T\otimes\Lambda^{\iota})}{\ker\mathrm{Col}_{T,J_{\mathfrak{p}}}}\oplus\frac{H^{1}(K_{{\mathfrak{p}}^{c}},T\otimes\Lambda^{\iota})}{\ker\mathrm{Col}_{T,J_{{\mathfrak{p}}^{c}}}}\to\mathcal{X}_{\underline{J}}(T^{\dagger}/K_{\infty})\to\mathcal{X}_{0}(T^{\dagger}/K_{\infty})\to 0,

where HIwi(K,T)H^{i}_{\text{Iw}}(K_{\infty},T) is the inverse limit of Hi(KΣ/Kn,T)H^{i}(K_{\Sigma}/K_{n},T) where KKnKK\subseteq K_{n}\subseteq K_{\infty} and the transition maps are the corestriction maps.

4. Coleman maps control the Mordell–Weil ranks

Let AA be an abelian variety over KK of dimension gg having supersingular reduction at all the primes in KK above pp. Let T=Tp(A)T=T_{p}(A) be the Tate module of AA and V=Tp(A)pV=T_{p}(A)\otimes\mathbb{Q}_{p}. Then V=pV=\mathcal{M}_{p} satisfies the hypotheses (H.crys), (H.HT), (H.Frob) and (H.P). Furthermore, T=A[p]T^{\dagger}=A^{\vee}[p^{\infty}] where AA^{\vee} denotes the dual abelian variety.

4.1. Ranks of Iwasawa modules

We say that VV satisfies weak Leopoldt's conjecture Leop(K,V)\mathrm{Leop}(K_{\infty},V) if the following equivalent statements are true.

Proposition 4.1.

The following statements are equivalent:

  1. (1)

    HIw2(K,T)H^{2}_{\mathrm{Iw}}(K_{\infty},T) is Λ\Lambda-torsion,

  2. (2)

    H2(KΣ/K,Tp/p)=0.H^{2}(K_{\Sigma}/K_{\infty},T\otimes{\mathbb{Q}}_{p}/{\mathbb{Z}}_{p})=0.

Proof.

See [22, Proposition A.5]. ∎

We recall the fine Selmer group over KK_{\infty} is given by

Selp0(A/K)=Ker(HΣ1(K,A[p])vH1(K,v,A[p])).{\mathrm{Sel}}_{p}^{0}(A^{\vee}/K_{\infty})={\mathrm{Ker}}\Big{(}H^{1}_{\Sigma}(K_{\infty},A^{\vee}[p^{\infty}])\rightarrow\prod_{v}H^{1}(K_{\infty,v},A^{\vee}[p^{\infty}])\Big{)}.

Let 𝒳0\mathcal{X}^{0} be the Pontryagin dual of this fine Selmer group over KK_{\infty}. Define

\colonequals{J¯=(J𝔭,J𝔭c):|J𝔭|+|J𝔭c|=2g},\mathcal{I}\colonequals\{\underline{J}=(J_{\mathfrak{p}},J_{{\mathfrak{p}}^{c}}):|J_{\mathfrak{p}}|+|J_{{\mathfrak{p}}^{c}}|=2g\},

where J𝔮{1,,2g}J_{\mathfrak{q}}\subseteq\{1,...,2g\} for 𝔮{𝔭,𝔭c}{\mathfrak{q}}\in\{{\mathfrak{p}},{\mathfrak{p}}^{c}\}.

Lemma 4.2.

Suppose that there exists a J¯\underline{J}\in\mathcal{I} such that SelJ¯(A/K){\mathrm{Sel}}_{\underline{J}}(A^{\vee}/K_{\infty}) is Λ\Lambda-cotorsion. Then HIw1(K,T)H^{1}_{\mathrm{Iw}}(K_{\infty},T) is of rank 2g2g over Λ\Lambda.

Proof.

By the Poitou–Tate exact sequence (3.2), Selp0(A/K){\mathrm{Sel}}_{p}^{0}(A^{\vee}/K_{\infty})^{\vee} is a quotient of SelJ¯(A/K){\mathrm{Sel}}_{\underline{J}}(A^{\vee}/K_{\infty})^{\vee} and hence is Λ\Lambda-torsion. By [24, Proposition A.3.2], we have the exact sequence

0Selp0(A/K)HIw2(K,T)vΣHIw2(K,v,T).0\rightarrow{\mathrm{Sel}}_{p}^{0}(A^{\vee}/K_{\infty})^{\vee}\rightarrow H^{2}_{\mathrm{Iw}}(K_{\infty},T)\rightarrow\prod_{v\in\Sigma}H^{2}_{\mathrm{Iw}}(K_{\infty,v},T).

Now HIw2(K,v,T)H^{2}_{\mathrm{Iw}}(K_{\infty,v},T) is Λ\Lambda-torsion for all vv (see [22, Theorem A.2]). Therefore HIw2(K,T)H^{2}_{\mathrm{Iw}}(K_{\infty},T) is Λ\Lambda-torsion and hence Leop(K,V)\mathrm{Leop}(K_{\infty},V) is true. Now since KK is totally imaginary, [22, Corollary A.8] shows that the rank over Λ\Lambda of HIw1(K,T)H^{1}_{\mathrm{Iw}}(K_{\infty},T) is 2g2g. ∎

We now discuss a partial result toward the converse of lemma 4.2. Suppose that Leop(K,V)\mathrm{Leop}(K_{\infty},V) is true. Then HIw1(K,T)H^{1}_{\mathrm{Iw}}(K_{\infty},T) is of rank 2g2g over Λ\Lambda by [22, Theorem A.4]. Assume that the images of ColT,𝔮,ik\mathrm{Col}_{T,{\mathfrak{q}},i}^{k_{\infty}} (1i2g,𝔮{𝔭,𝔭c}1\leq i\leq 2g,{\mathfrak{q}}\in\{{\mathfrak{p}},{\mathfrak{p}}^{c}\}) are nonzero. Thus, if vv and ww are the unique places above 𝔭{\mathfrak{p}} and 𝔭c{\mathfrak{p}}^{c} in KK_{\infty} respectively,

(4.1) HIw1(K,v,T)HIw1(K,w,T)kerColT,J𝔭kerColT,J𝔭c\frac{H^{1}_{\mathrm{Iw}}(K_{\infty,v},T)\bigoplus H^{1}_{\mathrm{Iw}}(K_{\infty,w},T)}{\ker\mathrm{Col}_{T,J_{{\mathfrak{p}}}}\bigoplus\ker\mathrm{Col}_{T,J_{{\mathfrak{p}}^{c}}}}

is of rank 2g2g since |J𝔭|+|J𝔭c|=2g|J_{\mathfrak{p}}|+|J_{{\mathfrak{p}}^{c}}|=2g and HIw1(K,v,T)HIw1(K,w,T)H^{1}_{\mathrm{Iw}}(K_{\infty,v},T)\bigoplus H^{1}_{\mathrm{Iw}}(K_{\infty,w},T) has rank 4g4g by [22, Theorem A.2]. By [25, proof of proposition 6], H1(K,w,A[p])Hur1(K,w,A[p])\frac{H^{1}(K_{\infty,w},A^{\vee}[p^{\infty}])}{H^{1}_{\mathrm{ur}}(K_{\infty,w},A^{\vee}[p^{\infty}])} is Λ\Lambda-cotorsion for wΣw\in\Sigma^{\prime} such that wpw\nmid p. We deduce, by local Tate duality, that the Λ\Lambda-module (4.1) has the same rank as the Pontryagin dual of 𝒫Σ,J¯(A[p]/K)\mathcal{P}_{\Sigma,\underline{J}}(A^{\vee}[p^{\infty}]/K_{\infty}). Thus 𝒫Σ,J¯(A[p]/K)\mathcal{P}_{\Sigma,\underline{J}}(A^{\vee}[p^{\infty}]/K_{\infty}) has Λ\Lambda-corank 2g2g and it follows that SelJ¯(A/K){\mathrm{Sel}}_{\underline{J}}(A^{\vee}/K_{\infty}) is the kernel of a morphism between two modules of the same corank. Being such a kernel is a necessary (but not sufficient) condition for SelJ¯(A/K){\mathrm{Sel}}_{\underline{J}}(A^{\vee}/K_{\infty}) to be Λ\Lambda-cotorsion.

Remark 4.3.

Suppose that A=EA=E is an elliptic curve. In that case, the Coleman maps defined in section 3.3 are the same as the /\sharp/\flat-Coleman maps in [8]

Col𝔮,Col𝔮:H1(K𝔮,Tp(E)Λι)Λ(Γ).\mathrm{Col}_{{\mathfrak{q}}}^{\sharp},\mathrm{Col}_{{\mathfrak{q}}}^{\flat}:H^{1}(K_{\mathfrak{q}},T_{p}(E)\otimes\Lambda^{\iota})\to\Lambda(\Gamma).

For ,{,}\bullet,\circ\in\{\sharp,\flat\}, it is then possible to construct signed pp-adic LL-functions 𝔏p,(E/K)\mathfrak{L}_{p}^{\bullet,\circ}(E/K). If we suppose that 𝔏p,(E/K)\mathfrak{L}_{p}^{\bullet,\circ}(E/K) is nonzero, then [26, Theorem 3.7] shows that the signed Selmer group Sel,(E/K){\mathrm{Sel}}^{\bullet,\circ}(E/K_{\infty}) is Λ\Lambda-cotorsion. Moreover, when ap=0a_{p}=0, the /\sharp/\flat-Selmer groups generalise Kim's ±/±\pm/\pm-Selmer groups [17]. The cotorsion-ness of those Selmer groups is known unconditionally by [27, Remark 8.5] in the case J¯=(1,1)\underline{J}=(1,1) and J¯=(2,2)\underline{J}=(2,2) corresponding to +/++/+ and /-/- Selmer groups.

In the next lemma, we give a condition over the cyclotomic extension which will ensure that HIw1(K,T)H^{1}_{\mathrm{Iw}}(K_{\infty},T) is of rank 2g2g over Λ\Lambda.

Lemma 4.4.

Suppose that there exists any J¯\underline{J}\in\mathcal{I} such that SelJ¯(A/Kcyc){\mathrm{Sel}}_{\underline{J}}(A^{\vee}/K^{\mathrm{cyc}}) is p[[Gal(Kcyc/K)]]{\mathbb{Z}}_{p}[[{\mathrm{Gal}}(K^{\mathrm{cyc}}/K)]]-cotorsion. Then HIw1(K,T)H^{1}_{\mathrm{Iw}}(K_{\infty},T) is of rank 2g2g over Λ\Lambda.

Proof.

Since KK is totally imaginary, in view of [22, Corollary A.8], it is sufficient to show that Leop(K,V)\mathrm{Leop}(K_{\infty},V) is true. By the discussion before Example A.7 of (loc.cit), Leop(Kcyc,V)\mathrm{Leop}(K^{\mathrm{cyc}},V) implies Leop(K,V)\mathrm{Leop}(K_{\infty},V). But [7, Lemma 2.4(2)] implies that SelJ¯(A/Kcyc){\mathrm{Sel}}_{\underline{J}}(A^{\vee}/K^{\mathrm{cyc}}) being cotorsion as a module over p[[Gal(Kcyc/K)]]{\mathbb{Z}}_{p}[[{\mathrm{Gal}}(K^{\mathrm{cyc}}/K)]] is a sufficient condition for Leop(Kcyc,V)\mathrm{Leop}(K^{\mathrm{cyc}},V) to hold (see Remark 4.5). ∎

Remark 4.5.

Note that [7, Lemma 2.4(2)] says that it needs to assume that SelJ¯(A/Kcyc){\mathrm{Sel}}_{\underline{J}}(A^{\vee}/K^{\mathrm{cyc}}) is cotorsion for all J¯\underline{J}\in\mathcal{I} in order to ensure Leop(Kcyc,V)\mathrm{Leop}(K^{\mathrm{cyc}},V) holds. But a careful analysis of their proof reveals that the cotorsionness of SelJ¯(A/Kcyc){\mathrm{Sel}}_{\underline{J}}(A^{\vee}/K^{\mathrm{cyc}}) for only one J¯\underline{J} is sufficient.

Remark 4.6.

[4, Proposition 3.28] gives a sufficient condition for SelJ¯(A/K(μp))+{\mathrm{Sel}}_{\underline{J}}(A^{\vee}/K(\mu_{p^{\infty}}))_{+} to be cotorsion (the plus notation is defined before proposition 3.14 of [4]).

Remark 4.7.

If the Bloch-Kato Selmer group SelBK(A/K){\mathrm{Sel}}_{\mathrm{BK}}(A^{\vee}/K) is finite, it is known that SelJ¯(A/Kcyc){\mathrm{Sel}}_{\underline{J}}(A^{\vee}/K^{\mathrm{cyc}}) is p[[Gal(Kcyc/K)]]{\mathbb{Z}}_{p}[[{\mathrm{Gal}}(K^{\mathrm{cyc}}/K)]]-cotorsion (cf. [28, Corollary 2.10 and Remark 2.2]). In this case, it is also known that SelJ¯(A/K){\mathrm{Sel}}_{\underline{J}}(A^{\vee}/K_{\infty}) is Λ\Lambda-cotorsion (cf. [29, Corollary 1.1]).

4.2. Bound using logarithmic matrices

Let FF be a pp-adic local field. Consider the Bloch–Kato's dual exponential map

exp:H1(F,T)F𝔻cris(T)\mathrm{exp}^{*}:H^{1}(F,T)\rightarrow F\otimes\mathbb{D}_{\mathrm{cris}}(T)

with kernel Hf1(F,T)H^{1}_{f}(F,T). Note that the image of the dual Bloch–Kato exponential map exp\mathrm{exp}^{*} lands inside FFil0𝔻cris(T)F\otimes\mathrm{Fil}^{0}\mathbb{D}_{\mathrm{cris}}(T) and Fil0𝔻cris(T)\mathrm{Fil}^{0}\mathbb{D}_{\mathrm{cris}}(T) has p{\mathbb{Z}}_{p}-rank g=g+=gg=g_{+}=g_{-} by (H.P). Denote

H/f1(F,T)=H1(F,T)/Hf1(F,T)H^{1}_{/f}(F,T)=H^{1}(F,T)/H^{1}_{f}(F,T)

and

H/f1(Kn,p,T)=vpH/f1(Kn,v,T).H^{1}_{/f}(K_{n,p},T)=\prod_{v\mid p}H^{1}_{/f}(K_{n,v},T).

For n0n\geq 0, we define

(4.2) 𝒴n\displaystyle\mathcal{Y}^{\prime}_{n} :=Coker(HIw1(K,T)ΓnvpH/f1(Kn,v,T)).\displaystyle:={\mathrm{Coker}}\Big{(}H^{1}_{\mathrm{Iw}}(K_{\infty},T)_{\Gamma_{n}}\rightarrow\prod_{v\mid p}H^{1}_{/f}(K_{n,v},T)\Big{)}.

