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

Kolyvagin’s Conjecture, bipartite Euler systems, and higher congruences of modular forms

Naomi Sweeting [email protected] Department of Mathematics, Harvard University
Abstract.

Let E/E/\mathbb{Q} be an elliptic curve and let KK be an imaginary quadratic field. Under a certain Heegner hypothesis, Kolyvagin constructed cohomology classes for EE using KK-CM points and conjectured they did not all vanish. Conditional on this conjecture, he described the Selmer rank of EE using his system of classes. We extend work of Wei Zhang to prove new cases of Kolyvagin’s conjecture by considering congruences of modular forms modulo large powers of pp. Additionally, we prove an analogous result, and give a description of the Selmer rank, in a complementary “definite” case (using certain modified LL-values rather than CM points). Similar methods are also used to improve known results on the Heegner point main conjecture of Perrin-Riou. One consequence of our results is a new converse theorem, that pp-Selmer rank one implies analytic rank one, when the residual representation has dihedral image.

1. Introduction

Let f:𝕋N𝒪ff:\mathbb{T}_{N}\to\mathcal{O}_{f} be a non-CM newform of level NN, weight 2, and trivial nebentypus, where 𝒪f\mathcal{O}_{f} is the ring of integers in an algebraic extension of \mathbb{Q}. The Birch and Swinnerton-Dyer Conjecture asserts the equality:

(1) r(A/)=ords=1L(A,s),r(A/\mathbb{Q})=\operatorname{ord}_{s=1}L(A,s),

where A=AfA=A_{f} is the associated abelian variety to ff and rr is its Mordell-Weil rank. In pioneering works on this problem, Perrin-Riou [39] and Kolyvagin [30, 31] studied ranks of elliptic curves over an auxiliary imaginary quadratic field KK through the theory of Heegner points on modular curves. We prove, in new cases, conjectures made by both authors.

Fix a quadratic imaginary field KK, and a prime 𝒪f\wp\subset\mathcal{O}_{f} of residue characteristic. We write TfT_{f} for (a lattice in) the \wp-adic Galois representation associated to ff. Assume the following generalized Heegner hypothesis:

(Heeg) N=N+N, where all |N+ are split in K, all |N are inert in K, and N is squarefree,N=N^{+}N^{-},\text{ where all }\ell|N^{+}\text{ are split in }K,\text{ all }\ell|N^{-}\text{ are inert in }K,\text{ and }N^{-}\text{ is squarefree,}

as well as:

(unr) p2Ndisc(K).p\nmid 2N\operatorname{disc}(K).

For purposes of exposition in this introduction, we also assume:

(sclr) The image of the GK action on T¯f contains a nonzero scalar.\text{The image of the $G_{K}$ action on }\overline{T}_{f}\text{ contains a nonzero scalar.}

To state Kolyvagin’s conjecture, assume that the number of prime factors ν(N)\nu(N^{-}) is even. If mm is a squarefree product of primes inert in KK, one can use Heegner points of conductor mm on the Shimura curve XN+,NX_{N^{+},N^{-}} to construct classes

c(m)H1(K,Tf/Im),c(m)\in H^{1}(K,T_{f}/I_{m}),

where ImI_{m} is the ideal of 𝒪=𝒪f,\mathcal{O}=\mathcal{O}_{f,\wp} generated by +1\ell+1 and aa_{\ell} for all |m\ell|m. (In the text, c(m)c(m) is denoted c¯(m,1)\overline{c}(m,1).) These classes are a mild generalization of the ones constructed by Kolyvagin [31]. We are able to prove the following result towards Kolyvagin’s conjecture on the nonvanishing of the system {c(m)}\left\{c(m)\right\}:

Theorem A.

[Corollary 8.3.7] Assume (Heeg), (unr), and (sclr) hold for f,,f,\wp, and KK, and ν(N)\nu(N^{-}) is even. Suppose the following conditions hold:

(){ The modulo representation ¯Tf associated to f is absolutely irreducible; if =p3, then ¯Tf is not induced from a character of GQ-3. If p is inert in K or ap is not a -adic unit, then there exists some prime ||N. If ap is not a -adic unit, then either may be chosen above so that Af has non-split toric reduction at , or the image of the Galois action on Tf contains a conjugate of SL2(Zp). (\diamondsuit)\hskip 14.22636pt\left\{\parbox{398.33858pt}{$\mathbin{\vbox{\hbox{\scalebox{0.75}{$\bullet$}}}}$ The modulo $\wp$ representation $\overline{T}_{f}$ associated to $f$ is absolutely irreducible; if $p=3,$ then $\overline{T}_{f}$ is not induced from a character of $G_{\mathbb{Q}\sqrt{-3}}$. \\ $\mathbin{\vbox{\hbox{\scalebox{0.75}{$\bullet$}}}}$ If $p$ is inert in $K$ or $a_{p}$ is not a $\wp$-adic unit, then there exists some prime $\ell||N$. \\ $\mathbin{\vbox{\hbox{\scalebox{0.75}{$\bullet$}}}}$ If $a_{p}$ is not a $\wp$-adic unit, then either $\ell$ may be chosen above so that $A_{f}$ has non-split toric reduction at $\ell$, or the image of the Galois action on $T_{f}$ contains a conjugate of $SL_{2}(\mathbb{Z}_{p})$.}\right.

Then there exists a nonzero Kolyvagin class

0c(m)H1(K,Tf/Im).0\neq c(m)\in H^{1}(K,T_{f}/I_{m}).

As Kolyvagin observed, Theorem A can be used to give a description of the Selmer ranks

r±=rk𝒪Sel(K,Tf)±,r^{\pm}=\operatorname{rk}_{\mathcal{O}}\operatorname{Sel}(K,T_{f})^{\pm},

where superscripts refer to the action of complex conjugation. Indeed, define the vanishing order of the system {c(m)}\left\{c(m)\right\} as

(2) νmin{ν(m):c(m)0}\nu\coloneqq\min\left\{\nu(m)\,:\,c(m)\neq 0\right\}

where as before ν\nu denotes the number of prime factors. Then we have:

Corollary B.

Under the assumptions of Theorem A,

max{r+,r}=ν+1.\max\left\{r^{+},r^{-}\right\}=\nu+1.

Moreover r++rr^{+}+r^{-} is odd, and the larger eigenspace has sign (1)ν+1ϵf(-1)^{\nu+1}\epsilon_{f}, where ϵf\epsilon_{f} is the global root number of ff.

Of course, the latter two assertions follow from the parity conjecture for ff, already proven by Nekovar [36].

Since c(1)Sel(K,Tf)c(1)\in\operatorname{Sel}(K,T_{f}) is the Kummer image of the classical Heegner point, the Gross-Zagier formula implies that L(f/K,1)0L^{\prime}(f/K,1)\neq 0 if and only if c(1)0.c(1)\neq 0. Hence Corollary B yields a so-called pp-converse theorem (in fact, under a slightly weaker hypothesis):

Corollary C.

Assume that (Heeg), (unr), and Condition \diamondsuit hold for ff, \wp, and KK, and ν(N)\nu(N^{-}) is even. Then

L(f/K,1)0rk𝒪Sel(K,Tf)=1rkAf(K)=[𝒪f:].L^{\prime}(f/K,1)\neq 0\iff\operatorname{rk}_{\mathcal{O}}\operatorname{Sel}(K,T_{f})=1\iff\operatorname{rk}_{\mathbb{Z}}A_{f}(K)=[\mathcal{O}_{f}:\mathbb{Z}].

Now suppose instead that ν(N)\nu(N^{-}) is odd; it turns out that Kolyvagin’s construction, suitably modified, may still be used to relate Selmer ranks and CM points. The Jacquet-Langlands correspondence associates to ff a quaternionic modular form

(3) ϕf:XN+,N𝒪f,\phi_{f}:X_{N^{+},N^{-}}\to\mathcal{O}_{f},

where XN+,NX_{N^{+},N^{-}} is a double coset space for a definite quaternion algebra, usually called a Shimura set. If mm is a squarefree product of primes inert in KK, there exist analogues of CM points of conductor mm on the Shimura set. Using the values of ϕf\phi_{f} at these points, we construct certain special elements (well-defined up to units)

(4) λ(m)𝒪/Im\lambda(m)\in\mathcal{O}/I_{m}

(λ(m,1)\lambda(m,1) in the text). Here the ideal Im𝒪I_{m}\subset\mathcal{O} is as before. The elements λ(m)\lambda(m) encode the same information about the Selmer ranks of Af/KA_{f}/K as Kolyvagin’s classes c(m)c(m).

Theorem D.

Suppose that (Heeg), (unr), (sclr), and Condition \diamondsuit hold for f,,f,\wp, and KK, and that ν(N)\nu(N^{-}) is odd. Then the vanishing order

νmin{ν(m):λ(m)0}\nu\coloneqq\min\left\{\nu(m)\,:\,\lambda(m)\neq 0\right\}

is finite and

ν=max{r+,r}.\nu=\max\left\{r^{+},r^{-}\right\}.

Moreover (1)ν=ϵf(-1)^{\nu}=\epsilon_{f} and r++rr^{+}+r^{-} is even.

As before, the final statement is a consequence of the parity conjecture; we include it only to emphasize that it follows from the non-vanishing of some λ(m)\lambda(m), in analogy to the indefinite case.

1.1. Comparison to previous results

In the indefinite case, the first results towards Kolyvagin’s conjecture were obtained by Zhang [53], under a number of additional assumptions: that p5,p\geq 5, that the Galois representation associated to T¯f\overline{T}_{f} is surjective, and additional hypotheses on the residual ramification. In particular, under the hypotheses of [53], there exists a class c(m)c(m) whose reduction in H1(K,T¯f)H^{1}(K,\overline{T}_{f}) is nonzero; this is not the case in general. In the definite case, the classes λ(m)\lambda(m) are a novel feature of this work and were not considered in [53].

The converse theorem we obtain (Corollary C) is new in several cases, most notably when the image of the Galois action on T¯f\overline{T}_{f} is dihedral, or when p=3.p=3. Previous results, under various additional hypotheses, were obtained by Zhang as a corollary of his work on Kolyvagin’s conjecture, and by Skinner [46] by a purely Iwasawa-theoretic method. For converse theorems in other settings, see Burungale [7] for the CM case, Castella-Grossi-Lee-Skinner [8] for the residually reducible case, Castella-Wan [9] for the supersingular case, and Skinner-Zhang [48] for the case of multiplicative reduction.

1.2. Iwasawa theory

Now suppose again that ν(N)\nu(N^{-}) is even. While the Kolyvagin classes are constructed by varying the conductor of CM points on XN+,NX_{N^{+},N^{-}} over squarefree integers, one may instead pp-adically interpolate CM points of pp-power conductor to obtain a class:

(5) 𝜿H1(K,TfΛ(Ψ)),\bm{\kappa}_{\infty}\in H^{1}(K,T_{f}\otimes\Lambda(\Psi)),

where Λ=𝒪Gal(K/K)\Lambda=\mathcal{O}\llbracket\operatorname{Gal}(K_{\infty}/K)\rrbracket is the anticyclotomic Iwasawa algebra, given GKG_{K}-action by the tautological character Ψ\Psi. (Note that the specialization of 𝜿\bm{\kappa}_{\infty} at the trivial character is a multiple of c(1)c(1).) The methods used to prove Theorem A also yield the following result towards Perrin-Riou’s Heegner point main conjecture.

Theorem E (Theorem 8.2.1).

Suppose that (Heeg), (unr), and Condition \diamondsuit hold for f,f,\wp, and KK, and that ν(N)\nu(N^{-}) is even. Suppose further that apa_{p} is a \wp-adic unit and pp splits in KK. Then there is a pseudo-isomorphism of Λ\Lambda-modules:

Sel(K,Af[])ΛMM\operatorname{Sel}(K_{\infty},A_{f}[\wp^{\infty}])^{\vee}\approx\Lambda\oplus M\oplus M

for some torsion Λ\Lambda-module MM, and

charΛ(Sel(K,TfΛ)Λ𝜿)=charΛ(M)\operatorname{char}_{\Lambda}\left(\frac{\operatorname{Sel}(K,T_{f}\otimes\Lambda)}{\Lambda\cdot\bm{\kappa}_{\infty}}\right)=\operatorname{char}_{\Lambda}(M)

as ideals of Λp\Lambda\otimes\mathbb{Q}_{p}. If (sclr) holds, then the equality is true in Λ.\Lambda.

For precise definitions of these Selmer groups and of 𝜿\bm{\kappa}_{\infty}, which is denoted 𝜿(1)\bm{\kappa}(1) in the text, see §6.1.

Finally, we have the following result on the anticyclotomic main conjecture for ff when ν(N)\nu(N^{-}) is odd. Evaluating the quaternionic modular form ϕf\phi_{f} on CM points of pp-power conductor on the Shimura set XN+,NX_{N^{+},N^{-}}, one constructs the algebraic pp-adic LL-function

(6) 𝝀Λ,\bm{\lambda}_{\infty}\in\Lambda,

denoted 𝝀(1)\bm{\lambda}(1) in the text. The square of 𝝀\bm{\lambda}_{\infty} has an interpolation property for twisted LL-values of ff.

Theorem F (Theorem 3.4.6, Proposition 6.1.5).

Suppose that (Heeg), (unr), and Condition \diamondsuit hold for f,f,\wp, and KK, and that ν(N)\nu(N^{-}) is odd. Suppose further that apa_{p} is a \wp-adic unit and pp splits in KK. Then there is a pseudo-isomorphism of Λ\Lambda-modules:

Sel(K,Af[])MM\operatorname{Sel}(K_{\infty},A_{f}[\wp^{\infty}])^{\vee}\approx M\oplus M

for some torsion Λ\Lambda-module MM, and

(𝝀)charΛ(M)(\bm{\lambda}_{\infty})\subset\operatorname{char}_{\Lambda}(M)

as ideals of Λp\Lambda\otimes\mathbb{Q}_{p}. If additionally (sclr) holds, then the inclusion is true in Λ.\Lambda.

The opposite inclusion of ideals may be deduced directly from Skinner-Urban’s proof of one divisibility in the three-variable main conjecture [47]; indeed, this is an essential ingredient in all of our results, as explained below.

1.3. Comparison to previous results

The technical hypotheses in Zhang’s proof of Kolyvagin’s conjecture were carried over to Burungale, Castella, and Kim’s proof [6] of the lower bound on the Selmer group in the Heegner point main conjecture, where it is also assumed that pp is not anomalous. While the methods used in this paper build on those of [6], Castella and Wan [10] have given a completely independent proof of a three-variable main conjecture when ν(N)\nu(N^{-}) is even. Their result also requires some hypotheses on residual ramification avoided here, and that NN be squarefree.

For upper bounds on the Selmer group in Theorem E and Theorem F, various technical assumptions on the residual representation and on the image of the Galois action were used in prior works by Howard [24, 25] and Chida-Hsieh [11].

1.4. Overview of the proofs

To prove Theorems A and D, we extend Kolyvagin’s construction to a larger system of classes

(7) c(m,Q1)H1(K,Tf/M),λ(m,Q2)𝒪/M,c(m,Q_{1})\in H^{1}(K,T_{f}/\wp^{M}),\;\;\lambda(m,Q_{2})\in\mathcal{O}/\wp^{M},

where MM is a fixed integer, and m,Q1,Q2m,Q_{1},Q_{2} are squarefree product of auxiliary primes satisfying certain congruence conditions, such that ν(NQ1)\nu(N^{-}Q_{1}) is even and ν(NQ2)\nu(N^{-}Q_{2}) is odd. The classes (7) form a bipartite Euler system in the sense of Howard [25] for each fixed mm and a Kolyvagin system for each fixed Q1.Q_{1}. If ν(N)\nu(N^{-}) itself is even, then the classes c(m,1)c(m,1) agree with Kolyvagin’s original construction. The Euler system relations are of the form:

(8) locqc(m,Q1)λ(m,Q1q)qc(m,Q1qq),\operatorname{loc}_{q}c(m,Q_{1})\sim\lambda(m,Q_{1}q)\sim\partial_{q^{\prime}}c(m,Q_{1}qq^{\prime}),

where q,qq,q^{\prime} are two additional auxiliary primes not dividing Q1Q_{1}; and

(9) loc±c(m,Q1)c(m,Q1),\operatorname{loc}_{\ell}^{\pm}c(m,Q_{1})\sim\partial_{\ell}^{\mp}c(m\ell,Q_{1}),

where \ell is an additional auxiliary prime not dividing mm. (Here locq\operatorname{loc}_{q}, q\partial_{q^{\prime}}, loc±\operatorname{loc}_{\ell}^{\pm}, ±\partial^{\pm}_{\ell} are certain localization maps landing in subspaces of the local cohomology free of rank one over 𝒪/M.\mathcal{O}/\wp^{M}.) The classes c(m,Q1)c(m,Q_{1}) were introduced by Zhang, although the λ(m,Q2)\lambda(m,Q_{2}) are only implicit in [53].

If c(m,Q1)0c(m,Q_{1})\neq 0, then one can use the Kolyvagin system relation to find an auxiliary \ell — either prime or equal to 11— such that qc(m,Q1)0\partial_{q}c(m\ell,Q_{1})\neq 0. By the bipartite Euler system relation, this implies λ(m,Q1/q)0.\lambda(m\ell,Q_{1}/q)\neq 0. On the other hand if λ(m,Q2)0\lambda(m,Q_{2})\neq 0 and q|Q2,q|Q_{2}, then c(m,Q2/q)0.c(m,Q_{2}/q)\neq 0. Combining these two observations, we reduce the non-vanishing of some class c(m,1)c(m,1) or λ(m,1)\lambda(m,1) — depending on the parity of ν(N)\nu(N^{-}) — to exhibiting a single Q2Q_{2} such that λ(1,Q2)0.\lambda(1,Q_{2})\neq 0.

Now, if there exists a newform gg of level NQ2NQ_{2} with a congruence to ff modulo M\wp^{M}, then λ(1,Q2)\lambda(1,Q_{2}) is essentially the reduction of the algebraic LL-value Lalg (g/K,1)L^{\text{alg }}(g/K,1) modulo M\wp^{M}, which is related to the length of the Selmer group of gg by the Iwasawa main conjecture [47, 51]. To complete the proof, it therefore suffices to choose a suitable Q2Q_{2} and construct such a gg with a small Selmer group. We remark that our results can only be obtained by working modulo M\wp^{M} for a large MM, since in general it will not be possible to choose gg such that Lalg(g/K,1)L^{\text{alg}}(g/K,1) is a \wp-adic unit; in [53], M=1M=1 is fixed throughout, and the need to show that the LL-value is a unit is responsible for most of the additional hypotheses.

To construct gg, we use the deformation-theoretic techniques developed by Ramakrishna [41]. Standard level-raising methods work by producing a modulo \wp eigenform of the desired level, and then using that all modulo \wp eigenforms lift to characteristic zero, but this is not the case modulo M\wp^{M}. Instead, we deform the representation Tf/MT_{f}/\wp^{M} to a \wp-adic Galois representation of a suitable auxiliary level, and then apply modularity lifting to ensure the resulting representation is modular. The auxiliary level Q2Q_{2} must be chosen to control two Selmer groups: the adjoint Selmer group governing the deformation problem, and the Selmer group Sel(K,Ag[])\operatorname{Sel}(K,A_{g}[\wp^{\infty}]) that is related to the LL-value.

We now make some remarks on the construction of the Euler system. The elements c(m,Q1)c(m,Q_{1}) (resp. λ(m,Q2)\lambda(m,Q_{2})) are constructed from CM points of conductor mm on the Shimura curve XN+,NQ1X_{N^{+},N^{-}Q_{1}} (resp. Shimura set XN+,NQ2X_{N^{+},N^{-}Q_{2}}). Similar Euler system constructions have been made by many authors, e.g. in [11, 3] as well as in [53], but all have relied on certain hypotheses ensuring an integral multiplicity one property for the space of algebraic modular forms on XN+,NQiX_{N^{+},N^{-}Q_{i}}, which we do not impose here. Instead, we obtain a control on the failure of multiplicity one, using the work of Helm [22] on maps between Jacobians of modular curves and Shimura curves. The construction of the Euler system is intimately related to level-raising, and so our method can also be viewed as improving results on level-raising of ff to algebraic eigenforms modulo M\wp^{M} new at multiple auxiliary primes, which had previously been restricted to the multiplicity one case.

The proof of Theorem E is similar to that of Theorem A: the pp-adically interpolated Heegner class 𝜿\bm{\kappa}_{\infty} is viewed as the bottom layer of an Euler system {𝜿(Q1),𝝀(Q2)}\left\{\bm{\kappa}(Q_{1}),\bm{\lambda}(Q_{2})\right\}. (The squarefree conductor mm no longer plays a role.) If gg, as above, is a newform of level NQ2NQ_{2} with a congruence to ff, then 𝝀(Q2)\bm{\lambda}(Q_{2}) is congruent to Bertolini and Darmon’s anticyclotomic pp-adic LL-function of gg [3]. Using this and an Euler system argument, we reduce the lower bound on the Selmer group in the Heegner point main conjecture to the anticyclotomic main conjecture for gg, which was proven in [47]. Finally, the upper bound on the Selmer group in the Heegner point main conjecture, as well as Theorem F, follow by standard arguments from the construction of the Euler system.

In the text, the arguments described above are phrased in the language of ultrapatching, which amounts to a formalism for letting MM tend to infinity; this also forces each prime factor of mm, Q1Q_{1}, Q2Q_{2} to tend to infinity in order to satisfy the congruence conditions. (The number of prime factors of mm, Q1,Q_{1}, and Q2Q_{2} remain bounded.) This method was inspired by [45], where ultrapatching was applied to the Taylor-Wiles construction. Our setting is different in that we patch Galois cohomology groups and Selmer groups rather than geometric étale cohomology groups. The benefit of ultrapatching is that it allows us to consider the Euler system classes as characteristic zero objects in patched Selmer groups, significantly streamlining the Euler system arguments. For instance, with patching, we are able to make precise the heuristic that the non-vanishing of each Euler system class c(m,Q1)c(m,Q_{1}) or λ(m,Q2)\lambda(m,Q_{2}) is equivalent to the (m,Qi)(m,Q_{i})-transverse Selmer group being rank one or zero, respectively, cf. Lemma 8.3.4.

Structure of the paper

In §2, we review basic properties of ultrafilters and introduce patched cohomology and Selmer groups. In §3, we present a simplified version of the theory of bipartite Euler systems that appeared in [25], using patched cohomology. In §4, we recall the geometric inputs that will be used to construct bipartite Euler systems: the work of Helm on maps between modular curves and Shimura curves, and the behavior of Heegner points on Shimura curves under reduction and specialization. In §5, we prove the modulo M\wp^{M} level-raising result and present a general framework for constructing bipartite Euler systems out of CM points, which we then specialize in §6 for our applications. In §7, we give the deformation-theoretic input to construct the newform gg (in fact a sequence gng_{n} satisfying increasingly deep congruence conditions). Finally, we prove the main results in §8. An additional calculation in cyclotomic Iwasawa theory is required for Kolyvagin’s conjecture when pp is non-ordinary or inert in KK; this is done in the appendix.

Acknowledgments

I am grateful to Mark Kisin for suggesting this problem and for his ongoing encouragement. Special thanks are additionally due to Francesc Castella, who first alerted me to the relation between Kolyvagin’s conjecture and the Heegner point main conjecture. It is also a pleasure to thank many other people with whom I had helpful conversations and correspondence over the course of this project: Robert Pollack, Xin Wan, Christopher Skinner, Alexander Petrov, Sam Marks, Aaron Landesman, and Alexander Smith. This work was supported by NSF Grant #DGE1745303.

2. Ultrafilters and patching

2.1. Ultraproducts

The facts recalled in this subsection are discussed in more detail in [32].

2.1.1.

A (non-principal) ultrafilter 𝔉\mathfrak{F} for the natural numbers ={0,1,}\mathbb{N}=\left\{0,1,\ldots\right\} is a collection of subsets of \mathbb{N} satisfying the following properties:

  1. (1)

    Every set S𝔉S\in\mathfrak{F} is infinite.

  2. (2)

    For every SS\subset\mathbb{N}, either S𝔉S\in\mathfrak{F} or S𝔉\mathbb{N}-S\in\mathfrak{F}.

  3. (3)

    If S1S2S_{1}\subset S_{2}\subset\mathbb{N} and S1𝔉S_{1}\in\mathfrak{F}, then S2𝔉S_{2}\in\mathfrak{F}.

  4. (4)

    If S1,S2𝔉S_{1},S_{2}\in\mathfrak{F}, then S1S2𝔉S_{1}\cap S_{2}\in\mathfrak{F}.

Throughout this paper, we fix once and for all a non-principal ultrafilter 𝔉\mathfrak{F} on \mathbb{N}, which is possible assuming the axiom of choice. We will say that a statement PP holds for 𝔉\mathfrak{F}-many nn\in\mathbb{N} if the set SS of nn for which PP holds lies in 𝔉\mathfrak{F}.

Proposition 2.1.2.

Suppose that 𝒞\mathcal{C} is a finite set and SS\subset\mathbb{N} lies in 𝔉\mathfrak{F}. Then for any function t:S𝒞t:S\to\mathcal{C}, there is a unique c𝒞c\in\mathcal{C} such that t(n)=ct(n)=c for 𝔉\mathfrak{F}-many nn.

Proof.

The function tt defines a finite partition of \mathbb{N}:

=(S)c𝒞t1(c).\mathbb{N}=(\mathbb{N}-S)\sqcup\bigsqcup_{c\in\mathcal{C}}t^{-1}(c).

An easy induction argument shows that, for any partition of \mathbb{N} into a finite number sets, exactly one of the sets lies in 𝔉\mathfrak{F}. Since S𝔉\mathbb{N}-S\not\in\mathfrak{F}, the result follows. ∎

2.1.3.

If ={Mn}n\mathcal{M}=\left\{M_{n}\right\}_{n\in\mathbb{N}} is a sequence of sets indexed by \mathbb{N}, then 𝔉\mathfrak{F} defines an equivalence relation \sim on Mn\prod M_{n}:

(mn)n(mn)n{n:mn=mn}𝔉.(m_{n})_{n\in\mathbb{N}}\sim(m^{\prime}_{n})_{n\in\mathbb{N}}\iff\left\{n\,:\,m_{n}=m^{\prime}_{n}\right\}\in\mathfrak{F}.

The quotient Mn/\prod M_{n}/\sim is called the ultraproduct of the sequence \mathcal{M} and is denoted 𝒰()\mathcal{U}(\mathcal{M}). The ultraproduct is functorial: let ={Mn}\mathcal{M}^{\prime}=\left\{M_{n}^{\prime}\right\} be another sequence of sets and suppose given, for 𝔉\mathfrak{F}-many nn, maps φn:MnMn\varphi_{n}:M_{n}\to M^{\prime}_{n}. Then there is a natural map φ𝒰:𝒰()𝒰()\varphi^{\mathcal{U}}:\mathcal{U}(\mathcal{M})\to\mathcal{U}(\mathcal{M}^{\prime}). Similarly, if RR is a fixed (topological) ring and each MnM_{n} is a (continuous) RR-module, then 𝒰()\mathcal{U}(\mathcal{M}) may be naturally endowed with the structure of a (continuous) RR-module; in particular, if each MnM_{n} is an abelian group, then 𝒰()\mathcal{U}(\mathcal{M}) has a natural abelian group structure as well.

The following basic properties are proven in [32, §I.1] using the functoriality of the ultraproduct.

Proposition 2.1.4.

Suppose that each MnM_{n} is a finite set, and that #Mn<C\#M_{n}<C for some constant CC and for 𝔉\mathfrak{F}-many nn. Then:

  1. (1)

    𝒰()\mathcal{U}(\mathcal{M}) is finite and #𝒰()=#Mn\#\mathcal{U}(\mathcal{M})=\#M_{n} for 𝔉\mathfrak{F}-many nn.

  2. (2)

    Suppose that each MnM_{n} is additionally endowed with the structure of a (continuous) RR-module, and let AA be another (continuous) RR-module. Given a family of isomorphisms φn:MnA\varphi_{n}:M_{n}\xrightarrow{\sim}A for 𝔉\mathfrak{F}-many nn, there is an induced isomorphism

    φ𝒰:𝒰()A.\varphi^{\mathcal{U}}:\mathcal{U}(\mathcal{M})\xrightarrow{\sim}A.
Remark 2.1.5.

If RR is (topologically) finitely generated, then there are only finitely many isomorphism classes of (continuous) RR-modules of a fixed cardinality. Hence, if \mathcal{M} is a sequence of finite RR-modules of bounded cardinality, Proposition 2.1.2 and Proposition 2.1.4 together imply that 𝒰()\mathcal{U}(\mathcal{M}) is non-canonically isomorphic to 𝔉\mathfrak{F}-many MnM_{n}.

Proposition 2.1.6.

Let 𝒞\mathcal{C} be the category of sequences of (continuous) RR-modules of uniformly bounded cardinality. Then 𝒰\mathcal{U} is exact as a functor from 𝒞\mathcal{C} to the category of (continuous) RR-modules.

Proof.

We wish to show that 𝒰\mathcal{U} preserves finite limits and colimits. Since any given finite limit or colimit in 𝒞\mathcal{C} involves only sequences of NN-torsion RR-modules, for some integer NN, the limit or colimit may be computed in the category of /N\mathbb{Z}/N\mathbb{Z}-modules, in which case [32, Proposition I.2.2] applies. ∎

2.2. Ultraprimes

2.2.1.

Fix a number field LL and let MLM_{L} be its set of places; for each vMLv\in M_{L}, fix as well an embedding L¯Lv¯\overline{L}\hookrightarrow\overline{L_{v}}. If L\mathcal{M}_{L} is the constant sequence of sets {ML}n\left\{M_{L}\right\}_{n\in\mathbb{N}}, then we define the set of ultraprimes of LL as

𝖬L=𝒰(L).\mathsf{M}_{L}=\mathcal{U}(\mathcal{M}_{L}).

By definition, an ultraprime 𝗏𝖬L\mathsf{v}\in\mathsf{M}_{L} is an equivalence class of sequences (vn)n(v_{n})_{n\in\mathbb{N}}, where each vnv_{n} is a place of LL; note that Gal(L/)\operatorname{Gal}(L/\mathbb{Q}) acts by set automorphisms on 𝖬L\mathsf{M}_{L}, compatibly with the natural projection 𝖬E𝖬L\mathsf{M}_{E}\to\mathsf{M}_{L} for a finite extension E/LE/L. The map v(v,v,)v\mapsto(v,v,\ldots) induces an embedding ML𝖬LM_{L}\hookrightarrow\mathsf{M}_{L}, written vv¯v\mapsto\underline{v}, and we say an ultraprime is constant if it lies in the image of this embedding.

Proposition 2.2.2.

Let 𝗏\mathsf{v} be a non-constant ultraprime. Then there exists a unique Frobenius element Frob𝗏Gal(L¯/L)\operatorname{Frob}_{\mathsf{v}}\in\operatorname{Gal}(\overline{L}/L) with the following property: for each finite Galois extension LEL¯L\subset E\subset\overline{L}, and for any representative (vn)(v_{n}) of 𝗏\mathsf{v}, there are 𝔉\mathfrak{F}-many nn such that vnv_{n} is unramified in E/LE/L and the Frobenius of vnv_{n} in Gal(E/L)\operatorname{Gal}(E/L) is the natural image of Frob𝗏\operatorname{Frob}_{\mathsf{v}}.

Proof.

Let (vn)n(v_{n})_{n\in\mathbb{N}} be a representative of 𝗏\mathsf{v}, and fix for the time being a finite extension E/LE/L. If vnv_{n} is archimedian or ramified in EE for 𝔉\mathfrak{F}-many nn, then Proposition 2.1.2 implies that 𝗏\mathsf{v} is constant. Thus the map that sends nn to the Frobenius of vnv_{n} in Gal(E/L)\operatorname{Gal}(E/L) is defined for 𝔉\mathfrak{F}-many nn; by Proposition 2.1.2, it sends 𝔉\mathfrak{F}-many nn to a (unique) common value gEGal(E/L)g_{E}\in\operatorname{Gal}(E/L). Note that gEg_{E} does not depend on the representative (vn)(v_{n}). By the uniqueness of gEg_{E}, the association EgEE\mapsto g_{E} is compatible with restriction to subextensions EEE^{\prime}\subset E, hence defines an element of the absolute Galois group. ∎

2.2.3.