For i[1,2g]i\in[1,2g], let us suppose that we fix a family of classes c1,,c2gHIw1(K,T)c_{1},...,c_{2g}\in H^{1}_{\mathrm{Iw}}(K_{\infty},T) such that HIw1(K,T)/c1,,c2gH^{1}_{\mathrm{Iw}}(K_{\infty},T)/\langle c_{1},\cdots,c_{2g}\rangle is Λ\Lambda-torsion. Their existence is guaranteed since HIw1(K,T)H^{1}_{\mathrm{Iw}}(K_{\infty},T) is of rank 2g2g (under the hypothesis of either lemma 4.2 or lemma 4.4). Let locp,n(ci)\mathrm{loc}_{p,n}(c_{i}) be the image of cic_{i} in H/f1(Kn,p,T)H^{1}_{/f}(K_{n,p},T).

We write W=μp×μpW=\mu_{p^{\infty}}\times\mu_{p^{\infty}} and p[w]=p[w1,w2]{\mathbb{Z}}_{p}[w]={\mathbb{Z}}_{p}[w_{1},w_{2}] for (w1,w2)W.(w_{1},w_{2})\in W. If FΛF\in\Lambda, we can evaluate FF at ww in the following way. We write FF as a power series in γ𝔭1\gamma_{\mathfrak{p}}-1 and γ𝔭c1\gamma_{{\mathfrak{p}}^{c}}-1, say F0(γ𝔭1,γ𝔭c1)F_{0}(\gamma_{{\mathfrak{p}}}-1,\gamma_{{\mathfrak{p}}^{c}}-1). We then define

F(w)=F0(w11,w21).F(w)=F_{0}(w_{1}-1,w_{2}-1).

For a Λ\Lambda-module MM, we write Mw=Mp[w]M_{w}=M\otimes{\mathbb{Z}}_{p}[w] for the p[w]{\mathbb{Z}}_{p}[w]-module induced by this evaluation map.

Let o(w1)o(w_{1}) (resp. o(w2)o(w_{2})) be the least r1r_{1} (resp. r2r_{2}) such that w1pr1=1w_{1}^{p^{r_{1}}}=1 (resp. w2pr2=1w_{2}^{p^{r_{2}}}=1). Then p[w]{\mathbb{Z}}_{p}[w] is a free p{\mathbb{Z}}_{p}-module of rank φ(pr)\varphi(p^{r}) where r=max{o(w1),o(w2)}r=\max\{o(w_{1}),o(w_{2})\}. Let n1n\geq 1 be an integer. There are exactly p2np^{2n} elements wWw\in W such that wpn=1w^{p^{n}}=1; two such elements ww are called conjugate if there is an automorphism of the algebraic closure of p{\mathbb{Q}}_{p} taking one to the other [30, Section 2, p. 238]. The number of conjugates of a given ww is the rank of p[w]{\mathbb{Z}}_{p}[w].

Lemma 4.8.

Let n1n\geq 1 be an integer. We have

rankp𝒴nrankp𝒴n1wrankp(H/f1(Kn,p,T)locp,n(c1),,locp,n(c2g))w\mathrm{rank}_{{\mathbb{Z}}_{p}}\mathcal{Y}^{\prime}_{n}-\mathrm{rank}_{{\mathbb{Z}}_{p}}\mathcal{Y}^{\prime}_{n-1}\leq\sum_{w}\mathrm{rank}_{{\mathbb{Z}}_{p}}\Big{(}\frac{H^{1}_{/f}(K_{n,p},T)}{\langle\mathrm{loc}_{p,n}(c_{1}),\cdots,\mathrm{loc}_{p,n}(c_{2g})\rangle}\Big{)}_{w}

where the sum runs over conjugacy classes of wWw\in W such that wpn=(1,1)w^{p^{n}}=(1,1), but wpn1(1,1)w^{p^{n-1}}\neq(1,1).

Proof.

We will essentially follow the line of sketch given in [8, Corollary 5.7] with some modifications. The Bloch–Kato dual exponential map implies

vpH/f1(Kn,v,T)ppvpKn,vg.\prod_{v\mid p}H^{1}_{/f}(K_{n,v},T)\otimes_{{\mathbb{Z}}_{p}}{\mathbb{Q}}_{p}\cong\bigoplus_{v\mid p}K_{n,v}^{\oplus g}.

This gives

(H/f1(Kn,p,T)pp)Γn1=H/f1(Kn1,p,T)pp\Big{(}H^{1}_{/f}(K_{n,p},T)\otimes_{{\mathbb{Z}}_{p}}{\mathbb{Q}}_{p}\Big{)}_{\Gamma_{n-1}}=H^{1}_{/f}(K_{n-1,p},T)\otimes_{{\mathbb{Z}}_{p}}{\mathbb{Q}}_{p}

which implies

(H/f1(Kn,p,T)pp)w=(H/f1(Kn1,p,T)pp)w\Big{(}H^{1}_{/f}(K_{n,p},T)\otimes_{{\mathbb{Z}}_{p}}{\mathbb{Q}}_{p}\Big{)}_{w}=\Big{(}H^{1}_{/f}(K_{n-1,p},T)\otimes_{{\mathbb{Z}}_{p}}{\mathbb{Q}}_{p}\Big{)}_{w}

for wpn1=(1,1)w^{p^{n-1}}=(1,1). Now suppose MM is a Λ\Lambda-module. Then [30, Lemma 2.7] implies that

rankpMΓn=wrankpMw\mathrm{rank}_{{\mathbb{Z}}_{p}}M_{\Gamma_{n}}=\sum_{w}\mathrm{rank}_{{\mathbb{Z}}_{p}}M_{w}

where the direct sum runs over the conjugacy classes of wWw\in W such that wpn=(1,1)w^{p^{n}}=(1,1). Applying this for M=H/f1(Kn,p,T)M=H^{1}_{/f}(K_{n,p},T), we deduce

rankpH/f1(Kn,p,T)=rankpH/f1(Kn1,p,T)+wrankp(H/f1(Kn,p,T))w,\mathrm{rank}_{{\mathbb{Z}}_{p}}H^{1}_{/f}(K_{n,p},T)=\mathrm{rank}_{{\mathbb{Z}}_{p}}H^{1}_{/f}(K_{n-1,p},T)+\sum_{w}\mathrm{rank}_{{\mathbb{Z}}_{p}}\Big{(}H^{1}_{/f}(K_{n,p},T)\Big{)}_{w},

where the sum runs over conjugacy classes of wWw\in W such that wpn=(1,1)w^{p^{n}}=(1,1), but wpn1(1,1)w^{p^{n-1}}\neq(1,1). We obtain a similar decomposition for HIw1(K,T)ΓnH^{1}_{\mathrm{Iw}}(K_{\infty},T)_{\Gamma_{n}}. Let us abbreviate H/f1(Kn,p,T)H^{1}_{/f}(K_{n,p},T) by Hn1H^{1}_{n} and HIw1(K,T)ΓnH^{1}_{\mathrm{Iw}}(K_{\infty},T)_{\Gamma_{n}} by HIw,n1H^{1}_{\mathrm{Iw},n}. Functoriality of coinvariance and tensor products gives the following commutative diagram

Hn1p{H^{1}_{n}\otimes{\mathbb{Q}}_{p}}{\cong}Hn11p{H^{1}_{n-1}\otimes{\mathbb{Q}}_{p}}{\oplus}(wW(Hn1)wp){\Big{(}\oplus_{w\in W}\big{(}H^{1}_{n}\big{)}_{w}\otimes{\mathbb{Q}}_{p}\Big{)}}HIw,n1p{H^{1}_{\mathrm{Iw},n}\otimes{\mathbb{Q}}_{p}}{\cong}HIw,n11p{H^{1}_{\mathrm{Iw},n-1}\otimes{\mathbb{Q}}_{p}}{\oplus}(wW(HIw,n1)wp){\Big{(}\oplus_{w\in W}\big{(}H^{1}_{\mathrm{Iw},n}\big{)}_{w}\otimes{\mathbb{Q}}_{p}\Big{)}}

where the sum on the right runs over conjugacy classes of wWw\in W such that wpn=(1,1)w^{p^{n}}=(1,1), but wpn1(1,1)w^{p^{n-1}}\neq(1,1). This implies

rankp𝒴nrankp𝒴n1=wrankp(𝒴n)w.\mathrm{rank}_{{\mathbb{Z}}_{p}}\mathcal{Y}^{\prime}_{n}-\mathrm{rank}_{{\mathbb{Z}}_{p}}\mathcal{Y}^{\prime}_{n-1}=\sum_{w}\mathrm{rank}_{{\mathbb{Z}}_{p}}(\mathcal{Y}^{\prime}_{n})_{w}.

The lemma follows from noting that 𝒴n\mathcal{Y}^{\prime}_{n} is a quotient of H/f1(Kn,p,T)locp,n(c1),,locp,n(c2g).\frac{H^{1}_{/f}(K_{n,p},T)}{\langle\mathrm{loc}_{p,n}(c_{1}),\cdots,\mathrm{loc}_{p,n}(c_{2g})\rangle}.

Suppose θ\theta is a character of Γ\Gamma sending (γ𝔭,γ𝔭c)(\gamma_{\mathfrak{p}},\gamma_{{\mathfrak{p}}^{c}}) to (w1,w2)W.(w_{1},w_{2})\in W. Then evaluating an element of p[[X,Y]]{\mathbb{Z}}_{p}[[X,Y]] at X=w11X=w_{1}-1 and Y=w21Y=w_{2}-1 is equivalent to evaluating an element of Λ\Lambda at θ\theta. Let w=(w1,w2)Ww=(w_{1},w_{2})\in W with w1w_{1} a primitive prp^{r}-th root of unity and w2w_{2} a primitive psp^{s}-th root of unity, then the corresponding character of Γ\Gamma has conductor 𝔭r+1(𝔭c)s+1.{\mathfrak{p}}^{r+1}({\mathfrak{p}}^{c})^{s+1}. Suppose θ\theta be a character of Γ\Gamma of conductor 𝔭r+1(𝔭c)s+1\mathfrak{p}^{r+1}(\mathfrak{p}^{c})^{s+1}. When restricted to Γ𝔭\Gamma_{\mathfrak{p}} (respectively Γ𝔭c\Gamma_{\mathfrak{p}^{c}}), the character θ\theta gives a character of Γp\Gamma_{p} whose cyclotomic part is of conductor pr+1p^{r+1} (respectively ps+1p^{s+1})).

Recall that J¯=(J𝔭,J𝔭c)\underline{J}=(J_{\mathfrak{p}},J_{{\mathfrak{p}}^{c}}) where |J𝔭|+|J𝔭c|=2g|J_{\mathfrak{p}}|+|J_{{\mathfrak{p}}^{c}}|=2g.
Let 𝐳=z1z2z2g2g(vpH1(Kv,Tp[[Γp]]ι)).\mathbf{z}=z_{1}\wedge z_{2}\wedge\cdots\wedge z_{2g}\in\bigwedge^{2g}\Big{(}\prod_{v\mid p}H^{1}(K_{v},T\otimes{\mathbb{Z}}_{p}[[\Gamma_{p}]]^{\iota})\Big{)}. Then we define

ColT,J¯k(𝐳)=det(ColT,v,ik(zj))v{𝔭,𝔭c},1j2g,iJv.\mathrm{Col}_{T,\underline{J}}^{k_{\infty}}(\mathbf{z})=\det\Big{(}\mathrm{Col}_{T,v,i}^{k_{\infty}}(z_{j})\Big{)}_{v\in\{{\mathfrak{p}},{\mathfrak{p}}^{c}\},1\leq j\leq 2g,i\in J_{v}.}
Lemma 4.9.

Let 𝐳=z1z2z2g2g(vpH1(Kv,Tp[[Γp]]ι))\mathbf{z}=z_{1}\wedge z_{2}\wedge\cdots\wedge z_{2g}\in\bigwedge^{2g}\Big{(}\prod_{v\mid p}H^{1}(K_{v},T\otimes{\mathbb{Z}}_{p}[[\Gamma_{p}]]^{\iota})\Big{)} and let θ\theta be a character of Γ\Gamma of conductor 𝔭r+1(𝔭c)s+1\mathfrak{p}^{r+1}(\mathfrak{p}^{c})^{s+1}. Put n\colonequalsmax{r,s}n\colonequals\max\{r,s\}. Write eθe_{\theta} for the idempotent corresponding to θ\theta. Then eθ𝐳e_{\theta}\cdot\mathbf{z} lies in 2g(vpHf1(Kn,v,T))\bigwedge^{2g}\Big{(}\prod_{v\mid p}H^{1}_{f}(K_{n,v},T)\Big{)} if and only if

J¯HI¯0,J¯,r,sColT,J¯k(𝐳)(θ)=0.\sum_{\underline{J}\in\mathcal{I}}H_{\underline{I}_{0},\underline{J},r,s}\mathrm{Col}_{T,\underline{J}}^{k_{\infty}}(\mathbf{z})(\theta)=0.
Proof.

If vv is a place above 𝔭\mathfrak{p} or 𝔭c\mathfrak{p}^{c} in KnK_{n}, we have the dual exponential maps

expv:H1(Kn,v,T)Kn,v𝔻cris,v(T).\mathrm{exp}_{v}^{*}:H^{1}(K_{n,v},T)\rightarrow K_{n,v}\otimes\mathds{D}_{{\mathrm{cris}},v}(T).

Define exp𝔭=v𝔭expv{\mathrm{exp}}_{{\mathfrak{p}}}^{*}=\prod_{v\mid{\mathfrak{p}}}{\mathrm{exp}}_{v}^{*} and exp𝔭c=v𝔭cexpv{\mathrm{exp}}_{{\mathfrak{p}}^{c}}^{*}=\prod_{v\mid{\mathfrak{p}}^{c}}{\mathrm{exp}}_{v}^{*} where the products run over the places dividing 𝔭{\mathfrak{p}} and 𝔭c{\mathfrak{p}}^{c} in KnK_{n}. Then eθ𝐳e_{\theta}\cdot\mathbf{z} lies in 2g(vpHf1(Kn,v,T))\bigwedge^{2g}\Big{(}\prod_{v\mid p}H^{1}_{f}(K_{n,v},T)\Big{)} if and only if

eθ1j2g(exp𝔭(zj)×exp𝔭c(zj))=0.e_{\theta}\cdot\bigwedge_{1\leq j\leq 2g}\Big{(}{\mathrm{exp}}^{*}_{\mathfrak{p}}(z_{j})\times{\mathrm{exp}}^{*}_{{\mathfrak{p}}^{c}}(z_{j})\Big{)}=0.

(Here exp𝔭(zj){\mathrm{exp}}^{*}_{\mathfrak{p}}(z_{j}) is the dual Bloch–Kato exponential map applied to the appropriate projection of zjz_{j}). Via the interpolation property of the Loeffler–Zerbes' big logarithm map , this is equivalent to

(4.3) 1j2g(Φr1T,𝔭k(zj)×Φs1T,𝔭ck(zj))(θ)=0.\bigwedge_{1\leq j\leq 2g}\Big{(}\Phi^{-r-1}\mathcal{L}_{T,{\mathfrak{p}}}^{k_{\infty}}(z_{j})\times\Phi^{-s-1}\mathcal{L}_{T,{\mathfrak{p}}^{c}}^{k_{\infty}}(z_{j})\Big{)}(\theta)=0.

Applying Φr1\Phi^{-r-1} to T,𝔭k(zj)\mathcal{L}_{T,{\mathfrak{p}}}^{k_{\infty}}(z_{j}) we obtain

Φr1T,𝔭k(zj)(θ)=(v1,,v2g)C𝔭,rC𝔭,1(θ)[ColT,𝔭,1k(zj)σr1(θ)ColT,𝔭,2gk(zj)σr1(θ)].\Phi^{-r-1}\mathcal{L}_{T,{\mathfrak{p}}}^{k_{\infty}}(z_{j})(\theta)=(v_{1},\cdots,v_{2g})C_{{\mathfrak{p}},r}\cdots C_{{\mathfrak{p}},1}(\theta)\begin{bmatrix}\mathrm{Col}_{T,{\mathfrak{p}},1}^{k_{\infty}}(z_{j})^{\sigma^{-r-1}}(\theta)\\ \vdots\\ \mathrm{Col}_{T,{\mathfrak{p}},2g}^{k_{\infty}}(z_{j})^{\sigma^{-r-1}}(\theta)\end{bmatrix}.

Similarly, applying Φs1\Phi^{-s-1} to T,𝔭ck(zj)\mathcal{L}_{T,{\mathfrak{p}}^{c}}^{k_{\infty}}(z_{j}) we obtain

Φs1T,𝔭ck(zj)(θ)=(v1,,v2g)C𝔭c,sC𝔭c,1(θ)[ColT,𝔭c,1k(zj)σs1(θ)ColT,𝔭c,2gk(zj)σs1(θ)].\Phi^{-s-1}\mathcal{L}_{T,{\mathfrak{p}}^{c}}^{k_{\infty}}(z_{j})(\theta)=(v_{1},\cdots,v_{2g})C_{{\mathfrak{p}}^{c},s}\cdots C_{{\mathfrak{p}}^{c},1}(\theta)\begin{bmatrix}\mathrm{Col}_{T,{\mathfrak{p}}^{c},1}^{k_{\infty}}(z_{j})^{\sigma^{-s-1}}(\theta)\\ \vdots\\ \mathrm{Col}_{T,{\mathfrak{p}}^{c},2g}^{k_{\infty}}(z_{j})^{\sigma^{-s-1}}(\theta)\end{bmatrix}.

Note that H𝔭,r(θ)=C𝔭,rC𝔭,1(θ)H_{{\mathfrak{p}},r}(\theta)=C_{{\mathfrak{p}},r}\cdots C_{{\mathfrak{p}},1}(\theta) and H𝔭c,s(θ)=C𝔭c,sC𝔭c,1(θ)H_{{\mathfrak{p}}^{c},s}(\theta)=C_{{\mathfrak{p}}^{c},s}\cdots C_{{\mathfrak{p}}^{c},1}(\theta). Therefore by taking wedge product we obtain that (4.3) is equivalent to the vanishing of

I¯,J¯HI¯,J¯,r,sColT,J¯k(𝐳)(θ).\sum_{\underline{I},\underline{J}\in\mathcal{I}}H_{\underline{I},\underline{J},r,s}\mathrm{Col}_{T,\underline{J}}^{k_{\infty}}(\mathbf{z})(\theta).

Now we claim that HI¯,J¯,r,s(θ)H_{\underline{I},\underline{J},r,s}(\theta) is zero unless I=I¯0I=\underline{I}_{0} and J¯\underline{J}\in\mathcal{I}.

If i{g+1,,2g}i\in\{g+1,...,2g\}, then the entire ii-th row of C𝔭,rC_{{\mathfrak{p}},r} (respectively C𝔭c,sC_{{\mathfrak{p}}^{c},s}) is divisible by Φpr(γ𝔭)\Phi_{p^{r}}(\gamma_{\mathfrak{p}}) (respectively Φps(γ𝔭c)\Phi_{p^{s}}(\gamma_{{\mathfrak{p}}^{c}})). That is, the lower half of H𝔭,rH_{{\mathfrak{p}},r} and H𝔭c,sH_{{\mathfrak{p}}^{c},s} are always zero. Hence, in order for a 2g×2g2g\times 2g minor to be nonzero we should take the upper halves of H𝔭,rH_{{\mathfrak{p}},r} and H𝔭c,sH_{{\mathfrak{p}}^{c},s}. Hence the claim follows. ∎

Lemma 4.10.

Let θ\theta be a character of Γ\Gamma of conductor 𝔭r+1(𝔭c)s+1{\mathfrak{p}}^{r+1}({\mathfrak{p}}^{c})^{s+1} and let w=(θ(γp),θ(γ𝔭c))Ww=(\theta(\gamma_{p}),\theta(\gamma_{{\mathfrak{p}}^{c}}))\in W. Let n\colonequalsmax{r,s}n\colonequals\max\{r,s\}. If

J¯HI¯0,J¯,r,sColT,J¯k(𝐜)(θ)0\sum_{\underline{J}\in\mathcal{I}}H_{\underline{I}_{0},\underline{J},r,s}\mathrm{Col}_{T,\underline{J}}^{k_{\infty}}(\mathbf{c})(\theta)\neq 0

then

rankp(H/f1(Kn,p,T)locp,n(c1),,locp,n(c2g))w=0.\mathrm{rank}_{{\mathbb{Z}}_{p}}\Big{(}\frac{H^{1}_{/f}(K_{n,p},T)}{\langle\mathrm{loc}_{p,n}(c_{1}),\cdots,\mathrm{loc}_{p,n}(c_{2g})\rangle}\Big{)}_{w}=0.

Let p(θ){\mathbb{Q}}_{p}(\theta) be the field obtained by adjoining the image of θ\theta. When rankp(H/f1(Kn,p,T)locp,n(c1),,locp,n(c2g))w\mathrm{rank}_{{\mathbb{Z}}_{p}}\Big{(}\frac{H^{1}_{/f}(K_{n,p},T)}{\langle\mathrm{loc}_{p,n}(c_{1}),\cdots,\mathrm{loc}_{p,n}(c_{2g})\rangle}\Big{)}_{w} is nonzero, it is always bounded by 2gdimpp(θ).2g\dim_{{\mathbb{Q}}_{p}}{\mathbb{Q}}_{p}(\theta).

Proof.

For 𝔮{𝔭,𝔭c}{\mathfrak{q}}\in\{{\mathfrak{p}},{\mathfrak{p}}^{c}\},

v𝔮H/f1(Kn,v,T)wpp(θ)g.\prod_{v\mid{\mathfrak{q}}}H^{1}_{/f}(K_{n,v},T)_{w}\otimes{\mathbb{Q}}_{p}\cong{\mathbb{Q}}_{p}(\theta)^{\oplus g}.

So

2g(H/f1(Kn,p,T))wp=2g(v𝔭H/f1(Kn,v,T)w×v𝔭cH/f1(Kn,v,T)w)p\bigwedge^{2g}\Big{(}H^{1}_{/f}(K_{n,p},T)\Big{)}_{w}\otimes{\mathbb{Q}}_{p}=\bigwedge^{2g}\Big{(}\prod_{v\mid{\mathfrak{p}}}H^{1}_{/f}(K_{n,v},T)_{w}\times\prod_{v\mid{\mathfrak{p}}^{c}}H^{1}_{/f}(K_{n,v},T)_{w}\Big{)}\otimes{\mathbb{Q}}_{p}

is a one dimensional p(θ){\mathbb{Q}}_{p}(\theta)-vector space. By lemma 4.9,

rankp(2gH/f1(Kn,p,T)locp,n(c1)locp,n(c2g))w>0\mathrm{rank}_{{\mathbb{Z}}_{p}}\Big{(}\frac{\bigwedge^{2g}H^{1}_{/f}(K_{n,p},T)}{\mathrm{loc}_{p,n}(c_{1})\wedge\cdots\wedge\mathrm{loc}_{p,n}(c_{2g})}\Big{)}_{w}>0

if and only if

J¯HI¯0,J¯,r,sColT,J¯k(𝐜)(θ)=0.\sum_{\underline{J}\in\mathcal{I}}H_{\underline{I}_{0},\underline{J},r,s}\mathrm{Col}_{T,\underline{J}}^{k_{\infty}}(\mathbf{c})(\theta)=0.

Therefore under the hypothesis of lemma 4.10,

rankp(2gH/f1(Kn,p,T)locp,n(c1)locp,n(c2g))w=0.\mathrm{rank}_{{\mathbb{Z}}_{p}}\Big{(}\frac{\bigwedge^{2g}H^{1}_{/f}(K_{n,p},T)}{\mathrm{loc}_{p,n}(c_{1})\wedge\cdots\wedge\mathrm{loc}_{p,n}(c_{2g})}\Big{)}_{w}=0.

Note that 2gH/f1(Kn,p,T)wp\bigwedge^{2g}H^{1}_{/f}(K_{n,p},T)_{w}\otimes{\mathbb{Q}}_{p} is a one dimensional p(θ){\mathbb{Q}}_{p}(\theta)-vector space and so the image of locp,n(c1)locp,n(c2g)\mathrm{loc}_{p,n}(c_{1})\wedge\cdots\wedge\mathrm{loc}_{p,n}(c_{2g}) in it is nonzero, since the rank of the quotient is zero. This implies that rankp(H/f1(Kn,p,T)locp,n(c1),,locp,n(c2g))w=0.\mathrm{rank}_{{\mathbb{Z}}_{p}}\Big{(}\frac{H^{1}_{/f}(K_{n,p},T)}{\langle\mathrm{loc}_{p,n}(c_{1}),\cdots,\mathrm{loc}_{p,n}(c_{2g})\rangle}\Big{)}_{w}=0.

We impose the following hypothesis whenever needed.

(H-large): For r,sr,s such that |rs|0|r-s|\gg 0, J¯HI¯0,J¯,r,sColT,J¯k(𝐜)(θ)0.\sum_{\underline{J}\in\mathcal{I}}H_{\underline{I}_{0},\underline{J},r,s}\mathrm{Col}_{T,\underline{J}}^{k_{\infty}}(\mathbf{c})(\theta)\neq 0.

Lemma 4.11.

Suppose the hypothesis (H-large) holds. Let n1n\geq 1 and Ξn\Upxi_{n} be the set of characters θ\theta of Γ\Gamma which factor through Γn\Gamma_{n} but not Γn1\Gamma_{n-1} such that

J¯HI¯0,J¯,r,sColT,J¯k(𝐜)(θ)=0\sum_{\underline{J}\in\mathcal{I}}H_{\underline{I}_{0},\underline{J},r,s}\mathrm{Col}_{T,\underline{J}}^{k_{\infty}}(\mathbf{c})(\theta)=0

(with the conductor of θ\theta being 𝔭r+1(𝔭c)s+1{\mathfrak{p}}^{r+1}({\mathfrak{p}}^{c})^{s+1}). Then the cardinality of Ξn\Upxi_{n} is bounded independent of nn.

Proof.

If θ\theta factors through Γn\Gamma_{n} but not Γn1\Gamma_{n-1}, then either rr or ss equals nn. By the hypothesis (H-large), there exists a fixed integer n0n_{0} (independent of nn) such that if

J¯HI¯0,J¯,r,sColT,J¯k(𝐜)(θ)=0,\sum_{\underline{J}\in\mathcal{I}}H_{\underline{I}_{0},\underline{J},r,s}\mathrm{Col}_{T,\underline{J}}^{k_{\infty}}(\mathbf{c})(\theta)=0,

then |rs|n0|r-s|\leq n_{0}. If either rr or ss equals nn, then the number of such (r,s)(r,s) is bounded and hence the result. ∎

Proposition 4.12.

Suppose the hypothesis (H-large) holds. Then rankp𝒴n=O(pn)\mathrm{rank}_{{\mathbb{Z}}_{p}}\mathcal{Y}^{\prime}_{n}=O(p^{n}).

Proof.

Let CnC_{n} be the cardinality of Ξn\Upxi_{n}. By lemmas 4.10 and 4.8,

rankp𝒴nrankp𝒴n12gCnφ(pn).\mathrm{rank}_{{\mathbb{Z}}_{p}}\mathcal{Y}^{\prime}_{n}-\mathrm{rank}_{{\mathbb{Z}}_{p}}\mathcal{Y}^{\prime}_{n-1}\leq 2gC_{n}\varphi(p^{n}).

Under the hypothesis (H-large), CnC_{n} is less than a certain constant independent of nn. Summing the inequality over nn, we deduce that

rankp𝒴nrankp𝒴1i=2nCφ(pi)=O(pn),\mathrm{rank}_{{\mathbb{Z}}_{p}}\mathcal{Y}^{\prime}_{n}-\mathrm{rank}_{{\mathbb{Z}}_{p}}\mathcal{Y}^{\prime}_{1}\leq\sum_{i=2}^{n}C\varphi(p^{i})=O(p^{n}),

where CC is some constant depending on gg and independent of nn. ∎

Lemma 4.13.

Let J¯\underline{J}\in\mathcal{I} and suppose that SelJ¯(A/K){\mathrm{Sel}}_{\underline{J}}(A^{\vee}/K_{\infty})^{\vee} is Λ\Lambda-torsion. Then ColT,J¯k(𝐜)0\mathrm{Col}_{T,\underline{J}}^{k_{\infty}}(\mathbf{c})\neq 0.

Proof.