Let 𝗏\mathsf{v} be an ultraprime. We define its abstract Galois group 𝖦𝗏\mathsf{G}_{\mathsf{v}} as Gal(Lv¯/Lv)\operatorname{Gal}(\overline{L_{v}}/L_{v}) if 𝗏=v¯\mathsf{v}=\underline{v} is constant, and as the semi direct product

^(1)Frob𝗏\widehat{\mathbb{Z}}(1)\rtimes\langle\operatorname{Frob}_{\mathsf{v}}\rangle

otherwise. Here, Frob𝗏\langle\operatorname{Frob}_{\mathsf{v}}\rangle denotes the free profinite group on one generator, acting on ^(1)\widehat{\mathbb{Z}}(1) by Frob𝗏\operatorname{Frob}_{\mathsf{v}}. We define the inertia group 𝖨𝗏𝖦𝗏\mathsf{I}_{\mathsf{v}}\subset\mathsf{G}_{\mathsf{v}} of 𝗏\mathsf{v} to be the usual inertia group if 𝗏\mathsf{v} is constant, and the normal subgroup ^(1)𝖦𝗏\widehat{\mathbb{Z}}(1)\subset\mathsf{G}_{\mathsf{v}} otherwise.

2.3. Local cohomology

2.3.1.

For any (continuous) Galois module AA defined over LL, and for any 𝗏𝖬L\mathsf{v}\in\mathsf{M}_{L}, there is a natural action of 𝖦𝗏\mathsf{G}_{\mathsf{v}} on AA (factoring through the quotient 𝖦𝗏Frob𝗏\mathsf{G}_{\mathsf{v}}\to\operatorname{Frob}_{\mathsf{v}} if 𝗏\mathsf{v} is nonconstant). We define local cohomology groups by:

𝖧i(L𝗏,A)Hctsi(𝖦𝗏,A),\displaystyle\mathsf{H}^{i}(L_{\mathsf{v}},A)\coloneqq H^{i}_{cts}(\mathsf{G}_{\mathsf{v}},A),
𝖧i(L𝗏nr,A)Hctsi(𝖨𝗏,A),i0.\displaystyle\mathsf{H}^{i}(L_{\mathsf{v}}^{nr},A)\coloneqq H^{i}_{cts}(\mathsf{I}_{\mathsf{v}},A),\;\;\;i\geq 0.

Note that the local cohomology commutes with direct limits and countable inverse limits of finite, discrete Galois modules; the former is essentially by definition of continuous cohomology and the latter is by [38, Corollary 2.6.7] applied to 𝖦𝗏,𝖨𝗏.\mathsf{G}_{\mathsf{v}},\mathsf{I}_{\mathsf{v}}.

Proposition 2.3.2.

Let 𝗏𝖬L\mathsf{v}\in\mathsf{M}_{L} be an ultraprime represented by a sequence (vn)n(v_{n})_{n\in\mathbb{N}}. If AA is a finite, discrete Galois module over LL, then for 𝔉\mathfrak{F}-many nn there are natural isomorphisms (compatible with the restriction maps and with the cup product):

Hi(Lvn,A)𝖧i(L𝗏,A),\displaystyle H^{i}(L_{v_{n}},A)\simeq\mathsf{H}^{i}(L_{\mathsf{v}},A),
Hi(Lvnnr,A)𝖧i(L𝗏nr,A),i0.\displaystyle H^{i}(L_{v_{n}}^{nr},A)\simeq\mathsf{H}^{i}(L^{nr}_{\mathsf{v}},A),\;\;\;i\geq 0.
Proof.

If 𝗏\mathsf{v} is the constant ultraprime v¯\underline{v}, then vn=vv_{n}=v for 𝔉\mathfrak{F}-many nn, and the desired isomorphisms are given by the identity maps; so suppose 𝗏\mathsf{v} is nonconstant. For 𝔉\mathfrak{F}-many nn, the action of the decomposition group GvnG_{v_{n}} at vnv_{n} on AA is unramified and the Frobenius of vnv_{n} acts by Frob𝗏\operatorname{Frob}_{\mathsf{v}}. Let n\ell_{n} be the prime of \mathbb{Q} lying under vnv_{n}; since L/L/\mathbb{Q} is a finite extension and AA is a finite Galois module, for 𝔉\mathfrak{F}-many nn we have n|A|\ell_{n}\nmid|A|. Restricting to these nn, the inflation map induces isomorphisms:

Hi(Gvnt,A)Hi(Lvn,A),Hi(Ivnt,A)Hi(Lvnnr,A),H^{i}(G_{v_{n}}^{t},A)\simeq H^{i}(L_{v_{n}},A),\;\;H^{i}(I_{v_{n}}^{t},A)\simeq H^{i}(L_{v_{n}}^{nr},A),

where GvntG_{v_{n}}^{t} and IvntI_{v_{n}}^{t} denote the tame quotients. The tame Galois group GvntG_{v_{n}}^{t} is identified with the semi direct product:

IvntFrobvn^(n)(1)Frobvn.I_{v_{n}}^{t}\rtimes\langle\operatorname{Frob}_{v_{n}}\rangle\simeq\widehat{\mathbb{Z}}^{(\ell_{n})}(1)\rtimes\langle\operatorname{Frob}_{v_{n}}\rangle.

Since Frobvn\operatorname{Frob}_{v_{n}} and Frob𝗏\operatorname{Frob}_{\mathsf{v}} may act differently on the Tate twist, GvntG_{v_{n}}^{t} and 𝖦𝗏\mathsf{G}_{\mathsf{v}} cannot be compared directly; we wish to show that the cohomologies are nonetheless canonically isomorphic for 𝔉\mathfrak{F}-many nn.

Let G=IFG=I\rtimes\langle F\rangle be an abstract group, where II is abelian and profinite, and F\langle F\rangle denotes the free profinite group on one generator, acting on II by an automorphism. If AA is a [F]\mathbb{Z}[F]-module, then the Galois cohomology groups Hi(G,A)H^{i}(G,A) and Hi(I,A)H^{i}(I,A) depend only on AA and Hom(I,A)\operatorname{Hom}(I,A) as [F]\mathbb{Z}[F]-modules; in particular, the cohomology groups for GG are canonically isomorphic to the cohomology groups for its quotient I/|A|FI/|A|\rtimes\langle F\rangle, and similarly for II and I/|A|I/|A|. Applying this to GvntG_{v_{n}}^{t} and 𝖦𝗏\mathsf{G}_{\mathsf{v}} completes the proof, since Frobvn\operatorname{Frob}_{v_{n}} and Frob𝗏\operatorname{Frob}_{\mathsf{v}} have the same action on the finite Tate twist /|A|(1)\mathbb{Z}/|A|(1) for 𝔉\mathfrak{F}-many nn.

2.4. Patched cohomology

2.4.1.

Let 𝖲𝖬L\mathsf{S}\subset\mathsf{M}_{L} be a finite set of ultraprimes {𝗌1,𝗌2,,𝗌r}\left\{\mathsf{s}_{1},\mathsf{s}_{2},\ldots,\mathsf{s}_{r}\right\}. A representative of 𝖲\mathsf{S} is a sequence of sets SnMLS^{n}\subset M_{L} such that Sn={s1n,,srn}S^{n}=\left\{s_{1}^{n},\cdots,s_{r}^{n}\right\} for some sequences (sin)n(s_{i}^{n})_{n\in\mathbb{N}} representing 𝗌i\mathsf{s}_{i}. If AA is a Gal(L¯/L)\operatorname{Gal}(\overline{L}/L) module, we say AA is unramified outside 𝖲𝖬L\mathsf{S}\subset\mathsf{M}_{L} if it is unramified outside 𝖲ML\mathsf{S}\cap M_{L}.

Definition 2.4.2.

Let AA be a topological Gal(L¯/L)\operatorname{Gal}(\overline{L}/L)-module unramified outside a finite set 𝖲𝖬L\mathsf{S}\subset\mathsf{M}_{L}, represented by a sequence SnMLS^{n}\subset M_{L}. If AA is profinite, then we define the iith unramified-outside-𝖲\mathsf{S} patched cohomology, for all i0,i\geq 0, by:

𝖧i(L𝖲/S,A)=limAA𝒰({Hi(LSn/L,A)})n,\mathsf{H}^{i}(L^{\mathsf{S}}/S,A)=\lim_{\begin{subarray}{c}\longleftarrow\\ A\twoheadrightarrow A^{\prime}\end{subarray}}\mathcal{U}\left(\left\{H^{i}(L^{S^{n}}/L,A^{\prime})\right\}\right)_{n\in\mathbb{N}},

where the inverse limit runs over continuous finite quotients of AA. If AA is ind-finite, then its unramified-outside-𝖲\mathsf{S} patched cohomology is defined as:

𝖧i(L𝖲/L,A)=limAA𝒰({Hi(LSn/L,A)}n),\mathsf{H}^{i}(L^{\mathsf{S}}/L,A)=\lim_{\begin{subarray}{c}\longrightarrow\\ A^{\prime}\subset A\end{subarray}}\mathcal{U}\left(\left\{H^{i}(L^{S^{n}}/L,A^{\prime})\right\}_{n\in\mathbb{N}}\right),

where the direct limit runs over finite submodules. If AA is either profinite or ind-finite, then the totally patched cohomology is defined as

𝖧i(L,A)=lim𝖲𝖬L𝖧i(L𝖲/L,A),\mathsf{H}^{i}(L,A)=\lim_{\begin{subarray}{c}\longrightarrow\\ \mathsf{S}\subset\mathsf{M}_{L}\end{subarray}}\mathsf{H}^{i}(L^{\mathsf{S}}/L,A),

where the direct limit is over finite subsets and the transition maps are induced by the functoriality of the ultraproduct.

Remark 2.4.3.
  1. (1)

    To see that these cohomology groups are well-defined, first note that they are independent of the choice of SnS^{n} since any two representatives of a finite set 𝖲𝖬L\mathsf{S}\subset\mathsf{M}_{L} agree for 𝔉\mathfrak{F}-many nn. Moreover, if AA is both profinite and ind-finite, then it is finite, and it is clear that either definition gives the same cohomology groups.

  2. (2)

    There is a canonical isomorphism 𝖧0(L𝖲/L,A)=H0(L,A)\mathsf{H}^{0}(L^{\mathsf{S}}/L,A)=H^{0}(L,A) for all finite 𝖲𝖬L\mathsf{S}\subset\mathsf{M}_{L} and all profinite or ind-finite AA.

  3. (3)

    The assignments

    A𝖧i(L𝖲/L,A),A𝖧i(L,A)A\mapsto\mathsf{H}^{i}(L^{\mathsf{S}}/L,A),\;\;\;A\mapsto\mathsf{H}^{i}(L,A)

    are functorial in AA. If AA is an RR-module for some ring RR, then each patched cohomology group 𝖧i(L𝖲/L,A),\mathsf{H}^{i}(L^{\mathsf{S}}/L,A), 𝖧i(L,A)\mathsf{H}^{i}(L,A) has a natural RR-module structure.

  4. (4)

    In practice, we will want our profinite Galois modules to be countably profinite, i.e. to have a presentation as a countable inverse limit of finite, discrete topological Galois modules. The significance of this technical hypothesis is that countable inverse limits of finite abelian groups are exact. For example, see [38, Corollary 2.7.6].

  5. (5)

    Suppose AA is ind-finite or countably profinite. If every ultraprime in 𝖲\mathsf{S} is constant, and SMLS\subset M_{L} is the corresponding finite set of places, then 𝖧i(L𝖲/L,A)\mathsf{H}^{i}(L^{\mathsf{S}}/L,A) is canonically isomorphic to Hi(LS/L,A)H^{i}(L^{S}/L,A).

  6. (6)

    Suppose AA is ind-finite or countably profinite. For each ultraprime 𝗏\mathsf{v}, there are natural localization maps

    Res𝗏:𝖧i(L,A)𝖧i(L𝗏,A)\operatorname{Res}_{\mathsf{v}}:\mathsf{H}^{i}(L,A)\to\mathsf{H}^{i}(L_{\mathsf{v}},A)

    deduced from Proposition 2.3.2 (and from [38, Corollary 2.7.6] applied to 𝖦𝗏\mathsf{G}_{\mathsf{v}} in the profinite case).

  7. (7)

    If the Galois action on AA is the restriction of an action of GKG_{K}, where L/KL/K is a Galois extension, then Gal(L/K)\operatorname{Gal}(L/K) acts naturally on 𝖧i(L,A)\mathsf{H}^{i}(L,A), again by functoriality of ultraproducts; this is compatible with the localization maps in the obvious way.

Lemma 2.4.4.

For any finite set of primes SMLS\subset M_{L}, and any finite Galois module AA over LL, the cardinality of Hi(LS/L,A)H^{i}(L^{S}/L,A) is uniformly bounded, with a bound depending only on AA, LL, and |S||S|. In particular, if 𝖲𝖬L\mathsf{S}\subset\mathsf{M}_{L} is finite, then the patched cohomology groups 𝖧i(L𝖲/L,A)\mathsf{H}^{i}(L^{\mathsf{S}}/L,A) are finite for each finite Galois module AA and each i0i\geq 0.

Proof.

The first claim is easily seen from [35, Theorem 4.10]; the second follows by Proposition 2.1.4. ∎

Proposition 2.4.5.

If AA is either countably profinite or ind-finite, then, for all ii, the natural map induces an isomorphism

𝖧i(L𝖲/L,A)ker(𝖧i(L,A)𝗏𝖬L𝖲𝖧i(L𝗏nr,A)).\mathsf{H}^{i}(L^{\mathsf{S}}/L,A)\simeq\ker\left(\mathsf{H}^{i}(L,A)\to\prod_{\mathsf{v}\in\mathsf{M}_{L}-\mathsf{S}}\mathsf{H}^{i}(L_{\mathsf{v}}^{nr},A)\right).
Proof.

It suffices to show that, for all finite sets 𝖳𝖬L𝖲\mathsf{T}\subset\mathsf{M}_{L}-\mathsf{S},

𝖧i(L𝖲/L,A)ker(𝖧i(L𝖲𝖳,A)𝗍𝖳𝖧i(L𝗍nr,A)).\mathsf{H}^{i}(L^{\mathsf{S}}/L,A)\simeq\ker\left(\mathsf{H}^{i}(L^{\mathsf{S}\cup\mathsf{T}},A)\to\prod_{\mathsf{t}\in\mathsf{T}}\mathsf{H}^{i}(L_{\mathsf{t}}^{nr},A)\right).

This holds when AA is finite by Lemma 2.4.4 and Proposition 2.1.6; the general case follows by taking limits. ∎

Lemma 2.4.6.

Let

0ABC00\to A\to B\to C\to 0

be an exact sequence of either countably profinite or ind-finite Galois modules unramified outside 𝖲.\mathsf{S}. Then there is an induced long exact sequence beginning:

0\displaystyle 0 𝖧0(L𝖲/L,A)𝖧0(L𝖲/L,B)𝖧0(L𝖲/L,C)\displaystyle\to\mathsf{H}^{0}(L^{\mathsf{S}}/L,A)\to\mathsf{H}^{0}(L^{\mathsf{S}}/L,B)\to\mathsf{H}^{0}(L^{\mathsf{S}}/L,C)\to
𝖧1(L𝖲/L,A)𝖧1(L𝖲/L,B)\displaystyle\to\mathsf{H}^{1}(L^{\mathsf{S}}/L,A)\to\mathsf{H}^{1}(L^{\mathsf{S}}/L,B)\to\cdots
Proof.

If AA, BB, and CC are all finite, then this follows from Proposition 2.1.6 and Lemma 2.4.4.

Now suppose that AA, BB, and CC are all profinite. Let II, JJ, and KK be directed sets indexing the finite quotients AAiA\twoheadrightarrow A_{i}, BBjB\twoheadrightarrow B_{j}, and CCkC\twoheadrightarrow C_{k}, respectively. We define morphisms of directed sets t:JIt:J\to I and s:JKs:J\to K by

At(j)=im(ABj),Cs(j)=Bj/At(j).A_{t(j)}=\operatorname{im}(A\to B_{j}),\;\;C_{s(j)}=B_{j}/A_{t(j)}.

Because the subgroup and quotient topologies on AA and CC agree with the profinite topologies, the images of tt and ss are cofinal in II and KK, respectively. We therefore have:

𝖧(L𝖲/L,A)=limjJ𝖧(L𝖲/L,At(j)),𝖧(L𝖲/L,C)=limjJ𝖧(L𝖲/L,Cs(j)).\mathsf{H}^{\ast}(L^{\mathsf{S}}/L,A)=\lim_{\begin{subarray}{c}\leftarrow\\ j\in J\end{subarray}}\mathsf{H}^{\ast}(L^{\mathsf{S}}/L,A_{t(j)}),\;\;\mathsf{H}^{\ast}(L^{\mathsf{S}}/L,C)=\lim_{\begin{subarray}{c}\leftarrow\\ j\in J\end{subarray}}\mathsf{H}^{\ast}(L^{\mathsf{S}}/L,C_{s(j)}).

For each jj, we have a long exact sequence associated to the short exact sequence of finite Galois modules

0At(j)BjCs(j)0;0\to A_{t(j)}\to B_{j}\to C_{s(j)}\to 0;

by Lemma 2.4.4, each term in the long exact sequence is finite. Since countable inverse limits of finite abelian groups are exact, taking limits completes the proof. The ind-finite case is completely analogous.∎

2.5. Selmer structures and patched Selmer groups

Definition 2.5.1.

Let AA be a countably profinite or ind-finite p[GL]\mathbb{Z}_{p}[G_{L}]-module. A generalized Selmer structure (,𝖲)(\mathcal{F},\mathsf{S}) for AA consists of:

  • a finite set 𝖲𝖬L\mathsf{S}\subset\mathsf{M}_{L} containing all Archimedian places, all places over pp, and all ramified places for AA;

  • for each 𝗏𝖬L\mathsf{v}\in\mathsf{M}_{L}, a closed p\mathbb{Z}_{p}-submodule (the local condition)

    𝖧1(L𝗏,A)𝖧1(L𝗏,A)\mathsf{H}^{1}_{\mathcal{F}}(L_{\mathsf{v}},A)\subset\mathsf{H}^{1}(L_{\mathsf{v}},A)

    such that

    𝖧1(L𝗏,A)=𝖧unr1(L𝗌,A)ker(𝖧1(L𝗌,A)𝖧1(L𝗌nr,A))\mathsf{H}^{1}_{\mathcal{F}}(L_{\mathsf{v}},A)=\mathsf{H}^{1}_{\operatorname{unr}}(L_{\mathsf{s}},A)\coloneqq\ker\left(\mathsf{H}^{1}(L_{\mathsf{s}},A)\to\mathsf{H}^{1}(L^{nr}_{\mathsf{s}},A)\right)

    for all 𝗏𝖲\mathsf{v}\not\in\mathsf{S}.

If AA is an RR-module for some ring RR and GLG_{L} acts on AA by RR-module automorphisms, a Selmer structure for AA over RR is a Selmer structure such that every local condition is an RR-submodule.

2.5.2.

If BAB\subset A is any closed Galois-stable submodule, then a Selmer structure (,𝖲)(\mathcal{F},\mathsf{S}) for AA induces Selmer structures on BB and A/BA/B defined in the usual way:

𝖧1(L𝗌,B)=ker(𝖧1(L𝗌,B)𝖧1(L𝗌,A)𝖧1(L𝗌,A)),\displaystyle\mathsf{H}^{1}_{\mathcal{F}}(L_{\mathsf{s}},B)=\ker\left(\mathsf{H}^{1}(L_{\mathsf{s}},B)\to\frac{\mathsf{H}^{1}(L_{\mathsf{s}},A)}{\mathsf{H}_{\mathcal{F}}^{1}(L_{\mathsf{s}},A)}\right),
𝖧1(L𝗌,A/B)=im(𝖧1(L𝗌,A)𝖧1(L𝗌,A/B)).\displaystyle\mathsf{H}^{1}_{\mathcal{F}}(L_{\mathsf{s}},A/B)=\operatorname{im}\left(\mathsf{H}^{1}_{\mathcal{F}}(L_{\mathsf{s}},A)\to\mathsf{H}^{1}(L_{\mathsf{s}},A/B)\right).

2.5.3.

To a generalized Selmer structure we associate the patched Selmer group, defined by the exact sequence:

(10) 0Sel(A)𝖧1(L𝖲/L,A)𝗌𝖲𝖧1(L𝗌,A)𝖧1(L𝗌,A),0\to\operatorname{Sel}_{\mathcal{F}}(A)\to\mathsf{H}^{1}(L^{\mathsf{S}}/L,A)\to\prod_{\mathsf{s}\in\mathsf{S}}\frac{\mathsf{H}^{1}(L_{\mathsf{s}},A)}{\mathsf{H}^{1}_{\mathcal{F}}(L_{\mathsf{s}},A)},

or equivalently (by Proposition 2.4.5):

(11) 0Sel(A)𝖧1(L,A)𝗌𝖲𝖧1(L𝗌,A)𝖧1(L𝗌,A)×𝗌𝖲𝖧1(L𝗌nr,A).0\to\operatorname{Sel}_{\mathcal{F}}(A)\to\mathsf{H}^{1}(L,A)\to\prod_{\mathsf{s}\in\mathsf{S}}\frac{\mathsf{H}^{1}(L_{\mathsf{s}},A)}{\mathsf{H}^{1}_{\mathcal{F}}(L_{\mathsf{s}},A)}\times\prod_{\mathsf{s}\not\in\mathsf{S}}\mathsf{H}^{1}(L_{\mathsf{s}}^{nr},A).

(Note that the Selmer group attached to a Selmer structure does not depend on the choice of set 𝖲\mathsf{S} but only on the local conditions; we will therefore sometimes omit 𝖲\mathsf{S} from the notation when there is no risk of confusion.)

2.5.4.

If BAB\subset A is Galois-stable, and B,A/BB,A/B are equipped with the induced Selmer structures, then by definition there are natural maps:

Sel(B)Sel(A)Sel(A/B).\operatorname{Sel}_{\mathcal{F}}(B)\to\operatorname{Sel}_{\mathcal{F}}(A)\to\operatorname{Sel}_{\mathcal{F}}(A/B).
Proposition 2.5.5.

Let (,𝖲)(\mathcal{F},\mathsf{S}) be a generalized Selmer structure for AA. If AA is countably profinite and each continuous finite quotient AAA\twoheadrightarrow A^{\prime} is equipped with the Selmer structure induced by \mathcal{F}, then:

limSel(A)Sel(A).\lim_{\longleftarrow}\operatorname{Sel}_{\mathcal{F}}(A^{\prime})\simeq\operatorname{Sel}_{\mathcal{F}}(A).

If instead AA is ind-finite and each finite submodule AAA^{\prime}\subset A is given its induced Selmer structure, then:

limSel(A)Sel(A).\lim_{\longrightarrow}\operatorname{Sel}_{\mathcal{F}}(A^{\prime})\simeq\operatorname{Sel}_{\mathcal{F}}(A).
Proof.

We show the countably profinite case; the ind-finite case is similar. By definition, Sel(A)\operatorname{Sel}_{\mathcal{F}}(A) is the kernel of

lim𝖧1(L𝖲/L,A)𝗌𝖲𝖧1(L𝗌,A)𝖧1(L𝗌,A),\lim_{\longleftarrow}\mathsf{H}^{1}(L^{\mathsf{S}}/L,A^{\prime})\to\prod_{\mathsf{s}\in\mathsf{S}}\frac{\mathsf{H}^{1}(L_{\mathsf{s}},A)}{\mathsf{H}^{1}_{\mathcal{F}}(L_{\mathsf{s}},A)},

whereas

limSel(A)\displaystyle\lim_{\longleftarrow}\operatorname{Sel}_{\mathcal{F}}(A^{\prime}) =limker(𝖧1(L𝖲/L,A)𝗌𝖲𝖧1(L𝗌,A)𝖧1(L𝗌,A))=ker(𝖧1(L𝖲/L,A)lim𝖧1(L𝗌,A)𝖧1(L𝗌,A)).\displaystyle=\lim_{\longleftarrow}\ker\left(\mathsf{H}^{1}(L^{\mathsf{S}}/L,A^{\prime})\to\prod_{\mathsf{s}\in\mathsf{S}}\frac{\mathsf{H}^{1}(L_{\mathsf{s}},A^{\prime})}{\mathsf{H}^{1}_{\mathcal{F}}(L_{\mathsf{s}},A^{\prime})}\right)=\ker\left(\mathsf{H}^{1}(L^{\mathsf{S}}/L,A)\to\lim_{\longleftarrow}\frac{\mathsf{H}^{1}(L_{\mathsf{s}},A^{\prime})}{\mathsf{H}^{1}_{\mathcal{F}}(L_{\mathsf{s}},A^{\prime})}\right).

Since 𝖧1(L𝗌,A)\mathsf{H}^{1}_{\mathcal{F}}(L_{\mathsf{s}},A) is closed, we have

lim𝖧1(L𝗌,A)=𝖧1(L𝗌,A),\lim_{\longleftarrow}\mathsf{H}^{1}_{\mathcal{F}}(L_{\mathsf{s}},A^{\prime})=\mathsf{H}^{1}_{\mathcal{F}}(L_{\mathsf{s}},A),

which implies the result. ∎

2.5.6.

Given two Selmer structures (,𝖲)(\mathcal{F},\mathsf{S}) and (𝒢,𝖳)(\mathcal{G},\mathsf{T}) for AA, we may define Selmer structures (+𝒢,𝖲𝖳)(\mathcal{F}+\mathcal{G},\mathsf{S}\cup\mathsf{T}) and (𝒢,𝖲𝖳)(\mathcal{F}\cap\mathcal{G},\mathsf{S}\cup\mathsf{T}) by the local conditions:

(12) 𝖧+𝒢1(L𝗏,A)=𝖧1(L𝗏,A)+𝖧𝒢1(L𝗏,A),𝖧𝒢1(L𝗏,A)=𝖧1(L𝗏,A)𝖧𝒢1(L𝗏,A).\mathsf{H}^{1}_{\mathcal{F}+\mathcal{G}}(L_{\mathsf{v}},A)=\mathsf{H}^{1}_{\mathcal{F}}(L_{\mathsf{v}},A)+\mathsf{H}^{1}_{\mathcal{G}}(L_{\mathsf{v}},A),\;\;\;\mathsf{H}^{1}_{\mathcal{F}\cap\mathcal{G}}(L_{\mathsf{v}},A)=\mathsf{H}^{1}_{\mathcal{F}}(L_{\mathsf{v}},A)\cap\mathsf{H}^{1}_{\mathcal{G}}(L_{\mathsf{v}},A).

2.6. Dual Selmer groups

2.6.1.

Fix an ultraprime 𝗏𝖬L\mathsf{v}\in\mathsf{M}_{L}. If AA is a countably profinite p\mathbb{Z}_{p}-Galois module and AA^{\ast} denotes the Cartier dual, then the cup product induces pairings:

(13) ,𝗏:𝖧i(L𝗏,A)×𝖧2i(L𝗏,A)p/p,i=0,1,2.\langle\cdot,\cdot\rangle_{\mathsf{v}}:\mathsf{H}^{i}(L_{\mathsf{v}},A)\times\mathsf{H}^{2-i}(L_{\mathsf{v}},A^{\ast})\to\mathbb{Q}_{p}/\mathbb{Z}_{p},\;\;\;i=0,1,2.
Proposition 2.6.2.

The pairing (13) is perfect if 𝗏\mathsf{v} is non-Archimedean. Moreover, the induced pairing

𝖧1(L𝖲/L,A)×𝖧1(L𝖲/L,A)𝗌𝖲𝖧1(L𝗌,A)×𝖧1(L𝗌,A)Σ,𝗌p/p\mathsf{H}^{1}(L^{\mathsf{S}}/L,A)\times\mathsf{H}^{1}(L^{\mathsf{S}}/L,A^{\ast})\to\prod_{\mathsf{s}\in\mathsf{S}}\mathsf{H}^{1}(L_{\mathsf{s}},A)\times\mathsf{H}^{1}(L_{\mathsf{s}},A^{\ast})\xrightarrow{\Sigma\langle\cdot,\cdot\rangle_{\mathsf{s}}}\mathbb{Q}_{p}/\mathbb{Z}_{p}

is identically zero.

Proof.

For the perfectness of (13), the usual proof of Poitou-Tate duality applies equally well to 𝖦𝗏\mathsf{G}_{\mathsf{v}}; alternatively, one may take limits using Proposition 2.3.2. The second claim is clear when AA is finite by functoriality of the ultraproduct, and the general case follows by taking limits. ∎

2.6.3.

Suppose that AA is either countably profinite or countably ind-finite, i.e. the Pontryagin dual of a countably profinite Galois module. If (,𝖲)(\mathcal{F},\mathsf{S}) is any Selmer structure for AA, then we define the dual Selmer structure (,𝖲)(\mathcal{F}^{\ast},\mathsf{S}) for AA^{\ast} by:

𝖧1(L𝗌,A)=𝖧1(L𝗌,A).\mathsf{H}^{1}_{\mathcal{F}^{\ast}}(L_{\mathsf{s}},A^{\ast})=\mathsf{H}^{1}_{\mathcal{F}}(L_{\mathsf{s}},A)^{\perp}.

Here \perp denotes the orthogonal complement under either the pairing of (2.6.1), or the usual modified Tate pairing of [14, Theorem 2.17] at Archimedian places. We observe that the dual Selmer structure to (,𝖲)(\mathcal{F}^{\ast},\mathsf{S}) is again (,𝖲)(\mathcal{F},\mathsf{S}). WhenAA is finite, the dual Selmer groups are related by the Greenberg-Wiles formula:

Proposition 2.6.4.

Let (,𝖲)(\mathcal{F},\mathsf{S}) be a Selmer structure for a finite p[GL]\mathbb{Z}_{p}[G_{L}]-module AA. We have:

#Sel(A)#Sel(A)=#𝖧0(L𝖲/L,A)#𝖧0(L𝖲/L,A)𝗌𝖲#𝖧1(L𝗌,A)#𝖧0(L𝗌,A).\frac{\#\operatorname{Sel}_{\mathcal{F}}(A)}{\#\operatorname{Sel}_{\mathcal{F}^{\ast}}(A^{\ast})}=\frac{\#\mathsf{H}^{0}(L^{\mathsf{S}}/L,A)}{\#\mathsf{H}^{0}(L^{\mathsf{S}}/L,A^{\ast})}\prod_{\mathsf{s}\in\mathsf{S}}\frac{\#\mathsf{H}^{1}_{\mathcal{F}}(L_{\mathsf{s}},A)}{\#\mathsf{H}^{0}(L_{\mathsf{s}},A)}.
Proof.

This follows from [14, Theorem 2.19] by the exactness of ultraproducts and Proposition 2.1.4(1). ∎

2.7. Selmer groups over discrete valuation rings

2.7.1.

Let RR be a discrete valuation ring with uniformizer π\pi which is a finite, flat extension of p\mathbb{Z}_{p}, and suppose that A=TA=T is a free RR-module of finite rank, with GLG_{L} action through RR-module automorphisms. In particular, TT is countably profinite. Suppose 𝖲𝖬L\mathsf{S}\subset\mathsf{M}_{L} is a finite set containing all Archimedian places and all places over pp, such that TT is unramified outside 𝖲\mathsf{S}. If T=HomR(T,R(1))T^{\dagger}=\operatorname{Hom}_{R}(T,R(1)) is the dual, then the cup product induces a local Tate pairing

(14) ,𝗏:𝖧1(L𝗏,T)×𝖧1(L𝗏,T)R.\langle\cdot,\cdot\rangle_{\mathsf{v}}:\mathsf{H}^{1}(L_{\mathsf{v}},T)\times\mathsf{H}^{1}(L_{\mathsf{v}},T^{\dagger})\to R.
Proposition 2.7.2.

The kernels on the left and right of (14) are the RR-torsion submodules; moreover, the induced pairing

𝖧1(L𝖲/L,T)×𝖧1(L𝖲/L,T)𝗏𝖲𝖧1(L𝗏,T)×𝖧1(L𝗏,T),𝗏R\mathsf{H}^{1}(L^{\mathsf{S}}/L,T)\times\mathsf{H}^{1}(L^{\mathsf{S}}/L,T^{\dagger})\to\prod_{\mathsf{v}\in\mathsf{S}}\mathsf{H}^{1}(L_{\mathsf{v}},T)\times\mathsf{H}^{1}(L_{\mathsf{v}},T^{\dagger})\xrightarrow{\sum\langle\cdot,\cdot\rangle_{\mathsf{v}}}R

is identically zero.

Proof.

This follows from Proposition 2.6.2. ∎

Given a Selmer structure (,𝖲)(\mathcal{F},\mathsf{S}) for TT over RR, taking the orthogonal complement of each local condition under (14) yields a Selmer structure (,𝖲)(\mathcal{F}^{\dagger},\mathsf{S}) for TT^{\dagger}. Note that \mathcal{F}^{\dagger\dagger}\neq\mathcal{F} in general, but we always have =\mathcal{F}^{\dagger\dagger\dagger}=\mathcal{F}^{\dagger}.

Proposition 2.7.3.

Let (,𝖲)(\mathcal{F},\mathsf{S}) be a Selmer structure such that

𝖧1(L𝗏,T)𝖧1(L𝗏,T)\frac{\mathsf{H}^{1}(L_{\mathsf{v}},T)}{\mathsf{H}^{1}_{\mathcal{F}}(L_{\mathsf{v}},T)}

is torsion-free. Then, for all jj and all 𝗏𝖬L\mathsf{v}\in\mathsf{M}_{L},

𝖧1(L𝗏,T[πj])=𝖧1(L𝗏,T/πj)\mathsf{H}^{1}_{\mathcal{F}^{\ast}}(L_{\mathsf{v}},T^{\ast}[\pi^{j}])=\mathsf{H}^{1}_{\mathcal{F}^{\dagger}}(L_{\mathsf{v}},T^{\dagger}/\pi^{j})

under an identification T[πj]T/πjT^{\ast}[\pi^{j}]\simeq T^{\dagger}/\pi^{j}, and in particular

Sel(T[πj])=Sel(T/πj).\operatorname{Sel}_{\mathcal{F}^{\ast}}(T^{\ast}[\pi^{j}])=\operatorname{Sel}_{\mathcal{F}^{\dagger}}(T^{\dagger}/\pi^{j}).
Proof.

Although this fact is presumably standard, we give a proof for lack of a reference. For ease of notation, we abbreviate 𝖧i(T)=𝖧i(L𝗏,T)\mathsf{H}^{i}(T^{\dagger})=\mathsf{H}^{i}(L_{\mathsf{v}},T^{\dagger}), etc. and Rj=R/πjR_{j}=R/\pi^{j}. The choice of uniformizer induces an identification TR(R[1/π]/R)TT^{\dagger}\otimes_{R}(R[1/\pi]/R)\simeq T^{\ast} and an embedding T/πjTT^{\dagger}/\pi^{j}\hookrightarrow T^{\ast}; let 𝖧1(T/πj)\mathsf{H}^{1}_{\mathcal{F}^{\ast}}(T^{\dagger}/\pi^{j}) be the induced local condition from this embedding. Consider the following commutative diagram with exact rows:

0{0}𝖧0(T)/div{\mathsf{H}^{0}(T^{\ast})_{/\operatorname{div}}}𝖧1(T){\mathsf{H}^{1}_{\mathcal{F}^{\dagger}}(T^{\dagger})}Hom(𝖧1(T),R){\operatorname{Hom}(\mathsf{H}^{1}(T),R)}Hom(𝖧1(T),R){\operatorname{Hom}(\mathsf{H}^{1}_{\mathcal{F}}(T),R)}0{0}0{0}𝖧0(T)/πj{\mathsf{H}^{0}(T^{\ast})/\pi^{j}}𝖧1(T/πj){\mathsf{H}^{1}_{\mathcal{F}^{\ast}}(T^{\dagger}/\pi^{j})}Hom(𝖧1(T),Rj){\operatorname{Hom}(\mathsf{H}^{1}(T),R_{j})}Hom(𝖧1(T),Rj){\operatorname{Hom}(\mathsf{H}^{1}_{\mathcal{F}}(T),R_{j})}0{0}α\scriptstyle{\alpha}β\scriptstyle{\beta}γ\scriptstyle{\gamma}δ\scriptstyle{\delta}

Here, the first horizontal map on each row is the Kummer map, and the subscript /div/\operatorname{div} refers to the quotient by the maximal divisible submodule. By the hypothesis on 𝖧1(T)\mathsf{H}^{1}_{\mathcal{F}}(T), the maps cokerγcokerδ\operatorname{coker}\gamma\to\operatorname{coker}\delta and kerγkerδ\ker\gamma\to\ker\delta are injective and surjective, respectively. Also, α\alpha is clearly surjective. Breaking the diagram into two and applying the snake lemma, it follows that β\beta is surjective. ∎

Proposition 2.7.4.

Let (,𝖲)(\mathcal{F},\mathsf{S}) be a Selmer structure for TT over RR. Then:

rkRSel(T)rkRSel(T)=rkRH0(L,T)rkRH0(L,T)+𝗌𝖲(rkR𝖧1(L𝗌,T)rkR𝖧0(L𝗌,T)).\operatorname{rk}_{R}\operatorname{Sel}_{\mathcal{F}}(T)-\operatorname{rk}_{R}\operatorname{Sel}_{\mathcal{F}^{\dagger}}(T^{\dagger})=\operatorname{rk}_{R}H^{0}(L,T)-\operatorname{rk}_{R}H^{0}(L,T^{\dagger})+\sum_{\mathsf{s}\in\mathsf{S}}\left(\operatorname{rk}_{R}\mathsf{H}^{1}_{\mathcal{F}}(L_{\mathsf{s}},T)-\operatorname{rk}_{R}\mathsf{H}^{0}(L_{\mathsf{s}},T)\right).
Proof.

Without loss of generality, we may assume that

𝖧1(L𝗏,T)𝖧1(L𝗏,T)\frac{\mathsf{H}^{1}(L_{\mathsf{v}},T)}{\mathsf{H}^{1}_{\mathcal{F}}(L_{\mathsf{v}},T)}

is torsion-free for all 𝗏\mathsf{v}. By Propositions 2.6.4 and 2.7.3, we then have for each jj:

lgSel(T/πj)lgSel(T/πj)=lg𝖧0(L,T/πj)lg𝖧0(L,T/πj)+𝗏𝖲(lg𝖧1(L𝗏,T/πj)lg𝖧0(L𝗏,T/πj)).\begin{split}\lg\operatorname{Sel}_{\mathcal{F}}(T/\pi^{j})-\lg\operatorname{Sel}_{\mathcal{F}^{\dagger}}(T^{\dagger}/\pi^{j})=\lg\mathsf{H}^{0}(L,T/\pi^{j})-\lg\mathsf{H}^{0}(L,T^{\dagger}/\pi^{j})\\ +\sum_{\mathsf{v}\in\mathsf{S}}\left(\lg\mathsf{H}^{1}_{\mathcal{F}}(L_{\mathsf{v}},T/\pi^{j})-\lg\mathsf{H}^{0}(L_{\mathsf{v}},T/\pi^{j})\right).\end{split}

Since Sel(T)\operatorname{Sel}_{\mathcal{F}}(T) is a finitely generated RR-module, it follows from [33, Lemma 3.7.1] that

lgSel(T/πj)=(rkRSel(T))lgR/πj+O(1)\lg\operatorname{Sel}_{\mathcal{F}}(T/\pi^{j})=(\operatorname{rk}_{R}\operatorname{Sel}_{\mathcal{F}}(T))\cdot\lg R/\pi^{j}+O(1)

as jj varies, and likewise for Sel(T)\operatorname{Sel}_{\mathcal{F}^{\dagger}}(T^{\dagger}) and each term on the right-hand side; the proposition follows. ∎

3. Bipartite Euler systems

3.1. Admissible primes

3.1.1.

Let ff be a modular form of weight two, trivial character, and level NN, and let 𝒪f\wp\subset\mathcal{O}_{f} be a prime ideal of the ring of integers of its field of coefficients. We assume the rational prime pp lying under \wp is odd, and write 𝒪\mathcal{O} for the completion of 𝒪f\mathcal{O}_{f} at \wp. Fix a Galois-stable 𝒪\mathcal{O}-lattice TfT_{f} in the \wp-adic Galois representation associated to ff, and let T¯f\overline{T}_{f} be the residual representation Tf/T_{f}/\wp; we also write WfW_{f} for Tfp/pT_{f}\otimes\mathbb{Q}_{p}/\mathbb{Z}_{p}. Also let K/K/\mathbb{Q} be an imaginary quadratic field. We assume throughout this section that T¯f\overline{T}_{f} is absolutely irreducible as a GKG_{K}-module. We will sometimes use the condition:

(sclr) The image of the GK action on T¯f contains a nonzero scalar.\text{The image of the $G_{K}$ action on }\overline{T}_{f}\text{ contains a nonzero scalar.}
Definition 3.1.2.

A nonconstant ultraprime 𝗊𝖬\mathsf{q}\in\mathsf{M}_{\mathbb{Q}} is said to be admissible with sign ϵ𝗊=±1\epsilon_{\mathsf{q}}=\pm 1 for ff if Frob𝗊\operatorname{Frob}_{\mathsf{q}} has nonzero image in Gal(K/)\operatorname{Gal}(K/\mathbb{Q}), χ(Frob𝗊)1(modp)\chi(\operatorname{Frob}_{\mathsf{q}})\not\equiv 1\pmod{p}, and there is a rank-one direct summand Fil𝗊,ϵ𝗊+TfTf\operatorname{Fil}^{+}_{\mathsf{q},\epsilon_{\mathsf{q}}}T_{f}\subset T_{f} on which Frob𝗊\operatorname{Frob}_{\mathsf{q}} acts as χ(Frob𝗊)ϵ𝗊\chi(\operatorname{Frob}_{\mathsf{q}})\epsilon_{\mathsf{q}}. (Equivalently, χ(Frob𝗊)1(modp)\chi(\operatorname{Frob}_{\mathsf{q}})\not\equiv 1\pmod{p} and TfT_{f} admits a basis of eigenvectors for Frob𝗊\operatorname{Frob}_{\mathsf{q}} with eigenvalues ϵ𝗊\epsilon_{\mathsf{q}} and χ(Frob𝗊)ϵ𝗊\chi(\operatorname{Frob}_{\mathsf{q}})\epsilon_{\mathsf{q}}.)

For example, if Frob𝗊G\operatorname{Frob}_{\mathsf{q}}\in G_{\mathbb{Q}} is a complex conjugation, then 𝗊\mathsf{q} is admissible with either choice of ϵ𝗊\epsilon_{\mathsf{q}}. We abusively write 𝗊\mathsf{q} for the unique ultraprime in 𝖬K\mathsf{M}_{K} lying over 𝗊𝖬\mathsf{q}\in\mathsf{M}_{\mathbb{Q}}, whose Frobenius is Frob𝗊2.\operatorname{Frob}_{\mathsf{q}}^{2}.

Definition 3.1.3.

If 𝗊\mathsf{q} is admissible with sign ϵ𝗊\epsilon_{\mathsf{q}} for ff, then we define the ordinary local condition (with sign ϵ𝗊\epsilon_{\mathsf{q}}) as:

𝖧ord,ϵ𝗊1(K𝗊,Tf)=im(𝖧1(K𝗊,Fil𝗊,ϵ𝗊+Tf)𝖧1(K𝗊,Tf)).\mathsf{H}^{1}_{\operatorname{ord},\epsilon_{\mathsf{q}}}(K_{\mathsf{q}},T_{f})=\operatorname{im}\left(\mathsf{H}^{1}(K_{\mathsf{q}},\operatorname{Fil}^{+}_{\mathsf{q},\epsilon_{\mathsf{q}}}T_{f})\to\mathsf{H}^{1}(K_{\mathsf{q}},T_{f})\right).

The subscript ϵq\epsilon_{q} will often be omitted (from this and future notation) when there is no risk of confusion.

3.1.4.

Note that the ordinary local condition is self-annihilating under the local Tate pairing

𝖧1(K𝗊,Tf)×𝖧1(K𝗊,Tf)𝒪\mathsf{H}^{1}(K_{\mathsf{q}},T_{f})\times\mathsf{H}^{1}(K_{\mathsf{q}},T_{f})\to\mathcal{O}

induced by (14) and the Weil pairing.

3.1.5.

For any finite set 𝖲𝖬K\mathsf{S}\subset\mathsf{M}_{K} such that TfT_{f} is unramified outside 𝖲\mathsf{S}, and any admissible 𝗊𝖲\mathsf{q}\not\in\mathsf{S} with sign ϵ𝗊\epsilon_{\mathsf{q}}, define a localization map

(15) loc𝗊,ϵ𝗊:𝖧1(K𝖲/K,Tf)𝖧unr1(K𝗊,Tf)𝖧unr1(K𝗊,Tf)𝖧unr1(K𝗊,Tf)𝖧ord,ϵ𝗊1(K𝗊,Tf)𝒪.\operatorname{loc}_{\mathsf{q},\epsilon_{\mathsf{q}}}:\mathsf{H}^{1}(K^{\mathsf{S}}/K,T_{f})\to\mathsf{H}^{1}_{\operatorname{unr}}(K_{\mathsf{q}},T_{f})\to\frac{\mathsf{H}^{1}_{\operatorname{unr}}(K_{\mathsf{q}},T_{f})}{\mathsf{H}^{1}_{\operatorname{unr}}(K_{\mathsf{q}},T_{f})\cap\mathsf{H}^{1}_{\operatorname{ord},\epsilon_{\mathsf{q}}}(K_{\mathsf{q}},T_{f})}\approx\mathcal{O}.

Define as well a residue map

(16) 𝗊,ϵ𝗊:𝖧1(K,Tf)𝖧1(K𝗊,Tf)𝖧ord,ϵ𝗊1(K𝗊,Tf)𝖧ord,ϵ𝗊1(K𝗊,Tf)𝖧unr1(K𝗊,Tf)𝖧ord,ϵ𝗊1(K𝗊,Tf)𝒪,\partial_{\mathsf{q},\epsilon_{\mathsf{q}}}:\mathsf{H}^{1}(K,T_{f})\to\mathsf{H}^{1}(K_{\mathsf{q}},T_{f})\to\mathsf{H}^{1}_{\operatorname{ord},\epsilon_{\mathsf{q}}}(K_{\mathsf{q}},T_{f})\to\frac{\mathsf{H}^{1}_{\operatorname{ord},\epsilon_{\mathsf{q}}}(K_{\mathsf{q}},T_{f})}{\mathsf{H}^{1}_{\operatorname{unr}}(K_{\mathsf{q}},T_{f})\cap\mathsf{H}^{1}_{\operatorname{ord},\epsilon_{\mathsf{q}}}(K_{\mathsf{q}},T_{f})}\approx\mathcal{O},

where the second map is given by the projection Tf(Frob𝗊ϵ𝗊)TfFil𝗊,ϵ𝗊+TfT_{f}\twoheadrightarrow(\operatorname{Frob}_{\mathsf{q}}-\epsilon_{\mathsf{q}})T_{f}\simeq\operatorname{Fil}^{+}_{\mathsf{q},\epsilon_{\mathsf{q}}}T_{f}. The maps loc𝗊,ϵq\operatorname{loc}_{\mathsf{q},\epsilon_{q}} and 𝗊,ϵ𝗊\partial_{\mathsf{q},\epsilon_{\mathsf{q}}} may be extended in the obvious way to the patched cohomology for WfW_{f} and all Tf/jT_{f}/\wp^{j}.

3.2. Euler systems for anticyclotomic twists

3.2.1.

Let RR be a complete flat Noetherian local 𝒪\mathcal{O}-algebra with finite residue field, equipped with an anticyclotomic character φ:GKR×\varphi:G_{K}\to R^{\times} which is trivial modulo the maximal ideal of RR. We write TφT_{\varphi} for the anticyclotomic twist Tf𝒪R(φ)T_{f}\otimes_{\mathcal{O}}R(\varphi), which is a countably profinite Galois module. If 𝗊\mathsf{q} is admissible with sign ϵ𝗊\epsilon_{\mathsf{q}}, then φ(Frob𝗊2)=1,\varphi(\operatorname{Frob}_{\mathsf{q}}^{2})=1, so

𝖧1(K𝗊,Tφ)=𝖧1(K𝗊,Tf)𝒪R.\mathsf{H}^{1}(K_{\mathsf{q}},T_{\varphi})=\mathsf{H}^{1}(K_{\mathsf{q}},T_{f})\otimes_{\mathcal{O}}R.

We extend the ordinary local condition of the previous subsection by linearity to define 𝖧ord,ϵ𝗊1(K𝗊,Tφ)\mathsf{H}^{1}_{\operatorname{ord},\epsilon_{\mathsf{q}}}(K_{\mathsf{q}},T_{\varphi}), and likewise the maps loc𝗊,ϵ𝗊,𝗊,ϵ𝗊\operatorname{loc}_{\mathsf{q},\epsilon_{\mathsf{q}}},\partial_{\mathsf{q},\epsilon_{\mathsf{q}}}.

3.2.2.

Suppose given a finite set 𝖲𝖬K\mathsf{S}\subset\mathsf{M}_{K} and a generalized Selmer structure (,𝖲)(\mathcal{F},\mathsf{S}) for TφT_{\varphi}. Let 𝖭=𝖭𝖲\mathsf{N}=\mathsf{N}_{\mathsf{S}} be the set of pairs {𝖰,ϵ𝖰}\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\} where 𝖰𝖬K𝖲\mathsf{Q}\subset\mathsf{M}_{K}-\mathsf{S} is a finite set of ultraprimes and ϵ𝖰:𝖰{±1}\epsilon_{\mathsf{Q}}:\mathsf{Q}\to\left\{\pm 1\right\} is a function such that 𝗊\mathsf{q} is admissible with sign ϵ𝖰(𝗊)\epsilon_{\mathsf{Q}}(\mathsf{q}) for all 𝗊𝖰\mathsf{q}\in\mathsf{Q}. (We will drop the subscript 𝖲\mathsf{S} when it is clear from context, or when 𝖲\mathsf{S} contains only constant ultraprimes.) Given a pair {𝖰,ϵ𝖰}𝖭\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}, define a generalized Selmer structure ((𝖰,ϵ𝖰),𝖲𝖰)(\mathcal{F}(\mathsf{Q},\epsilon_{\mathsf{Q}}),\mathsf{S}\cup\mathsf{Q}) for TφT_{\varphi} by the local conditions:

(17) 𝖧(𝖰,ϵ𝖰)1(K𝗏,Tφ)={𝖧1(K𝗏,Tφ),𝗏𝖰𝖧ord,ϵ𝖰(𝗊)1(K𝗊,Tφ),𝗏=𝗊𝖰.\mathsf{H}^{1}_{\mathcal{F}(\mathsf{Q},\epsilon_{\mathsf{Q}})}(K_{\mathsf{v}},T_{\varphi})=\begin{cases}\mathsf{H}^{1}_{\mathcal{F}}(K_{\mathsf{v}},T_{\varphi}),&\mathsf{v}\not\in\mathsf{Q}\\ \mathsf{H}^{1}_{\operatorname{ord},\epsilon_{\mathsf{Q}}(\mathsf{q})}(K_{\mathsf{q}},T_{\varphi}),&\mathsf{v}=\mathsf{q}\in\mathsf{Q}.\end{cases}

For δ/2\delta\in\mathbb{Z}/2\mathbb{Z}, let 𝖭δ𝖭\mathsf{N}^{\delta}\subset\mathsf{N} be the collection of pairs {𝖰,ϵ𝖰}𝖭\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N} such that |𝖰|δ(mod2).|\mathsf{Q}|\equiv\delta\pmod{2}. Also, given two pairs {𝖰,ϵ𝖰}𝖭δ\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}^{\delta} and {𝖰,ϵ𝖰}𝖭δ\left\{\mathsf{Q}^{\prime},\epsilon_{\mathsf{Q}^{\prime}}\right\}\in\mathsf{N}^{\delta^{\prime}} such that 𝖰𝖰=,\mathsf{Q}\cap\mathsf{Q}^{\prime}=\emptyset, write

{𝖰𝖰,ϵ𝖰𝖰}𝖭δ+δ\left\{\mathsf{Q}\mathsf{Q}^{\prime},\epsilon_{\mathsf{Q}\mathsf{Q}^{\prime}}\right\}\in\mathsf{N}^{\delta+\delta^{\prime}}

for the pair formed in the obvious way from 𝖰𝖰\mathsf{Q}\cup\mathsf{Q}^{\prime} and the sign functions ϵ𝖰,ϵ𝖰\epsilon_{\mathsf{Q}},\epsilon_{\mathsf{Q}^{\prime}}. The pair {,}𝖭\left\{\emptyset,\emptyset\right\}\in\mathsf{N} will be abbreviated as 1.

Definition 3.2.3.

A bipartite system (κ,λ)(\kappa,\lambda) for (Tφ,,𝖲)(T_{\varphi},\mathcal{F},\mathsf{S}) of parity δ/2\delta\in\mathbb{Z}/2\mathbb{Z} consists of the following data:

  1. (1)

    for each pair {𝖰,ϵ𝖰}𝖭δ\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}^{\delta}, a principal submodule

    (κ(𝖰,ϵ𝖰))Sel(𝖰)(Tφ);(\kappa(\mathsf{Q},\epsilon_{\mathsf{Q}}))\subset\operatorname{Sel}_{\mathcal{F}(\mathsf{Q})}(T_{\varphi});
  2. (2)

    for each pair {𝖰,ϵ𝖰}𝖭δ+1\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}^{\delta+1}, a principal ideal

    (λ(𝖰,ϵ𝖰))R.(\lambda(\mathsf{Q},\epsilon_{\mathsf{Q}}))\subset R.

A bipartite Euler system is a bipartite system satisfying the “reciprocity laws”:

  1. (1)

    For each {𝖰𝗊,ϵ𝖰𝗊}𝖭δ+1\left\{\mathsf{Q}\mathsf{q},\epsilon_{\mathsf{Q}\mathsf{q}}\right\}\in\mathsf{N}^{\delta+1},

    loc𝗊((κ(𝖰))=(λ(𝖰𝗊))R.\operatorname{loc}_{\mathsf{q}}((\kappa(\mathsf{Q}))=(\lambda(\mathsf{Q}\mathsf{q}))\subset R.
  2. (2)

    For each {𝖰𝗊,ϵ𝖰𝗊}𝖭δ\left\{\mathsf{Q}\mathsf{q},\epsilon_{\mathsf{Q}\mathsf{q}}\right\}\in\mathsf{N}^{\delta},

    𝗊((κ(𝖰𝗊))=(λ(𝖰))R.\partial_{\mathsf{q}}((\kappa(\mathsf{Q}\mathsf{q}))=(\lambda(\mathsf{Q}))\subset R.

We say (κ,λ)(\kappa,\lambda) is nontrivial if there exists some {𝖰,ϵ𝖰}𝖭\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N} such that either λ(𝖰,ϵ𝖰)0\lambda(\mathsf{Q},\epsilon_{\mathsf{Q}})\neq 0 or κ(𝖰,ϵ𝖰)0\kappa(\mathsf{Q},\epsilon_{\mathsf{Q}})\neq 0 depending on the parity of |𝖰|+δ|\mathsf{Q}|+\delta.

3.3. Euler systems over discrete valuation rings

3.3.1.

Suppose that RR is a discrete valuation ring with uniformizer π\pi, and let Wφ=Tφp/pW_{\varphi}=T_{\varphi}\otimes\mathbb{Q}_{p}/\mathbb{Z}_{p}. Exactly as in [24], there is a perfect pairing Tφ×TφR(1)T_{\varphi}\times T_{\varphi}\to R(1), GKG_{K}-equivariant up to a twist, which induces local pairings:

𝖧1(K𝗏,Tφ)×𝖧1(K𝗏¯,Wφ)\displaystyle\mathsf{H}^{1}(K_{\mathsf{v}},T_{\varphi})\times\mathsf{H}^{1}(K_{\overline{\mathsf{v}}},W_{\varphi}) Rp/p,\displaystyle\to R\otimes\mathbb{Q}_{p}/\mathbb{Z}_{p},
𝖧1(K𝗏,Tφ/πj)×𝖧1(K𝗏¯,Wφ[πj])\displaystyle\mathsf{H}^{1}(K_{\mathsf{v}},T_{\varphi}/\pi^{j})\times\mathsf{H}^{1}(K_{\overline{\mathsf{v}}},W_{\varphi}[\pi^{j}]) R/πj,\displaystyle\to R/\pi^{j},
𝖧1(K𝗏,Tφ)×𝖧1(K𝗏¯,Tφ)\displaystyle\mathsf{H}^{1}(K_{\mathsf{v}},T_{\varphi})\times\mathsf{H}^{1}(K_{\overline{\mathsf{v}}},T_{\varphi}) R.\displaystyle\to R.

Here 𝗏¯𝖬K\overline{\mathsf{v}}\in\mathsf{M}_{K} is the complex conjugate of 𝗏\mathsf{v}; the first two pairings are perfect. A Selmer structure (,𝖲)(\mathcal{F},\mathsf{S}) for TφT_{\varphi} induces a Selmer structure for WφW_{\varphi}, denoted the same way, by taking orthogonal complement local conditions.

Definition 3.3.2.

We say (,𝖲)(\mathcal{F},\mathsf{S}) is self-dual if, for all 𝗏𝖬K\mathsf{v}\in\mathsf{M}_{K}, 𝖧1(K𝗏,Tφ)\mathsf{H}^{1}_{\mathcal{F}}(K_{\mathsf{v}},T_{\varphi}) and 𝖧1(K𝗏¯,Tφ)\mathsf{H}^{1}_{\mathcal{F}}(K_{\overline{\mathsf{v}}},T_{\varphi}) are exact annihilators under the local pairing.

Proposition 3.3.3.

Suppose that (,𝖲)(\mathcal{F},\mathsf{S}) is a self-dual Selmer structure for TφT_{\varphi}. Then, for each {𝖰,ϵ𝖰}𝖭\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}:

  1. (1)

    ((𝖰),𝖲)(\mathcal{F}(\mathsf{Q}),\mathsf{S}) is self-dual and

    Sel(𝖰)(Wφ)(Rp/p)r𝖰M𝖰M𝖰\operatorname{Sel}_{\mathcal{F}(\mathsf{Q})}(W_{\varphi})\approx(R\otimes\mathbb{Q}_{p}/\mathbb{Z}_{p})^{r_{\mathsf{Q}}}\oplus M_{\mathsf{Q}}\oplus M_{\mathsf{Q}}

    for some torsion RR-module M𝖰M_{\mathsf{Q}} and an integer r𝖰r_{\mathsf{Q}}.

  2. (2)

    r𝖰=rkRSel(𝖰)(Tφ)r_{\mathsf{Q}}=\operatorname{rk}_{R}\operatorname{Sel}_{\mathcal{F}(\mathsf{Q})}(T_{\varphi}).

  3. (3)

    For any {𝖰𝗊,ϵ𝖰𝗊}𝖭\left\{\mathsf{Q}\mathsf{q},\epsilon_{\mathsf{Q}\mathsf{q}}\right\}\in\mathsf{N}, one of the following holds:

    1. (a)

      loc𝗊(Sel(𝖰)(Tφ))=0,\operatorname{loc}_{\mathsf{q}}(\operatorname{Sel}_{\mathcal{F}(\mathsf{Q})}(T_{\varphi}))=0, 𝗊(Sel(𝖰𝗊)(Tφ))0,\partial_{\mathsf{q}}(\operatorname{Sel}_{\mathcal{F}(\mathsf{Q}\mathsf{q})}(T_{\varphi}))\neq 0, r𝖰𝗊=r𝖰+1,r_{\mathsf{Q}\mathsf{q}}=r_{\mathsf{Q}}+1, and there exists an exact sequence of RR-modules:

      0M𝖰𝗊M𝖰loc𝗊(Sel(𝖰))(Wφ)0.0\to M_{\mathsf{Q}\mathsf{q}}\to M_{\mathsf{Q}}\to\operatorname{loc}_{\mathsf{q}}(\operatorname{Sel}_{\mathcal{F}(\mathsf{Q})})(W_{\varphi})\to 0.

      Moreover,

      lgloc𝗊(Sel(𝖰))(Wφ)=lgcoker𝗊(Sel(𝖰𝗊)(Tφ)).\lg\operatorname{loc}_{\mathsf{q}}(\operatorname{Sel}_{\mathcal{F}(\mathsf{Q})})(W_{\varphi})=\lg\operatorname{coker}\partial_{\mathsf{q}}(\operatorname{Sel}_{\mathcal{F}(\mathsf{Q}\mathsf{q})}(T_{\varphi})).
    2. (b)

      loc𝗊(Sel(𝖰)(Tφ))0,\operatorname{loc}_{\mathsf{q}}(\operatorname{Sel}_{\mathcal{F}(\mathsf{Q})}(T_{\varphi}))\neq 0, 𝗊(Sel(𝖰𝗊)(Tφ))=0,\partial_{\mathsf{q}}(\operatorname{Sel}_{\mathcal{F}(\mathsf{Q}\mathsf{q})}(T_{\varphi}))=0, r𝖰𝗊=r𝖰1,r_{\mathsf{Q}\mathsf{q}}=r_{\mathsf{Q}}-1, and there exists an exact sequence of RR-modules:

      0M𝖰M𝖰𝗊𝗊(Sel(𝖰𝗊)(Wφ))0.0\to M_{\mathsf{Q}}\to M_{\mathsf{Q}\mathsf{q}}\to\partial_{\mathsf{q}}(\operatorname{Sel}_{\mathcal{F}(\mathsf{Q}\mathsf{q})}(W_{\varphi}))\to 0.

      Moreover,

      lg𝗊(Sel(𝖰𝗊)(Wφ))=lgcokerloc𝗊(Sel(𝖰)(Tφ)).\lg\partial_{\mathsf{q}}(\operatorname{Sel}_{\mathcal{F}(\mathsf{Q}\mathsf{q})}(W_{\varphi}))=\lg\operatorname{coker}\operatorname{loc}_{\mathsf{q}}(\operatorname{Sel}_{\mathcal{F}(\mathsf{Q})}(T_{\varphi})).
Proof.
  1. (1)

    The self-duality claim is clear since Hord1(K𝗊,Tf)H^{1}_{\operatorname{ord}}(K_{\mathsf{q}},T_{f}) is self-dual. Now, for all j0,j\geq 0,

    (\ast) Sel(𝖰)(Wφ[πj])Sel(𝖰)(Wφ)[πj]\operatorname{Sel}_{\mathcal{F}(\mathsf{Q})}(W_{\varphi}[\pi^{j}])\simeq\operatorname{Sel}_{\mathcal{F}(\mathsf{Q})}(W_{\varphi})[\pi^{j}]

    by Lemma 2.4.6 and the definition of the induced Selmer structure on Wφ[πj]W_{\varphi}[\pi^{j}]. (Note H0(K,T¯f)=0H^{0}(K,\overline{T}_{f})=0 since we have assumed T¯f\overline{T}_{f} is an absolutely irreducible GKG_{K}-module.) Since Sel(𝖰)(Wφ)\operatorname{Sel}_{\mathcal{F}(\mathsf{Q})}(W_{\varphi}) is co-finitely generated, we may conclude by [24, Theorem 1.4.2] (or its proof).

  2. (2)

    As explained in [24, 25], the cohomological pairings deduced from Tφ×TφR(1)T_{\varphi}\times T_{\varphi}\to R(1) behave “exactly like” the Tate pairing; in particular, by the self-duality of the local conditions and Proposition 2.7.3, Sel(𝖰)(Tφ/πj)=Sel(𝖰)(Wφ[πj])\operatorname{Sel}_{\mathcal{F}(\mathsf{Q})}(T_{\varphi}/\pi^{j})=\operatorname{Sel}_{\mathcal{F}(\mathsf{Q})}(W_{\varphi}[\pi^{j}]) and the result follows as in the proof of Proposition 2.7.4.

  3. (3)

    Consider the Selmer structures 𝗊(𝖰)=(𝖰)+(𝖰𝗊)\mathcal{F}^{\mathsf{q}}(\mathsf{Q})=\mathcal{F}(\mathsf{Q})+\mathcal{F}(\mathsf{Q}\mathsf{q}) and 𝗊(𝖰)=(𝖰)(𝖰𝗊)\mathcal{F}_{\mathsf{q}}(\mathsf{Q})=\mathcal{F}(\mathsf{Q})\cap\mathcal{F}(\mathsf{Q}\mathsf{q}). By Proposition 2.7.4,

    rkRSel𝗊(𝖰)(Tφ)=rkRSel𝗊(𝖰)(Tφ)+1.\operatorname{rk}_{R}\operatorname{Sel}_{\mathcal{F}^{\mathsf{q}}(\mathsf{Q})}(T_{\varphi})=\operatorname{rk}_{R}\operatorname{Sel}_{\mathcal{F}_{\mathsf{q}}(\mathsf{Q})}(T_{\varphi})+1.

    Moreover, because (𝖰)\mathcal{F}(\mathsf{Q}) is self-dual, Proposition 2.7.2 implies that the image of

    Sel𝗊(𝖰)(Tφ)Sel𝗊(𝖰)(Tφ)\displaystyle\frac{\operatorname{Sel}_{\mathcal{F}^{\mathsf{q}}(\mathsf{Q})}(T_{\varphi})}{\operatorname{Sel}_{\mathcal{F}_{\mathsf{q}}(\mathsf{Q})}(T_{\varphi})} 𝖧𝗊(𝖰)1(K𝗊,Tφ)𝖧𝗊(𝖰)1(K𝗊,Tφ)=𝖧unr1(K𝗊,Tφ)𝖧𝗊(𝖰)1(K𝗊,Tφ)𝖧ord1(K𝗊,Tφ)𝖧𝗊(𝖰)1(K𝗊,Tφ)R2\displaystyle\hookrightarrow\frac{\mathsf{H}^{1}_{\mathcal{F}^{\mathsf{q}}(\mathsf{Q})}(K_{\mathsf{q}},T_{\varphi})}{\mathsf{H}^{1}_{\mathcal{F}_{\mathsf{q}}(\mathsf{Q})}(K_{\mathsf{q}},T_{\varphi})}=\frac{\mathsf{H}^{1}_{\operatorname{unr}}(K_{\mathsf{q}},T_{\varphi})}{\mathsf{H}^{1}_{\mathcal{F}_{\mathsf{q}}(\mathsf{Q})}(K_{\mathsf{q}},T_{\varphi})}\oplus\frac{\mathsf{H}^{1}_{\operatorname{ord}}(K_{\mathsf{q}},T_{\varphi})}{\mathsf{H}^{1}_{\mathcal{F}_{\mathsf{q}}(\mathsf{Q})}(K_{\mathsf{q}},T_{\varphi})}\approx R^{2}

    is self-annihilating under the induced local pairing, hence is contained either in the ordinary or unramified part.