The fact that SelJ¯(A/K){\mathrm{Sel}}_{\underline{J}}(A^{\vee}/K_{\infty})^{\vee} is Λ\Lambda-torsion combined with (3.2) and lemma 4.2 imply the desired result. ∎

4.3. Block anti-diagonal matrices

Suppose we assume that SelJ¯(A/K){\mathrm{Sel}}_{\underline{J}}(A^{\vee}/K_{\infty})^{\vee} is Λ\Lambda-torsion. Then it will imply that ColT,J¯k(𝐜)(θ)0\mathrm{Col}_{T,\underline{J}}^{k_{\infty}}(\mathbf{c})(\theta)\neq 0 for |rs|0|r-s|\gg 0. But this is not enough to ensure that J¯HI¯0,J¯,r,sColT,J¯k(𝐜)(θ)0\sum_{\underline{J}\in\mathcal{I}}H_{\underline{I}_{0},\underline{J},r,s}\mathrm{Col}_{T,\underline{J}}^{k_{\infty}}(\mathbf{c})(\theta)\neq 0 since HI¯0,J¯,r,s(θ)H_{\underline{I}_{0},\underline{J},r,s}(\theta) could be zero. In general, this is complicated to verify unless we are in some special cases where we can explicitly write down the matrices Cφ,vC_{\varphi,v}. This special case that we will be looking at can be thought of as an analogue for the case ap=0a_{p}=0 for supersingular elliptic curves as explained below.

In this section, for each vv, let us suppose that there exist a basis of 𝔻cris,v(T)\mathbb{D}_{{\mathrm{cris}},v}(T) such that the matrix CvC_{v} is of the form [00],\left[\begin{array}[]{c|c}0&*\\ \hline\cr*&0\end{array}\right], where * is a g×gg\times g matrix defined over p{\mathbb{Z}}_{p}. This means φ(vi)Fil0𝔻cris,v(T)\varphi(v_{i})\notin\mathrm{Fil}^{0}\mathbb{D}_{{\mathrm{cris}},v}(T) and φ2(vi)Fil0𝔻cris,v(T)\varphi^{2}(v_{i})\in\mathrm{Fil}^{0}\mathbb{D}_{{\mathrm{cris}},v}(T) for all i{1,,g}i\in\{1,...,g\}. Therefore, Cv,nC_{v,n} is of the form

Cv,n=[0Bv,1Φpn(1+X)Bv,20],C_{v,n}=\left[\begin{array}[]{c|c}0&B_{v,1}\\ \hline\cr\Phi_{p^{n}}(1+X)B_{v,2}&0\end{array}\right],

for some invertible g×gg\times g matrices Bv,1B_{v,1} and Bv,2B_{v,2} defined over p{\mathbb{Z}}_{p}. In the context of elliptic curve with ap=0a_{p}=0, we may choose a basis {ω}\{\omega\} for Fil0𝔻cris,v(T)\mathrm{Fil}^{0}\mathbb{D}_{{\mathrm{cris}},v}(T) and extend it to a basis {ω,φ(ω)}\{\omega,\varphi(\omega)\} of 𝔻cris,v(T)\mathbb{D}_{{\mathrm{cris}},v}(T). With this choice, the matrix CvC_{v} is equal to the anti-diagonal matrix [0110]\begin{bmatrix}0&-1\\ 1&0\end{bmatrix} [4, Appendix 4]. In particular, the matrix Cv,nC_{v,n} will have the same structure as the matrix CnC_{n} defined in [8, p. 192]. In this case, it can be shown [8, Proposition 5.12] that the hypothesis (H-large) is satisfied for such elliptic curves.

For all n1n\geq 1, we fix a compatible system {ζpn:ζpn+1p=ζpn}\{\zeta_{p^{n}}:\zeta_{p^{n+1}}^{p}=\zeta_{p^{n}}\} of primitive pnp^{n}-th roots of unity and we write εn=ζpn1\varepsilon_{n}=\zeta_{p^{n}}-1.

Recall that I¯0=(I𝔭,0,I𝔭c,0)\underline{I}_{0}=(I_{{\mathfrak{p}},0},I_{{{\mathfrak{p}}^{c}},0}) and I𝔮,0={1,,g}I_{{\mathfrak{q}},0}=\{1,...,g\} for 𝔮{𝔭,𝔭c}.{\mathfrak{q}}\in\{{\mathfrak{p}},{\mathfrak{p}}^{c}\}. Let I¯1\underline{I}_{1} be the complement of I¯0\underline{I}_{0}. That is I¯1=(I𝔭,1,I𝔭c,1)\underline{I}_{1}=(I_{{\mathfrak{p}},1},I_{{{\mathfrak{p}}^{c}},1}) and I𝔮,1={g+1,,2g}I_{{\mathfrak{q}},1}=\{g+1,...,2g\} for 𝔮{𝔭,𝔭c}.{\mathfrak{q}}\in\{{\mathfrak{p}},{\mathfrak{p}}^{c}\}. Let I¯mix0,1=(I𝔭,0,I𝔭c,1)\underline{I}_{\mathrm{mix}_{0,1}}=(I_{{\mathfrak{p}},0},I_{{\mathfrak{p}}^{c},1}) and I¯mix1,0=(I𝔭,1,I𝔭c,0)\underline{I}_{\mathrm{mix}_{1,0}}=(I_{{\mathfrak{p}},1},I_{{\mathfrak{p}}^{c},0}).

Proposition 4.14.

Suppose C𝔭C_{\mathfrak{p}} and C𝔭cC_{{\mathfrak{p}}^{c}} are block anti-diagonal matrices and the Selmer groups SelI¯0(A/K){\mathrm{Sel}}_{\underline{I}_{0}}(A^{{\vee}}/K_{\infty}), SelI¯1(A/K){\mathrm{Sel}}_{\underline{I}_{1}}(A^{{\vee}}/K_{\infty}), SelI¯mix0,1(A/K){\mathrm{Sel}}_{\underline{I}_{\mathrm{mix}_{0,1}}}(A^{{\vee}}/K_{\infty}) and SelI¯mix1,0(A/K){\mathrm{Sel}}_{\underline{I}_{\mathrm{mix}_{1,0}}}(A^{{\vee}}/K_{\infty}) are all Λ\Lambda-cotorsion. Then the hypothesis (H-large) holds.

Proof.

For any k1k\geq 1, let δk\delta_{k} be the constant

δk={ε1ε2ε3ε4εk2εk1if k is oddε1ε2ε3ε4εk1εkif k is even.\delta_{k}=\begin{cases}\frac{\varepsilon_{1}}{\varepsilon_{2}}\cdot\frac{\varepsilon_{3}}{\varepsilon_{4}}\cdots\frac{\varepsilon_{k-2}}{\varepsilon_{k-1}}&\text{if $k$ is odd}\\ \frac{\varepsilon_{1}}{\varepsilon_{2}}\cdot\frac{\varepsilon_{3}}{\varepsilon_{4}}\cdots\frac{\varepsilon_{k-1}}{\varepsilon_{k}}&\text{if $k$ is even.}\end{cases}

Let θ\theta be a character of Γ\Gamma of conductor 𝔭r+1(𝔭c)s+1{\mathfrak{p}}^{r+1}({\mathfrak{p}}^{c})^{s+1}. By [7, Lemma 3.5],

H𝔮,k(ζpk1)={[0δk(B𝔮,1B𝔮,2)k12B𝔮,100]if k is odd,[δk(B𝔮,1B𝔮,2)k2000]if k is even,H_{{\mathfrak{q}},k}(\zeta_{p^{k}}-1)=\begin{cases}\left[\begin{array}[]{c|c}0&\delta_{k}(B_{{\mathfrak{q}},1}B_{{\mathfrak{q}},2})^{\frac{k-1}{2}}B_{{\mathfrak{q}},1}\\ \hline\cr 0&0\end{array}\right]&\text{if $k$ is odd,}\\ &\\ \left[\begin{array}[]{c|c}\delta_{k}(B_{{\mathfrak{q}},1}B_{{\mathfrak{q}},2})^{\frac{k}{2}}&0\\ \hline\cr 0&0\end{array}\right]&\text{if $k$ is even,}\end{cases}

where k=rk=r if 𝔮=𝔭{\mathfrak{q}}={\mathfrak{p}} and k=sk=s if 𝔮=𝔭c{\mathfrak{q}}={\mathfrak{p}}^{c}. Hence,

(4.4) HI¯0,J¯,r,s(θ)=0 unless {J¯=I¯0if r is even and s is even,J¯=I¯1if r is odd and s is odd,J¯=(J𝔭,J𝔭c),J𝔭=I𝔭,1,J𝔭c=I𝔭c,0if r is odd and s is even,J¯=(J𝔭,J𝔭c),J𝔭=I𝔭,0,J𝔭c=I𝔭c,1if r is even and s is odd.H_{\underline{I}_{0},\underline{J},r,s}(\theta)=0\text{ unless }\begin{cases}\underline{J}=\underline{I}_{0}&\text{if $r$ is even and $s$ is even,}\\ \underline{J}=\underline{I}_{1}&\text{if $r$ is odd and $s$ is odd,}\\ \underline{J}=(J_{\mathfrak{p}},J_{{\mathfrak{p}}^{c}}),J_{\mathfrak{p}}=I_{{\mathfrak{p}},1},J_{{\mathfrak{p}}^{c}}=I_{{\mathfrak{p}}^{c},0}&\text{if $r$ is odd and $s$ is even,}\\ \underline{J}=(J_{\mathfrak{p}},J_{{\mathfrak{p}}^{c}}),J_{\mathfrak{p}}=I_{{\mathfrak{p}},0},J_{{\mathfrak{p}}^{c}}=I_{{\mathfrak{p}}^{c},1}&\text{if $r$ is even and $s$ is odd.}\end{cases}

Under the hypotheses on the Selmer groups being cotorsion Λ\Lambda-modules, we obtain ColT,J¯k(𝐜)0\mathrm{Col}_{T,\underline{J}}^{k_{\infty}}(\mathbf{c})\neq 0 for J¯=I¯0,J¯=I¯1,J¯=I¯mix0,1\underline{J}=\underline{I}_{0},\underline{J}=\underline{I}_{1},\underline{J}=\underline{I}_{\mathrm{mix}_{0,1}} and J¯=I¯mix1,0\underline{J}=\underline{I}_{\mathrm{mix}_{1,0}} depending on the parity of rr and ss. Therefore, when |rs|0|r-s|\gg 0, by [8, Corollary 5.10], there exists integers a,b1,b2,c1,c2a,b_{1},b_{2},c_{1},c_{2} such that

(4.5) valp(ColT,J¯k(𝐜)(θ))=a+biφ(pr)+ciφ(ps)\operatorname{val}_{p}\Big{(}\mathrm{Col}_{T,\underline{J}}^{k_{\infty}}(\mathbf{c})(\theta)\Big{)}=a+\frac{b_{i}}{\varphi(p^{r})}+\frac{c_{i}}{\varphi(p^{s})}

where i=1i=1 if r>sr>s and i=2i=2 if s>rs>r. (One can see the proof of [8, Proposition 5.9] on how to obtain these integers a,b1,b2,c1a,b_{1},b_{2},c_{1} and c2c_{2}.) In particular, for |rs|0|r-s|\gg 0, ColT,J¯k(𝐜)(θ)0\mathrm{Col}_{T,\underline{J}}^{k_{\infty}}(\mathbf{c})(\theta)\neq 0 and hence

J¯HI¯0,J¯,r,sColT,J¯k(𝐜)(θ)=HI¯0,J¯,r,sColT,J¯k(𝐜)(θ)0,\sum_{\underline{J}\in\mathcal{I}}H_{\underline{I}_{0},\underline{J},r,s}\mathrm{Col}_{T,\underline{J}}^{k_{\infty}}(\mathbf{c})(\theta)=H_{\underline{I}_{0},\underline{J},r,s}\mathrm{Col}_{T,\underline{J}}^{k_{\infty}}(\mathbf{c})(\theta)\neq 0,

where J¯{I¯0,I¯1,I¯mix0,1,I¯mix1,0}\underline{J}\in\{\underline{I}_{0},\underline{I}_{1},\underline{I}_{\mathrm{mix}_{0,1}},\underline{I}_{\mathrm{mix}_{1,0}}\} depending on the parity of rr and ss given in (4.4). This shows that hypothesis (H-large) is satisfied.

Remark 4.15.

Proposition 4.14 is a generalization of [7, Lemma 3.6] and it was already assumed (and conjectured) there that the Selmer groups are cotorsion [7, Conjecture 2.2].

As mentioned previously, the Selmer groups SelJ¯(A[p]/K){\mathrm{Sel}}_{\underline{J}}(A^{\vee}[p^{\infty}]/K_{\infty}) not only depends on the choice of subset J¯\underline{J} but also on the choice of p\mathbb{Z}_{p}-bases for 𝔻cris,𝔭(T)\mathbb{D}_{{\mathrm{cris}},{\mathfrak{p}}}(T) and 𝔻cris,𝔭c(T)\mathbb{D}_{{\mathrm{cris}},{\mathfrak{p}}^{c}}(T). However, as we will see, the Selmer groups SelI¯0(A/K){\mathrm{Sel}}_{\underline{I}_{0}}(A^{{\vee}}/K_{\infty}), SelI¯1(A/K){\mathrm{Sel}}_{\underline{I}_{1}}(A^{{\vee}}/K_{\infty}), SelI¯mix0,1(A/K){\mathrm{Sel}}_{\underline{I}_{\mathrm{mix}_{0,1}}}(A^{{\vee}}/K_{\infty}) and SelI¯mix1,0(A/K){\mathrm{Sel}}_{\underline{I}_{\mathrm{mix}_{1,0}}}(A^{{\vee}}/K_{\infty}) are canonically defined when C𝔭C_{\mathfrak{p}} and C𝔭cC_{{\mathfrak{p}}^{c}} are block anti-diagonal matrices.

Definition 4.16.

Fix a Hodge-compatible p\mathbb{Z}_{p}-basis 𝔮={v1,,v2g}\mathcal{B}_{\mathfrak{q}}=\{v_{1},\ldots,v_{2g}\} of 𝔻cris,𝔮(T)\mathbb{D}_{{\mathrm{cris}},{\mathfrak{q}}}(T). Hypothesis (H.HT) implies that Fil0𝔻cris,𝔮(T)\mathrm{Fil}^{0}\mathbb{D}_{{\mathrm{cris}},{\mathfrak{q}}}(T) is a direct summand of 𝔻cris,𝔮(T)\mathbb{D}_{{\mathrm{cris}},{\mathfrak{q}}}(T) [4, Remark 2.3]. Let N𝔮𝔻cris,𝔮(T)N_{\mathfrak{q}}\subseteq\mathbb{D}_{{\mathrm{cris}},{\mathfrak{q}}}(T) denote the free p\mathbb{Z}_{p}-module complementary to Fil0𝔻cris,𝔮(T)\mathrm{Fil}^{0}\mathbb{D}_{{\mathrm{cris}},{\mathfrak{q}}}(T) and generated by the set {vg+1,,v2g}\{v_{g+1},\ldots,v_{2g}\}. We say that a p\mathbb{Z}_{p}-basis 𝔮={w1,,w2g}\mathcal{B}_{\mathfrak{q}}^{\prime}=\{w_{1},\ldots,w_{2g}\} is Hodge-compatible with the basis 𝔮\mathcal{B}_{\mathfrak{q}} if {w1,,wg}\{w_{1},\ldots,w_{g}\} (resp. {wg+1,,w2g}\{w_{g+1},\ldots,w_{2g}\}) generated the submodule Fil0𝔻cris,𝔮(T)\mathrm{Fil}^{0}\mathbb{D}_{{\mathrm{cris}},{\mathfrak{q}}}(T) (resp. N𝔮N_{\mathfrak{q}}).

Lemma 4.17.

Let 𝔮\mathcal{B}_{\mathfrak{q}} and 𝔮\mathcal{B}_{\mathfrak{q}}^{\prime} be two Hodge-compatible bases in the sense of definition 2.1. Suppose that C𝔮C_{\mathfrak{q}} is a block anti-diagonal matrix with respect to the basis 𝔮\mathcal{B}_{\mathfrak{q}}. Then 𝔮\mathcal{B}_{\mathfrak{q}}^{\prime} is Hodge-compatible with 𝔮\mathcal{B}_{\mathfrak{q}} in the sense of definition 4.16.

Proof.

Since Fil0𝔻cris,𝔮(T)\mathrm{Fil}^{0}\mathbb{D}_{{\mathrm{cris}},{\mathfrak{q}}}(T) is a direct summand of 𝔻cris,𝔮(T)\mathbb{D}_{{\mathrm{cris}},{\mathfrak{q}}}(T), we get decompositions 𝔻cris,𝔮(T)=Fil0𝔻cris,𝔮(T)N𝔮\mathbb{D}_{{\mathrm{cris}},{\mathfrak{q}}}(T)=\mathrm{Fil}^{0}\mathbb{D}_{{\mathrm{cris}},{\mathfrak{q}}}(T)\bigoplus N_{\mathfrak{q}} and 𝔻cris,𝔮(T)=Fil0𝔻cris,𝔮(T)N𝔮\mathbb{D}_{{\mathrm{cris}},{\mathfrak{q}}}(T)=\mathrm{Fil}^{0}\mathbb{D}_{{\mathrm{cris}},{\mathfrak{q}}}(T)\bigoplus N_{\mathfrak{q}}^{\prime} where N𝔮N_{\mathfrak{q}} is generated by {vg+1,,v2g}\{v_{g+1},\ldots,v_{2g}\} and N𝔮N_{\mathfrak{q}}^{\prime} is generated by {wg+1,,w2g}\{w_{g+1},\ldots,w_{2g}\}. Since C𝔮C_{{\mathfrak{q}}} is block anti-diagonal with respect to 𝔮\mathcal{B}_{\mathfrak{q}}, we have that Span{φ(v1),,φ(vg)}N𝔮\mathrm{Span}\{\varphi(v_{1}),\ldots,\varphi(v_{g})\}\subseteq N_{\mathfrak{q}} and since φ\varphi is injective we in fact have equality. By a similar argument, Span{φ(v1),,φ(vg)}=N𝔮\mathrm{Span}\{\varphi(v_{1}),\ldots,\varphi(v_{g})\}=N_{\mathfrak{q}}^{\prime}. Thus {wg+1,,w2g}\{w_{g+1},\ldots,w_{2g}\} generates N𝔮N_{\mathfrak{q}}. ∎

Let I¯BDP2g,=({1,,2g},)\underline{I}_{\text{BDP}_{2g,\emptyset}}=(\{1,...,2g\},\emptyset) and I¯BDP,2g=(,{1,,2g})\underline{I}_{\text{BDP}_{\emptyset,2g}}=(\emptyset,\{1,...,2g\}). These indices should correspond to Selmer groups related with BDP type pp-adic LL-functions.

Proposition 4.18.

Suppose that there exist Hodge-compatible bases 𝔭\mathcal{B}_{\mathfrak{p}} and 𝔭c\mathcal{B}_{{\mathfrak{p}}^{c}} of 𝔻cris,𝔭(T)\mathbb{D}_{{\mathrm{cris}},{\mathfrak{p}}}(T) and 𝔻cris,𝔭c(T)\mathbb{D}_{{\mathrm{cris}},{\mathfrak{p}}^{c}}(T) such that C𝔭C_{\mathfrak{p}} and C𝔭cC_{{\mathfrak{p}}^{c}} are block anti-diagonal with respect to these bases. Suppose that we have J¯{I¯0,I¯1,I¯mix0,1,I¯mix1,0,I¯BDP2g,,I¯BDP,2g}\underline{J}\in\{\underline{I}_{0},\underline{I}_{1},\underline{I}_{\mathrm{mix}_{0,1}},\underline{I}_{\mathrm{mix}_{1,0}},\underline{I}_{\text{BDP}_{2g,\emptyset}},\underline{I}_{\text{BDP}_{\emptyset,2g}}\}. Then SelJ¯(A[p]/K){\mathrm{Sel}}_{\underline{J}}(A^{\vee}[p^{\infty}]/K_{\infty}) does not depend on the choice of Hodge-compatible bases for 𝔻cris,𝔭(T)\mathbb{D}_{{\mathrm{cris}},{\mathfrak{p}}}(T) and 𝔻cris,𝔭c(T)\mathbb{D}_{{\mathrm{cris}},{\mathfrak{p}}^{c}}(T).

In other words, once there exist Hodge-compatible bases yielding block anti-diagonal matrices, SelJ¯(A[p]/K){\mathrm{Sel}}_{\underline{J}}(A^{\vee}[p^{\infty}]/K_{\infty}) becomes independent of the choice of Hodge-compatible bases for J¯\underline{J} as in the statement of the proposition.

Proof.

Write 𝔭={v1,,v2g}\mathcal{B}_{\mathfrak{p}}=\{v_{1},\ldots,v_{2g}\} and let 𝔭={w1,,w2g}\mathcal{B}_{\mathfrak{p}}^{\prime}=\{w_{1},\ldots,w_{2g}\} be another Hodge-compatible basis of 𝔻cris,𝔭(T)\mathbb{D}_{{\mathrm{cris}},{\mathfrak{p}}}(T). Recall that Hodge-compatible means that {v1,,vg}\{v_{1},\ldots,v_{g}\} and {w1,,wg}\{w_{1},\ldots,w_{g}\} are bases for the p\mathbb{Z}_{p}-submodule Fil0𝔻cris,𝔭(T)\mathrm{Fil}^{0}\mathbb{D}_{{\mathrm{cris}},{\mathfrak{p}}}(T). Let B𝔭B_{\mathfrak{p}} be the change of basis matrix from 𝔭\mathcal{B}_{\mathfrak{p}}^{\prime} to 𝔭\mathcal{B}_{\mathfrak{p}}. Then the hypothesis on both basis forces B𝔭B_{\mathfrak{p}} to be block diagonal. Write

B𝔭=[B1,100B2,2]B_{\mathfrak{p}}=\left[\begin{array}[]{c|c}B_{1,1}&0\\ \hline\cr 0&B_{2,2}\end{array}\right]

where B1,1,B2,2GLg(p)B_{1,1},B_{2,2}\in{\mathrm{GL}}_{g}(\mathbb{Z}_{p}). Let ColT,𝔭,vik\mathrm{Col}_{T,{\mathfrak{p}},v_{i}}^{k_{\infty}} be the ii-th Coleman map as defined in section 3.3 with respect to the basis 𝔭\mathcal{B}_{\mathfrak{p}}. Let ColT,𝔭,𝔭k\mathrm{Col}_{T,{\mathfrak{p}},\mathcal{B}_{\mathfrak{p}}}^{k_{\infty}} denotes the vector of Coleman maps (ColT,𝔭,vik)i=12g(\mathrm{Col}_{T,{\mathfrak{p}},v_{i}}^{k_{\infty}})_{i=1}^{2g} which we see as a column vector. Define similarly ColT,𝔭,𝔭k\mathrm{Col}_{T,{\mathfrak{p}},\mathcal{B}_{\mathfrak{p}}^{\prime}}^{k_{\infty}} the column vector of Coleman maps defined with respect to 𝔭\mathcal{B}_{\mathfrak{p}}^{\prime}. Since C𝔭C_{\mathfrak{p}} is block anti-diagonal, lemma 4.17 gives us that 𝔭\mathcal{B}_{\mathfrak{p}}^{\prime} is Hodge-compatible with 𝔭\mathcal{B}_{\mathfrak{p}}. Hence we can use [4, Lemma 2.16] to deduce that both vectors of Coleman maps are related by the linear transformation ColT,𝔭,𝔭k=BColT,𝔭,𝔭k.\mathrm{Col}_{T,{\mathfrak{p}},\mathcal{B}_{\mathfrak{p}}^{\prime}}^{k_{\infty}}=B\cdot\mathrm{Col}_{T,{\mathfrak{p}},\mathcal{B}_{\mathfrak{p}}}^{k_{\infty}}. Thus,

[ColT,𝔭,w1kColT,𝔭,wgk]=B1,1[ColT,𝔭,v1kColT,𝔭,vgk],[ColT,𝔭,wg+1kColT,𝔭,w2gk]=B2,2[ColT,𝔭,vg+1kColT,𝔭,v2gk].\begin{bmatrix}\mathrm{Col}_{T,{\mathfrak{p}},w_{1}}^{k_{\infty}}\\ \vdots\\ \mathrm{Col}_{T,{\mathfrak{p}},w_{g}}^{k_{\infty}}\end{bmatrix}=B_{1,1}\begin{bmatrix}\mathrm{Col}_{T,{\mathfrak{p}},v_{1}}^{k_{\infty}}\\ \vdots\\ \mathrm{Col}_{T,{\mathfrak{p}},v_{g}}^{k_{\infty}}\end{bmatrix},\quad\begin{bmatrix}\mathrm{Col}_{T,{\mathfrak{p}},w_{g+1}}^{k_{\infty}}\\ \vdots\\ \mathrm{Col}_{T,{\mathfrak{p}},w_{2g}}^{k_{\infty}}\end{bmatrix}=B_{2,2}\begin{bmatrix}\mathrm{Col}_{T,{\mathfrak{p}},v_{g+1}}^{k_{\infty}}\\ \vdots\\ \mathrm{Col}_{T,{\mathfrak{p}},v_{2g}}^{k_{\infty}}\end{bmatrix}.

This implies that (ColT,𝔭,wik(z))i=1g=0(\mathrm{Col}_{T,{\mathfrak{p}},w_{i}}^{k_{\infty}}(z))_{i=1}^{g}=0 if and only if (ColT,𝔭,vik(z))i=1g=0(\mathrm{Col}_{T,{\mathfrak{p}},v_{i}}^{k_{\infty}}(z))_{i=1}^{g}=0 since B1,1B_{1,1} is invertible. The same goes for (ColT,𝔭,wik(z))i=g+12g(\mathrm{Col}_{T,{\mathfrak{p}},w_{i}}^{k_{\infty}}(z))_{i=g+1}^{2g} and (ColT,𝔭,vik(z))i=g+12g(\mathrm{Col}_{T,{\mathfrak{p}},v_{i}}^{k_{\infty}}(z))_{i=g+1}^{2g}. We conclude that kerColT,J𝔭\ker\mathrm{Col}_{T,J_{\mathfrak{p}}} is independent of the choice of basis if J𝔭J_{{\mathfrak{p}}} is I𝔭,0I_{{\mathfrak{p}},0}, I𝔭,1I_{{\mathfrak{p}},1} or {1,,2g}\{1,...,2g\}. The same argument holds for 𝔭{\mathfrak{p}} replaced by 𝔭c{\mathfrak{p}}^{c} and so the result follows. ∎

Remark 4.19.

It is easy to see that C𝔭C_{\mathfrak{p}} and C𝔭cC_{{\mathfrak{p}}^{c}} remain block anti-diagonal matrices upon a change of basis. This follows because B𝔮C𝔮B𝔮1B_{\mathfrak{q}}C_{{\mathfrak{q}}}B_{{\mathfrak{q}}}^{-1} is block anti-diagonal if B𝔮B_{\mathfrak{q}} is block diagonal and C𝔮C_{\mathfrak{q}} is block anti-diagonal for 𝔮{𝔭,𝔭c}{\mathfrak{q}}\in\{{\mathfrak{p}},{\mathfrak{p}}^{c}\}.

5. The bound on the Mordell–Weil rank

Let Σ\Sigma be a finite set of primes of KK dividing pp, the archimedean primes and the primes of bad reduction of AA^{\vee}. Let KΣK_{\Sigma} be the maximal extension of KK unramified outside Σ\Sigma. Let HΣ1(Kn,T)H^{1}_{\Sigma}(K_{n},T) be the Galois cohomology group H1(Gal(KΣ/Kn),T)H^{1}({\mathrm{Gal}}(K_{\Sigma}/K_{n}),T).

For n0n\geq 0, we define

𝒴n\displaystyle\mathcal{Y}_{n} :=Coker(HΣ1(Kn,T)vpH/f1(Kn,v,T)),\displaystyle:={\mathrm{Coker}}\Big{(}H^{1}_{\Sigma}(K_{n},T)\rightarrow\prod_{v\mid p}H^{1}_{/f}(K_{n,v},T)\Big{)},
Selp0(A/Kn)\displaystyle{\mathrm{Sel}}_{p}^{0}(A^{\vee}/K_{n}) :=Ker(HΣ1(Kn,A[p])vH1(Kn,v,A[p])).\displaystyle:={\mathrm{Ker}}\Big{(}H^{1}_{\Sigma}(K_{n},A^{\vee}[p^{\infty}])\rightarrow\prod_{v}H^{1}(K_{n,v},A^{\vee}[p^{\infty}])\Big{)}.

Let 𝒳n0\mathcal{X}_{n}^{0} and 𝒳n\mathcal{X}_{n} be the dual of the fine Selmer group Selp0(A/Kn){\mathrm{Sel}}_{p}^{0}(A^{\vee}/K_{n}) and the classical pp-Selmer group Selp(A/Kn){\mathrm{Sel}}_{p}(A^{\vee}/K_{n}) respectively. The Poitou–Tate exact sequence [24, Proposition A.3.2] gives

HΣ1(Kn,T)vpH1(Kn,v,T)Hf1(Kn,v,T)𝒳n𝒳n00.H^{1}_{\Sigma}(K_{n},T)\rightarrow\prod_{v\mid p}\frac{H^{1}(K_{n,v},T)}{H^{1}_{f}(K_{n,v},T)}\rightarrow\mathcal{X}_{n}\rightarrow\mathcal{X}_{n}^{0}\rightarrow 0.

For vpv\mid p, as Hf1(Kn,v,T)A(Kn,v)pH^{1}_{f}(K_{n,v},T)\cong A(K_{n,v})\otimes{\mathbb{Z}}_{p} (ref. [4, Remark 3.27]), we obtain the exact sequence,

(5.1) 0𝒴n𝒳n𝒳n00.0\rightarrow\mathcal{Y}_{n}\rightarrow\mathcal{X}_{n}\rightarrow\mathcal{X}_{n}^{0}\rightarrow 0.
Theorem 5.1.