    For the relation between M𝖰M_{\mathsf{Q}} and M𝖰𝗊M_{\mathsf{Q}\mathsf{q}}, we suppose we are in case (a), because the two arguments are identical. Using the perfect pairing between WφW_{\varphi} and TφT_{\varphi}, we see by Proposition 2.6.2 that loc𝗊(Sel(𝖰)(Wφ))𝗊(Sel(𝖰)(Wφ))\operatorname{loc}_{\mathsf{q}}(\operatorname{Sel}_{\mathcal{F}(\mathsf{Q})}(W_{\varphi}))\oplus\partial_{\mathsf{q}}(\operatorname{Sel}_{\mathcal{F}(\mathsf{Q})}(W_{\varphi})) is the exact annihilator of 𝗊(Sel(𝖰𝗊)(Tφ))\partial_{\mathsf{q}}(\operatorname{Sel}_{\mathcal{F}(\mathsf{Q}\mathsf{q})}(T_{\varphi})) under the perfect induced local pairing

    𝖧𝗊(𝖰)1(K𝗊,Tφ)𝖧𝗊(𝖰)1(K𝗊,Tφ)×𝖧𝗊(K𝗊,𝖰)1(Wφ)𝖧𝗊(𝖰)1(K𝗊,Wφ)Rp/p.\frac{\mathsf{H}^{1}_{\mathcal{F}^{\mathsf{q}}(\mathsf{Q})}(K_{\mathsf{q}},T_{\varphi})}{\mathsf{H}^{1}_{\mathcal{F}_{\mathsf{q}}(\mathsf{Q})}(K_{\mathsf{q}},T_{\varphi})}\times\frac{\mathsf{H}^{1}_{\mathcal{F}^{\mathsf{q}}(K_{\mathsf{q}},\mathsf{Q})}(W_{\varphi})}{\mathsf{H}^{1}_{\mathcal{F}_{\mathsf{q}}(\mathsf{Q})}(K_{\mathsf{q}},W_{\varphi})}\to R\otimes\mathbb{Q}_{p}/\mathbb{Z}_{p}.

    This implies that 𝗊(Sel(𝖰𝗊))(Wφ)\partial_{\mathsf{q}}(\operatorname{Sel}_{\mathcal{F}(\mathsf{Q}\mathsf{q})})(W_{\varphi}) is divisible and

    lgloc𝗊(Sel(𝖰))(Wφ)=lgcoker𝗊(Sel(𝖰𝗊)(Tφ)).\lg\operatorname{loc}_{\mathsf{q}}(\operatorname{Sel}_{\mathcal{F}(\mathsf{Q})})(W_{\varphi})=\lg\operatorname{coker}\partial_{\mathsf{q}}(\operatorname{Sel}_{\mathcal{F}(\mathsf{Q}\mathsf{q})}(T_{\varphi})).

    Furthermore, the direct sum decomposition of Sel(𝖰𝗊)(Wφ)\operatorname{Sel}_{\mathcal{F}(\mathsf{Q}\mathsf{q})}(W_{\varphi}) may be chosen so that 𝗊(M𝖰𝗊M𝖰𝗊)=0,\partial_{\mathsf{q}}(M_{\mathsf{Q}\mathsf{q}}\oplus M_{\mathsf{Q}\mathsf{q}})=0, and in particular

    M𝖰𝗊M𝖰𝗊(M𝖰M𝖰)Sel𝗊(𝖰)(Sel(𝖰))(Wφ).M_{\mathsf{Q}\mathsf{q}}\oplus M_{\mathsf{Q}\mathsf{q}}\simeq(M_{\mathsf{Q}}\oplus M_{\mathsf{Q}})\cap\operatorname{Sel}_{\mathcal{F}_{\mathsf{q}}(\mathsf{Q})}\subset(\operatorname{Sel}_{\mathcal{F}(\mathsf{Q})})(W_{\varphi}).

    Since loc𝗊(M𝖰M𝖰)\operatorname{loc}_{\mathsf{q}}(M_{\mathsf{Q}}\oplus M_{\mathsf{Q}}) must generate the image of loc𝗊(Sel(𝖰))(Wφ)\operatorname{loc}_{\mathsf{q}}(\operatorname{Sel}_{\mathcal{F}(\mathsf{Q})})(W_{\varphi}), the desired exact sequence follows.

The following result will allow us to control the alternative in Proposition 3.3.3(3).

Theorem 3.3.4.

Let c𝖧1(K𝖳/K,Tφ)c\in\mathsf{H}^{1}(K^{\mathsf{T}}/K,T_{\varphi}) be any nonzero element, where 𝖳𝖲\mathsf{T}\supset\mathsf{S} is a finite set. Then there are infinitely many admissible ultraprimes 𝗊𝖳\mathsf{q}\not\in\mathsf{T}, with associated signs ϵ𝗊\epsilon_{\mathsf{q}}, such that loc𝗊c0.\operatorname{loc}_{\mathsf{q}}c\neq 0.

The proof is via a series of lemmas.

Lemma 3.3.5.

There is an integer jj such that, for all n0,n\geq 0,

πjH1(K(Tφ)/K,Tφ/πn)=0.\pi^{j}H^{1}(K(T_{\varphi})/K,T_{\varphi}/\pi^{n})=0.

If (sclr) holds, then we may take j=0.j=0.

Proof.

Let G=Gal(K(Tφ)/K)G=\operatorname{Gal}(K(T_{\varphi})/K), and let ZGZ\subset G be its center; since TfT_{f} is absolutely irreducible over KK, ZZ acts on TφT_{\varphi} by scalars. We claim:

(18) Z{1}.Z\neq\left\{1\right\}.

Assuming (18), the lemma follows from the inflation-restriction exact sequence

H1(G/Z,H0(Z,Tφ/πn))H1(G,Tφ/πn)H1(Z,Tφ/πn).H^{1}(G/Z,H^{0}(Z,T_{\varphi}/\pi^{n}))\hookrightarrow H^{1}(G,T_{\varphi}/\pi^{n})\to H^{1}(Z,T_{\varphi}/\pi^{n}).

Let us now prove (18). Let G=Gal(K(Tf)/K)G^{\prime}=\operatorname{Gal}(K(T_{f})/K), and let L/KL/K be the Galois subfield of K(Tf)K(T_{f}) cut out by the center Z=Z(G)GZ^{\prime}=Z(G^{\prime})\subset G^{\prime}. By a result of Momose [43], ZZ^{\prime} is nontrivial. Let E/KE/K be the Galois extension determined by the kernel of φ\varphi; then it suffices to show that EL/LEL/L and K(Tf)/LK(T_{f})/L are linearly disjoint. Both ELEL and K(Tf)K(T_{f}) are Galois over \mathbb{Q}, so GG_{\mathbb{Q}} acts on Gal(EL/L)\operatorname{Gal}(EL/L) and Gal(K(Tf)/L)\operatorname{Gal}(K(T_{f})/L) by conjugation. If τG\tau\in G_{\mathbb{Q}} is a complex conjugation, then τ\tau acts trivially on Gal(K(Tf)/L)\operatorname{Gal}(K(T_{f})/L) but nontrivially on Gal(EL/L)\operatorname{Gal}(EL/L), so the two groups have no nontrivial common quotient compatible with the GG_{\mathbb{Q}}-action; hence ELK(Tf)=LEL\cap K(T_{f})=L. ∎

Lemma 3.3.6.

Suppose given a cocycle

cH1(K,Tφ/πn)c\in H^{1}(K,T_{\varphi}/\pi^{n})

such that πjc0,\pi^{j}c\neq 0, where jj is as in Lemma 3.3.5. Then, for any integer NnN\geq n, there exists a sign ϵ=±1\epsilon=\pm 1 and infinitely many rational primes qq such that:

  1. (1)

    qq is inert in KK and unramified in the splitting field (Tf,c)\mathbb{Q}(T_{f},c).

  2. (2)

    FrobqGal((Tf)/)\operatorname{Frob}_{q}\in\operatorname{Gal}(\mathbb{Q}(T_{f})/\mathbb{Q}) has distinct eigenvalues ±1\pm 1 on TfR/πNT_{f}\otimes R/\pi^{N} (where RR has trivial Galois action).

  3. (3)

    For any cocycle representative, c(Frobq2)c(\operatorname{Frob}_{q}^{2}) has nonzero component in the ϵ\epsilon eigenspace for Frobq\operatorname{Frob}_{q}.

Proof.

Abbreviate L=K(Tφ/πN)L=K(T_{\varphi}/\pi^{N}), and let ϕHomGK(GL,Tφ/πn)\phi\in\operatorname{Hom}_{G_{K}}(G_{L},T_{\varphi}/\pi^{n}) be the image of cc under restriction; by hypothesis ϕ0.\phi\neq 0. Without loss of generality, we may suppose that the image of ϕ\phi is contained in Tφ/πn[π]Tφ/πT_{\varphi}/\pi^{n}[\pi]\simeq T_{\varphi}/\pi, which, since φ\varphi is residually trivial, is an extension of scalars T¯f𝒪/k\overline{T}_{f}\otimes_{\mathcal{O}/\wp}k. Now,

HomGK(GL,T¯fk)\operatorname{Hom}_{G_{K}}(G_{L},\overline{T}_{f}\otimes k)

has a natural action of Gal(K/)\operatorname{Gal}(K/\mathbb{Q}), and we may assume without loss of generality that ϕ\phi lies in the ϵ\epsilon eigenspace for some ϵ{±1}.\epsilon\in\left\{\pm 1\right\}. Fix a complex conjugation τG\tau\in G_{\mathbb{Q}}. Since T¯f\overline{T}_{f} is absolutely irreducible over GKG_{K}, there exists gGLg\in G_{L} such that ϕ(g)\phi(g) has nonzero component in the ϵ\epsilon eigenspace of τ\tau. Then

ϕ(τgτg)=ϵτϕ(g)+ϕ(g)\phi(\tau g\tau g)=\epsilon\tau\phi(g)+\phi(g)

has nonzero component in the ϵ\epsilon eigenspace as well. Any qq with Frobenius τg\tau g in L(ϕ)L(\phi) satisfies the desired conditions. ∎

Proof of Theorem 3.3.4.

Since H0(K,Tφ/π)=0,H^{0}(K,T_{\varphi}/\pi)=0, Lemma 2.4.6 implies that

𝖧1(K𝖳/K,Tφ)[π]=0.\mathsf{H}^{1}(K^{\mathsf{T}}/K,T_{\varphi})[\pi]=0.

Thus there exists some nn such that the image c¯\overline{c} of cc in 𝖧1(K𝖳/K,Tφ/πn)\mathsf{H}^{1}(K^{\mathsf{T}}/K,T_{\varphi}/\pi^{n}) satisfies πjc¯0,\pi^{j}\overline{c}\neq 0, for some jj as in Lemma 3.3.5. By definition, c¯\overline{c} is represented by a sequence of classes cmH1(KTm/K,Tφ/πn)c_{m}\in H^{1}(K^{T_{m}}/K,T_{\varphi}/\pi^{n}) such that πjcm0\pi^{j}c_{m}\neq 0 for 𝔉\mathfrak{F}-many mm, where {Tm}m\left\{T_{m}\right\}_{m\in\mathbb{N}} represents 𝖳\mathsf{T}. For each mm, apply Lemma 3.3.6 (with N=mN=m) to obtain a prime qmTmq_{m}\not\in T_{m} and a sign ϵm\epsilon_{m}. If 𝗊𝖬\mathsf{q}\in\mathsf{M}_{\mathbb{Q}} is the equivalence class of the sequence {qm}m\left\{q_{m}\right\}_{m\in\mathbb{N}}, and ϵ𝒰({±1}m){±1}\epsilon\in\mathcal{U}(\left\{\pm 1\right\}_{m\in\mathbb{N}})\simeq\left\{\pm 1\right\} is the equivalence class of the sequence {ϵm}m,\left\{\epsilon_{m}\right\}_{m\in\mathbb{N}}, then the pair {𝗊,ϵ}\left\{\mathsf{q},\epsilon\right\} has the desired properties. Since there are infinitely many choices for each qmq_{m}, there are also infinitely many choices for 𝗊\mathsf{q}. ∎

Corollary 3.3.7.

For any {𝖰,ϵ𝖰}𝖭\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}, there exists some {𝖰𝖰,ϵ𝖰𝖰}𝖭\left\{\mathsf{Q}\mathsf{Q}^{\prime},\epsilon_{\mathsf{Q}\mathsf{Q}^{\prime}}\right\}\in\mathsf{N} such that r𝖰𝖰=0.r_{\mathsf{Q}\mathsf{Q}^{\prime}}=0.

Proof.

This is an obvious induction argument using Theorem 3.3.4 and Proposition 3.3.3. ∎

Combining Proposition 3.3.3 and Theorem 3.3.4 allows us to prove the main result of this subsection.

Theorem 3.3.8.

Suppose that (,𝖲)(\mathcal{F},\mathsf{S}) is self-dual and that (κ,λ)(\kappa,\lambda) is a nontrivial bipartite Euler system with sign δ\delta for (Tφ,,𝖲)(T_{\varphi},\mathcal{F},\mathsf{S}). Then there exists a nonzero fractional ideal II of RR such that:

  1. (1)

    For all {𝖰,ϵ𝖰}𝖭δ\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}^{\delta}, r𝖰r_{\mathsf{Q}} is odd, r𝖰=1r_{\mathsf{Q}}=1 if and only if κ(𝖰)0,\kappa(\mathsf{Q})\neq 0, and in that case

    charR(M𝖰)I=charR(Sel(𝖰)(Tφ)(κ(𝖰))).\operatorname{char}_{R}(M_{\mathsf{Q}})\cdot I=\operatorname{char}_{R}\left(\frac{\operatorname{Sel}_{\mathcal{F}(\mathsf{Q})}(T_{\varphi})}{(\kappa(\mathsf{Q}))}\right).
  2. (2)

    For all {𝖰,ϵ𝖰}𝖭δ+1\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}^{\delta+1}, r𝖰r_{\mathsf{Q}} is even, r𝖰=0r_{\mathsf{Q}}=0 if and only if λ(𝖰)0,\lambda(\mathsf{Q})\neq 0, and in that case

    charR(M𝖰)I=(λ(𝖰)).\operatorname{char}_{R}(M_{\mathsf{Q}})\cdot I=(\lambda(\mathsf{Q})).

In particular,

δ=rkRSel(Tφ)+1(mod2).\delta=\operatorname{rk}_{R}\operatorname{Sel}_{\mathcal{F}}(T_{\varphi})+1\pmod{2}.
Proof.

The proof will be in several steps.

Step 1.

If λ(𝖰)0\lambda(\mathsf{Q})\neq 0 for some {𝖰,ϵ𝖰}𝖭δ+1\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}^{\delta+1}, then r𝖰=0.r_{\mathsf{Q}}=0.

Proof.

If 0cSel(𝖰)(Tφ)0\neq c\in\operatorname{Sel}_{\mathcal{F}(\mathsf{Q})}(T_{\varphi}), then by Theorem 3.3.4, there exists an admissible ultraprime 𝗊\mathsf{q} with sign ϵ𝗊\epsilon_{\mathsf{q}} such that loc𝗊c0.\operatorname{loc}_{\mathsf{q}}c\neq 0. By Proposition 3.3.3, 𝗊(κ(𝖰𝗊))=0,\partial_{\mathsf{q}}(\kappa(\mathsf{Q}\mathsf{q}))=0, which contradicts the reciprocity laws. ∎

Step 2.

If κ(𝖰)0\kappa(\mathsf{Q})\neq 0 for some {𝖰,ϵ𝖰}𝖭δ\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}^{\delta}, then r𝖰=1.r_{\mathsf{Q}}=1.

Proof.

Choose an admissible ultraprime 𝗊\mathsf{q} with sign ϵ𝗊\epsilon_{\mathsf{q}} such that loc𝗊κ(𝖰)0.\operatorname{loc}_{\mathsf{q}}\kappa(\mathsf{Q})\neq 0. Then by the reciprocity laws, λ(𝖰𝗊)0,\lambda(\mathsf{Q}\mathsf{q})\neq 0, so by Step 1 r𝖰𝗊=0.r_{\mathsf{Q}\mathsf{q}}=0. Proposition 3.3.3 implies r𝖰=1.r_{\mathsf{Q}}=1.

Step 3.

For all {𝖰,ϵ𝖰}𝖭\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}, r𝖰δ+|𝖰|+1(mod2).r_{\mathsf{Q}}\equiv\delta+|\mathsf{Q}|+1\pmod{2}.

Proof.

If {𝖰𝖰,ϵ𝖰𝖰}𝖭\left\{\mathsf{Q}\mathsf{Q}^{\prime},\epsilon_{\mathsf{Q}\mathsf{Q}^{\prime}}\right\}\in\mathsf{N}, then by Proposition 3.3.3

r𝖰r𝖰𝖰|𝖰|(mod2).r_{\mathsf{Q}}-r_{\mathsf{Q}\mathsf{Q}^{\prime}}\equiv|\mathsf{Q}^{\prime}|\pmod{2}.

So Steps 1 and 2 imply Step 3. ∎

Step 4.

Suppose r𝖰=0r_{\mathsf{Q}}=0 for some {𝖰,ϵ𝖰}\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}. Then, for all admissible ultraprimes 𝗊𝖰𝖲\mathsf{q}\not\in\mathsf{Q}\cup\mathsf{S} with sign ϵ𝗊\epsilon_{\mathsf{q}}, r𝖰𝗊=1r_{\mathsf{Q}\mathsf{q}}=1 and

charR(M𝖰𝗊)(λ(𝖰))=charR(M𝖰)charR(Sel(𝖰𝗊)(Tφ)(κ(𝖰𝗊))).\operatorname{char}_{R}(M_{\mathsf{Q}\mathsf{q}})\cdot(\lambda(\mathsf{Q}))=\operatorname{char}_{R}(M_{\mathsf{Q}})\cdot\operatorname{char}_{R}\left(\frac{\operatorname{Sel}_{\mathcal{F}(\mathsf{Q}\mathsf{q})}(T_{\varphi})}{(\kappa(\mathsf{Q}\mathsf{q}))}\right).
Proof.

By Step 3, λ(𝖰)\lambda(\mathsf{Q}) and κ(𝖰𝗊)\kappa(\mathsf{Q}\mathsf{q}) are well-defined. Then Step 4 follows from Proposition 3.3.3, since

charR(Sel(𝖰𝗊)(Tφ)(κ(𝖰𝗊)))(𝗊(Sel(𝖰𝗊)(Tφ))=(λ(𝖰))R.\operatorname{char}_{R}\left(\frac{\operatorname{Sel}_{\mathcal{F}(\mathsf{Q}\mathsf{q})}(T_{\varphi})}{(\kappa(\mathsf{Q}\mathsf{q}))}\right)\cdot(\partial_{\mathsf{q}}(\operatorname{Sel}_{\mathcal{F}(\mathsf{Q}\mathsf{q})}(T_{\varphi}))=(\lambda(\mathsf{Q}))\subset R.

The exact same reasoning implies:

Step 5.

Suppose that r𝖰=1r_{\mathsf{Q}}=1 and 𝗊𝖰𝖲\mathsf{q}\not\in\mathsf{Q}\cup\mathsf{S} is an admissible ultraprime with sign ϵ𝗊\epsilon_{\mathsf{q}} such that r𝖰𝗊=0.r_{\mathsf{Q}\mathsf{q}}=0. Then

charR(M𝖰𝗊)(Sel(𝖰)(Tφ)(κ(𝖰)))=charR(M𝖰)(λ(𝖰𝗊)).\operatorname{char}_{R}(M_{\mathsf{Q}\mathsf{q}})\cdot\left(\frac{\operatorname{Sel}_{\mathcal{F}(\mathsf{Q})}(T_{\varphi})}{(\kappa(\mathsf{Q}))}\right)=\operatorname{char}_{R}(M_{\mathsf{Q}})\cdot(\lambda(\mathsf{Q}\mathsf{q})).

Now consider the graph 𝒳\mathcal{X} [25] whose vertices are the elements of 𝖭\mathsf{N}, and where the edges are between vertices of the form {𝖰,ϵ𝖰}\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\} and {𝖰𝗊,ϵ𝖰𝗊}\left\{\mathsf{Q}\mathsf{q},\epsilon_{\mathsf{Q}\mathsf{q}}\right\}, for some admissible ultraprime 𝗊\mathsf{q} with sign ϵ𝗊\epsilon_{\mathsf{q}}. We say {𝖰,ϵQT}\left\{\mathsf{Q},\epsilon_{Q}T\right\} is a core vertex if r𝖰1.r_{\mathsf{Q}}\leq 1. The core subgraph 𝒳0\mathcal{X}_{0} of 𝒳\mathcal{X} is the full subgraph on core vertices. Applying Steps 4 and 5, it suffices to show that 𝒳0\mathcal{X}_{0} is path-connected to complete the proof of the theorem.

Step 6.

If v={𝖰,ϵ𝖰}v=\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\} and v={𝖰𝖰,ϵ𝖰𝖰}v^{\prime}=\left\{\mathsf{Q}\mathsf{Q}^{\prime},\epsilon_{\mathsf{Q}\mathsf{Q}^{\prime}}\right\} are core vertices, then they are connected by a path in 𝒳0\mathcal{X}_{0}.

Proof.

We proceed by induction on |𝖰||\mathsf{Q}^{\prime}|, where the base case is trivial. If r𝖰𝖰/𝗊1r_{\mathsf{Q}\mathsf{Q}^{\prime}/\mathsf{q}}\leq 1 for any 𝗊𝖰\mathsf{q}\in\mathsf{Q}^{\prime}, then we may apply the inductive hypothesis, so assume otherwise. By Proposition 3.3.3, r𝖰𝖰=1r_{\mathsf{Q}\mathsf{Q}^{\prime}}=1 and 𝗊(Sel(𝖰𝖰)(Tφ))=0\partial_{\mathsf{q}}(\operatorname{Sel}_{\mathcal{F}(\mathsf{Q}\mathsf{Q}^{\prime})}(T_{\varphi}))=0 for all 𝗊𝖰\mathsf{q}\in\mathsf{Q}^{\prime}. Hence

Sel(𝖰𝖰)(Tφ)Sel(𝖰)(Tφ).\operatorname{Sel}_{\mathcal{F}(\mathsf{Q}\mathsf{Q}^{\prime})}(T_{\varphi})\subset\operatorname{Sel}_{\mathcal{F}(\mathsf{Q})}(T_{\varphi}).

Then, by Theorem 3.3.4 and Proposition 3.3.3, there exists an admissible ultraprime 𝗊𝖰𝖰𝖲\mathsf{q}\not\in\mathsf{Q}\cup\mathsf{Q}^{\prime}\cup\mathsf{S} with sign ϵ𝗊\epsilon_{\mathsf{q}} such that r𝖰𝗊=r𝖰𝖰𝗊=0.r_{\mathsf{Q}\mathsf{q}}=r_{\mathsf{Q}\mathsf{Q}^{\prime}\mathsf{q}}=0. If 𝗊𝖰\mathsf{q}^{\prime}\in\mathsf{Q}^{\prime} is any factor, then {𝖰𝖰𝗊/𝗊,ϵ𝖰𝖰𝗊/𝗊}𝖭\left\{\mathsf{Q}\mathsf{Q}^{\prime}\mathsf{q}/\mathsf{q}^{\prime},\epsilon_{\mathsf{Q}\mathsf{Q}^{\prime}\mathsf{q}/\mathsf{q}^{\prime}}\right\}\in\mathsf{N} is a core vertex, which is connected to vv^{\prime} in 𝒳0\mathcal{X}_{0}. By the inductive hypothesis, {𝖰𝖰𝗊/𝗊,ϵ𝖰𝖰𝗊/𝗊}\left\{\mathsf{Q}\mathsf{Q}^{\prime}\mathsf{q}/\mathsf{q}^{\prime},\epsilon_{\mathsf{Q}\mathsf{Q}^{\prime}\mathsf{q}/\mathsf{q}^{\prime}}\right\} is also connected to the core vertex {𝖰𝗊,ϵ𝖰𝗊}\left\{\mathsf{Q}\mathsf{q},\epsilon_{\mathsf{Q}\mathsf{q}}\right\}, hence to vv, by a path in 𝒳0.\mathcal{X}_{0}. This completes the inductive step. ∎

Step 7.

If v={𝖰,ϵ𝖰}v=\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\} is a core vertex and 𝖳𝖬\mathsf{T}\subset\mathsf{M}_{\mathbb{Q}} is any finite set, then there exists a core vertex v={𝖰,ϵ𝖰}v^{\prime}=\left\{\mathsf{Q}^{\prime},\epsilon_{\mathsf{Q}^{\prime}}\right\} such that vv and vv^{\prime} are connected by a path in 𝒳0\mathcal{X}_{0} and 𝖰𝖳=\mathsf{Q}^{\prime}\cap\mathsf{T}=\emptyset.

Proof.

By iterating, it suffices to assume that 𝖰𝖳\mathsf{Q}\cap\mathsf{T} consists of exactly one ultraprime 𝗊𝖰\mathsf{q}\in\mathsf{Q}. If r𝖰/𝗊1,r_{\mathsf{Q}/\mathsf{q}}\leq 1, then the conclusion is obvious, so suppose otherwise. As in the proof of Step 6, choose an admissible ultraprime 𝗊𝖰𝖲𝖳\mathsf{q}^{\prime}\not\in\mathsf{Q}\cup\mathsf{S}\cup\mathsf{T} with associated sign ϵ𝗊\epsilon_{\mathsf{q}^{\prime}} such that r𝖰𝗊=0,r_{\mathsf{Q}\mathsf{q}^{\prime}}=0, which implies r𝖰𝗊/𝗊=1r_{\mathsf{Q}\mathsf{q}^{\prime}/\mathsf{q}}=1. The core vertex v={𝖰𝗊/𝗊,ϵ𝖰𝗊/𝗊}v^{\prime}=\left\{\mathsf{Q}\mathsf{q}^{\prime}/\mathsf{q},\epsilon_{\mathsf{Q}\mathsf{q}^{\prime}/\mathsf{q}}\right\} has the desired properties. ∎

Finally, we have:

Step 8.

The core subgraph 𝒳0\mathcal{X}_{0} is path-connected.

Proof.

Let {𝖰1,ϵ𝖰1}\left\{\mathsf{Q}_{1},\epsilon_{\mathsf{Q}_{1}}\right\} and {𝖰2,ϵ𝖰2}\left\{\mathsf{Q}_{2},\epsilon_{\mathsf{Q}_{2}}\right\} be two core vertices. Without loss of generality, by Step 7, we may assume 𝖰1𝖰2=\mathsf{Q}_{1}\cap\mathsf{Q}_{2}=\emptyset. (This step is necessary because the sign functions ϵ𝖰1\epsilon_{\mathsf{Q}_{1}} and ϵ𝖰2\epsilon_{\mathsf{Q}_{2}} need not agree on 𝖰1𝖰2.\mathsf{Q}_{1}\cap\mathsf{Q}_{2}.) Consider {𝖰1𝖰2,ϵ𝖰1𝖰2}𝖭\left\{\mathsf{Q}_{1}\mathsf{Q}_{2},\epsilon_{\mathsf{Q}_{1}\mathsf{Q}_{2}}\right\}\in\mathsf{N}. This may not be a core vertex, but, by repeatedly applying Theorem 3.3.4 and Proposition 3.3.3, there exists {𝖰3,ϵ𝖰3}𝖭\left\{\mathsf{Q}_{3},\epsilon_{\mathsf{Q}_{3}}\right\}\in\mathsf{N} such that {𝖰1𝖰2𝖰3,ϵ𝖰1𝖰2𝖰3}\left\{\mathsf{Q}_{1}\mathsf{Q}_{2}\mathsf{Q}_{3},\epsilon_{\mathsf{Q}_{1}\mathsf{Q}_{2}\mathsf{Q}_{3}}\right\} is a core vertex. We may then conclude by Step 6. ∎

Proposition 3.3.9.

Under the hypotheses of Theorem 3.3.8, there exists a constant CC depending on |𝖲||\mathsf{S}|, TfT_{f}, and the ramification index of R/𝒪R/\mathcal{O}, but not on φ\varphi, such that IπCRI\pi^{C}\subset R. If (sclr) holds, then we may take C=0.C=0.

Proof.

By Theorem 3.3.8, it suffices to show that there exists a constant with the desired dependencies and a pair {𝖰,ϵ𝖰}𝖭δ\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}^{\delta} such that πCcharR(M𝖰)\pi^{C}\in\operatorname{char}_{R}(M_{\mathsf{Q}}). We first note that the constant jj in Lemma 3.3.6 depends only on TfT_{f} and the ramification index of R/𝒪R/\mathcal{O}, and can be taken to be 0 under (sclr).

Moreover, if kk is the residue field of RR, then d=dimk𝖧1(K𝖲/K,Wφ[π])d=\dim_{k}\mathsf{H}^{1}(K^{\mathsf{S}}/K,W_{\varphi}[\pi]) is also bounded with the desired uniformity. We now construct a sequence {𝖰i,ϵ𝖰i}\left\{\mathsf{Q}_{i},\epsilon_{\mathsf{Q}_{i}}\right\} recursively (starting from 𝖰1=1\mathsf{Q}_{1}=1) by the following rules:

  • If r𝖰i>0,r_{\mathsf{Q}_{i}}>0, then choose any 𝗊i+1𝖰i\mathsf{q}_{i+1}\not\in\mathsf{Q}_{i} with sign ϵ𝗊i+1\epsilon_{\mathsf{q}_{i+1}} such that

    lgcoker(loc𝗊Sel(𝖰i)(Tφ))j.\lg\operatorname{coker}(\operatorname{loc}_{\mathsf{q}}\operatorname{Sel}_{\mathcal{F}(\mathsf{Q}_{i})}(T_{\varphi}))\leq j.
  • If r𝖰i=0r_{\mathsf{Q}_{i}}=0 and the exponent of Sel(𝖰i)(Wφ)0\operatorname{Sel}_{\mathcal{F}(\mathsf{Q}_{i})}(W_{\varphi})\neq 0 is ni>ijn_{i}>i\cdot j, then choose any 𝗊i+1𝖰i\mathsf{q}_{i+1}\not\in\mathsf{Q}_{i} with sign ϵ𝗊i+1\epsilon_{\mathsf{q}_{i+1}} such that the exponent of loc𝗊(Sel(𝖰i)(Wφ))\operatorname{loc}_{\mathsf{q}}(\operatorname{Sel}_{\mathcal{F}(\mathsf{Q}_{i})}(W_{\varphi})) is at least nijn_{i}-j.

These choices are possible by Lemma 3.3.6. In either of the above two cases, set

{𝖰i+1,ϵ𝖰i+1}={𝖰i𝗊i+1,ϵ𝖰i𝗊i+1};\left\{\mathsf{Q}_{i+1},\epsilon_{\mathsf{Q}_{i+1}}\right\}=\left\{\mathsf{Q}_{i}\mathsf{q}_{i+1},\epsilon_{\mathsf{Q}_{i}\mathsf{q}_{i+1}}\right\};

if neither holds, then end the construction. For each ii, let r𝖰ir_{\mathsf{Q}_{i}}^{\prime} be the minimal number of generators of the RR-module πijM𝖰i.\pi^{i\cdot j}M_{\mathsf{Q}_{i}}. In the first case of the construction, r𝖰i+1r𝖰ir_{\mathsf{Q}_{i+1}}^{\prime}\leq r_{\mathsf{Q}_{i}}^{\prime}; in the second case, r𝖰i+1<r𝖰ir_{\mathsf{Q}_{i+1}}^{\prime}<r_{\mathsf{Q}_{i}}^{\prime} (by Proposition 3.3.3(3b,a) respectively). After r1dr_{1}\leq d steps, we alternate between the two cases of the construction, taking at most 2r12d2r_{1}^{\prime}\leq 2d more steps. Hence for some i3di\leq 3d, r𝖰i=0r^{\prime}_{\mathsf{Q}_{i}}=0 and r𝖰i=0r_{\mathsf{Q}_{i}}=0, and the construction halts. For this ii,

lgM𝖰iijdimkSel(𝖰i)(Wφ)[π]3dj(d+3d),\lg M_{\mathsf{Q}_{i}}\leq ij\dim_{k}\operatorname{Sel}_{\mathcal{F}(\mathsf{Q}_{i})}(W_{\varphi})[\pi]\leq 3dj(d+3d),

the last inequality by the reasoning of [25, Corollary 2.2.10]. (Less precisely, we could deduce the bound 3dj(d+6d)3dj(d+6d) directly from Proposition 3.3.3(3).)

Since dd and jj have bounds of the desired sort, the claim follows. ∎

3.4. Euler systems over Λ\Lambda

Let Λ\Lambda be the anticyclotomic Iwasawa algebra 𝒪T\mathcal{O}\llbracket T\rrbracket with canonical character

Ψ:GKΛ×.\Psi:G_{K}\to\Lambda^{\times}.

For each height-one prime 𝔓Λ\mathfrak{P}\subset\Lambda, let S𝔓S_{\mathfrak{P}} be the integral closure of Λ/𝔓\Lambda/\mathfrak{P} in its field of fractions, so that Ψ\Psi induces a character GKΛ×S𝔓×G_{K}\to\Lambda^{\times}\to S_{\mathfrak{P}}^{\times}. We write T𝔓T_{\mathfrak{P}} for the twist T𝒪S𝔓(Ψ)T\otimes_{\mathcal{O}}S_{\mathfrak{P}}(\Psi) and 𝐓\mathbf{T} for the interpolated twist T𝒪Λ(Ψ)T\otimes_{\mathcal{O}}\Lambda(\Psi). Also let 𝐖=𝐓\mathbf{W}=\mathbf{T}^{\ast} be the Cartier dual with Λ\Lambda action twisted by the canonical involution ι\iota, so that for each 𝔓\mathfrak{P} there is a natural map

W𝔓𝐖W_{\mathfrak{P}}\to\mathbf{W}

of Λ[GK]\Lambda[G_{K}]-modules (see, e.g., [24]). The following definition is motivated by [33, Lemma 5.3.13] and its applications in [24, 25].

Definition 3.4.1.

An interpolated self-dual Selmer structure

(𝖲,Λ,𝔓,ΣΛ)(\mathsf{S},\mathcal{F}_{\Lambda},\mathcal{F}_{\mathfrak{P}},\Sigma_{\Lambda})

for 𝐓\mathbf{T} consists of the following data:

  • A finite set 𝖲𝖬K\mathsf{S}\subset\mathsf{M}_{K}.

  • For each height-one prime 𝔓Λ\mathfrak{P}\subset\Lambda, a self-dual Selmer structure (𝔓,𝖲)(\mathcal{F}_{\mathfrak{P}},\mathsf{S}) for T𝔓T_{\mathfrak{P}}.

  • A finite set ΣΛ\Sigma_{\Lambda} of height-one primes 𝔓Λ\mathfrak{P}\subset\Lambda.

  • A Selmer structure (Λ,𝖲)(\mathcal{F}_{\Lambda},\mathsf{S}) for 𝐓\mathbf{T} such that, for all 𝗏𝖬K\mathsf{v}\in\mathsf{M}_{K} and all 𝔓Λ\mathfrak{P}\subset\Lambda, there are well-defined maps induced in the obvious way:

    (19) 𝖧Λ1(K𝗏,𝐓/𝔓)𝖧𝔓1(K𝗏,T𝔓),𝖧𝔓1(K𝗏,W𝔓)𝖧Λ1(K𝗏,𝐖[𝔓]).\begin{split}\mathsf{H}^{1}_{\mathcal{F}_{\Lambda}}(K_{\mathsf{v}},\mathbf{T}/\mathfrak{P})&\to\mathsf{H}^{1}_{\mathcal{F}_{\mathfrak{P}}}(K_{\mathsf{v}},T_{\mathfrak{P}}),\\ \mathsf{H}^{1}_{\mathcal{F}_{\mathfrak{P}}}(K_{\mathsf{v}},W_{\mathfrak{P}})&\to\mathsf{H}^{1}_{\mathcal{F}_{\Lambda}}(K_{\mathsf{v}},\mathbf{W}[\mathfrak{P}]).\end{split}

    Moreover, for all 𝔓ΣΛ\mathfrak{P}\not\in\Sigma_{\Lambda}, the maps (19) have finite kernel and cokernel with order bounded by a constant depending only on [S𝔓:Λ/𝔓][S_{\mathfrak{P}}:\Lambda/\mathfrak{P}] as 𝔓\mathfrak{P} varies.

Proposition 3.4.2 ([33], Proposition 5.3.14).

Suppose (𝖲,𝔓,Λ,ΣΛ)(\mathsf{S},\mathcal{F}_{\mathfrak{P}},\mathcal{F}_{\Lambda},\Sigma_{\Lambda}) is an interpolated self-dual Selmer structure for 𝐓\mathbf{T}. Then for all 𝔓Λ\mathfrak{P}\subset\Lambda, there are well-defined maps induced in the obvious way:

SelΛ(𝐓)/𝔓\displaystyle\operatorname{Sel}_{\mathcal{F}_{\Lambda}}(\mathbf{T})/\mathfrak{P} Sel𝔓(T)\displaystyle\to\operatorname{Sel}_{\mathcal{F}_{\mathfrak{P}}}(T_{\wp})
Sel𝔓(W𝔓)\displaystyle\operatorname{Sel}_{\mathcal{F}_{\mathfrak{P}}}(W_{\mathfrak{P}}) SelΛ(𝐖)[𝔓].\displaystyle\to\operatorname{Sel}_{\mathcal{F}_{\Lambda}}(\mathbf{W})[\mathfrak{P}].

For all 𝔓ΣΛ\mathfrak{P}\not\in\Sigma_{\Lambda}, these maps have finite kernel and cokernel with a bound depending on \mathcal{F} and on [S𝔓:Λ/𝔓][S_{\mathfrak{P}}:\Lambda/\mathfrak{P}], but not on 𝔓\mathfrak{P} itself.

3.4.3.

Recall that, for any finitely generated Λ\Lambda-module MM, there exists a unique Λ\Lambda-module NN of the form ΛrΛ/𝔓iei\Lambda^{r}\oplus\bigoplus\Lambda/\mathfrak{P}_{i}^{e_{i}} such that MM admits a map to NN with finite kernel and cokernel; we denote this relationship by MNM\sim N. The characteristic ideal charΛ(M)\operatorname{char}_{\Lambda}(M) is zero if r1,r\geq 1, and equal to 𝔓iei\prod\mathfrak{P}_{i}^{e_{i}} otherwise. The following easy lemma is implicit in [33, p. 66].

Lemma 3.4.4.

Let 𝔓Λ\mathfrak{P}\subset\Lambda be a height-one prime. Then there exists an integer dd and a sequence of height-one primes 𝔓m\mathfrak{P}_{m} such that, for all finitely generated torsion Λ\Lambda-modules MM,

lg𝒪(M/𝔓m)=mdord𝔓charΛ(M)+O(1)\lg_{\mathcal{O}}(M/{\mathfrak{P}_{m}})=md\operatorname{ord}_{\mathfrak{P}}\operatorname{char}_{\Lambda}(M)+O(1)

as mm varies (holding MM fixed). Moreover [S𝔓m:Λ/𝔓m][S_{\mathfrak{P}_{m}}:\Lambda/\mathfrak{P}_{m}] is constant for large enough mm, and if 𝔓(),\mathfrak{P}\neq(\wp), then the rings Λ/𝔓m\Lambda/\mathfrak{P}_{m} are abstractly isomorphic.

Proof.

If 𝔓()\mathfrak{P}\neq(\wp) is generated by a distinguished polynomial fΛf\in\Lambda, and π\pi is a uniformizer for 𝒪\mathcal{O}, then we may take 𝔓m=f+πm\mathfrak{P}_{m}=f+\pi^{m} (for sufficiently large mm) and d=[S𝔓:𝒪]d=[S_{\mathfrak{P}}:\mathcal{O}]. If 𝔓=()=(π)\mathfrak{P}=(\wp)=(\pi), then we may take 𝔓m=Tm+π\mathfrak{P}_{m}=T^{m}+\pi and d=1.d=1.

Proposition 3.4.5.

Suppose that (𝖲,𝔓,Λ,ΣΛ)(\mathsf{S},\mathcal{F}_{\mathfrak{P}},\mathcal{F}_{\Lambda},\Sigma_{\Lambda}) is an interpolated self-dual Selmer structure for 𝐓\mathbf{T}. Then for all {𝖰,ϵ𝖰}𝖭\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}:

  1. (1)

    (𝖲𝖰,𝔓(𝖰),Λ(𝖰),ΣΛ)(\mathsf{S}\cup\mathsf{Q},\mathcal{F}_{\mathfrak{P}}(\mathsf{Q}),\mathcal{F}_{\Lambda}(\mathsf{Q}),\Sigma_{\Lambda}) is an interpolated self-dual Selmer structure for 𝐓\mathbf{T} and

    SelΛ(𝖰)(𝐖)Λr𝖰M𝖰M𝖰\operatorname{Sel}_{\mathcal{F}_{\Lambda}(\mathsf{Q})}(\mathbf{W})^{\vee}\sim\Lambda^{r_{\mathsf{Q}}}\oplus M_{\mathsf{Q}}\oplus M_{\mathsf{Q}}

    for some torsion Λ\Lambda-module M𝖰M_{\mathsf{Q}} and an integer r𝖰r_{\mathsf{Q}}.

  2. (2)

    r𝖰=rkΛSelΛ(𝖰)(𝐓)r_{\mathsf{Q}}=\operatorname{rk}_{\Lambda}\operatorname{Sel}_{\mathcal{F}_{\Lambda}(\mathsf{Q})}(\mathbf{T}).

Proof.

At places 𝗊𝖰\mathsf{q}\in\mathsf{Q},

𝖧Λ(𝖰)1(K𝗊,𝐓)=𝖧ord1(K𝗊,Tf)Λ\mathsf{H}^{1}_{\mathcal{F}_{\Lambda}(\mathsf{Q})}(K_{\mathsf{q}},\mathbf{T})=\mathsf{H}^{1}_{\operatorname{ord}}(K_{\mathsf{q}},T_{f})\otimes\Lambda

and

𝖧𝔓(𝖰)1(K𝗊,𝐓)=𝖧ord1(K𝗊,Tf)S𝔓,\mathsf{H}^{1}_{\mathcal{F}_{\mathfrak{P}}(\mathsf{Q})}(K_{\mathsf{q}},\mathbf{T})=\mathsf{H}^{1}_{\operatorname{ord}}(K_{\mathsf{q}},T_{f})\otimes S_{\mathfrak{P}},

so we clearly have local maps with kernel and cokernels bounded as desired (and similarly for 𝐖f\mathbf{W}_{f} and W𝔓W_{\mathfrak{P}}); so indeed (𝖲𝖰,𝔓(𝖰),Λ(𝖰),ΣΛ)(\mathsf{S}\cup\mathsf{Q},\mathcal{F}_{\mathfrak{P}}(\mathsf{Q}),\mathcal{F}_{\Lambda}(\mathsf{Q}),\Sigma_{\Lambda}) is an interpolated self-dual Selmer structure. The rest of the claims are deduced from Proposition 3.4.2 and Proposition 3.3.3(1,2) exactly as in [24, Theorem 2.2.10]. ∎

Theorem 3.4.6.

Suppose that (𝖲,𝔓,Λ,ΣΛ)(\mathsf{S},\mathcal{F}_{\mathfrak{P}},\mathcal{F}_{\Lambda},\Sigma_{\Lambda}) is an interpolated self-dual Selmer structure for 𝐓\mathbf{T} and {𝛋,𝛌}\left\{\bm{\kappa},\bm{\lambda}\right\} is a nontrivial bipartite Euler system with parity δ\delta for the triple (𝐓,Λ,𝖲)(\mathbf{T},\mathcal{F}_{\Lambda},\mathsf{S}). Then there exists a nonzero fractional ideal IΛpI\subset\Lambda\otimes\mathbb{Q}_{p} such that:

  1. (1)

    For all {𝖰,ϵ𝖰}𝖭δ\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}^{\delta}, r𝖰r_{\mathsf{Q}} is odd, r𝖰=1r_{\mathsf{Q}}=1 if and only if 𝜿(𝖰)0,\bm{\kappa}(\mathsf{Q})\neq 0, and in that case

    charΛ(M𝖰)I=charΛ(SelΛ(𝖰)(𝐓)(𝜿(𝖰))).\operatorname{char}_{\Lambda}(M_{\mathsf{Q}})\cdot I=\operatorname{char}_{\Lambda}\left(\frac{\operatorname{Sel}_{\mathcal{F}_{\Lambda}(\mathsf{Q})}(\mathbf{T})}{(\bm{\kappa}(\mathsf{Q}))}\right).
  2. (2)

    For all {𝖰,ϵ𝖰}𝖭δ+1\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}^{\delta+1}, r𝖰r_{\mathsf{Q}} is even, r𝖰=0r_{\mathsf{Q}}=0 if and only if 𝝀(𝖰)0,\bm{\lambda}(\mathsf{Q})\neq 0, and in that case

    charΛ(M𝖰)I=(𝝀(𝖰)).\operatorname{char}_{\Lambda}(M_{\mathsf{Q}})\cdot I=(\bm{\lambda}(\mathsf{Q})).

In particular,

δ=rkRSel(𝐓)+1(mod2).\delta=\operatorname{rk}_{R}\operatorname{Sel}_{\mathcal{F}}(\mathbf{T})+1\pmod{2}.

If (sclr) holds, then IΛI\subset\Lambda.

Proof.

Let 𝔓Λ\mathfrak{P}\subset\Lambda be any height-one prime; via the natural maps SelΛ(𝐓)Sel𝔓(T𝔓)\operatorname{Sel}_{\mathcal{F}_{\Lambda}}(\mathbf{T})\to\operatorname{Sel}_{\mathcal{F}_{\mathfrak{P}}}(T_{\mathfrak{P}}) and ΛS𝔓\Lambda\to S_{\mathfrak{P}}, the Euler system (𝜿,𝝀)(\bm{\kappa},\bm{\lambda}) defines an Euler system (κ𝔓,λ𝔓)(\kappa_{\mathfrak{P}},\lambda_{\mathfrak{P}}) of parity δ\delta for the triple (T𝔓,𝔓,𝖲)(T_{\mathfrak{P}},\mathcal{F}_{\mathfrak{P}},\mathsf{S}). In particular, Theorem 3.3.8 applies.

If 𝜿(𝖰)0,\bm{\kappa}(\mathsf{Q})\neq 0, then by Proposition 3.4.2 κ𝔓(𝖰)0\kappa_{\mathfrak{P}}(\mathsf{Q})\neq 0 for all but finitely many 𝔓\mathfrak{P} (since SelΛ(𝖰)(𝐓)\operatorname{Sel}_{\mathcal{F}_{\Lambda}(\mathsf{Q})}(\mathbf{T}) is torsion-free), and similarly for 𝝀(𝖰)\bm{\lambda}(\mathsf{Q}). Because

rkΛSelΛ(𝐓)rkS𝔓Sel𝔓(T𝔓)\operatorname{rk}_{\Lambda}\operatorname{Sel}_{\mathcal{F}_{\Lambda}}(\mathbf{T})\leq\operatorname{rk}_{S_{\mathfrak{P}}}\operatorname{Sel}_{\mathcal{F}_{\mathfrak{P}}}(T_{\mathfrak{P}})

with equality for all but finitely many 𝔓\mathfrak{P}, the claims about r𝖰r_{\mathsf{Q}} follow from Theorem 3.3.8.

For any 𝔓\mathfrak{P} and {𝖰,ϵ𝖰}𝖭δ+1\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}^{\delta+1} such that 𝝀(𝖰)0\bm{\lambda}(\mathsf{Q})\neq 0, by Proposition 3.4.5 and Lemma 3.4.4 we have

e𝔓(𝖰)\displaystyle e_{\mathfrak{P}}(\mathsf{Q}) ord𝔓(𝝀(𝖰))ord𝔓charΛ(M𝖰)\displaystyle\coloneqq\operatorname{ord}_{\mathfrak{P}}(\bm{\lambda}(\mathsf{Q}))-\operatorname{ord}_{\mathfrak{P}}\operatorname{char}_{\Lambda}(M_{\mathsf{Q}})
=limmlg𝒪(S𝔓m/λ𝔓m(𝖰))lg𝒪M𝖰,𝔓mmd.\displaystyle=\lim_{m\to\infty}\frac{\lg_{\mathcal{O}}(S_{\mathfrak{P}_{m}}/\lambda_{\mathfrak{P}_{m}}(\mathsf{Q}))-\lg_{\mathcal{O}}M_{\mathsf{Q},\mathfrak{P}_{m}}}{md}.

Applying Theorem 3.3.8, this quantity does not depend on {𝖰,ϵ𝖰}\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\} (as long as 𝝀(𝖰)0\bm{\lambda}(\mathsf{Q})\neq 0); it is also clearly zero for almost all 𝔓\mathfrak{P}, so that 𝔓𝔓e𝔓\prod_{\mathfrak{P}}\mathfrak{P}^{e_{\mathfrak{P}}} defines a fractional ideal II of Λ\Lambda satisfying (2). The same calculation shows that II satisfies (1) as well, and the integrality properties follow from Proposition 3.3.9. ∎

4. Geometry of modular Jacobians

4.1. Multiplicity one

4.1.1.

Let N1N_{1} and N2N_{2} be coprime positive integers. Consider the Hecke algebra 𝕋=𝕋N1,N2\mathbb{T}=\mathbb{T}_{N_{1},N_{2}} generated over \mathbb{Z} by operators TT_{\ell} for all primes N=N1N2\ell\nmid N=N_{1}N_{2} and UU_{\ell} for all |N,\ell|N, acting on the modular forms of weight two and level Γ0(N)\Gamma_{0}(N) which are new at all factors |N2.\ell|N_{2}. If II is the kernel of the projection 𝕋N1N2,1𝕋\mathbb{T}_{N_{1}N_{2},1}\to\mathbb{T}, then we set

(20) JminN1,N2J0(N)/IJ0(N),J_{\min}^{N_{1},N_{2}}\coloneqq J_{0}(N)/IJ_{0}(N),

an abelian variety with a (faithful) action of 𝕋\mathbb{T}. If N1,N2N_{1},N_{2} are clear from context, we will omit the superscript.

For any abelian variety AA with an action of 𝕋\mathbb{T}, and any maximal ideal 𝔪𝕋\mathfrak{m}\subset\mathbb{T}, the 𝔪\mathfrak{m}-adic Tate module is defined to be the localization

(21) T𝔪ATpA𝕋𝕋𝔪,T_{\mathfrak{m}}A\coloneqq T_{p}A\otimes_{\mathbb{T}}\mathbb{T}_{\mathfrak{m}},

where pp is the residue characteristic of 𝔪\mathfrak{m}. (Note that this is dual to the notation of [22].) For any 𝔪\mathfrak{m} which is non-Eisenstein with odd residue characteristic pNp\nmid N, it follows from [49] that T𝔪JminT_{\mathfrak{m}}J_{\min} is free of rank two over 𝕋𝔪\mathbb{T}_{\mathfrak{m}}; by [22, Corollary 4.7], the natural map then induces an isomorphism

(22) 𝕋𝔪End𝕋(Jmin)𝔪.\mathbb{T}_{\mathfrak{m}}\xrightarrow{\sim}\operatorname{End}_{\mathbb{T}}(J_{\min})_{\mathfrak{m}}.

4.1.2.

Now suppose that AA is an abelian variety with a Hecke-equivariant isogeny to JminJ_{\min}. For any ||N2,\ell||N_{2}, let 𝒜/\mathcal{A}_{/\mathbb{Z}_{\ell}} be the Néron model of AA. The connected component 𝒜𝔽0\mathcal{A}^{0}_{\mathbb{F}_{\ell}} of the special fiber of 𝒜\mathcal{A} is a torus, and we write 𝒳(A)=Hom(𝒜𝔽0,𝔾m)\mathcal{X}_{\ell}(A)=\operatorname{Hom}(\mathcal{A}^{0}_{\mathbb{F}_{\ell}},\mathbb{G}_{m}) for its character group. The association A𝒳(A)A\mapsto\mathcal{X}_{\ell}(A) is contravariantly functorial.

Proposition 4.1.3 (Helm).

Let 𝔪𝕋\mathfrak{m}\subset\mathbb{T} be non-Eisenstein of residue characteristic p2Np\nmid 2N. Then the natural maps induce 𝕋𝔪\mathbb{T}_{\mathfrak{m}}-module isomorphisms:

T𝔪JminHom(Jmin,A)𝔪T𝔪A,T_{\mathfrak{m}}J_{\min}\otimes\operatorname{Hom}(J_{\min},A)_{\mathfrak{m}}\xrightarrow{\sim}T_{\mathfrak{m}}A,
𝒳(Jmin)Hom(Jmin,A)𝔪𝒳(A),\mathcal{X}_{\ell}(J_{\min}^{\vee})\otimes\operatorname{Hom}(J_{\min},A)_{\mathfrak{m}}\xrightarrow{\sim}\mathcal{X}_{\ell}(A^{\vee}),
Hom(A,Jmin)𝔪Hom𝕋𝔪(Hom(Jmin,A)𝔪,End(Jmin)𝔪).\operatorname{Hom}(A,J_{\min})_{\mathfrak{m}}\xrightarrow{\sim}\operatorname{Hom}_{\mathbb{T}_{\mathfrak{m}}}\left(\operatorname{Hom}(J_{\min},A)_{\mathfrak{m}},\operatorname{End}(J_{\min})_{\mathfrak{m}}\right).

Here, all Hom-sets are understood to be 𝕋\mathbb{T}-equivariant morphisms, and tensor products are taken modulo \mathbb{Z}-torsion.

Proof.

This follows by duality from [22, Corollary 4.1, Theorem 4.11, Proposition 4.14]. ∎

We record the following elementary lemma for later use.

Lemma 4.1.4.

Let 𝒳=𝒳(Jmin)𝔪\mathcal{X}=\mathcal{X}_{\ell}(J_{\min}^{\vee})_{\mathfrak{m}} for some |N2\ell|N_{2} and 𝔪𝕋\mathfrak{m}\subset\mathbb{T}, where 𝔪\mathfrak{m} is non-Eisenstein of odd residue characteristic pp. If the associated residual representation ρ¯𝔪\overline{\rho}_{\mathfrak{m}} is ramified at \ell, then 𝒳\mathcal{X} is free of rank one over 𝕋𝔪\mathbb{T}_{\mathfrak{m}}. In general, there exist 𝕋𝔪\mathbb{T}_{\mathfrak{m}}-module maps

ϕi:𝒳𝕋𝔪,ψi:𝕋𝔪𝒳,i=1,2\phi_{i}:\mathcal{X}\to\mathbb{T}_{\mathfrak{m}},\;\;\psi_{i}:\mathbb{T}_{\mathfrak{m}}\to\mathcal{X},\;\;i=1,2

such that

ϕiψi=ψiϕi=ti𝕋𝔪End(𝒳)\phi_{i}\circ\psi_{i}=\psi_{i}\circ\phi_{i}=t_{i}\in\mathbb{T}_{\mathfrak{m}}\subset\operatorname{End}(\mathcal{X})

and

t1+t2=1𝕋𝔪.t_{1}+t_{2}=\ell-1\in\mathbb{T}_{\mathfrak{m}}.
Proof.

If 1\ell-1 is a pp-adic unit, or if ρ¯𝔪\overline{\rho}_{\mathfrak{m}} is ramified, then this follows from [22, Lemma 6.5]. In general, we have

(23) 𝒳=Hom((𝒥min)𝔽0[𝔪],μp)\mathcal{X}=\operatorname{Hom}((\mathcal{J}^{\vee}_{\min})_{\mathbb{F}_{\ell}}^{0}[\mathfrak{m}^{\infty}],\mu_{p^{\infty}})

so that 𝒳\mathcal{X} may be identified with a 𝕋𝔪[G]\mathbb{T}_{\mathfrak{m}}[G_{\mathbb{Q}_{\ell}}]-module quotient

π:T𝔪Jmin𝒳;\pi:T_{\mathfrak{m}}J_{\min}\to\mathcal{X};

the Galois action on 𝒳\mathcal{X} is unramified and Frobenius acts as UU_{\ell}, which is a constant ±\pm1 because the residue characteristic of 𝔪\mathfrak{m} is p>2.p>2.

Because T𝔪JminT_{\mathfrak{m}}J_{\min} is free of rank two over 𝕋𝔪\mathbb{T}_{\mathfrak{m}}, it may be equipped with a basis {e1,e2}\left\{e_{1},e_{2}\right\}, and moreover an alternating 𝕋𝔪\mathbb{T}_{\mathfrak{m}}-module pairing

(24) ,:T𝔪Jmin×T𝔪Jmin𝕋𝔪\langle\cdot,\cdot\rangle:T_{\mathfrak{m}}J_{\min}\times T_{\mathfrak{m}}J_{\min}\to\mathbb{T}_{\mathfrak{m}}

such that

(25) y=e1,ye2e2,ye1y=\langle e_{1},y\rangle e_{2}-\langle e_{2},y\rangle e_{1}

for all yT𝔪Jmin.y\in T_{\mathfrak{m}}J_{\min}. Define maps

ϕi:T𝔪Jmin\displaystyle\phi_{i}:T_{\mathfrak{m}}J_{\min} 𝕋𝔪,i=1,2\displaystyle\to\mathbb{T}_{\mathfrak{m}},\;\;i=1,2
ϕ1:y\displaystyle\phi_{1}:y y,(FU)e2\displaystyle\mapsto\langle y,(F-U_{\ell})e_{2}\rangle
ϕ2:y\displaystyle\phi_{2}:y y,(FU)e1,\displaystyle\mapsto\langle y,(F-U_{\ell})e_{1}\rangle,

where FGF\in G_{\mathbb{Q}_{\ell}} is any lift of Frobenius. We first claim that the maps ϕi\phi_{i} factor through π\pi. Since 𝕋𝔪\mathbb{T}_{\mathfrak{m}} is pp-torsion-free, it suffices to check this after inverting pp. On T𝔪JminpT_{\mathfrak{m}}J_{\min}\otimes\mathbb{Q}_{p}, FF acts with distinct eigenvalues UU_{\ell} and U\ell U_{\ell}, and πp:T𝔪Jminp𝒳p\pi\otimes\mathbb{Q}_{p}:T_{\mathfrak{m}}J_{\min}\otimes\mathbb{Q}_{p}\to\mathcal{X}\otimes\mathbb{Q}_{p} coincides with the projection onto the UU_{\ell}-eigenspace. Since ,\langle\cdot,\cdot\rangle is alternating and 𝕋𝔪\mathbb{T}_{\mathfrak{m}}-linear, it follows that each ϕi\phi_{i} does indeed descend to a 𝕋𝔪\mathbb{T}_{\mathfrak{m}}-module map 𝒳𝕋𝔪\mathcal{X}\to\mathbb{T}_{\mathfrak{m}}. Now define maps

ψi:𝕋𝔪\displaystyle\psi_{i}:\mathbb{T}_{\mathfrak{m}} 𝒳,i=1,2\displaystyle\to\mathcal{X},\;\;i=1,2
ψ1:1\displaystyle\psi_{1}:1 Uπ(e1)\displaystyle\mapsto U_{\ell}\pi(e_{1})
ψ2:1\displaystyle\psi_{2}:1 Uπ(e2).\displaystyle\mapsto-U_{\ell}\pi(e_{2}).

We claim that ψi\psi_{i} and ϕi\phi_{i} satisfy the conclusion of the lemma. One readily calculates:

ϕ1ψ1(1)\displaystyle\phi_{1}\circ\psi_{1}(1) =Ue1,(FU)e2\displaystyle=U_{\ell}\langle e_{1},(F-U_{\ell})e_{2}\rangle
ψ1ϕ1(e1)\displaystyle\psi_{1}\circ\phi_{1}(e_{1}) =Ue1,(FU)e2π(e1)\displaystyle=U_{\ell}\langle e_{1},(F-U_{\ell})e_{2}\rangle\pi(e_{1})
ψ1ϕ1(e2)\displaystyle\psi_{1}\circ\phi_{1}(e_{2}) =Ue2,(FU)e2π(e1)\displaystyle=U_{\ell}\langle e_{2},(F-U_{\ell})e_{2}\rangle\pi(e_{1})
=Ue1,(FU)e2π(e2)U(FU)π(e2)\displaystyle=U_{\ell}\langle e_{1},(F-U_{\ell})e_{2}\rangle\pi(e_{2})-U_{\ell}(F-U_{\ell})\pi(e_{2})
=Ue1,(FU)e2π(e2),\displaystyle=U_{\ell}\langle e_{1},(F-U_{\ell})e_{2}\rangle\pi(e_{2}),

where in the last two steps we have used (25) and the fact that F=UF=U_{\ell} on 𝒳\mathcal{X}. Similarly,

ϕ2ψ2=ψ2ϕ2=Ue2,(FU)e1,\phi_{2}\circ\psi_{2}=\psi_{2}\circ\phi_{2}=-U_{\ell}\langle e_{2},(F-U_{\ell})e_{1}\rangle,

and

Ue1,(FU)e2Ue2,(FU)e1=trT𝔪JminU(FU)=1.U_{\ell}\langle e_{1},(F-U_{\ell})e_{2}\rangle-U_{\ell}\langle e_{2},(F-U_{\ell})e_{1}\rangle=\operatorname{tr}_{T_{\mathfrak{m}}J_{\min}}U_{\ell}(F-U_{\ell})=\ell-1.

4.2. Shimura curves

4.2.1.

If ν(N2)\nu(N_{2}) is even, then there exists a Shimura curve XN1,N2,X_{N_{1},N_{2}}, with Γ0(N1)\Gamma_{0}(N_{1}) level structure, associated to the indefinite quaternion algebra over \mathbb{Q} of discriminant N2.N_{2}. Let

JN1,N2J(XN1,N2),J^{N_{1},N_{2}}\coloneqq J(X_{N_{1},N_{2}}),

an abelian variety with a natural action of 𝕋\mathbb{T} by correspondences (induced by Picard functoriality). When N1N_{1} and N2N_{2} are understood, we abbreviate J=JN1,N2J=J^{N_{1},N_{2}}. There is a noncanonical Hecke-equivariant isogeny JJminJ\to J_{\min}. Consider the following technical hypothesis on the residual representation ρ¯𝔪:GGL2(𝕋/𝔪)\overline{\rho}_{\mathfrak{m}}:G_{\mathbb{Q}}\to GL_{2}(\mathbb{T}/\mathfrak{m}) associated to 𝔪\mathfrak{m}:

(\ast) If p=3 and ρ¯𝔪 is induced from a character of G(3),||N2such that either 1(mod3) or ρ¯𝔪 is ramified at .\begin{split}&\text{If }p=3\text{ and }\overline{\rho}_{\mathfrak{m}}\text{ is induced from a character of }G_{\mathbb{Q}(\sqrt{-3})},\;\exists\;\ell||N_{2}\\ &\text{such that either }\ell\equiv-1\pmod{3}\text{ or }\overline{\rho}_{\mathfrak{m}}\text{ is ramified at }\ell.\end{split}
Theorem 4.2.2 (Helm).

Let 𝔪𝕋\mathfrak{m}\subset\mathbb{T} be a non-Eisenstein maximal ideal of residue characteristic p2Np\nmid 2N satisfying (\ast). Then there is an isomorphism of 𝕋𝔪\mathbb{T}_{\mathfrak{m}}-modules:

Hom(Jmin,J)|N2𝒳(Jmin)𝔪,\operatorname{Hom}(J_{\min},J)\simeq\otimes_{\ell|N_{2}}\mathcal{X}_{\ell}(J_{\min}^{\vee})_{\mathfrak{m}},

modulo \mathbb{Z}-torsion on the right-hand side.

Proof.

This is essentially [22, Theorem 8.7]; to complete the case p=3,p=3, by [22, Remark 8.12] one only needs a level-raising input that is provided by [16]. ∎

4.3. Shimura sets

Now suppose that ν(N2)\nu(N_{2}) is odd, and consider the finite double coset space (often called a Shimura set):

(26) XN1,N2R(𝔸)×\B(𝔸)×/B()×,X_{N_{1},N_{2}}\coloneqq R(\mathbb{A}_{\mathbb{Q}})^{\times}\backslash B(\mathbb{A}_{\mathbb{Q}})^{\times}/B(\mathbb{Q})^{\times},

where BB is a definite quaternion algebra over \mathbb{Q} ramified at N2N_{2} and \infty, and RR is an Eichler order in BB of level Γ0(N1).\Gamma_{0}(N_{1}). When N1N_{1} and N2N_{2} are clear from context, the subscripts may be omitted.

4.3.1.