Let MM be a finitely generated Λ\Lambda-module of rank rr. Then,

rankpMΓn=rp2n+O(pn).\mathrm{rank}_{\mathbb{Z}_{p}}M_{\Gamma_{n}}=rp^{2n}+O(p^{n}).
Proof.

This is [31, Theorem 1.10]. ∎

Lemma 5.2.

Let ww be a place above 𝔮{\mathfrak{q}} in KK_{\infty}. We have that A(K,w)[p]=0A^{\vee}(K_{\infty,w})[p^{\infty}]=0.

Proof.

We follow the arguments in the proof of [7, Lemma 1.1]. Let κ𝔮\kappa_{\mathfrak{q}} denote the residue field of K𝔮K_{\mathfrak{q}}. By [1, Lemma 5.11], the reduction map A(K𝔮)A(κ𝔮)A^{\vee}(K_{\mathfrak{q}})\to A^{\vee}(\kappa_{\mathfrak{q}}) induces an isomorphism on pp-torsion points. Thus, the supersingularity of AA^{\vee} at 𝔮{\mathfrak{q}} implies that A(K𝔮)[p]=0A^{\vee}(K_{\mathfrak{q}})[p^{\infty}]=0. Furthermore, for a place ww above 𝔮{\mathfrak{q}} in KK_{\infty}, the decomposition group Gal(K,w/K𝔮){\mathrm{Gal}}(K_{\infty,w}/K_{\mathfrak{q}}) is a pro-pp group. We deduce that A(K,w)A^{\vee}(K_{\infty,w}) also has no pp-torsion by an application of the orbit-stabilizer theorem as shown by the following argument.

Let vv be a prime above 𝔮{\mathfrak{q}} in KnK_{n}. Suppose for the sake of contradiction that A(Kn,v)[p]A^{\vee}(K_{n,v})[p] is not trivial. Write

A(Kn,v)[p]=iOrb(xi)A^{\vee}(K_{n,v})[p]=\bigcup_{i}\mathrm{Orb}(x_{i})

as a disjoint union of orbits under the action of Gal(Kn,v/K𝔮){\mathrm{Gal}}(K_{n,v}/K_{\mathfrak{q}}). If xi0x_{i}\neq 0, Gal(Kn,v/K𝔮){\mathrm{Gal}}(K_{n,v}/K_{\mathfrak{q}}) cannot fix xix_{i} or else xix_{i} would be a pp-torsion point in A(K𝔮)A^{\vee}(K_{\mathfrak{q}}). So, if xi0x_{i}\neq 0, then the orbit-stabilizer theorem

|Orb(xi)|=|Gal(Kn,v/K𝔮)||Stab(xi)||\mathrm{Orb}(x_{i})|=\frac{|{\mathrm{Gal}}(K_{n,v}/K_{\mathfrak{q}})|}{|\mathrm{Stab}(x_{i})|}

implies that pp divides |Orb(xi)||\mathrm{Orb}(x_{i})| since Gal(Kn,v/K𝔮){\mathrm{Gal}}(K_{n,v}/K_{\mathfrak{q}}) is a pp-group. Thus, all the |Orb(xi)||\mathrm{Orb}(x_{i})| are divisible by pp except Orb(0)\mathrm{Orb}(0) which is of cardinality 11. Hence, |A(Kn,v)[p]|1modp|A^{\vee}(K_{n,v})[p]|\equiv 1\bmod p. This is a contradiction, since A(Kn,v)[p]A^{\vee}(K_{n,v})[p] must contain a subgroup of order pp. We conclude that A(Kn,v)[p]=0A^{\vee}(K_{n,v})[p]=0 and thus A(K,w)[p]=nA(Kn,v)[p]=0A^{\vee}(K_{\infty,w})[p]=\bigcup_{n}A^{\vee}(K_{n,v})[p]=0. ∎

Remark 5.3.

If ww is a place above 𝔮{\mathfrak{q}} in K(μp)K(\mu_{p^{\infty}}), then the fact that the torsion part of A(K(μp)w)A^{\vee}(K(\mu_{p^{\infty}})_{w}) is finite was first showed by Imai in [32].

Proposition 5.4.

Suppose SelJ¯(A/K){\mathrm{Sel}}_{\underline{J}}(A^{\vee}/K_{\infty})^{\vee} is Λ\Lambda-torsion for some J¯\underline{J}\in\mathcal{I}. Then, the p{\mathbb{Z}}_{p}-rank of 𝒳n0\mathcal{X}_{n}^{0} is O(pn).O(p^{n}).

Proof.

We follow the outline of the proof of [8, Proposition 5.5]. We begin by considering the commutative diagram with exact rows

0{0}Selp0(A/Kn){{\mathrm{Sel}}_{p}^{0}(A^{\vee}/K_{n})}H1(Kn,A[p]){H^{1}(K_{n},A^{\vee}[p^{\infty}])}vnH1(Kn,vn,A[p]){\prod_{v_{n}}H^{1}(K_{n,v_{n}},A^{\vee}[p^{\infty}])}0{0}Selp0(A/K)Γn{{\mathrm{Sel}}_{p}^{0}(A^{\vee}/K_{\infty})^{\Gamma_{n}}}H1(K,A[p])Γn{H^{1}(K_{\infty},A^{\vee}[p^{\infty}])^{\Gamma_{n}}}wH1(K,w,A[p])Γn,{\prod_{w}H^{1}(K_{\infty,w},A^{\vee}[p^{\infty}])^{\Gamma_{n}},}αn\scriptstyle{\alpha_{n}}βn\scriptstyle{\beta_{n}}γn\scriptstyle{\prod_{\gamma_{n}}}

where the vertical maps are restriction maps, the product in the first row runs over all the places vnv_{n} of KnK_{n} and the product in the second row over all the places ww of KK_{\infty}. By lemma 5.2, A(K,w)[p]=0A^{\vee}(K_{\infty,w})[p^{\infty}]=0 and hence γn\prod\gamma_{n} is an isomorphism for all vn|pv_{n}|p by the inflation-restriction exact sequence. As the pp-torsion global points inject into the pp-torsion local points, by the same argument, βn\beta_{n} is also an isomorphism. If vnpv_{n}\nmid p, [33, page 270] shows that kerγn\ker\gamma_{n} is finite and of bounded order as KnK_{n} varies. The snake lemma then shows that kerαn=0\ker\alpha_{n}=0 and cokerαn\operatorname{coker}\alpha_{n} is finite with order bounded independently of nn. By the proof of lemma 4.2, 𝒳0\mathcal{X}^{0} is Λ\Lambda-torsion. It follows from theorem 5.1 and the control on αn\alpha_{n} that

rankp𝒳n0=rankp(𝒳0)Γn=O(pn).\mathrm{rank}_{\mathbb{Z}_{p}}\mathcal{X}_{n}^{0}=\mathrm{rank}_{\mathbb{Z}_{p}}(\mathcal{X}^{0})_{\Gamma_{n}}=O(p^{n}).

Theorem 5.5.

Suppose that SelJ¯(A/K){\mathrm{Sel}}_{\underline{J}}(A^{\vee}/K_{\infty}) is Λ\Lambda-cotorsion for some J¯\underline{J}\in\mathcal{I} and hypothesis (H-large) hold. Then the bound for the Mordell–Weil rank is given by rankA(Kn)=O(pn).\mathrm{rank}A^{\vee}(K_{n})=O(p^{n}).

Proof.

Under hypothesis (H-large), proposition 4.12 implies rankp𝒴n=O(pn)\mathrm{rank}_{{\mathbb{Z}}_{p}}\mathcal{Y}^{\prime}_{n}=O(p^{n}). It is easy to see that there is a natural surjection 𝒴n𝒴n\mathcal{Y}^{\prime}_{n}\rightarrow\mathcal{Y}_{n}. In particular rankZp𝒴nrankp𝒴n\mathrm{rank}_{Z_{p}}\mathcal{Y}_{n}\leq\mathrm{rank}_{{\mathbb{Z}}_{p}}\mathcal{Y}^{\prime}_{n}. Then the theorem follows from the exact sequence (5.1) and proposition 5.4. ∎

Theorem 5.6.

Suppose C𝔭C_{\mathfrak{p}} and C𝔭cC_{{\mathfrak{p}}^{c}} are block anti-diagonal matrices and the multi-signed Selmer groups SelI¯0(A/K){\mathrm{Sel}}_{\underline{I}_{0}}(A^{{\vee}}/K_{\infty}), SelI¯1(A/K){\mathrm{Sel}}_{\underline{I}_{1}}(A^{{\vee}}/K_{\infty}), SelI¯mix0,1(A/K){\mathrm{Sel}}_{\underline{I}_{\mathrm{mix}_{0,1}}}(A^{{\vee}}/K_{\infty}) and SelI¯mix1,0(A/K){\mathrm{Sel}}_{\underline{I}_{\mathrm{mix}_{1,0}}}(A^{{\vee}}/K_{\infty}) are all cotorsion over Λ\Lambda. Then rankp(𝒳n)=O(pn)\mathrm{rank}_{{\mathbb{Z}}_{p}}(\mathcal{X}_{n})=O(p^{n}). Hence the bound for the Mordell–Weil rank is given by rankA(Kn)=O(pn).\mathrm{rank}A^{\vee}(K_{n})=O(p^{n}).

Proof.

The hypotheses imply that the hypothesis (H-large) holds by proposition 4.14. Rest of the proof follows from that of theorem 5.5. ∎

Examples. Suppose that AA is an abelian variety defined over {\mathbb{Q}} with good supersingular reduction at both prime over pp in KK such that the algebra of {\mathbb{Q}}-endomorphisms of AA contains a number field EE with [E:]=dimA[E:{\mathbb{Q}}]=\dim A. Such abelian varieties are said to be of GL2{\mathrm{GL}}_{2}-type. Suppose further that the ring of integers 𝒪E\mathcal{O}_{E} of EE is the ring of {\mathbb{Q}}-endomorphisms of AA and that pp is unramified in EE. It follows that the pp-adic Tate module of AA splits into

T|pT(A)T\cong\bigoplus_{\ell|p}T_{\ell}(A)

where the direct sum runs over all the prime \ell of EE above pp and T(A)T_{\ell}(A) is a free 𝒪\mathcal{O}_{\ell}-module of rank 22 where 𝒪\mathcal{O}_{\ell} is the completion of 𝒪E\mathcal{O}_{E} at \ell. Then, it was proved in [7, Section 3.3] that there exists a basis of 𝔻cris,𝔮(T(A))\mathbb{D}_{{\mathrm{cris}},{\mathfrak{q}}}(T_{\ell}(A)) where the action of φ\varphi is given a matrix of the form [0b,𝔮p1a,𝔮p]\begin{bmatrix}0&\frac{b_{\ell,{\mathfrak{q}}}}{p}\\ 1&\frac{a_{\ell,{\mathfrak{q}}}}{p}\end{bmatrix} for some a,𝔮p𝒪a_{\ell,{\mathfrak{q}}}\in p\mathcal{O}_{\ell} and b,𝔮𝒪×b_{\ell,{\mathfrak{q}}}\in\mathcal{O}_{\ell}^{\times}. If we assume that a,𝔮=0a_{\ell,{\mathfrak{q}}}=0 for all \ell and 𝔮{𝔭,𝔭c}{\mathfrak{q}}\in\{{\mathfrak{p}},{\mathfrak{p}}^{c}\}, both matrices C𝔮C_{\mathfrak{q}} will be block anti-diagonal and theorem 5.6 holds.

Remark 5.7.

Let AA be an abelian variety over a number field FF with good ordinary reduction at all the primes above pp in FF. Let FF_{\infty} be a uniform admissible pp-adic extension of FF of dimension d2d\geq 2. In this context, assume that the Selmer group over FF_{\infty} is cotorison over the corresponding Iwasawa algebra. Under the stronger hypothesis of the 𝔐H(G)\mathfrak{M}_{H}(G)-conjecture, Hung and Lim give an explicit bound on the growth of the Mordell–Weil rank of AA along FF_{\infty} (see [34, Theorem 3.1]). In the case of a supersingular elliptic curve, they also give an explicit bound for the Mordell–Weil rank of EE along the p2\mathbb{Z}_{p}^{2}-extension of a quadratic imaginary field where pp splits (see [34, Theorem 6.3, Conjecture 2]). Under the same assumptions, in the recent work [35], A. Ray proved stronger bounds for the Mordell–Weil rank of an elliptic curve with good ordinary reduction at all the primes above pp along the noncommutative extension F(E[p])F(E[p^{\infty}]) (see [35, Theorem 2.5 and Remark 2.6]. Therefore, it is a natural question to generalize those result in the supersingular case when FF_{\infty} is a pd{\mathbb{Z}}_{p}^{d}-extension over FF containing the cyclotomic extension FcycF_{\mathrm{cyc}}. One might also ask if it is possible to remove this hypothesis on the 𝔐H(G)\mathfrak{M}_{H}(G)-conjecture, or at least replace it with something weaker. Another avenue to explore is the case of mixed reduction type. Let EE be an elliptic curve defined over a number field FF where pp splits completely. Suppose that EE has good reduction at primes above pp. In [36], Lei and Lim constructed multi-signed Selmer groups where they allow both ordinary or supersingular reduction at primes above pp. Under the assumption that at least one prime is supersingular for EE and that the dual of the aforementioned Selmer groups are torsion over the appropriate Iwasawa algebra, they show [36, Theorem 5.9] that the Mordell-Weil rank of EE stays bounded along the cyclotomic p\mathbb{Z}_{p}-extension of FF. One may ask if this can be generalized to abelian varieties with mixed reduction type at primes above pp over more general extensions. These are our future projects and needs further research.

6. Speculative remarks on multi-signed pp-adic LL-functions for GSp(4)\mathrm{GSp}(4) by Chris Williams

We thank Chris Williams222CW would like to thank Antonio Lei and David Loeffler for very informative conversations on this topic, though any and all misconceptions are his own. for allowing us to include this in our paper.

There are a number of works in the literature proving signed Iwasawa main conjectures relating signed Selmer groups to signed pp-adic LL-functions. As such, it is natural to ask if the Selmer groups of the present paper have corresponding pp-adic LL-functions. We finally make some (very) speculative remarks in this direction.

6.1. Elliptic curves

In the case of elliptic curves over {\mathbb{Q}} with good supersingular reduction, there are two signed Selmer groups, and two `standard' pp-adic LL-functions, i.e. those interpolating LL-values in the standard (and only) critical region. Via [37] these give rise to two signed pp-adic LL-functions, and the signed Iwasawa main conjectures [26] relate all possible Selmer groups and all possible pp-adic LL-functions.