The \mathbb{Z}-module [X]0\mathbb{Z}[X]^{0} of formal degree-zero divisors in XX has two natural actions of 𝕋=𝕋N1,N2\mathbb{T}=\mathbb{T}_{N_{1},N_{2}} by correspondences: an “Albanese” action induced by viewing an element of [X]0\mathbb{Z}[X]^{0} as a formal sum of points in a double coset space, and a “Picard” action induced by identifying [X]=HomSet(X,)\mathbb{Z}[X]=\operatorname{Hom}_{\text{Set}}(X,\mathbb{Z}). We will consider [X]0\mathbb{Z}[X]^{0} as a 𝕋\mathbb{T}-module through the latter action. The analogue of Theorem 4.2.2 is:

Theorem 4.3.2.

Let 𝔪𝕋\mathfrak{m}\subset\mathbb{T} be a non-Eisenstein maximal ideal of residue characteristic p2Np\nmid 2N satisfying (\ast). Then there is an isomorphism of 𝕋𝔪\mathbb{T}_{\mathfrak{m}}-modules:

[X]0|N2𝒳(Jmin)𝔪,\mathbb{Z}[X]^{0}\simeq\otimes_{\ell|N_{2}}\mathcal{X}_{\ell}(J_{\min}^{\vee})_{\mathfrak{m}},

modulo \mathbb{Z}-torsion on the right-hand side.

Proof.

Choose any prime q|N2,q|N_{2}, so that ν(N2/q)\nu(N_{2}/q) is even. Let 𝕋=𝕋N1q,N2/q,\mathbb{T}^{\prime}=\mathbb{T}_{N_{1}q,N_{2}/q}, and write 𝔪\mathfrak{m} as well for the maximal ideal of 𝕋\mathbb{T}^{\prime} induced by the map 𝕋𝕋\mathbb{T}^{\prime}\to\mathbb{T}.

Applying Theorem 4.2.2 to the pair N1q,N2/qN_{1}q,N_{2}/q, we obtain an isomorphism of 𝕋𝔪\mathbb{T}^{\prime}_{\mathfrak{m}}-modules (modulo \mathbb{Z}-torsion)

(27) Hom(JminN1q,N2/q,JN1q,N2/q)𝔪|N2/q𝒳(JminN1q,N2/q,).\operatorname{Hom}(J_{\min}^{N_{1}q,N_{2}/q},J^{N_{1}q,N_{2}/q})_{\mathfrak{m}}\simeq\otimes_{\ell|N_{2}/q}\mathcal{X}_{\ell}(J_{\min}^{N_{1}q,N_{2}/q,\vee}).

By [22, Corollary 5.3, Lemma 8.2], this implies an isomorphism of 𝕋𝔪\mathbb{T}_{\mathfrak{m}}-modules

(28) Hom(JminN1,N2,JqnewN1,N2/q)𝔪|N2/q𝒳(JminN1,N2,),\operatorname{Hom}(J_{\min}^{N_{1},N_{2}},J^{N_{1},N_{2}/q}_{q\operatorname{-new}})_{\mathfrak{m}}\simeq\otimes_{\ell|N_{2}/q}\mathcal{X}_{\ell}(J_{\min}^{N_{1},N_{2},\vee}),

where JqnewN1q,N2/qJ^{N_{1}q,N_{2}/q}_{q\operatorname{-new}} is the qq-new quotient of JN1q,N2/qJ^{N_{1}q,N_{2}/q}. Then, by Proposition 4.1.3, we have

(29) 𝒳q(JqnewqN1,N2/q,)𝔪𝒳q(JminN1,N2,)𝔪|N2/q𝒳(JminN1,N2,)𝔪.\mathcal{X}_{q}(J^{qN_{1},N_{2}/q,\vee}_{q\operatorname{-new}})_{\mathfrak{m}}\simeq\mathcal{X}_{q}(J_{\min}^{N_{1},N_{2},\vee})_{\mathfrak{m}}\otimes_{\ell|N_{2}/q}\mathcal{X}_{\ell}(J_{\min}^{N_{1},N_{2},\vee})_{\mathfrak{m}}.

By [3, Proposition 5.3], 𝒳q(JqN1,N2/q,)\mathcal{X}_{q}(J^{qN_{1},N_{2}/q,\vee}) is canonically identified with [XN1,N2]0.\mathbb{Z}[X_{N_{1},N_{2}}]^{0}. It remains to show that the inclusion JqnewN1q,N2/q,JN1q,N2/q,J^{N_{1}q,N_{2}/q,\vee}_{q\operatorname{-new}}\hookrightarrow J^{N_{1}q,N_{2}/q,\vee} induces an isomorphism on character groups at qq. Indeed, since JqnewN1q,N2/q,J^{N_{1}q,N_{2}/q,\vee}_{q\operatorname{-new}} has purely toric reduction at qq, there is a surjection of character groups 𝒳q(JN1q,N2/q,)𝒳q(JqnewN1q,N2/q,)\mathcal{X}_{q}(J^{N_{1}q,N_{2}/q,\vee})\twoheadrightarrow\mathcal{X}_{q}(J^{N_{1}q,N_{2}/q,\vee}_{q\operatorname{-new}}), which is clearly an isomorphism after tensoring both sides with \mathbb{Q}, hence also before.

4.4. CM points

4.4.1.

Let us now fix an imaginary quadratic field K/K/\mathbb{Q}, and, once and for all, an embedding KGL2()K\hookrightarrow GL_{2}(\mathbb{Q}) such that KM2()=𝒪KK\cap M_{2}(\mathbb{Z})=\mathcal{O}_{K}.

For any integer N=N+NN=N^{+}N^{-} such that every prime factor of N+N^{+} is (unramified and) split in KK, and NN^{-} is a squarefree product of primes (unramified and) inert in KK, let B=BNB=B_{N^{-}} be the quaternion algebra over \mathbb{Q} ramified exactly at the factors of NN^{-} (and possibly \infty), and RBR\subset B an Eichler order of level N+N^{+}. For each such BB, we fix an optimal embedding KBK\hookrightarrow B such that KR=𝒪KK\cap R=\mathcal{O}_{K}. Then we may define the space of KK-CM points:

𝒞N+,N=K×\B(𝔸f)/R^×.\mathscr{C}_{N^{+},N^{-}}=K^{\times}\backslash B(\mathbb{A}_{f})/\widehat{R}^{\times}.

The Galois group GKabG_{K}^{\operatorname{ab}} acts on 𝒞N+,N\mathscr{C}_{N^{+},N^{-}} through the reciprocity map

(30) rec:GKabK×\K^×,\operatorname{rec}:G_{K}^{\operatorname{ab}}\xrightarrow{\sim}K^{\times}\backslash\widehat{K}^{\times},

i.e. σ[b]=[rec(σ)b].\sigma[b]=[\operatorname{rec}(\sigma)b].

4.4.2.

If SMS\subset M_{\mathbb{Q}} is a finite set of primes, consider the set of SS-CM points:

𝒞S=K×\K^×GL2(𝔸S)/O^K×GL2(^S).\mathscr{C}^{S}=K^{\times}\backslash\widehat{K}^{\times}GL_{2}(\mathbb{A}_{S})/\widehat{O}^{\times}_{K}GL_{2}(\widehat{\mathbb{Z}}_{S}).

We again have an action of GKabG_{K}^{\operatorname{ab}}, and by [37, Proposition 2.5] the field of definition K[y]K[y] of any y𝒞Sy\in\mathscr{C}^{S} is contained in the compositum K[S]K[S] of all ring class fields unramified outside SS. If SS is disjoint from the factors of NN, then there is a unique map GL2(𝔸S)B(𝔸S)GL_{2}(\mathbb{A}_{S})\hookrightarrow B(\mathbb{A}_{S}) so that the composite

K𝔸SGL2(𝔸S)B(𝔸S)K\otimes_{\mathbb{Q}}\mathbb{A}_{S}\hookrightarrow GL_{2}(\mathbb{A}_{S})\hookrightarrow B(\mathbb{A}_{S})

agrees with the embedding deduced from KBK\hookrightarrow B. This embedding identifies 𝒞S\mathscr{C}^{S} with a Galois-stable subset of 𝒞N+,N\mathscr{C}_{N^{+},N^{-}}.

4.4.3.

If ν(N)\nu(N^{-}) is even, then the Shimura curve X=XN+,NX=X_{N^{+},N^{-}} of (4.2.1) admits a complex uniformization:

(31) X()=B×\±×B(𝔸f)×/R^×,X(\mathbb{C})=B^{\times}\backslash\mathcal{H}^{\pm}\times B(\mathbb{A}_{f})^{\times}/\widehat{R}^{\times},

where ±=\mathcal{H}^{\pm}=\mathbb{C}\setminus\mathbb{R}. We have an injective map:

(32) CMN+,N:𝒞N+,NX()y[(h0,y)],\begin{split}\operatorname{CM}_{N^{+},N^{-}}:\mathscr{C}_{N^{+},N^{-}}&\to X(\mathbb{C})\\ y&\mapsto[(h_{0},y)],\end{split}

where h0h_{0} is the unique fixed point on +\mathcal{H}^{+} of the action of KK through the embedding KBK\hookrightarrow B. By Shimura’s reciprocity law, the image of CMN+,N\operatorname{CM}_{N^{+},N^{-}} is contained in X(Kab),X(K^{\operatorname{ab}}), and CMN+,N\operatorname{CM}_{N^{+},N^{-}} is Galois-equivariant for the action on 𝒞N+,N\mathscr{C}_{N^{+},N^{-}} defined above.

4.4.4.

If instead ν(N)\nu(N^{-}) is odd, then there is a natural projection

CMN+,N:𝒞N+,NXN+,N,\operatorname{CM}_{N^{+},N^{-}}:\mathscr{C}_{N^{+},N^{-}}\to X_{N^{+},N^{-}},

where XN+,NX_{N^{+},N^{-}} is the Shimura set of (4.3).

4.4.5.

With notation as above in (4.4.1), let qNq\nmid N be a prime inert in KK and not in SS; note that the unique prime of KK above qq splits completely in K[S]/KK[S]/K. We fix a prime 𝔮\mathfrak{q} of K[S]K[S] lying above qq according to a fixed embedding GqGG_{\mathbb{Q}_{q}}\hookrightarrow G_{\mathbb{Q}}. If ν(N)\nu(N^{-}) is even, then all points of CMN+,N(𝒞S)\operatorname{CM}_{N^{+},N^{-}}(\mathscr{C}^{S}) have supersingular reduction modulo 𝔮\mathfrak{q}, and the set of supersingular points X(𝔽q2)ssX(\mathbb{F}_{q^{2}})^{ss} may be identified with the Shimura set Xq=XN+,NqX_{q}=X_{N^{+},N^{-}q} associated to the definite quaternion algebra BqB_{q} ramified at NqN^{-}q\infty. Consider the following hypothesis on the residual representation ρ¯𝔪\overline{\rho}_{\mathfrak{m}} associated to a non-Eisenstein maximal ideal 𝔪𝕋N1,N2\mathfrak{m}\subset\mathbb{T}_{N_{1},N_{2}} of residue characteristic pp (for any Hecke algebra 𝕋N1,N2\mathbb{T}_{N_{1},N_{2}}):

(TW) if p=3, then ρ¯𝔪 is absolutely irreducible over 3.\text{if }p=3,\text{ then }\overline{\rho}_{\mathfrak{m}}\text{ is absolutely irreducible over }\mathbb{Q}{\sqrt{-3}}.

Note that this is strictly stronger than condition (\ast) above.

Proposition 4.4.6.

With identifications chosen compatibly, there is a commutative diagram:

[𝒞S]0{\mathbb{Z}\left[\mathscr{C}^{S}\right]^{0}}[CMN+,N(𝒞S)]0{\mathbb{Z}\left[\operatorname{CM}_{N^{+},N^{-}}(\mathscr{C}^{S})\right]^{0}}J(K[S]){J(K[S])}[𝒞S]0{\mathbb{Z}\left[\mathscr{C}^{S}\right]^{0}}[Xq]0{\mathbb{Z}[X_{q}]^{0}}J(𝔽q2){J(\mathbb{F}_{q^{2}})}Red𝔮\scriptstyle{\operatorname{Red}_{\mathfrak{q}}}Red𝔮\scriptstyle{\operatorname{Red}_{\mathfrak{q}}}

Moreover, the map [Xq]0J(𝔽q2)\mathbb{Z}[X_{q}]^{0}\to J(\mathbb{F}_{q^{2}}) is compatible with the action of 𝕋N+q,NQ\mathbb{T}_{N^{+}q,N^{-}Q} where UqU_{q} acts on J(𝔽q2)J(\mathbb{F}_{q^{2}}) through Frobq\operatorname{Frob}_{q}, and is surjective after localizing at a non-Eisenstein maximal ideal 𝔪𝕋N+q,NQ\mathfrak{m}\subset\mathbb{T}_{N^{+}q,N^{-}Q} satisfying condition (TW).

Proof.

See [52, Lemma 5.4.3] for the commutativity of the diagram and [42] for the UqU_{q} action. The surjectivity is an application of Ihara’s Lemma which can be deduced from the argument in [3, Proposition 9.2]: we add auxiliary level of the form Γ1()\Gamma_{1}(\ell), where N\ell\nmid N is a prime such that 1,T1𝔪\ell-1,T_{\ell}-\ell-1\not\in\mathfrak{m}. That such a prime exists follows from condition (TW) by [16, Lemma 3]. ∎

4.4.7.

Now suppose instead that ν(N)\nu(N^{-}) is odd, and let qNq\nmid N^{-} be a prime inert in KK and not lying in SS. The Shimura curve Xq=XN+,NqX_{q}=X_{N^{+},N^{-}q} has a canonical, semistable integral model over q\mathbb{Z}_{q}, whose irreducible components are identified with two copies of X=XN+,NX=X_{N^{+},N^{-}}. We denote this set by X±X^{\pm}, where the positive copy is the one containing the reduction of the point [(h0,1)][(h_{0},1)] in the uniformization of (31). We define a map 𝒞SX±\mathscr{C}^{S}\to X^{\pm} by the composition 𝒞SCMN+,NXX+X±\mathscr{C}^{S}\xrightarrow{\operatorname{CM}_{N^{+},N^{-}}}X\simeq X^{+}\subset X^{\pm}.

4.4.8.

The Néron model 𝒥q\mathcal{J}_{q} of the Jacobian Jq=JN+,NqJ_{q}=J^{N^{+},N^{-}q} has purely toric reduction, and we write 𝒳\mathcal{X} and Φ\Phi for the character group and the group of connected components, respectively, of its special fiber. Recall the rigid-analytic uniformization of JqJ_{q}, which gives rise to an exact sequence:

(33) 0𝒳𝒳¯qJq(¯q)0.0\to\mathcal{X}\to\mathcal{X}^{\dagger}\otimes\overline{\mathbb{Q}}_{q}\to J_{q}(\overline{\mathbb{Q}}_{q})\to 0.

Here, 𝒳=Hom(𝒳,)\mathcal{X}^{\dagger}=\operatorname{Hom}(\mathcal{X},\mathbb{Z}), and the maps are Hecke-equivariant if 𝒳\mathcal{X} is given Hecke action through Albanese functoriality, and the actions on 𝒳\mathcal{X}^{\dagger} and Jq(¯q)J_{q}(\overline{\mathbb{Q}}_{q}) are induced by Picard functoriality. Importantly for our later applications, (33) is compatible with the Galois action of GqG_{\mathbb{Q}_{q}} [5], where the action on 𝒳\mathcal{X} is unramified and Frobenius acts through UqU_{q} [42, Proposition 3.8]. The rigid analytic uniformization is related to the monodromy pairing j:𝒳𝒳j:\mathcal{X}\to\mathcal{X}^{\dagger} of Grothendieck [21] by the commutative diagram with exact rows:

0{0}𝒳{\mathcal{X}}𝒳q2{\mathcal{X}^{\dagger}\otimes\mathbb{Q}_{q^{2}}}Jq(q2){J_{q}(\mathbb{Q}_{q^{2}})}0{0}0{0}𝒳{\mathcal{X}}𝒳{\mathcal{X}^{\dagger}}Φ{\Phi}0{0}ord\scriptstyle{\operatorname{ord}}Spq\scriptstyle{\operatorname{Sp}_{q}}j\scriptstyle{j}

In particular, the specialization map is well-defined on Jq(K[S])J_{q}(K[S]).

Proposition 4.4.9.

With identifications chosen compatibly, there is a commutative diagram:

[𝒞S]0{\mathbb{Z}\left[\mathscr{C}^{S}\right]^{0}}[CMN+,Nq(𝒞S)]0{\mathbb{Z}\left[\operatorname{CM}_{N^{+},N^{-}q}(\mathscr{C}^{S})\right]^{0}}Jq(K[S]){J_{q}(K[S])}[𝒞S]0{\mathbb{Z}\left[\mathscr{C}^{S}\right]^{0}}[X±]0{\mathbb{Z}[X^{\pm}]^{0}}Φ{\Phi}Sp𝔮\scriptstyle{\operatorname{Sp}_{\mathfrak{q}}}Sp𝔮\scriptstyle{\operatorname{Sp}_{\mathfrak{q}}}

Moreover, the map [X±]0Φ\mathbb{Z}[X^{\pm}]^{0}\to\Phi is compatible with the action of 𝕋q=𝕋N+q,N\mathbb{T}_{q}=\mathbb{T}_{N^{+}q,N^{-}}, where UqU_{q} acts on [X±]0\mathbb{Z}[X^{\pm}]^{0} by the matrix

(Tqq10).\begin{pmatrix}T_{q}&q\\ -1&0\end{pmatrix}.

After localizing at a non-Eisenstein maximal ideal 𝔪𝕋q\mathfrak{m}\subset\mathbb{T}_{q}, the map [X±]0Φ\mathbb{Z}[X^{\pm}]^{0}\to\Phi induces an isomorphism

[X±]𝔪0𝕋q𝕋q/(Uq21)Φ𝔪.\mathbb{Z}[X^{\pm}]^{0}_{\mathfrak{m}}\otimes_{\mathbb{T}_{q}}\mathbb{T}_{q}/(U_{q}^{2}-1)\xrightarrow{\sim}\Phi_{\mathfrak{m}}.
Proof.

For the commutativity of the diagram, see [52, Lemma 5.4.6]; for the rest, see [3, Proposition 5.5]. ∎

5. CM classes in cohomology

5.1. Level raising

5.1.1.

Fix notation as in (3.1.1), and suppose that N=N+NN=N^{+}N^{-} is coprime to pDKpD_{K}, where all factors of N+N^{+} are split in KK and NN^{-} is a squarefree product of primes inert in KK. (In particular, H0(GK,T¯f)=0.H^{0}(G_{K},\overline{T}_{f})=0.) We shall denote by π\pi a uniformizer of 𝒪\mathcal{O}. From now on, we additionally assume that the maximal ideal 𝔪\mathfrak{m} associated to TfT_{f} satisfies (TW) above, i.e.:

(TW) if p=3, then T¯f is absolutely irreducible over 3.\text{if }p=3,\text{ then }\overline{T}_{f}\text{ is absolutely irreducible over }\mathbb{Q}{\sqrt{-3}}.

5.1.2.

We say a prime qNq\nmid N is weakly admissible with sign ϵq=±1\epsilon_{q}=\pm 1 if qq is inert in KK, aqϵq(q+1)(mod),a_{q}\equiv\epsilon_{q}(q+1)\pmod{\wp}, and q1(modp).q\not\equiv 1\pmod{p}. A weakly admissible pair {Q,ϵQ}\left\{Q,\epsilon_{Q}\right\} is an ordered pair of a squarefree number QQ and a function ϵQ:{q|Q}{±1}\epsilon_{Q}:\left\{q|Q\right\}\to\left\{\pm 1\right\} such that qq is weakly admissible with sign ϵQ(q)\epsilon_{Q}(q) for all q|Qq|Q. If {Q,ϵQ}\left\{Q,\epsilon_{Q}\right\} is a weakly admissible pair, then for all q|Qq|Q, there is a unique root uq𝒪u_{q}\in\mathcal{O} of the polynomial y2yaq+qy^{2}-ya_{q}+q such that uqϵQ(q)(mod)u_{q}\equiv\epsilon_{Q}(q)\pmod{\wp}. We may view 𝒪\mathcal{O} as a 𝕋N+Q,N\mathbb{T}_{N^{+}Q,N^{-}}-algebra by letting UqU_{q} act through uqu_{q}; let 𝔪QϵQ\mathfrak{m}_{Q}^{\epsilon_{Q}} be the associated maximal ideal (we will usually drop the superscript).

If Q=QQ′′Q=Q^{\prime}Q^{\prime\prime}, then we abbreviate 𝕋Q′′Q=𝕋N+Q,NQ′′\mathbb{T}^{Q^{\prime}}_{Q^{\prime\prime}}=\mathbb{T}_{N^{+}Q^{\prime},N^{-}Q^{\prime\prime}}, omitting any superscript or subscript which is equal to 1.

5.1.3.

In light of the structural similarity of Theorems 4.2.2 and 4.3.2, let

(34) MQ={Hom(JminN+,NQ,JN+,NQ),ν(NQ) even,[XN+,NQ]0,ν(NQ) odd.M_{Q}=\begin{cases}\operatorname{Hom}(J_{\min}^{N^{+},N^{-}Q},J^{N^{+},N^{-}Q}),&\nu(N^{-}Q)\text{ even},\\ \mathbb{Z}[X_{N^{+},N^{-}Q}]^{0},&\nu(N^{-}Q)\text{ odd}.\end{cases}

It is well-known that MQM_{Q} is a faithful 𝕋Q\mathbb{T}_{Q}-module, and indeed MQM_{Q}\otimes\mathbb{Q} is free of rank one over 𝕋Q\mathbb{T}_{Q}\otimes\mathbb{Q}.

Lemma 5.1.4.

Suppose {Q,ϵQ}\left\{Q,\epsilon_{Q}\right\} is a weakly admissible pair, and let

C=|NT¯f unram at ordπ(1).C=\sum_{\begin{subarray}{c}\ell|N^{-}\\ \overline{T}_{f}\text{ unram at }\ell\end{subarray}}\operatorname{ord}_{\pi}(\ell-1).

Then there exists an 𝒪\mathcal{O}-module map

MQ𝕋Q𝒪𝕋Q𝕋Q𝒪M_{Q}\otimes_{\mathbb{T}^{Q}}\mathcal{O}\to\mathbb{T}_{Q}\otimes_{\mathbb{T}^{Q}}\mathcal{O}

with kernel and cokernel annihilated by πC\pi^{C}; in particular, πC(MQ𝒪)\pi^{C}(M_{Q}\otimes\mathcal{O}) is principal of length at least lg(𝕋Q𝒪)2C.\lg(\mathbb{T}_{Q}\otimes\mathcal{O})-2C.

Proof.

We may assume that 𝔪Q𝕋Q\mathfrak{m}_{Q}\subset\mathbb{T}^{Q} descends to 𝕋Q\mathbb{T}_{Q}. Now, by Theorems 4.2.2 and 4.3.2, we have

MQ,𝔪Q|NQ𝒳(JminN+,NQ,)𝔪Q,M_{Q,\mathfrak{m}_{Q}}\simeq\otimes_{\ell|N^{-}Q}\mathcal{X}_{\ell}\left(J_{\min}^{N^{+},N^{-}Q,\vee}\right)_{\mathfrak{m}_{Q}},

modulo \mathbb{Z}-torsion on the right. Lemma 4.1.4 implies that there exist a collection of 𝕋Q\mathbb{T}_{Q}-module maps

ϕi:MQ,𝔪Q𝕋Q,𝔪Q,ψi:𝕋Q,𝔪QMQn,𝔪Qn,i=1,,r\phi_{i}:M_{Q,\mathfrak{m}_{Q}}\to\mathbb{T}_{Q,\mathfrak{m}_{Q}},\;\;\psi_{i}:\mathbb{T}_{Q,\mathfrak{m}_{Q}}\to M_{Q_{n},\mathfrak{m}_{Q_{n}}},\;\;i=1,\ldots,r

such that

ϕiψi=ψiϕi=ti𝕋Q,𝔪QEnd(MQn,𝔪Qn)\phi_{i}\circ\psi_{i}=\psi_{i}\circ\phi_{i}=t_{i}\in\mathbb{T}_{Q,\mathfrak{m}_{Q}}\subset\operatorname{End}(M_{Q_{n},\mathfrak{m}_{Q_{n}}})

and

t1++tr=|NT¯f unram at (1)𝕋Q,𝔪Q.t_{1}+\ldots+t_{r}=\prod_{\begin{subarray}{c}\ell|N^{-}\\ \overline{T}_{f}\text{ unram at }\ell\end{subarray}}(\ell-1)\in\mathbb{T}_{Q,\mathfrak{m}_{Q}}.

Since 𝒪\mathcal{O} is principal, we may choose some ii such that the image of tit_{i} in 𝕋Q𝒪\mathbb{T}_{Q}\otimes\mathcal{O} divides πC\pi^{C}. Then ϕi\phi_{i} and ψi\psi_{i} induce 𝒪\mathcal{O}-module maps

MQ𝒪𝕋Q𝒪MQ𝒪M_{Q}\otimes\mathcal{O}\to\mathbb{T}_{Q}\otimes\mathcal{O}\to M_{Q}\otimes\mathcal{O}

whose composition in either direction is multiplication by a divisor of πC\pi^{C}, which implies the result. ∎

Theorem 5.1.5.

If {Qq,ϵQq}\left\{Qq,\epsilon_{Qq}\right\} is a weakly admissible pair, then

lg(𝕋Qq𝕋Qq𝒪)lg(𝕋Q𝕋Q𝒪aqϵq(q+1))C,\lg(\mathbb{T}_{Qq}\otimes_{\mathbb{T}^{Qq}}\mathcal{O})\geq\lg\left(\frac{\mathbb{T}_{Q}\otimes_{\mathbb{T}^{Q}}\mathcal{O}}{a_{q}-\epsilon_{q}(q+1)}\right)-C,

where CC is the number of Lemma 5.1.4.

Proof.

The proof depends on the parity of ν(NQ)\nu(N^{-}Q).

Case 1.

ν(NQ)\nu(N^{-}Q) is even.

Let us abbreviate JQ=JN+,NQJ^{Q}=J^{N^{+},N^{-}Q} and JminQ=JminN+,NQJ^{Q}_{\min}=J^{N^{+},N^{-}Q}_{\min}. Consider the composite

MQqJQ(𝔽q2)H1(𝔽q2,T𝔪QJQ)𝔪QqMQT𝔪QJminQ(Uqϵq)M_{Qq}\to J^{Q}(\mathbb{F}_{q^{2}})\to H^{1}\left(\mathbb{F}_{q}^{2},T_{\mathfrak{m}_{Q}}J^{Q}\right)_{\mathfrak{m}_{Qq}}\simeq M_{Q}\otimes\frac{T_{\mathfrak{m}_{Q}}J^{Q}_{\min}}{(U_{q}-\epsilon_{q})}

induced from the diagram of Proposition 4.4.6, the Kummer map, and Proposition 4.1.3. These are surjective maps of 𝕋Qq\mathbb{T}_{Q}^{q}-modules, where UqU_{q} acts on the three latter modules through Frobq\operatorname{Frob}_{q}. Since T𝔪QJminQT_{\mathfrak{m}_{Q}}J_{\min}^{Q} is free of rank two over 𝕋Q,𝔪Q\mathbb{T}_{Q,\mathfrak{m}_{Q}} and Frobq\operatorname{Frob}_{q} acts with the characteristic polynomial Frobq2TqFrobq+q\operatorname{Frob}_{q}^{2}-T_{q}\operatorname{Frob}_{q}+q (whose roots are distinct modulo 𝔪Q\mathfrak{m}_{Q}), we may fix an identification

T𝔪QJminQUqϵq𝕋Q,𝔪QTqϵq(q+1),\frac{T_{\mathfrak{m}_{Q}}J^{Q}_{\min}}{U_{q}-\epsilon_{q}}\simeq\frac{\mathbb{T}_{Q,\mathfrak{m}_{Q}}}{T_{q}-\epsilon_{q}(q+1)},

considered as a 𝕋Q,𝔪Qqq\mathbb{T}^{q}_{Q,\mathfrak{m}_{Qq}}-module again through UqU_{q} acting by ϵq\epsilon_{q}. Tensoring with 𝒪\mathcal{O}, we obtain a surjective map

MQqMQ𝕋Qq𝒪aqϵq(q+1),M_{Qq}\to M_{Q}\otimes_{\mathbb{T}^{Qq}}\frac{\mathcal{O}}{a_{q}-\epsilon_{q}(q+1)},

hence (by Lemma 5.1.4) a map of 𝕋Qq\mathbb{T}^{Qq}-modules MQq𝕋Q𝒪aqϵq(q+1)M_{Qq}\to\frac{\mathbb{T}_{Q}\otimes\mathcal{O}}{a_{q}-\epsilon_{q}(q+1)} with cokernel annihilated by πC\pi^{C}. Since the action of 𝕋Qq\mathbb{T}^{Qq} on MQqM_{Qq} factors through 𝕋Qq\mathbb{T}_{Qq}, we obtain, by taking eigenvalues, a surjection 𝕋Qq𝒪/πj\mathbb{T}_{Qq}\to\mathcal{O}/\pi^{j} for some

jlg(𝕋Q𝕋Q𝒪aqϵq(q+1))C.j\geq\lg\left(\frac{\mathbb{T}_{Q}\otimes_{\mathbb{T}^{Q}}\mathcal{O}}{a_{q}-\epsilon_{q}(q+1)}\right)-C.
Case 2.

ν(NQ)\nu(N^{-}Q) is odd.

By Proposition 4.4.9, the action of 𝕋Q,𝔪Qqq\mathbb{T}^{q}_{Q,\mathfrak{m}_{Qq}} on

MQ,𝔪Q𝕋Q(𝕋Q2/im(Tqϵqq1ϵq)),M_{Q,\mathfrak{m}_{Q}}\otimes_{\mathbb{T}_{Q}}\left(\mathbb{T}_{Q}^{2}/\operatorname{im}\begin{pmatrix}T_{q}-\epsilon_{q}&q\\ -1&-\epsilon_{q}\end{pmatrix}\right),

with UqU_{q} acting by ϵq\epsilon_{q}, factors through 𝕋Qq,𝔪Qq\mathbb{T}_{Qq,\mathfrak{m}_{Qq}}. Hence the action of 𝕋Qq\mathbb{T}^{q}_{Q} on

A=MQ𝕋Q𝒪aqϵq(q+1)A=M_{Q}\otimes_{\mathbb{T}^{Q}}\frac{\mathcal{O}}{a_{q}-\epsilon_{q}(q+1)}

likewise factors through 𝕋Qq\mathbb{T}_{Qq} (again with UqU_{q} acting by ϵq\epsilon_{q}). The conclusion of Lemma 5.1.4 implies that AA has a 𝕋Qq\mathbb{T}^{q}_{Q}-module map to

𝕋Q𝕋Q𝒪aqϵq(q)\frac{\mathbb{T}_{Q}\otimes_{\mathbb{T}^{Q}}\mathcal{O}}{a_{q}-\epsilon_{q}(q)}

with cokernel annihilated by πC\pi^{C}, from which the result follows. ∎

Remark 5.1.6.

If {𝖰,ϵ𝖰}𝖭\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}, then for 𝔉\mathfrak{F}-many nn there is a corresponding weakly admissible pair {Qn,ϵQn}\left\{Q_{n},\epsilon_{Q_{n}}\right\}, where QnQ_{n} is a sequence representing 𝖰\mathsf{Q}. To be precise, if 𝖰={𝗊1,,𝗊r}\mathsf{Q}=\left\{\mathsf{q}_{1},\ldots,\mathsf{q}_{r}\right\}, we choose sequences qinq_{i}^{n} representing each qiq_{i}; for 𝔉\mathfrak{F}-many nn, the product Qn=q1nqrnQ_{n}=q_{1}^{n}\cdots q_{r}^{n}, equipped with sign function ϵQn(qin)=ϵ𝖰(𝗊i)\epsilon_{Q_{n}}(q_{i}^{n})=\epsilon_{\mathsf{Q}}(\mathsf{q}_{i}), forms a weakly admissible pair {Qn,ϵQn}\left\{Q_{n},\epsilon_{Q_{n}}\right\}. It follows from the definition of 𝖭\mathsf{N} and from the theorem that, for any j0,j\geq 0, there exist 𝔉\mathfrak{F}-many nn such that 𝕋Qn𝒪𝒪/πj\mathbb{T}^{Q_{n}}\to\mathcal{O}\to\mathcal{O}/\pi^{j} factors through 𝕋Qn\mathbb{T}_{Q_{n}}. We say that a sequence of weakly admissible pairs {Qn,ϵQn}\left\{Q_{n},\epsilon_{Q_{n}}\right\} (defined for 𝔉\mathfrak{F}-many nn) represents the pair {𝖰,ϵ𝖰}\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\} if it is obtained from this construction for some choice of representatives qinq_{i}^{n}.

5.2. The CM class construction

5.2.1.

Let SMS\subset M_{\mathbb{Q}} be a finite set of primes, and fix an element

y[𝒞S)]0;y\in\mathbb{Z}[\mathscr{C}^{S})]^{0};

let K[y]K[S]K[y]\subset K[S] be its field of definition and Gy=Gal(K[y]/K)G_{y}=\operatorname{Gal}(K[y]/K) the corresponding Galois group. A weakly admissible prime qq with sign ϵq\epsilon_{q} is called jj-admissible if aqϵq(q+1)(modπj)a_{q}\equiv\epsilon_{q}(q+1)\pmod{\pi^{j}}; in this case, TjTf/πjT_{j}\coloneqq T_{f}/\pi^{j} has a unique subspace Filq,ϵq+Tj\operatorname{Fil}_{q,\epsilon_{q}}^{+}T_{j}, free of rank one over 𝒪j𝒪/πj\mathcal{O}_{j}\coloneqq\mathcal{O}/\pi^{j}, on which Frobq\operatorname{Frob}_{q} acts as qϵqq\epsilon_{q}. We will omit the subscript ϵq\epsilon_{q} when there is no risk of confusion. A weakly admissible set {Q,ϵQ}\left\{Q,\epsilon_{Q}\right\} is called jj-admissible if lg(𝕋Q𝕋Q𝒪)j+2C\lg(\mathbb{T}_{Q}\otimes_{\mathbb{T}^{Q}}\mathcal{O})\geq j+2C; note that each q|Qq|Q is then necessarily jj-admissible (since Uq=ϵqU_{q}=\epsilon_{q} in 𝕋Q𝒪\mathbb{T}_{Q}\otimes\mathcal{O}). Let NjN_{j} be the collection of jj-admissible sets.

For any jj-admissible prime with sign ϵq\epsilon_{q}, we define the ordinary subspace:

(35) Hord,ϵq1(Kq,Tj)=im(H1(Kq,Filq,ϵq+Tj)H1(Kq,Tj)).H^{1}_{\operatorname{ord},\epsilon_{q}}(K_{q},T_{j})=\operatorname{im}\left(H^{1}(K_{q},\operatorname{Fil}^{+}_{q,\epsilon_{q}}T_{j})\to H^{1}(K_{q},T_{j})\right).

Using the map obtained from Schapiro’s Lemma (e.g. [47, §3.1.2])

(36) Resq:H1(K[y],Tj)HomSet(Gy,H1(Kq,Tj)),\operatorname{Res}_{q}:H^{1}(K[y],T_{j})\to\operatorname{Hom}_{\text{Set}}(G_{y},H^{1}(K_{q},T_{j})),

we also have maps:

q,ϵq:H1(K[y],Tj)\displaystyle\partial_{q,\epsilon_{q}}:H^{1}(K[y],T_{j}) HomSet(Gy,H1(Iq,Filq+Tj))𝒪j[Gy],\displaystyle\to\operatorname{Hom}_{\text{Set}}(G_{y},H^{1}(I_{q},\operatorname{Fil}_{q}^{+}T_{j}))\approx\mathcal{O}_{j}[G_{y}],
locq,ϵq:H1(K[y]Σ/K[y],Tj)\displaystyle\operatorname{loc}_{q,\epsilon_{q}}:H^{1}(K[y]^{\Sigma}/K[y],T_{j}) HomSet(Gy,Tj/Filq+Tj),𝒪j[Gy],qΣ,\displaystyle\to\operatorname{Hom}_{\text{Set}}(G_{y},T_{j}/\operatorname{Fil}_{q}^{+}T_{j}),\approx\mathcal{O}_{j}[G_{y}],\;\;q\not\in\Sigma,

defined as in (15,16).

Construction 5.2.2.

If ΣM\Sigma\subset M_{\mathbb{Q}} is the set of places dividing NpNp\infty, then for all {Q,ϵQ}Nj\left\{Q,\epsilon_{Q}\right\}\in N_{j}, there exist principal sub-𝒪j\mathcal{O}_{j}-modules:

(κj(y,Q,ϵQ))\displaystyle(\kappa_{j}(y,Q,\epsilon_{Q})) H1(K[y]ΣQ/K[y],Tj),\displaystyle\subset H^{1}(K[y]^{\Sigma\cup Q}/K[y],T_{j}), ν(NQ) even,\displaystyle\nu(N^{-}Q)\text{ even},
(λj(y,Q,ϵQ))\displaystyle(\lambda_{j}(y,Q,\epsilon_{Q})) 𝒪j[Gy],\displaystyle\subset\mathcal{O}_{j}[G_{y}], ν(NQ) odd,\displaystyle\nu(N^{-}Q)\text{ odd},

compatible under the natural reduction maps for jjj^{\prime}\leq j, and satisfying the following properties.

  1. (1)

    If {Qq,ϵQ}Nj\left\{Qq,\epsilon_{Q}\right\}\in N_{j} where ν(NQ)\nu(N^{-}Q) is even, then for all q|Qq|Q and all gGyg\in G_{y},

    Resq(κj(y,Q,ϵQ))(g)Hord,ϵQ(q)1(K𝗊,Tj).\operatorname{Res}_{q}(\kappa_{j}(y,Q,\epsilon_{Q}))(g)\subset H^{1}_{\operatorname{ord},\epsilon_{Q}(q)}(K_{\mathsf{q}},T_{j}).
  2. (2)

    If {Qq,ϵQq},{Q,ϵQ}Nj\left\{Qq,\epsilon_{Qq}\right\},\left\{Q,\epsilon_{Q}\right\}\in N_{j} where ϵQ=ϵQq|Q\epsilon_{Q}=\epsilon_{Qq}|_{Q} and ν(NQq)\nu(N^{-}Qq) is even, then

    q,ϵQq(q)(κj(y,Qq,ϵQq))=(λj(y,Q,ϵQ))𝒪j[Gy].\partial_{q,\epsilon_{Qq}(q)}(\kappa_{j}(y,Qq,\epsilon_{Qq}))=(\lambda_{j}(y,Q,\epsilon_{Q}))\subset\mathcal{O}_{j}[G_{y}].
  3. (3)

    If {Qq,ϵQq},{Q,ϵQ}Nj\left\{Qq,\epsilon_{Qq}\right\},\left\{Q,\epsilon_{Q}\right\}\in N_{j} where ϵQ=ϵQq|Q\epsilon_{Q}=\epsilon_{Qq}|_{Q} and ν(NQq)\nu(N^{-}Qq) is odd, then

    locq,ϵQq(q)(κj(y,Q,ϵQ))=(λj(y,Qq,ϵQq))𝒪j[Gy].\operatorname{loc}_{q,\epsilon_{Qq}(q)}(\kappa_{j}(y,Q,\epsilon_{Q}))=(\lambda_{j}(y,Qq,\epsilon_{Qq}))\subset\mathcal{O}_{j}[G_{y}].
Proof.

The specifications ϵQ\epsilon_{Q} will be dropped to ease notation. Suppose first that ν(NQ)\nu(N^{-}Q) is odd. By Lemma 5.1.4, there is a unique map (up to scalars) MQ𝒪jM_{Q}\to\mathcal{O}_{j} of 𝕋Q\mathbb{T}^{Q}-modules that factors through multiplication by πC\pi^{C} and is surjective after 𝒪\mathcal{O}-linearization. We define λj(y,Q)(g)\lambda_{j}(y,Q)(g) to be the image of gy[𝒞S]0gy\in\mathbb{Z}[\mathscr{C}^{S}]^{0} by the composite [𝒞S]0MQ𝒪j\mathbb{Z}[\mathscr{C}^{S}]^{0}\to M_{Q}\to\mathcal{O}_{j}.

Now suppose that ν(NQ)\nu(N^{-}Q) is even. Adopting the abbreviations of Case 1 of Theorem 5.1.5, CMN+,NQ(y)CM_{N^{+},N^{-}Q}(y) is a formal divisor on XQ=XN+,NQX_{Q}=X_{N^{+},N^{-}Q} for each nn, and its image in JQJ^{Q} is defined over K[y]K[y]. Let

d(y,Q)H1(K[y]ΣQ/K[y],T𝔪QJQ)d(y,Q)\in H^{1}(K[y]^{\Sigma\cup Q}/K[y],T_{\mathfrak{m}_{Q}}J^{Q})

be the Kummer image. By Lemma 5.1.4 and Proposition 4.1.3, there is a unique (up to scalars) map of 𝕋Q[G]\mathbb{T}^{Q}[G_{\mathbb{Q}}]-modules T𝔪QJQTjT_{\mathfrak{m}_{Q}}J^{Q}\to T_{j} that is surjective after 𝒪\mathcal{O}-linearization. We define κj(y,Q)\kappa_{j}(y,Q) to be the image of d(y,Q)d(y,Q) under the induced map

H1(K[y]ΣQ/K[y],T𝔪QJQ)H1(K[y]ΣQ/K[y],Tj).H^{1}(K[y]^{\Sigma\cup Q}/K[y],T_{\mathfrak{m}_{Q}}J^{Q})\to H^{1}(K[y]^{\Sigma\cup Q}/K[y],T_{j}).

Note that, for each gGyg\in G_{y}, Resqκj(y,Q)(g)\operatorname{Res}_{q}\kappa_{j}(y,Q)(g) is the local Kummer image of gCMN+,N(y)=CMN+,N(gy)g\cdot\operatorname{CM}_{N^{+},N^{-}}(y)=\operatorname{CM}_{N^{+},N^{-}}(gy) by Shimura’s reciprocity law.

  1. (1)

    This follows from the rigid analytic uniformization (33) by the argument of [20, p. 15]. Indeed, the argument there shows that the image of the Kummer map

    JQq(q2)H1(q2,T𝔪QJQ)J^{Qq}(\mathbb{Q}_{q^{2}})\to H^{1}(\mathbb{Q}_{q^{2}},T_{\mathfrak{m}_{Q}}J^{Q})

    agrees with the image of the map

    H1(q2,𝒳(JQ)(1))H1(qn2,T𝔪QJQ),H^{1}(\mathbb{Q}_{q^{2}},\mathcal{X}(J^{Q})(1))\to H^{1}(\mathbb{Q}_{q_{n}^{2}},T_{\mathfrak{m}_{Q}}J^{Q}),

    where 𝒳(JQ)(1)(T𝔪QJQ)[FrobqUqq]\mathcal{X}(J^{Q})(1)\to(T_{\mathfrak{m}_{Q}}J^{Q})[\operatorname{Frob}_{q}-U_{q}q] is a canonical isomorphism. Since qp1,q\nmid p-1, the surjection

    T𝔪QJQTjT_{\mathfrak{m}_{Q}}J^{Q}\twoheadrightarrow T_{j}

    induces

    𝒳(JQ)(1)Filq+Tj,\mathcal{X}(J^{Q})(1)\twoheadrightarrow\operatorname{Fil}^{+}_{q}T_{j},

    and the claim follows.

  2. (2)

    We claim that the composite

    JQq(q2)H1(q2,T𝔪QJQ)H1(Iq,Tj)J^{Qq}(\mathbb{Q}_{q^{2}})\to H^{1}(\mathbb{Q}_{q^{2}},T_{\mathfrak{m}_{Q}}J^{Q})\to H^{1}(I_{q},T_{j})

    factors through

    Spq:JQq(q2)ΦQ,𝔪Q.\operatorname{Sp}_{q}:J^{Qq}(\mathbb{Q}_{q^{2}})\to\Phi_{Q,\mathfrak{m}_{Q}}.

    Indeed, the target of the composite map has Frobenius eigenvalue UqU_{q}, and, because pq1p\nmid q-1, a diagram chase using (33) shows that the pro-pp part of the kernel of Spq\operatorname{Sp}_{q} has Frobenius eigenvalue Uq-U_{q} (if it is nontrivial at all).

    Using the commutativity of the diagram in Proposition 4.4.9 (and the fixed embedding K[S]q2K[S]\hookrightarrow\mathbb{Q}_{q^{2}}), q(κj(y,𝖰))(g)\partial_{q}(\kappa_{j}(y,\mathsf{Q}))(g) is therefore the image of gygy under the composite of the canonical map

    [𝒞S]0MQq\mathbb{Z}[\mathscr{C}^{S}]^{0}\to M_{Qq}

    with some surjective map of Hecke modules MQq𝒪jM_{Qq}\to\mathcal{O}_{j}, which factors through multiplication by πC\pi^{C} by the choice of map T𝔪QJQTjT_{\mathfrak{m}_{Q}}J^{Q}\to T_{j}. We may conclude by Lemma 5.1.4.

  3. (3)

    The proof is very similar to (2), invoking instead Proposition 4.4.6.

6. From CM classes to bipartite Euler systems

6.1. pp-adic interpolation

6.1.1.

Suppose for this subsection that:

(spl) p splits in Kp\text{ splits in }K

and

(ord) ap.a_{p}\not\in\wp.

6.1.2.

Fix an auxiliary prime 0Np\ell_{0}\nmid Np such that a001𝒪×a_{\ell_{0}}-\ell_{0}-1\in\mathcal{O}^{\times} (one exists by the irreducibility of T¯f\overline{T}_{f}), and set S={0,p}S=\left\{\ell_{0},p\right\}. For each m0,m\geq 0, consider the point

ym,0=(pm001)GL2(p)𝒞S;y_{m,0}=\begin{pmatrix}p^{m}&0\\ 0&1\end{pmatrix}\in GL_{2}(\mathbb{Q}_{p})\subset\mathscr{C}^{S};

note that ym,0y_{m,0} is defined over K[pm]K[p^{m}]. If T0T_{\ell_{0}} is the usual adelic Hecke operator, then set ym=(T001)TrK[pm]/Kmym,0[𝒞S]0y_{m}=(T_{\ell_{0}}-\ell_{0}-1)\operatorname{Tr}_{K[p^{m}]/K_{m}}y_{m,0}\in\mathbb{Z}[\mathscr{C}^{S}]^{0}, where KmK_{m} is the mmth layer of the anticyclotomic p\mathbb{Z}_{p}-extension.

6.1.3.

Now suppose given any {𝖰,ϵ𝖰}\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}, and let {Qn,ϵQn}\left\{Q_{n},\epsilon_{Q_{n}}\right\} be a representative sequence of weakly admissible pairs as in Remark 5.1.6. Since Tp𝔪T_{p}\not\in\mathfrak{m}, Hensel’s Lemma implies that the Hecke algebras 𝕋Qn,𝔪Qn\mathbb{T}_{Q_{n},\mathfrak{m}_{Q_{n}}} contain a (unique) element u𝔪Qnu\not\in\mathfrak{m}_{Q_{n}} such that u2uap+p=0.u^{2}-ua_{p}+p=0. Suppose first that |𝖰|+ν(N)|\mathsf{Q}|+\nu(N^{-}) is even. By the usual Heegner point norm relations (cf. e.g. [13, Proposition 3.10]), the classes

d(ym,Qn)um+1d(ym,Qn)umResKm/Km1dn(ym1,Qn)d(y_{m},Q_{n})^{\prime}\coloneqq u^{-m+1}d(y_{m},Q_{n})-u^{-m}\operatorname{Res}_{K_{m}/K_{m-1}}d_{n}(y_{m-1},Q_{n})

of Construction 5.2.2 are compatible under the corestriction maps

H1(Km,T𝔪QnJQn)H1(Km1,T𝔪QnJQn).H^{1}(K_{m},T_{\mathfrak{m}_{Q_{n}}}J^{Q_{n}})\to H^{1}(K_{m-1},T_{\mathfrak{m}_{Q_{n}}}J^{Q_{n}}).

If we replace d(ym,Qn)d(y_{m},Q_{n}) by d(ym,Qn)d(y_{m},Q_{n})^{\prime}, then, for any jj and for 𝔉\mathfrak{F}-many nn, we obtain classes κj(ym,Qn)𝖧1(KmΣQn/Km,Tj)\kappa_{j}(y_{m},Q_{n})^{\prime}\in\mathsf{H}^{1}(K_{m}^{\Sigma\cup Q_{n}}/K_{m},T_{j}) that are compatible under corestriction (as long the choices in the construction are made compatibly as mm varies, which is clearly possible). We let

𝜿(𝖰)limm,j𝖧1(KmΣ𝖰/Km,Tj)𝖧1(K,TfΛ(Ψ))\bm{\kappa}(\mathsf{Q})\in\lim_{\begin{subarray}{c}\longleftarrow\\ {m,j}\end{subarray}}\mathsf{H}^{1}(K_{m}^{\Sigma\cup\mathsf{Q}}/K_{m},T_{j})\simeq\mathsf{H}^{1}(K,T_{f}\otimes\Lambda(\Psi))

be the class represented by the family κj(ym,Qn)\kappa_{j}(y_{m},Q_{n})^{\prime}.

6.1.4.

Similarly, if |𝖰|+ν(N)|\mathsf{Q}|+\nu(N^{-}) is odd, the elements

λj(ym,Qn)αpm+1λ(ym,Qn)αpm+1λ(ym1,Qn)𝒪j[Gal(Km/K)]\lambda_{j}(y_{m},Q_{n})^{\prime}\coloneqq\alpha_{p}^{-m+1}\lambda(y_{m},Q_{n})-\alpha_{p}^{-m+1}\lambda(y_{m-1},Q_{n})\in\mathcal{O}_{j}[\operatorname{Gal}(K_{m}/K)]

are compatible under the natural projection maps

𝒪j[Gal(Km/K)]𝒪j[Gal(Km1/K)].\mathcal{O}_{j}[\operatorname{Gal}(K_{m}/K)]\to\mathcal{O}_{j}[\operatorname{Gal}(K_{m-1}/K)].

We then obtain an element

𝝀(𝖰)limm,j𝒰({𝒪j[Gal(Km/K)]}n)𝒪Gal(K/K)Λ.\bm{\lambda}(\mathsf{Q})\in\lim_{\begin{subarray}{c}\leftarrow\\ m,j\end{subarray}}\mathcal{U}\left(\left\{\mathcal{O}_{j}[\operatorname{Gal}(K_{m}/K)]\right\}_{n\in\mathbb{N}}\right)\simeq\mathcal{O}\llbracket\operatorname{Gal}(K_{\infty}/K)\rrbracket\simeq\Lambda.

Let 𝖲𝖬K\mathsf{S}\subset\mathsf{M}_{K} be the set of constant ultraprimes v¯\underline{v} such that v|Npv|Np\infty. We define a Selmer structure (Λ,𝖲)(\mathcal{F}_{\Lambda},\mathsf{S}) for 𝐓TfΛ\mathbf{T}\coloneqq T_{f}\otimes\Lambda in the usual way (see e.g. [24, 6]):

𝖧Λ1(K𝗏,𝐓)={im(H1(Kv,Filv+𝐓)H1(Kv,𝐓)),𝗏=v¯,v|p,H1(Kv,𝐓),𝗏=v¯,vp,𝖧unr1(K𝗏,𝐓),otherwise.\mathsf{H}^{1}_{\mathcal{F}_{\Lambda}}(K_{\mathsf{v}},\mathbf{T})=\begin{cases}\operatorname{im}\left(H^{1}(K_{v},\operatorname{Fil}_{v}^{+}\mathbf{T})\to H^{1}(K_{v},\mathbf{T})\right),&\mathsf{v}=\underline{v},v|p,\\ H^{1}(K_{v},\mathbf{T}),&\mathsf{v}=\underline{v},v\nmid p,\\ \mathsf{H}^{1}_{\operatorname{unr}}(K_{\mathsf{v}},\mathbf{T}),&\text{otherwise}.\end{cases}

Here Filv+𝐓𝐓\operatorname{Fil}_{v}^{+}\mathbf{T}\subset\mathbf{T} is the unique free, rank-one direct summand on which IvI_{v} acts by the cyclotomic character. By [25, proposition 3.3.1], Λ\mathcal{F}_{\Lambda} extends naturally to an interpolated self-dual Selmer structure (𝖲,Λ,𝔓,ΣΛ)(\mathsf{S},\mathcal{F}_{\Lambda},\mathcal{F}_{\mathfrak{P}},\Sigma_{\Lambda}) for 𝐓\mathbf{T}.

Proposition 6.1.5.

The pair (𝛋,𝛌)(\bm{\kappa},\bm{\lambda}) is a nontrivial bipartite Euler system for the triple (𝐓,Λ,𝖲)(\mathbf{T},\mathcal{F}_{\Lambda},\mathsf{S}).

Proof.

We first show that 𝜿(𝖰)\bm{\kappa}(\mathsf{Q}) lies in SelΛ(𝖰)(𝐓)\operatorname{Sel}_{\mathcal{F}_{\Lambda}(\mathsf{Q})}(\mathbf{T}) for all {𝖰,ϵ𝖰}𝖭ν(N)\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}^{\nu(N^{-})}. The only local conditions to verify are those at v|pv|p; the local conditions for 𝗊𝖰\mathsf{q}\in\mathsf{Q} follow from Construction 5.2.2(1), and the rest are trivial. If 𝖰\mathsf{Q} is represented by the sequence QnQ_{n}, let Filv+T𝔪QnJQn\operatorname{Fil}_{v}^{+}T_{\mathfrak{m}_{Q_{n}}}J^{Q_{n}} be the maximal 𝕋𝔪Qn\mathbb{T}_{\mathfrak{m}_{Q_{n}}} submodule on which IvI_{v} acts by the cyclotomic character (adopting the notation of Construction 5.2.2 and if necessary restricting our attention to 𝔉\mathfrak{F}-many nn). As in [11, Proposition 4.7], it suffices to show that, for all mm and nn and a fixed extension of vv to KK_{\infty}, the image dn,md_{n,m} of the class d(ym,Qn)d(y_{m},Q_{n})^{\prime} under the composite

H1(Km,T𝔪QnJQn)H1(Km,v,T𝔪QnJQn/Fil+T𝔪QnJQn)H^{1}(K_{m},T_{\mathfrak{m}_{Q_{n}}}J^{Q_{n}})\to H^{1}(K_{m,v},T_{\mathfrak{m}_{Q_{n}}}J^{Q_{n}}/\operatorname{Fil}^{+}T_{\mathfrak{m}_{Q_{n}}}J^{Q_{n}})

is trivial. Since dn(ym)d_{n}(y_{m})^{\prime} is a 𝕋𝔪\mathbb{T}_{\mathfrak{m}}-linear combination of Kummer images over KmK_{m}, by [4, Example 3.11] and [36, Proposition 12.5.8] dn,md_{n,m} lies in the kernel of

H1(Km,v,T𝔪QnJQn/Fil+T𝔪QnJQn)H1(Km,v,pT𝔪QnJQn/Fil+T𝔪QnJQn).H^{1}(K_{m,v},T_{\mathfrak{m}_{Q_{n}}}J^{Q_{n}}/\operatorname{Fil}^{+}T_{\mathfrak{m}_{Q_{n}}}J^{Q_{n}})\to H^{1}(K_{m,v},\mathbb{Q}_{p}\otimes T_{\mathfrak{m}_{Q_{n}}}J^{Q_{n}}/\operatorname{Fil}^{+}T_{\mathfrak{m}_{Q_{n}}}J^{Q_{n}}).

Since the classes dn,md_{n,m} are corestriction-compatible as mm varies, the argument of [26, Proposition 2.4.5] shows that indeed dn,m=0d_{n,m}=0 for all n,mn,m.

The explicit reciprocity laws are a consequence of Construction 5.2.2(2,3), and the nonvanishing of either 𝜿(1)\bm{\kappa}(1) or 𝝀(1)\bm{\lambda}(1) (according to the parity of ν(N)\nu(N^{-})) is due to the work of Cornut [12] and Vatsal [50]. ∎

6.2. Kolyvagin classes

6.2.1.

Before defining the Kolyvagin classes in patched cohomology, we begin by recalling a calculation explained in [19].

Let mm be a squarefree product of primes \ell inert in KK, and let

y(m)0|mGL2()y(m)_{0}\in\prod_{\ell|m}GL_{2}(\mathbb{Q}_{\ell})

be the element with \ellth component (001)\begin{pmatrix}\ell&0\\ 0&1\end{pmatrix}. If 0\ell_{0} is the auxiliary prime of the previous subsection, and S{0}{|m}S\left\{\ell_{0}\right\}\cup\left\{\ell|m\right\}, then we define

y(m)=(T0(0+1))y(m)0[𝒞S]0.y(m)=(T_{\ell_{0}}-(\ell_{0}+1))y(m)_{0}\in\mathbb{Z}[\mathscr{C}^{S}]^{0}.

6.2.2.

Note that K[y(m)]=K[m]K[y(m)]=K[m] and that Gal(K[m]/K[1])|mGal(K[]/K[1]);\operatorname{Gal}(K[m]/K[1])\simeq\prod_{\ell|m}\operatorname{Gal}(K[\ell]/K[1]); each Gal(K[]/K[1])\operatorname{Gal}(K[\ell]/K[1]) is cyclic of order +1.\ell+1. For any place λ\lambda of ¯\overline{\mathbb{Q}} over \ell, fixed for the time being, let FrobλG\operatorname{Frob}_{\lambda}\in G_{\mathbb{Q}} be a lift of absolute Frobenius, and σλIλGK\sigma_{\lambda}\in I_{\lambda}\subset G_{K} a generator of Gal(K[]/K[1])\operatorname{Gal}(K[\ell]/K[1]). Recall the Kolyvagin derivative operators [19]:

D=i=1iσλi[Gal(K[]/K[1])],Dm=|mD.D_{\ell}=\sum_{i=1}^{\ell}i\sigma_{\lambda}^{i}\in\mathbb{Z}[\operatorname{Gal}(K[\ell]/K[1])],\;\;D_{m}=\prod_{\ell|m}D_{\ell}.

Let

P(m)=Dmy(m)[𝒞S]0.P(m)=D_{m}y(m)\in\mathbb{Z}[\mathscr{C}^{S}]^{0}.

Finally, let QQ be a set of primes inert in KK that is disjoint from SS, pp, and the factors of NN, and let P(m,Q)=CMN+,NQ(P(m)),P(m,Q)=\operatorname{CM}_{N^{+},N^{-}Q}(P(m)), y(m,Q)=CMN+,NQ(y(m)).y(m,Q)=\operatorname{CM}_{N^{+},N^{-}Q}(y(m)).

Proposition 6.2.3 ([19], Proposition 3.7).

For all |m\ell|m, we have:

  1. (1)

    (σ1)P(m)=(+1)Dm/y(m)TP(m/).(\sigma_{\ell}-1)P(m)=(\ell+1)D_{m/\ell}y(m)-T_{\ell}P(m/\ell).

  2. (2)

    Suppose ν(NQ)\nu(N^{-}Q) is even. If λ\lambda lies over \ell, then

    Dm/y(m,Q)FrobλP(m/,Q)(modλ).D_{m/\ell}y(m,Q)\equiv\operatorname{Frob}_{\lambda}P({m/\ell},Q)\pmod{\lambda}.
Proposition 6.2.4.

Suppose 𝔪Q𝕋Q=𝕋N+,NQ\mathfrak{m}_{Q}\subset\mathbb{T}_{Q}=\mathbb{T}_{N^{+},N^{-}Q} is a maximal ideal whose associated residual representation has no GK[m]G_{K[m]}-fixed points, and let Im𝕋QI_{m}\subset\mathbb{T}_{Q} be the ideal generated by +1\ell+1 and TT_{\ell} for all |m\ell|m. Then if ν(NQ)\nu(N^{-}Q) is even:

  1. (1)

    Restriction induces an isomorphism

    Resm:H1(K[1],T𝔪QJQ/Im)H1(K[m],T𝔪QJQ/Im)Gal(K[m]/K[1]).\operatorname{Res}_{m}:H^{1}(K[1],T_{\mathfrak{m}_{Q}}J^{Q}/I_{m})\xrightarrow{\sim}H^{1}(K[m],T_{\mathfrak{m}_{Q}}J^{Q}/I_{m})^{\operatorname{Gal}(K[m]/K[1])}.
  2. (2)

    The Kummer image d(m,Q)d^{\prime}(m,Q) of P(m,Q)P(m,Q) in H1(K[m],T𝔪QJQ/Im)H^{1}(K[m],T_{\mathfrak{m}_{Q}}J^{Q}/I_{m}) lies in the image of Resm\operatorname{Res}_{m}.

  3. (3)

    If c(m,Q)=CoresK[1]/KResm1d(m,Q)c(m,Q)=\operatorname{Cores}_{K[1]/K}\operatorname{Res}_{m}^{-1}d^{\prime}(m,Q), then for all |m\ell|m and any choices of representatives,

    c(m,Q)(σλ)=Frobλ1d(m/,Q)(Frobλ2)(modI)m.c(m,Q)(\sigma_{\lambda})=\operatorname{Frob}_{\lambda}^{-1}d^{\prime}(m/\ell,Q)(\operatorname{Frob}_{\lambda}^{2})\pmod{I}_{m}.
  4. (4)

    The class c(m,Q)c(m,Q) is unramified at any place vNpmQ.v\nmid NpmQ\infty.

Proof.

(1) follows from the inflation restriction exact sequence as in [19], and (2) is immediate from Proposition 6.2.3. Also (4) is clear from the construction. For (3), it suffices to check the corresponding statement for c(m,Q)=Resm1d(m,Q)c^{\prime}(m,Q)=\operatorname{Res}_{m}^{-1}d^{\prime}(m,Q). The proof is a modification of the argument in [34]. Fix division points P(m,Q)+1\frac{P(m,Q)}{\ell+1} and P(m/,Q)+1\frac{P(m/\ell,Q)}{\ell+1}; one may verify that c(m,Q)(σ)c^{\prime}(m,Q)(\sigma_{\ell}) is the unique element AT𝔪QJQ/ImA\in T_{\mathfrak{m}_{Q}}J^{Q}/I_{m} such that, for all gGK[m/]g\in G_{K[m/\ell]},

(g1)A\displaystyle(g-1)A (g1)(σ~1)P(m,Q)+1T𝔪QJQ/Im.\displaystyle\equiv(g-1)(\widetilde{\sigma}_{\ell}-1)\frac{P(m,Q)}{\ell+1}\in T_{\mathfrak{m}_{Q}}J^{Q}/I_{m}.

By Proposition 6.2.3(1), AA is also the image of the (unique) point TJQ[+1]T\in J^{Q}[\ell+1] such that

TDm/CMN+,NQ(P(m,Q))TP(m/,Q)+1(modλ).T\equiv D_{m/\ell}\operatorname{CM}_{N^{+},N^{-}Q}(P(m,Q))-T_{\ell}\frac{P(m/\ell,Q)}{\ell+1}\pmod{\lambda}.

But by Proposition 6.2.3(2), this is equivalent to

(37) TFrobλ1P(m/,Q)TP(m/,Q)+1(modλ).T\equiv\operatorname{Frob}_{\lambda}^{-1}P(m/\ell,Q)-T_{\ell}\frac{P(m/\ell,Q)}{\ell+1}\pmod{\lambda}.

By the Eichler-Shimura relation, and the fact that \ell splits completely in K[m/]K[m/\ell], the image of TT in T𝔪QJQ/ImT_{\mathfrak{m}_{Q}}J^{Q}/I_{m} is precisely

Frobλ1d(m/,Q)(Frobλ2).\operatorname{Frob}_{\lambda}^{-1}d^{\prime}(m/\ell,Q)(\operatorname{Frob}_{\lambda}^{2}).

Definition 6.2.5.

For a squarefree product mm of primes inert in KK, let Im(f)𝒪I_{m}(f)\subset\mathcal{O} be the ideal generated by a(f)a_{\ell}(f) and +1\ell+1 for all |m.\ell|m. Suppose {Q,ϵQ}𝒩j\left\{Q,\epsilon_{Q}\right\}\in\mathcal{N}_{j} is jj-admissible and jv(Im(f)).j\geq v_{\wp}(I_{m}(f)). If ν(NQ)\nu(N^{-}Q) is even, then the Kolyvagin class

(38) c¯(m,Q)H1(KΣQm/K,Tf/Im(f))\overline{c}(m,Q)\in H^{1}(K^{\Sigma\cup Q\cup m}/K,T_{f}/I_{m}(f))

is defined to be the image of c(m,Q)c(m,Q). If ν(NQ)\nu(N^{-}Q) is odd, then the reduction of λ(P(m),Q)\lambda(P(m),Q) in

(𝒪/Im)[Gal(K[m]/K)](\mathcal{O}/I_{m})[\operatorname{Gal}(K[m]/K)]

is constant on cosets of Gal(K[m]/K[1])\operatorname{Gal}(K[m]/K[1]) by Proposition 6.2.3(1) and therefore descends to

(39) λ(m,Q)(𝒪/Im)[Gal(K[1]/K)].\lambda^{\prime}(m,Q)\in(\mathcal{O}/I_{m})[\operatorname{Gal}(K[1]/K)].

The Kolyvagin element is then defined as:

(40) λ(m,Q)=trK[1]/Kλ(P(m),Q)𝒪/Im.\lambda(m,Q)=\operatorname{tr}_{K[1]/K}\lambda^{\prime}(P(m),Q)\in\mathcal{O}/I_{m}.
Remark 6.2.6.

When Q=1Q=1 and ν(N)\nu(N^{-}) is even, this agrees with Kolyvagin’s construction [31].

For applications to the parity conjecture for ff, we will require the following:

Proposition 6.2.7.