Let KK be an imaginary quadratic field in which pp splits. Let EE be an elliptic curve over KK with good supersingular reduction at both primes above pp. Via modularity – recently proved in many cases in [38] – we expect that if it is not a \mathbb{Q}-curve, E/KE/K corresponds to a weight 2 cuspidal Bianchi newform \mathcal{F} of level Γ0(𝔫)\Gamma_{0}(\mathfrak{n}), for some ideal 𝔫𝒪K\mathfrak{n}\subset\mathcal{O}_{K} prime to p𝒪Kp\mathcal{O}_{K}. It thus makes sense to discuss the theory on the automorphic side, for {\mathcal{F}} rather than EE.

In this setting, the set \mathcal{I} has size 6, containing the pairs

(J𝔭,J𝔭c)=(,{1,2}),({1},{1}),({1},{2}),({2},{1}),({2},{2}),({1,2},).(J_{\mathfrak{p}},J_{\mathfrak{p}^{c}})=(\varnothing,\{1,2\}),\ (\{1\},\{1\}),\ (\{1\},\{2\}),\ (\{2\},\{1\}),\ (\{2\},\{2\}),\ (\{1,2\},\varnothing).

These are the examples of Proposition 4.18, and thus all 6 correspond to canonical multi-signed Selmer groups in this case. On the analytic side, there are only 4 multi-signed pp-adic LL-functions, constructed in [39, §5] from the 4 standard pp-adic LL-functions of §4 op. cit. These correspond to the 4 Selmer groups where |J𝔭|=|J𝔭c|=1|J_{\mathfrak{p}}|=|J_{\mathfrak{p}^{c}}|=1. The discrepancy here arises as the standard pp-adic LL-functions do not account for all possible regions where the LL-function has critical values. This phenomenon is discussed in [40, Fig. 4.1] or [41, Fig. 6.1]. There should be two additional two-variable pp-adic LL-functions of `BDP type', interpolating the critical LL-values in the `non-standard' region. When EE can be defined over {\mathbb{Q}}, i.e. when {\mathcal{F}} is the base-change of a classical modular form, one- and two-variable pp-adic LL-functions interpolating values in this region were studied in [40] and [42] respectively. The Selmer groups where J𝔭J_{\mathfrak{p}} or J𝔭cJ_{\mathfrak{p}^{c}} is empty should be related to these BDP type pp-adic LL-functions. Their construction in the non-base-change setting (that is, for `genuine' Bianchi modular forms) remains mysterious.

We briefly elaborate why there are four standard pp-adic LL-functions. Recall the Bianchi modular form {\mathcal{F}} has level Γ0(𝔫)\Gamma_{0}(\mathfrak{n}). There are four pp-refinements to level Γ0(p𝔫)\Gamma_{0}(p\mathfrak{n}): there are independently two refinements at 𝔭\mathfrak{p} and two at 𝔭c\mathfrak{p}^{c}, corresponding to choices of roots of the Hecke polynomials at 𝔭\mathfrak{p} and 𝔭c\mathfrak{p}^{c} (see [43, §2.1]). This amounts to choosing one of the (four) Hecke eigenspaces in S2(Γ0(p𝔫))S_{2}(\Gamma_{0}(p\mathfrak{n})) where the prime-to-pp Hecke operators act as they do on \mathcal{F}. Each pp-refinement yields a standard pp-adic LL-function.

Rephrasing, \mathcal{F} generates an automorphic representation π\pi of GL2(𝔸K)\mathrm{GL}_{2}(\mathbb{A}_{K}) whose local components π𝔭\pi_{\mathfrak{p}} and π𝔭c\pi_{\mathfrak{p}^{c}} are both unramified principal series. A pp-refinement is a pair of (independent) choices of U𝔭U_{\mathfrak{p}} and U𝔭cU_{\mathfrak{p}^{c}} eigenvalues in the 2-dimensional spaces π𝔭Iw𝔭\pi_{\mathfrak{p}}^{\mathrm{Iw}_{\mathfrak{p}}} and π𝔭cIw𝔭c\pi_{\mathfrak{p}^{c}}^{\mathrm{Iw}_{\mathfrak{p}^{c}}} of Iwahori-invariant vectors.

Summarising, in the case of elliptic curves, all of the 6 elements of the set \mathcal{I} have attached multi-signed Selmer groups and pp-adic LL-functions, which we expect to be related by suitable Iwasawa main conjectures.

6.2. Abelian surfaces

Suppose now 2g=22g=2, i.e. AA is an abelian surface. Under the paramodularity conjecture a suitably `generic' abelian surface A/KA/K should correspond to a cuspidal automorphic representation Π\Pi of GSp4(𝔸K)\mathrm{GSp}_{4}({\mathbb{A}}_{K}) (and this is known after base-change to some finite extension by [44]). If AA has good reduction at the primes above pp, then Π𝔭\Pi_{\mathfrak{p}} and Π𝔭c\Pi_{\mathfrak{p}^{c}} will be unramified.

In this case, the set \mathcal{I} has size 70. By analogy to the elliptic curves case, one might naively expect that there are 70 multi-signed Selmer groups with 70 corresponding multi-signed pp-adic LL-functions. A key difference in this case, however, is that unlike weight 2 classical and Bianchi modular forms, where it exists Π\Pi is not cohomological. This causes some degeneration, which on the algebraic side is reflected in Proposition 4.18 of the main text: only 6 elements of \mathcal{I} lead to canonically defined multi-signed Coleman maps that are independent of the choice of Hodge-compatible basis. We now indicate how this degeneration occurs on the analytic side.

6.2.1. Counting pp-adic LL-functions: cohomological case

First consider the cohomological case, which is the closest direct analogue to the elliptic curves setting. Let Π\Pi be a cohomological cuspidal automorphic representation of GSp4(𝔸K)\mathrm{GSp}_{4}({\mathbb{A}}_{K}) unramified at the primes above pp. As above, there should be standard pp-adic LL-functions attached to all pp-refinements of Π\Pi to sufficiently deep level at pp. In this case, the Panchishkin condition [45] predicts that we should look for Hecke eigensystems that arise in the Klingen-parahoric invariants of Π𝔭\Pi_{\mathfrak{p}} and Π𝔭c\Pi_{\mathfrak{p}^{c}}. One may compute that the Klingen-parahoric invariants of Π𝔭\Pi_{\mathfrak{p}} are 4-dimensional (e.g. [46, Prop. 3.15]). Given Π\Pi is cohomological, we expect that this 4-dimensional space should be the direct sum of four 1-dimensional Hecke eigenspaces, each with different Hecke eigenvalues. There should then be 16 standard pp-adic LL-functions for Π\Pi, corresponding to the 16 distinct pairs of choices of Klingen-invariant Hecke eigensystems in Π𝔭\Pi_{\mathfrak{p}} and Π𝔭c\Pi_{\mathfrak{p}^{c}}.

It is natural to expect higher-weight analogues of the constructions of the present article, upon which we may ask which elements of \mathcal{I} these pp-adic LL-functions should correspond to. An alternative version of this theory can be described after transferring Π\Pi to an automorphic representation π\pi of GL4(𝔸K)\mathrm{GL}_{4}({\mathbb{A}}_{K}) via [47], as considered in [48, §3.3] or [49, §5–7]. One sees that, for the Asgari–Shahidi conventions considered in [49], the 16 standard pp-adic LL-functions should correspond to Selmer groups where J𝔭J_{\mathfrak{p}} and J𝔭cJ_{\mathfrak{p}^{c}} are independently allowed to be one of {1,2},{1,3},{4,2},{4,3}\{1,2\},\{1,3\},\{4,2\},\{4,3\} (i.e. each can contain precisely one of {1,4}\{1,4\} and precisely one of {2,3}\{2,3\}). It is reasonable to expect that from these, one can construct 16 multi-signed pp-adic LL-functions (see §6.2.3). This accounts for only 16 of the 70 elements in \mathcal{I}:

  • (a)

    There are a further 20 elements of \mathcal{I} where each of J𝔭J_{\mathfrak{p}} and J𝔭cJ_{\mathfrak{p}^{c}} both have size 2, but are not of the special shape above: these should correspond to pp-refinements of the GL4\mathrm{GL}_{4}-representation π\pi that are not of `Shalika type'. This classification of the pp-refinements is studied in detail in [50]; in the language op. cit., the 16 `good/Shalika-type' elements above exactly correspond to the refinements that are `QQ-spin' at both primes above pp, and the other 20 are not QQ spin at one or both primes above pp.

  • (b)

    There are 16 elements of \mathcal{I} where |J𝔭|=1|J_{\mathfrak{p}}|=1 (so |J𝔭c|=3|J_{\mathfrak{p}^{c}}|=3), and 16 when |J𝔭c|=1|J_{\mathfrak{p}^{c}}|=1.

  • (c)

    Finally, there are 2 cases where either J𝔭=J_{\mathfrak{p}}=\varnothing or J𝔭c=J_{\mathfrak{p}^{c}}=\varnothing.

All of this is reflected in the fact that L(Π,s)L(\Pi,s) now has many more critical regions, analogous for example to [51, §2.3]. The standard pp-adic LL-functions above and in (a) will only see the standard critical region. The 2 cases of type (c) should correspond to BDP-style pp-adic LL-functions, and will interpolate values in another of the critical regions. One might expect more types of pp-adic LL-functions, corresponding to type (b) elements of \mathcal{I} and interpolating LL-values in the other regions. This hints at a tantalising, but at present very mysterious, picture of Selmer groups and pp-adic LL-functions in this setting.

6.2.2. Counting pp-adic LL-functions: non-cohomological case

Now suppose Π\Pi is not cohomological, attached to an abelian variety AA with supersingular reduction at each prime above pp. This causes degeneracy in the above picture.

In the picture of [51, §2.3], a number of the critical regions degenerate away to nothing: that is, some of the critical regions are empty in the non-cohomological case. (This is an analogue, for example, of the fact that the standard critical region is empty for weight 1 modular forms). In particular, in the type (b) cases above there is no hope of constructing pp-adic LL-functions with interpolative properties for non-cohomological Π\Pi.

There is also degeneracy in the 16 standard pp-adic LL-functions. In the supersingular case, the Frobenius eigenvalues at pp can only be ±p\pm\sqrt{p}, each appearing twice. The 4-dimensional Klingen-invariants in Π𝔭\Pi_{\mathfrak{p}} hence cannot split into four disjoint 1-dimensional Hecke eigenspaces, but only as a direct sum of two 2-dimensional Hecke eigenspaces, putting us in the so-called `irregular' setting. For GL2\mathrm{GL}_{2} over {\mathbb{Q}} and KK, the irregular setting was for example studied in [52]. As a result, there should only be 4 standard pp-adic LL-functions in this case, corresponding to independent choices of eigenvalue ±p\pm\sqrt{p} at 𝔭\mathfrak{p} and 𝔭c\mathfrak{p}^{c}.

These 4 standard pp-adic LL-functions should then have multi-signed analogues, corresponding to the 4 canonical Selmer groups in Proposition 4.18 with |J𝔭|=|J𝔭c|=2|J_{\mathfrak{p}}|=|J_{\mathfrak{p}^{c}}|=2.

This picture holds in higher dimensions: for cohomological Π\Pi, the number of standard pp-adic LL-functions for GSp2g\mathrm{GSp}_{2g} will grow with gg, but in the case of supersingular abelian varieties, there will only ever be 2 choices (again either ±p\pm\sqrt{p}) of pp-refinement for each of 𝔭\mathfrak{p} and 𝔭c\mathfrak{p}^{c}, reflected on the algebraic side in the four `standard' cases in Proposition 4.18.

The other cases of Proposition 4.18, where J𝔭J_{\mathfrak{p}} or J𝔭cJ_{\mathfrak{p}^{c}} is empty, provide good evidence that there exist two-variable BDP-style pp-adic LL-functions attached to abelian varieties.

6.2.3. What is known?

For cohomological automorphic representations of GSp4\mathrm{GSp}_{4} over KK, the existence of standard pp-adic LL-functions is now known for appropriate non-critical slope refinements (those of `Shalika type' above), but only after transfer to GL4\mathrm{GL}_{4} (see [53]). The signed theory has not yet been explored in this setting.

We briefly also mention some related work when the base field is {\mathbb{Q}}, and where more is known. For cohomological representations Π\Pi on GSp(4)/\mathrm{GSp}(4)/{\mathbb{Q}}, standard pp-adic LL-functions for Π\Pi have been constructed (after transfer to π\pi on GL(4)\mathrm{GL}(4)) for Klingen-invariant eigensystems that correspond to non-ordinary, non-critical pp-refinements for GL(4)\mathrm{GL}(4) (see [54]). When π\pi is `good supersingular' at pp, and there are two non-critical slope pp-refinements π~1\tilde{\pi}_{1} and π~2\tilde{\pi}_{2} of π\pi, the recent papers [55] and [56] construct two signed pp-adic LL-functions attached to π\pi. They use the unbounded pp-adic LL-functions Lp(π~1)L_{p}(\tilde{\pi}_{1}) and Lp(π~2)L_{p}(\tilde{\pi}_{2}) of [54], and prove that

Lp(π~1)±Lp(π~2)=Lp±×logπ±,L_{p}(\tilde{\pi}_{1})\pm L_{p}(\tilde{\pi}_{2})=L_{p}^{\pm}\times\mathrm{log}_{\pi}^{\pm},

where Lp±L_{p}^{\pm} is a measure and logπ±\mathrm{log}_{\pi}^{\pm} is Pollack's (unbounded) ±\pm-logarithmic distribution. The signed pp-adic LL-functions are then the measures Lp±L_{p}^{\pm}.

One might expect the existence of two more (critical slope) signed pp-adic LL-functions in this case. This is presently not known, but one can guess at the shape of this theory. There are two further critical slope pp-refinements of Π\Pi, transferring to critical slope pp-refinements π~3\tilde{\pi}_{3} and π~4\tilde{\pi}_{4} of π\pi via the process in [49, §5]. In the `supersingular' setting above, these refinements can be shown to have the same slope for the UpU_{p} operator defined by diag(p1n,1n)\mathrm{diag}(p1_{n},1_{n}); hence, if π~3\tilde{\pi}_{3} and π~4\tilde{\pi}_{4} are both non-critical (cf. [54, Rem. 3.15]), then the corresponding pp-adic LL-functions constructed in Theorem 6.23 op. cit. have the same growth property. One might reasonably expect (in line with [55, 56]) that Lp(π~3)±Lp(π~4)L_{p}(\tilde{\pi}_{3})\pm L_{p}(\tilde{\pi}_{4}) is of the form [measure]± ×\times [explicit unbounded logarithmic distribution]±. However, unlike in [55, 56], it is not possible to deduce this solely using the interpolative properties of Lp(π~3)L_{p}(\tilde{\pi}_{3}) and Lp(π~4)L_{p}(\tilde{\pi}_{4}) alone: the growth is too big, so there are insufficient critical LL-values to uniquely determine Lp(π~3)±Lp(π~4).L_{p}(\tilde{\pi}_{3})\pm L_{p}(\tilde{\pi}_{4}).

The above constructions fundamentally use that Π\Pi (hence π\pi) appears in the Betti cohomology of locally symmetric spaces, and break down for non-cohomological Π\Pi. To construct pp-adic LL-functions for such Π\Pi, one strategy is to deform from cohomological to non-cohomological weight along an eigenvariety. In the present setting, this appears extremely difficult, as the points are non-regular, non-cohomological, and non-ordinary. In particular:

  • One must obtain good control over the GSp4\mathrm{GSp}_{4}-eigenvariety at non-regular, non-cohomological points, analogous to the study done in [57] for GL2/\mathrm{GL}_{2}/{\mathbb{Q}}.

  • In the GSp4\mathrm{GSp}_{4} setting, the relevant points will not be ordinary, meaning one cannot study them using Hida families. Instead one must use Coleman families, which do not exist over the whole weight space but only over smaller neighbourhoods.

  • For GSp4/K\mathrm{GSp}_{4}/K, since GSp4()\mathrm{GSp}_{4}(\mathbb{C}) does not admit discrete series, the picture is even worse. In general, it is not expected that the pp-refinements of A/KA/K will vary in classical families in the eigenvariety; this is a folklore conjecture [58, Intro.], a general analogue of [59]. In this scenario deformation from cohomological weight is impossible.

For A/A/{\mathbb{Q}} ordinary, a pp-adic LL-function attached to its (unique) ordinary pp-refinement is given in [60, Prop. 3.3]. However, the method of proof – using higher Hida theory, via the coherent cohomology of Shimura varieties – will not apply to the analogous setting over KK, where the corresponding locally symmetric spaces are not even algebraic varieties.

Summarising, for general supersingular AA defined over an imaginary quadratic field KK with pp split, constructing the 4 conjectured standard pp-adic LL-functions attached to A/KA/K seems completely out of reach with present methods.

Finally, consider the degenerate case where A/KA/K is a product of elliptic curves E1××EgE_{1}\times\cdots\times E_{g}, each EiE_{i} defined over KK. The Galois representation attached to AA will just be a direct sum of the Tate modules of the EiE_{i}, and its LL-function will be a product of the L(Ei,s)L(E_{i},s). Good candidates for the multi-signed pp-adic LL-functions in this case would be products of the multi-signed pp-adic LL-functions attached to EiE_{i} constructed in [39, Prop. 5.2], where we take the same choice in {++,+,+,}\{++,+-,-+,--\} at each of the EiE_{i}.

References

  • [1] Barry Mazur. Rational points of abelian varieties with values in towers of number fields. Invent. Math., 18:183–266, 1972.
  • [2] Kazuya Kato. pp-adic Hodge theory and values of zeta functions of modular forms. Astérisque, (295):ix, 117–290, 2004. Cohomologies pp-adiques et applications arithmétiques. III.
  • [3] David E. Rohrlich. On LL-functions of elliptic curves and cyclotomic towers. Invent. Math., 75(3):409–423, 1984.
  • [4] Kâzım Büyükboduk and Antonio Lei. Integral Iwasawa theory of Galois representations for non-ordinary primes. Mathematische Zeitschrift, 286:361–398, 2017.
  • [5] Shin-ichi Kobayashi. Iwasawa theory for elliptic curves at supersingular primes. Invent. Math., 152(1):1–36, 2003.
  • [6] Florian E. Ito Sprung. Iwasawa theory for elliptic curves at supersingular primes: a pair of main conjectures. J. Number Theory, 132(7):1483–1506, 2012.
  • [7] Antonio Lei and Gautier Ponsinet. On the Mordell-Weil ranks of supersingular abelian varieties in cyclotomic extensions. Proceedings of the American Mathematical Society, Series B, 7(1):1–16, 2020.
  • [8] Antonio Lei and Florian Sprung. Ranks of elliptic curves over p2\mathbb{Z}_{p}^{2}-extensions. Israel J. Math., 236(1):183–206, 2020.
  • [9] Florian Ito Sprung. Chromatic Selmer groups and arithmetic invariants of elliptic curves. J. Théor. Nombres Bordeaux, 33(3):1103–1114, 2021.
  • [10] Massimo Bertolini. Selmer groups and Heegner points in anticyclotomic p\mathbb{Z}_{p}-extensions. Compositio Math., 99(2):153–182, 1995.
  • [11] Jan Nekovář. Level raising and anticyclotomic Selmer groups for Hilbert modular forms of weight two. Canad. J. Math., 64(3):588–668, 2012.
  • [12] Benjamin Howard. Iwasawa theory of Heegner points on abelian varieties of GL2\rm GL_{2} type. Duke Math. J., 124(1):1–45, 2004.
  • [13] Christophe Cornut. Mazur's conjecture on higher Heegner points. Invent. Math., 148(3):495–523, 2002.
  • [14] V. Vatsal. Special values of anticyclotomic LL-functions. Duke Math. J., 116(2):219–261, 2003.
  • [15] B. Mazur. Modular curves and arithmetic. In Proceedings of the International Congress of Mathematicians, Vol. 1, 2 (Warsaw, 1983), pages 185–211. PWN, Warsaw, 1984.
  • [16] Matteo Longo and Stefano Vigni. Plus/minus Heegner points and Iwasawa theory of elliptic curves at supersingular primes. Boll. Unione Mat. Ital., 12(3):315–347, 2019.
  • [17] Byoung Du Kim. Signed-Selmer groups over the p2\mathbb{Z}_{p}^{2}-extension of an imaginary quadratic field. Canad. J. Math., 66(4):826–843, 2014.
  • [18] Kâzım Büyükboduk and Antonio Lei. Coleman-adapted Rubin-Stark Kolyvagin systems and supersingular Iwasawa theory of CM abelian varieties. Proc. Lond. Math. Soc. (3), 111(6):1338–1378, 2015.
  • [19] Kâzım Büyükboduk and Antonio Lei. Iwasawa theory of elliptic modular forms over imaginary quadratic fields at non-ordinary primes. Int. Math. Res. Not. IMRN, (14):10654–10730, 2021.
  • [20] Jean-Marc Fontaine and Bernadette Perrin-Riou. Autour des conjectures de Bloch et Kato: cohomologie galoisienne et valeurs de fonctions LL. In Motives (Seattle, WA, 1991), volume 55 of Proc. Sympos. Pure Math., pages 599–706. Amer. Math. Soc., Providence, RI, 1994.
  • [21] Laurent Berger. Limites de représentations cristallines. Compos. Math., 140(6):1473–1498, 2004.
  • [22] David Loeffler and Sarah Livia Zerbes. Iwasawa theory and pp-adic L-functions over p2\mathbb{Z}_{p}^{2}-extensions. International Journal of Number Theory, 10(08):2045–2095, 2014.
  • [23] Antonio Lei and Luochen Zhao. On the BDP Iwasawa main conjecture for modular forms. Manuscripta Math., 173(3-4):867–888, 2024.
  • [24] Bernadette Perrin-Riou. Fonctions LL pp-adiques des représentations pp-adiques. Astérisque, (229):198, 1995.
  • [25] Ralph Greenberg. Iwasawa theory for pp-adic representations. In Algebraic number theory, volume 17 of Adv. Stud. Pure Math., pages 97–137. Academic Press, Boston, MA, 1989.
  • [26] Francesc Castella, Mirela Çiperiani, Christopher Skinner, and Florian Sprung. On the Iwasawa main conjectures for modular forms at non-ordinary primes. https://arxiv.org/abs/1804.10993, 2018.
  • [27] Antonio Lei and Bharathwaj Palvannan. Codimension two cycles in Iwasawa theory and elliptic curves with supersingular reduction. Forum Math. Sigma, 7:Paper No. e25, 81, 2019.
  • [28] Gautier Ponsinet. On the structure of signed Selmer groups. Math. Z., 294(3-4):1635–1658, 2020.
  • [29] Jishnu Ray and Florian Sprung. On the signed selmer groups for motives at non-ordinary primes in p2\mathbb{Z}_{p}^{2}-extensions. ArXiv 2309.02016, 2023.
  • [30] Albert A. Cuoco and Paul Monsky. Class numbers in 𝐙pd{\bf Z}^{d}_{p}-extensions. Math. Ann., 255(2):235–258, 1981.
  • [31] Michael Harris. Correction to: ``pp-adic representations arising from descent on abelian varieties'' [Compositio Math. 39 (1979), no. 2, 177–245; MR0546966 (80j:14035)]. Compositio Math., 121(1):105–108, 2000.
  • [32] Hideo Imai. A remark on the rational points of abelian varieties with values in cyclotomic ZpZ_{p}-extensions. Proc. Japan Acad., 51:12–16, 1975.
  • [33] Ralph Greenberg. Galois theory for the Selmer group of an abelian variety. Compositio Math., 136(3):255–297, 2003.
  • [34] Pin-Chi Hung and Meng Fai Lim. On the growth of Mordell-Weil ranks in pp-adic Lie extensions. Asian J. Math., 24(4):549–570, 2020.
  • [35] Anwesh Ray. Asymptotic growth of Mordell-Weil ranks of elliptic curves in noncommutative towers. Canad. Math. Bull., 65(4):1050–1062, 2022.
  • [36] Antonio Lei and Meng Fai Lim. Mordell-Weil ranks and Tate-Shafarevich groups of elliptic curves with mixed-reduction type over cyclotomic extensions. Int. J. Number Theory, 18(2):303–330, 2022.
  • [37] Robert Pollack. On the pp-adic LL-function of a modular form at a supersingular prime. Duke Math. J., 118(3):523–558, 2003.
  • [38] Ana Caraiani and James Newton. On the modularity of elliptic curves over imaginary quadratic fields. Preprint: https://arxiv.org/abs/2301.10509.
  • [39] David Loeffler. PP-adic integration on ray class groups and non-ordinary pp-adic LL-functions. In T. Bouganis and O. Venjakob, editors, Iwasawa 2012: State of the art and recent advances, volume 7 of Contributions in Mathematical and Computational Sciences, pages 357 – 378. Springer, 2014.
  • [40] Massimo Bertolini, Henri Darmon, and Kartik Prasanna. Generalized Heegner cycles and pp-adic Rankin LL-series. Duke Math. J., 162(6):1033–1148, 2013.
  • [41] Antonio Lei, David Loeffler, and Sarah Livia Zerbes. Euler systems for modular forms over imaginary quadratic fields. Compos. Math., 151(9):1585–1625, 2015.
  • [42] Xin Wan. Iwasawa main conjecture for Rankin-Selberg pp-adic LL-functions. Algebra Number Theory, 14(2):383–483, 2020.
  • [43] Daniel Barrera Salazar and Chris Williams. Families of Bianchi modular symbols: critical base-change pp-adic LL-functions and pp-adic Artin formalism. Selecta Math. (N.S.), 27(5):Paper No. 82, 45, 2021.
  • [44] George Boxer, Frank Calegari, Toby Gee, and Vincent Pilloni. Abelian surfaces over totally real fields are potentially modular. Publ. Math. Inst. Hautes Études Sci., 134:153–501, 2021.
  • [45] Alexei A. Panchishkin. Motives over totally real fields and pp-adic LL-functions. Ann. Inst. Fourier (Grenoble), 44(4):989–1023, 1994.
  • [46] Masao Oi, Ryotaro Sakamoto, and Hiroyoshi Tamori. Iwahori-Hecke algebra and unramified local LL-functions. Math. Z., 303(3):Paper No. 59, 42, 2023.
  • [47] Mahdi Asgari and Freydoon Shahidi. Generic transfer for general spin groups. Duke Math. J., 132(1):137–190, 2006.
  • [48] Mladen Dimitrov, Fabian Januszewski, and A. Raghuram. L{L}-functions of GL(2n)\mathrm{GL}(2n): pp-adic properties and nonvanishing of twists. Compositio Math., 156(12):2437–2468, 2020.
  • [49] Daniel Barrera Salazar, Mladen Dimitrov, Andrew Graham, Andrei Jorza, and Chris Williams. On pp-adic interpolation of branching laws for GL2n\mathrm{{GL}}_{2n}: pp-adic L{L}-functions in Shalika families. Preprint: https://arxiv.org/abs/2211.08126.
  • [50] Daniel Barrera Salazar, Andrew Graham, and Chris Williams. On pp-refined Friedberg–Jacquet integrals and the classical symplectic locus in the GL2n\mathrm{GL}_{2n}-eigenvariety. Preprint: https://arxiv.org/abs/2308.02649.
  • [51] David Loeffler and Sarah Livia Zerbes. P{P}-adic L{L}-functions and diagonal cycles for GSp(4) ×\times GL(2) ×\times GL(2). Preprint, https://arxiv.org/abs/2011.15064.
  • [52] Adel Betina and Chris Williams. Arithmetic of pp-irregular modular forms: families and pp-adic LL-functions. Mathematika, 67(4):917–948, 2021.
  • [53] Chris Williams. On pp-adic LL-functions for symplectic representations of GL(N)\mathrm{GL}(N) over number fields. Preprint: https://arxiv.org/abs/2305.07809.
  • [54] Daniel Barrera Salazar, Mladen Dimitrov, and Chris Williams. On pp-adic LL-functions for GL(2n)\mathrm{{GL}}(2n) in finite slope Shalika families. Preprint: https://arxiv.org/abs/2103.10907.
  • [55] Rob Rockwood. Plus/minus pp-adic LL-functions for GL2n{\rm GL}_{2n}. Ann. Math. Qué., 47(1):177–193, 2023.
  • [56] Antonio Lei and Jishnu Ray. Iwasawa theory of automorphic representations of GL2n{\rm GL}_{2n} at non-ordinary primes. Res. Math. Sci., 10(1):Paper No. 1, 25, 2023.
  • [57] Adel Betina, Mladen Dimitrov, and Alice Pozzi. On the failure of Gorensteinness at weight 1 Eisenstein points of the eigencurve. Amer. J. Math., 144(1):227–265, 2022.
  • [58] Eric Urban. Eigenvarieties for reductive groups. Ann. of Math. (2), 174(3):1685–1784, 2011.
  • [59] Frank Calegari and Barry Mazur. Nearly ordinary Galois deformations over arbitrary number fields. J. Inst. Math. Jussieu, 8(1):99–177, 2009.
  • [60] David Loeffler and Sarah Livia Zerbes. On the Birch–Swinnerton-Dyer conjecture for modular abelian surfaces. Preprint, https://arxiv.org/abs/2110.13102.