If ν(N)\nu(N^{-}) is even, then c¯(m,1)\overline{c}(m,1) lies in the ϵf(1)ν(m)+1{\epsilon_{f}}\cdot(-1)^{\nu(m)+1}-eigenspace for the action of τ\tau. If ν(N)\nu(N^{-}) is odd and λ(m,1)0,\lambda(m,1)\neq 0, then ϵf=(1)ν(m)\epsilon_{f}=(-1)^{\nu(m)}.

Proof.

The maps T𝔪JN+,NTf/πjT_{\mathfrak{m}}J^{N^{+},N^{-}}\to T_{f}/\pi^{j} or [XN+,N]0𝒪/πj\mathbb{Z}[X_{N^{+},N^{-}}]^{0}\to\mathcal{O}/\pi^{j} used in Construction 5.2.2 are equivariant for the action of the Atkin-Lehner involution because of the uniqueness property derived from Lemma 5.1.4. (Note this is not necessarily true of the corresponding maps at level N+NQN^{+}N^{-}Q, which are not necessarily reductions of genuine modular parameterizations.) Since the Atkin-Lehner eigenvalue of ff is ϵf-\epsilon_{f}, the proposition follows exactly as in [19, Proposition 5.4]. ∎

Definition 6.2.8.

An ultraprime 𝗅\mathsf{l} is called Kolyvagin-admissible if

Frob𝗅Gal(K(Tf)/)\operatorname{Frob}_{\mathsf{l}}\in\operatorname{Gal}(K(T_{f})/\mathbb{Q})

is a complex conjugation. A Kolyvagin-admissible set is a finite set of Kolyvagin-admissible ultraprimes, and the collection of all Kolyvagin-admissible sets is denoted 𝖪\mathsf{K}.

6.2.9.

If 𝗅\mathsf{l} is Kolyvagin-admissible, then the local cohomology

𝖧1(K𝗅,Tf)\mathsf{H}^{1}(K_{\mathsf{l}},T_{f})

is free of rank four over 𝒪\mathcal{O}, and carries a natural action of the complex conjugation τGal(K/)\tau\in\operatorname{Gal}(K/\mathbb{Q}). It has a canonical splitting of the finite-singular exact sequence:

𝖧1(K𝗅,Tf)=𝖧unr1(K𝗅,Tf)𝖧tr1(K𝗅,Tf),\mathsf{H}^{1}(K_{\mathsf{l}},T_{f})=\mathsf{H}^{1}_{\operatorname{unr}}(K_{\mathsf{l}},T_{f})\oplus\mathsf{H}^{1}_{\operatorname{tr}}(K_{\mathsf{l}},T_{f}),

defined as follows. If the sequence n\ell_{n} represents 𝗅\mathsf{l}, then for any jj and for 𝔉\mathfrak{F}-many nn, Frobn\operatorname{Frob}_{\ell_{n}} acts as complex conjugation on Tf/πjT_{f}/\pi^{j}, and

Htr1(Kn,Tf/πj)=ker(H1(Kn,Tf/πj)H1(K[n]λn,Tf/πj))H^{1}_{\operatorname{tr}}(K_{\ell_{n}},T_{f}/\pi^{j})=\ker\left(H^{1}(K_{\ell_{n}},T_{f}/\pi^{j})\to H^{1}(K[\ell_{n}]_{\lambda_{n}},T_{f}/\pi^{j})\right)

is isomorphic to H1(In,Tf/πj)Frob2=1,H^{1}(I_{\ell_{n}},T_{f}/\pi^{j})^{\operatorname{Frob}_{\ell}^{2}=1}, where λn\lambda_{n} is a prime of K[n]K[\ell_{n}] over n\ell_{n}. Then

Htr1(K𝗅,Tf)=lim𝒰({Htr1(Kn,Tf/πj)}n)𝖧1(K𝗅,Tf)H^{1}_{\operatorname{tr}}(K_{\mathsf{l}},T_{f})=\lim_{\leftarrow}\mathcal{U}\left(\left\{H^{1}_{\operatorname{tr}}(K_{\ell_{n}},T_{f}/\pi^{j})\right\}_{n\in\mathbb{N}}\right)\subset\mathsf{H}^{1}(K_{\mathsf{l}},T_{f})

is our transverse subspace. We denote by loc𝗅±\operatorname{loc}_{\mathsf{l}}^{\pm} and 𝗅±\partial_{\mathsf{l}}^{\pm} the composites 𝖧1(K,Tf)𝖧unr1(K𝗅,Tf)±\mathsf{H}^{1}(K,T_{f})\to\mathsf{H}^{1}_{\operatorname{unr}}(K_{\mathsf{l}},T_{f})^{\pm} and 𝖧1(K,Tf)𝖧tr1(K𝗅,Tf)±\mathsf{H}^{1}(K,T_{f})\to\mathsf{H}^{1}_{\operatorname{tr}}(K_{\mathsf{l}},T_{f})^{\pm}, respectively, where ±\pm is the Frobenius eigenvalue. The codomain of each is free of rank one over 𝒪\mathcal{O}.

Let 𝖲𝖬K\mathsf{S}\subset\mathsf{M}_{K} be the set of constant ultraprimes v¯\underline{v} such that vNpv\nmid Np\infty. We will consider the Kolyvagin-transverse Selmer structure ((𝗆),𝖲𝗆)(\mathcal{F}(\mathsf{m}),\mathsf{S}\cup\mathsf{m}) on TfT_{f}, for any 𝗆𝖪\mathsf{m}\in\mathsf{K}:

(41) 𝖧(𝗆)1(K𝗏,Tf)={im(A(Kv)𝒪f𝒪H1(Kv,Tf)),𝗏=v¯,𝖧tr1(K𝗅,Tf),𝗏=𝗅𝗆,𝖧unr1(K𝗏,Tf),otherwise.\mathsf{H}^{1}_{\mathcal{F}(\mathsf{m})}(K_{\mathsf{v}},T_{f})=\begin{cases}\operatorname{im}\left(A(K_{v})\otimes_{\mathcal{O}_{f}}\mathcal{O}\to H^{1}(K_{v},T_{f})\right),&\mathsf{v}=\underline{v},\\ \mathsf{H}^{1}_{\operatorname{tr}}(K_{\mathsf{l}},T_{f}),&\mathsf{v}=\mathsf{l}\in\mathsf{m},\\ \mathsf{H}^{1}_{\operatorname{unr}}(K_{\mathsf{v}},T_{f}),&\text{otherwise}.\end{cases}

Here AA is an (optimal) abelian variety with real multiplication by 𝒪f\mathcal{O}_{f} associated to ff. If {𝖰,ϵ𝖰}𝖭𝗆\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}_{\mathsf{m}}, then we denote by ((𝗆,𝖰),𝖲𝗆𝖰)(\mathcal{F}(\mathsf{m},\mathsf{Q}),\mathsf{S}\cup\mathsf{m}\cup\mathsf{Q}) the modified Selmer structure of (3.2.2).

6.2.10.

Let {𝖰,ϵ𝖰}𝖭𝗆ν(N)\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}_{\mathsf{m}}^{\nu(N^{-})}, and fix representatives QnQ_{n} and mnm_{n}, which we may assume to be disjoint. Our patched Kolyvagin class is the element

κ(𝗆,𝖰)𝖧1(K𝖲𝗆𝖰/K,Tf)\kappa(\mathsf{m},\mathsf{Q})\in\mathsf{H}^{1}(K^{\mathsf{S}\cup\mathsf{m}\cup\mathsf{Q}}/K,T_{f})

whose image in TjT_{j} is represented by the sequence of images of the classes c¯(mn,Qn)\overline{c}(m_{n},Q_{n}), well-defined for 𝔉\mathfrak{F}-many nn.

If {𝖰,ϵ𝖰}𝖭ν(N)+1,\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}^{\nu(N^{-})+1}, then we similarly set

λ(𝗆,𝖰)𝒪lim𝒰({𝒪/πj})\lambda(\mathsf{m},\mathsf{Q})\in\mathcal{O}\simeq\lim_{\longleftarrow}\mathcal{U}\left(\left\{\mathcal{O}/\pi^{j}\right\}\right)

to be the element whose image in 𝒪/πj\mathcal{O}/\pi^{j} is represented by the sequence λ(mn,Qn)\lambda(m_{n},Q_{n}).

Proposition 6.2.11.

For any 𝗆𝖪\mathsf{m}\in\mathsf{K} and {𝖰,ϵ𝖰}𝖭𝗆ν(N)\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}_{\mathsf{m}}^{\nu(N^{-})},

(κ(𝗆,𝖰))Sel(𝗆,𝖰)(Tf).(\kappa(\mathsf{m},\mathsf{Q}))\subset\operatorname{Sel}_{\mathcal{F}(\mathsf{m},\mathsf{Q})}(T_{f}).

Moreover:

  1. (1)

    For all 𝗅𝗆\mathsf{l}\in\mathsf{m},

    (loc𝗅±(κ(𝗆/𝗅,𝖰)))=(𝗅(κ(𝗆,𝖰))(\operatorname{loc}_{\mathsf{l}}^{\pm}(\kappa(\mathsf{m}/\mathsf{l},\mathsf{Q})))=(\partial_{\mathsf{l}}^{\mp}(\kappa(\mathsf{m},\mathsf{Q}))

    as submodules of 𝒪.\mathcal{O}.

  2. (2)

    For all 𝗊𝖰\mathsf{q}\in\mathsf{Q},

    (𝗊(κ(𝗆,𝖰)))=(λ(𝗆,𝖰/𝗊))(\partial_{\mathsf{q}}(\kappa(\mathsf{m},\mathsf{Q})))=(\lambda(\mathsf{m},\mathsf{Q}/\mathsf{q}))

    as submodules of 𝒪.\mathcal{O}.

  3. (3)

    For all 𝗊𝖰\mathsf{q}\not\in\mathsf{Q}, admissible with sign ϵ𝗊\epsilon_{\mathsf{q}},

    (loc𝗊(κ(𝗆,𝖰)))=(λ(𝗆,𝖰𝗊))(\operatorname{loc}_{\mathsf{q}}(\kappa(\mathsf{m},\mathsf{Q})))=(\lambda(\mathsf{m},\mathsf{Q}\mathsf{q}))

    as submodules of 𝒪\mathcal{O}.

In particular, for any fixed 𝗆\mathsf{m}, (κ(𝗆,),λ(𝗆,))(\kappa(\mathsf{m},\cdot),\lambda(\mathsf{m},\cdot)) forms a bipartite Euler system with sign ν(N)\nu(N^{-}) for the triple (Tf,(𝗆),𝖲𝗆).(T_{f},\mathcal{F}(\mathsf{m}),\mathsf{S}\cup\mathsf{m}).

Proof.

We verify the local conditions for each 𝗏𝖲𝗆𝖰\mathsf{v}\in\mathsf{S}\cup\mathsf{m}\cup\mathsf{Q}. If 𝗏=v¯\mathsf{v}=\underline{v} for a prime v|Nv|N\infty, then the local condition is all of H1(Kv,Tf)H^{1}(K_{v},T_{f}), so there is nothing to show. If v|pv|p, then we show that, for all jj, the image cjc_{j} of Resvκ(𝗆,𝖰)\operatorname{Res}_{v}\kappa(\mathsf{m},\mathsf{Q}) in H1(Kv,Tf/πj)H^{1}(K_{v},T_{f}/\pi^{j}) is a Kummer image. Recalling the notation used to construct κ(𝗆,𝖰)\kappa(\mathsf{m},\mathsf{Q}), the proof of [20, Lemma 7] implies that:

δv(A(Kv))\displaystyle\delta_{v}(A(K_{v})) =Hfl1(Kv,Tf/πj)H1(Kv,Tf/πj)\displaystyle=H^{1}_{\text{fl}}(K_{v},T_{f}/\pi^{j})\subset H^{1}(K_{v},T_{f}/\pi^{j})
κv(JQn(K[mn]v))\displaystyle\kappa_{v}(J^{Q_{n}}(K[m_{n}]_{v})) =Hfl1(K[mn]v,JQn[pM])H1(K[mn]v,JQn[pM])\displaystyle=H^{1}_{\text{fl}}(K[m_{n}]_{v},J^{Q_{n}}[p^{M}])\subset H^{1}(K[m_{n}]_{v},J^{Q_{n}}[p^{M}])

where we have extended vv to a place of K[n]K[n]. For 𝔉\mathfrak{F}-many nn, the restriction of cjc_{j} to K[mn]vK[m_{n}]_{v} is the image of a Kummer class in H1(K[mn]v,JQn[pM])H^{1}(K[m_{n}]_{v},J^{Q_{n}}[p^{M}]) by a map of Galois representations JQn[pM]Tf/πjA[πj]J^{Q_{n}}[p^{M}]\to T_{f}/\pi^{j}\simeq A[\pi^{j}], which extends to a map of finite flat group schemes. As a consequence, the image of cjc_{j} in H1(Kv,A)H^{1}(K_{v},A) is inflated from a class in H1(K[mn]v/Kv,A)H^{1}(K[m_{n}]_{v}/K_{v},A), which is trivial by [35, Proposition I.3.8]. (This argument is essentially [19, Proposition 6.2(1)].)

If 𝗏=𝗅|𝗆\mathsf{v}=\mathsf{l}|\mathsf{m}, then, adopting as well the notation of (6.2.1), the class c(P(mn),Qn)c(P(m_{n}),Q_{n}) is zero when restricted to K[mn]λnK[m_{n}]_{\lambda_{n}} because Dn=n(n+1)D_{\ell_{n}}=\ell_{n}(\ell_{n}+1) on 𝔽λn\mathbb{F}_{\lambda_{n}}; hence Res𝗏κ(𝗆,𝖰)𝖧tr1(K𝗅,Tf).\operatorname{Res}_{\mathsf{v}}\kappa(\mathsf{m},\mathsf{Q})\in\mathsf{H}^{1}_{\operatorname{tr}}(K_{\mathsf{l}},T_{f}). The local conditions at 𝗊𝖰\mathsf{q}\in\mathsf{Q} are satisfied because every factor of QnQ_{n} splits completely in K[mn]K[m_{n}]; for the same reason, (2, 3) follow from Construction 5.2.2(2, 3). (Note that the projection step in Definition 6.2.5 makes no difference to these identities.) Moreover (1) is clear from Proposition 6.2.4(3). ∎

Remark 6.2.12.

The Euler system (κ(1,),λ(1,))(\kappa(1,\cdot),\lambda(1,\cdot)) may be viewed as a specialization of (𝜿,𝝀)(\bm{\kappa},\bm{\lambda}). Indeed, by the usual Heegner point norm relations [13, Proposition 3.10], if pp splits in KK, 𝟙(𝝀(𝖰))=(αp1)2(λ(1,𝖰))\mathbbm{1}(\bm{\lambda}(\mathsf{Q}))=(\alpha_{p}-1)^{2}(\lambda(1,\mathsf{Q})) and 𝟙(𝜿(𝖰))=(αp1)2(κ(1,𝖰))\mathbbm{1}(\bm{\kappa}(\mathsf{Q}))=(\alpha_{p}-1)^{2}(\kappa(1,\mathsf{Q})) when {𝖰,ϵ𝖰}𝖭ν(N)+1\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}^{\nu(N^{-})+1} and {𝖰,ϵ𝖰}𝖭ν(N)\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}^{\nu(N^{-})}, respectively. (Here 𝟙:Λ𝒪\mathbbm{1}:\Lambda\to\mathcal{O} is specialization at the trivial character.)

7. Deformation theory

Theorem 5.1.5 allows us to produce weak eigenforms (i.e. ring maps) 𝕋N+,NQ𝒪/πj\mathbb{T}_{N^{+},N^{-}Q}\to\mathcal{O}/\pi^{j} for arbitrarily large jj, simply by requiring sufficiently deep congruence conditions on all q|Qq|Q. However, in general these maps do not lift to characteristic zero. To prove the main results, we also need to be able to \wp-adically approximate ff by genuine level-raised newforms. In this section, we provide this input via the relative deformation theory of Fakhruddin-Khare-Patrikis [17].

7.1. Patched adjoint Selmer groups

7.1.1.

Continuing to fix notation as in (3.1.1) and (5.1.1), we now assume moreover that ff is non-CM. Consider the (irreducible) adjoint representation

L=ad0TfL=\operatorname{ad}^{0}T_{f}

and its 𝒪\mathcal{O}-dual, LL(1),L^{\dagger}\simeq L(1), and let L¯\overline{L} and L¯L/π\overline{L}^{\ast}\simeq L^{\dagger}/\pi be the associated residual representations. For all v|Np,v|Np\infty, a choice of framing for TfT_{f} defines a smooth point of the generic fiber of an appropriate framed universal deformation ring (of fixed determinant, and fixed Hodge type if v|pv|p) by [1, Theorem D]. Taking this smooth point as the input, the construction of [17, Proposition 4.7] yields, for all jj, certain orthogonal local conditions

H𝒮1(v,L/πj)H1(v,L/πj),H𝒮1(v,L[πj])H1(v,L[πj]).H^{1}_{\mathcal{S}}(\mathbb{Q}_{v},L/\pi^{j})\subset H^{1}(\mathbb{Q}_{v},L/\pi^{j}),\;\;H^{1}_{\mathcal{S}^{\ast}}(\mathbb{Q}_{v},L^{\ast}[\pi^{j}])\subset H^{1}(\mathbb{Q}_{v},L^{\ast}[\pi^{j}]).

By [17, Lemma 6.1], taking inverse limits yields dual local conditions

H𝒮1(v,L)H1(v,L),H𝒮1(v,L)H1(v,L).H^{1}_{\mathcal{S}}(\mathbb{Q}_{v},L)\subset H^{1}(\mathbb{Q}_{v},L),\;\;H^{1}_{\mathcal{S}^{\dagger}}(\mathbb{Q}_{v},L^{\dagger})\subset H^{1}(\mathbb{Q}_{v},L^{\dagger}).

We use these to define generalized Selmer structures (𝒮,𝖲)(\mathcal{S},\mathsf{S}) and (𝒮,𝖲)(\mathcal{S}^{\dagger},\mathsf{S}) for LL and LL^{\dagger}, where 𝖲𝖬\mathsf{S}\subset\mathsf{M}_{\mathbb{Q}} is the set of constant ultraprimes v¯\underline{v} for v|Npv|Np\infty.

7.1.2.

Now suppose that 𝗊\mathsf{q} is an admissible ultraprime with sign ϵ𝗊\epsilon_{\mathsf{q}}. Using the exact sequence of 𝒪[G𝗊]\mathcal{O}[G_{\mathsf{q}}]-modules in Definition 3.1.2,

0Fil𝗊+TfTfTf/Fil𝗊+Tf0,0\to\operatorname{Fil}^{+}_{\mathsf{q}}T_{f}\to T_{f}\to T_{f}/\operatorname{Fil}^{+}_{\mathsf{q}}T_{f}\to 0,

we define

Fil𝗊+L=Hom(Tf/Fil𝗊+Tf,Fil𝗊+Tf)L\operatorname{Fil}^{+}_{\mathsf{q}}L=\operatorname{Hom}(T_{f}/\operatorname{Fil}^{+}_{\mathsf{q}}T_{f},\operatorname{Fil}^{+}_{\mathsf{q}}T_{f})\subset L

and

𝖧ord1(𝗊,L)=im(𝖧1(𝗊,Fil𝗊+L)𝖧1(𝗊,L)).\mathsf{H}^{1}_{\operatorname{ord}}(\mathbb{Q}_{\mathsf{q}},L)=\operatorname{im}\left(\mathsf{H}^{1}(\mathbb{Q}_{\mathsf{q}},\operatorname{Fil}^{+}_{\mathsf{q}}L)\to\mathsf{H}^{1}(\mathbb{Q}_{\mathsf{q}},L)\right).

We also define 𝖧ord1(𝗊,L)\mathsf{H}^{1}_{\operatorname{ord}}(\mathbb{Q}_{\mathsf{q}},L^{\dagger}) as the orthogonal complement of 𝖧ord1(𝗊,L)\mathsf{H}^{1}_{\operatorname{ord}}(\mathbb{Q}_{\mathsf{q}},L) under the local Tate pairing; note that, since 𝖧1(𝗊,L)\mathsf{H}^{1}(\mathbb{Q}_{\mathsf{q}},L) is torsion-free, 𝖧ord1(𝗊,L)\mathsf{H}^{1}_{\operatorname{ord}}(\mathbb{Q}_{\mathsf{q}},L^{\dagger}) and 𝖧ord1(𝗊,L)\mathsf{H}^{1}_{\operatorname{ord}}(\mathbb{Q}_{\mathsf{q}},L) are exact annihilators. We will require the restriction maps

(42) loc𝗊:𝖧1(,L)𝖧1(𝗊,L)𝖧ord1(𝗊,L)𝖧unr1(𝗊,L),loc𝗊:𝖧1(,L)𝖧1(𝗊,L)𝖧ord1(𝗊,L)𝖧unr1(𝗊,L).\begin{split}\operatorname{loc}_{\mathsf{q}}:\mathsf{H}^{1}(\mathbb{Q},L)&\to\frac{\mathsf{H}^{1}(\mathbb{Q}_{\mathsf{q}},L)}{\mathsf{H}^{1}_{\operatorname{ord}}(\mathbb{Q}_{\mathsf{q}},L)\cap\mathsf{H}^{1}_{\operatorname{unr}}(\mathbb{Q}_{\mathsf{q}},L)},\\ \operatorname{loc}_{\mathsf{q}}^{\dagger}:\mathsf{H}^{1}(\mathbb{Q},L^{\dagger})&\to\frac{\mathsf{H}^{1}(\mathbb{Q}_{\mathsf{q}},L^{\dagger})}{\mathsf{H}^{1}_{\operatorname{ord}}(\mathbb{Q}_{\mathsf{q}},L^{\dagger})\cap\mathsf{H}^{1}_{\operatorname{unr}}(\mathbb{Q}_{\mathsf{q}},L^{\dagger})}.\end{split}

Analogously, if qMq\in M_{\mathbb{Q}} is jj-admissible with sign ϵq\epsilon_{q}, then we may define Hord1(q,L/πj)H^{1}_{\operatorname{ord}}(\mathbb{Q}_{q},L/\pi^{j}), Hord1(q,L/πj)H^{1}_{\operatorname{ord}}(\mathbb{Q}_{q},L^{\dagger}/\pi^{j}), and the localization maps locq,\operatorname{loc}_{q}, locq\operatorname{loc}_{q}^{\dagger}.

7.1.3.

For any {𝖯𝖰𝖱,ϵ𝖯𝖰𝖱}𝖭\left\{\mathsf{P}\mathsf{Q}\mathsf{R},\epsilon_{\mathsf{P}\mathsf{Q}\mathsf{R}}\right\}\in\mathsf{N}, define the modified Selmer structure (𝒮(𝖰),𝖲𝖰)(\mathcal{S}(\mathsf{Q}),\mathsf{S}\cup\mathsf{Q}) for LL:

(43) 𝖧𝒮𝖱𝖯(𝖰)1(𝗏,L)={𝖧𝒮1(𝗏,L),𝗏𝖯𝖰𝖱𝖧ord1(𝗊,L),𝗏=𝗊𝖰,𝖧ord1(𝗊,L)+𝖧unr1(𝗊,L),𝗏=𝗊𝖯,𝖧ord1(𝗊,L)𝖧unr1(𝗊,L),𝗏=𝗊𝖱.\mathsf{H}^{1}_{\mathcal{S}^{\mathsf{P}}_{\mathsf{R}}(\mathsf{Q})}(\mathbb{Q}_{\mathsf{v}},L)=\begin{cases}\mathsf{H}^{1}_{\mathcal{S}}(\mathbb{Q}_{\mathsf{v}},L),&\mathsf{v}\not\in\mathsf{P}\mathsf{Q}\mathsf{R}\\ \mathsf{H}^{1}_{\operatorname{ord}}(\mathbb{Q}_{\mathsf{q}},L),&\mathsf{v}=\mathsf{q}\in\mathsf{Q},\\ \mathsf{H}^{1}_{\operatorname{ord}}(\mathbb{Q}_{\mathsf{q}},L)+\mathsf{H}^{1}_{\operatorname{unr}}(\mathbb{Q}_{\mathsf{q}},L),&\mathsf{v}=\mathsf{q}\in\mathsf{P},\\ \mathsf{H}^{1}_{\operatorname{ord}}(\mathbb{Q}_{\mathsf{q}},L)\cap\mathsf{H}^{1}_{\operatorname{unr}}(\mathbb{Q}_{\mathsf{q}},L),&\mathsf{v}=\mathsf{q}\in\mathsf{R}.\end{cases}

The corresponding dual Selmer structure for LL^{\dagger} will be written 𝒮𝖯𝖱,(𝖰)\mathcal{S}^{\mathsf{R},\dagger}_{\mathsf{P}}(\mathsf{Q}). Finally, define, for any finite set of places Σ\Sigma containing all v|Npv|Np\infty:

(44) (L¯)Σ1=ker(H1(Σ/,L¯)vΣH1(v,L¯)).{}^{1}_{\Sigma}(\overline{L}^{\ast})=\ker\left(H^{1}(\mathbb{Q}^{\Sigma}/\mathbb{Q},\overline{L}^{\ast})\to\prod_{v\in\Sigma}H^{1}(\mathbb{Q}_{v},\overline{L}^{\ast})\right).
Proposition 7.1.4.

For all {𝖰,ϵ𝖰}𝖭\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N},

d𝖰rk𝒪Sel𝒮(𝖰)(L)=rk𝒪Sel𝒮(𝖰)(L).d_{\mathsf{Q}}\coloneqq\operatorname{rk}_{\mathcal{O}}\operatorname{Sel}_{\mathcal{S}(\mathsf{Q})}(L)=\operatorname{rk}_{\mathcal{O}}\operatorname{Sel}_{\mathcal{S}^{\dagger}(\mathsf{Q})}(L^{\dagger}).
Proof.

It follows from the construction [17, Proposition 4.7] that

rk𝒪H𝒮1(v,L)=rk𝒪Hunr1(v,L)\operatorname{rk}_{\mathcal{O}}H^{1}_{\mathcal{S}}(\mathbb{Q}_{v},L)=\operatorname{rk}_{\mathcal{O}}H^{1}_{\operatorname{unr}}(\mathbb{Q}_{v},L)

for all v|Nv|N and

rk𝒪H𝒮1(v,L)=rk𝒪H0(v,L)+2\operatorname{rk}_{\mathcal{O}}H^{1}_{\mathcal{S}}(\mathbb{Q}_{v},L)=\operatorname{rk}_{\mathcal{O}}H^{0}(\mathbb{Q}_{v},L)+2

if v|pv|p. Since

rk𝒪𝖧ord1(𝗊,L)=rk𝒪𝖧0(𝗊,L)=1\operatorname{rk}_{\mathcal{O}}\mathsf{H}^{1}_{\operatorname{ord}}(\mathbb{Q}_{\mathsf{q}},L)=\operatorname{rk}_{\mathcal{O}}\mathsf{H}^{0}(\mathbb{Q}_{\mathsf{q}},L)=1

for all 𝗊𝖰\mathsf{q}\in\mathsf{Q}, the claim results from Proposition 2.7.4. ∎

The “relative deformation theory” developed in [17] may be summarized (for our context) as follows.

Theorem 7.1.5 (Fakhruddin-Khare-Patrikis, Kisin).

Suppose given {𝖰,ϵ𝖰}𝖭\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N} such that d𝖰=0,d_{\mathsf{Q}}=0, and a finite set of places Σ\Sigma containing all v|Npv|Np\infty such that (,L¯)Σ1=0{}^{1}_{\Sigma}(\mathbb{Q},\overline{L}^{\ast})=0. Fix a sequence {Qn,ϵQn}\left\{Q_{n},\epsilon_{Q_{n}}\right\} of weakly admissible pairs representing {𝖰,ϵ𝖰}\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\} and an integer j0.j\geq 0. Then there is a sequence (defined for 𝔉\mathfrak{F}-many nn) of newforms gng_{n} of weight two, level NQnNQ_{n}, and trivial nebentypus, with a prime gn\wp_{g_{n}} of the ring of integers of its coefficient field 𝒪gn\mathcal{O}_{g_{n}}, such that:

  • The completion 𝒪gn,gn\mathcal{O}_{g_{n},\wp_{g_{n}}} is isomorphic to 𝒪\mathcal{O}.

  • The associated Galois representations satisfy Tf|ITgn|IT_{f}|_{I_{\ell}}\simeq T_{g_{n}}|_{I_{\ell}} for all Qn\ell\nmid Q_{n}, and ρgn|Iqn\rho_{g_{n}}|_{I_{q_{n}}} is a Steinberg representation twisted by the unramified character FrobqnϵQn(qn)\operatorname{Frob}_{q_{n}}\mapsto\epsilon_{Q_{n}}(q_{n}) for all qn|Qnq_{n}|Q_{n}.

  • For any fixed jj, there is a congruence of Galois representations (in some basis)

    TfTgn(modπj)T_{f}\equiv T_{g_{n}}\pmod{\pi^{j}}

    for 𝔉\mathfrak{F}-many nn. In particular, the maps

    𝕋Qn=𝕋N+,NQn𝒪/πj\mathbb{T}_{Q_{n}}=\mathbb{T}_{N^{+},N^{-}Q_{n}}\to\mathcal{O}/\pi^{j}

    of Remark 5.1.6 admit 𝒪\mathcal{O}-valued lifts for 𝔉\mathfrak{F}-many nn.

Proof.

Since d𝖰=0,d_{\mathsf{Q}}=0, Proposition 2.5.5 implies that there exists some k0k\geq 0 such that the natural maps

(45) Sel𝒮(𝖰)(L/πk)Sel𝒮(𝖰)(L¯),Sel𝒮(𝖰)(L/πk)Sel𝒮(𝖰)(L¯)\operatorname{Sel}_{\mathcal{S}(\mathsf{Q})}(L/\pi^{k})\to\operatorname{Sel}_{\mathcal{S}(\mathsf{Q})}(\overline{L}),\;\;\operatorname{Sel}_{\mathcal{S}(\mathsf{Q})}(L^{\dagger}/\pi^{k})\to\operatorname{Sel}_{\mathcal{S}(\mathsf{Q})}(\overline{L}^{\ast})

are identically zero.

Now, for any jj and for 𝔉\mathfrak{F}-many nn, we have the weakly admissible pair {Qn,ϵQn}Nj\left\{Q_{n},\epsilon_{Q_{n}}\right\}\in N_{j} of Remark 5.1.6. Consider the (non-patched) Selmer groups

SelQn(L/πk)=ker(H1(SQn,L/πk)v|NpH1(v,L/πk)H𝒮1(v,L/πk)×qn|QnH1(qn,L/πk)Hord1(qn,L/πk)),\operatorname{Sel}_{Q_{n}}(L/\pi^{k})=\ker\left(H^{1}(\mathbb{Q}^{S\cup Q_{n}},L/\pi^{k})\to\prod_{v|Np\infty}\frac{H^{1}(\mathbb{Q}_{v},L/\pi^{k})}{H^{1}_{\mathcal{S}}(\mathbb{Q}_{v},L/\pi^{k})}\times\prod_{q_{n}|Q_{n}}\frac{H^{1}(\mathbb{Q}_{q_{n}},L/\pi^{k})}{H^{1}_{\operatorname{ord}}(\mathbb{Q}_{q_{n}},L/\pi^{k})}\right),

where SS is the set of places dividing NpQnNpQ_{n}\infty. For each qn|Qnq_{n}|Q_{n}, Hord1(qn,L/πk)H^{1}_{\operatorname{ord}}(\mathbb{Q}_{q_{n}},L/\pi^{k}) is exactly the local condition obtained from the smooth, Steinberg-with-sign-ϵQn(qn)\epsilon_{Q_{n}}(q_{n}) quotient of the framed local deformation ring at qnq_{n} via [17, Proposition 4.7] (as long as qnq_{n} is kk-admissible). There are also dual Selmer groups SelQn(L/πk)=SelQn(L[πk])\operatorname{Sel}_{Q_{n}}(L^{\dagger}/\pi^{k})=\operatorname{Sel}_{Q_{n}}(L^{\ast}[\pi^{k}]) defined in the same way (see Proposition 2.7.3 for the equality). By (45), for 𝔉\mathfrak{F}-many nn, the maps

SelQn(L/πk)SelQn(L¯),SelQn(L[k])SelQn(L¯)\operatorname{Sel}_{Q_{n}}(L/\pi^{k})\to\operatorname{Sel}_{Q_{n}}(\overline{L}),\;\;\operatorname{Sel}_{Q_{n}}(L^{\ast}[\wp^{k}])\to\operatorname{Sel}_{Q_{n}}(\overline{L}^{\ast})

are identically zero. Moreover, for all nn,

ΣQn1=Σ10.{}_{\Sigma\cup Q_{n}}^{1}\subset{}_{\Sigma}^{1}=0.

The proof of [17, Claim 6.12] now implies that, for j3kj\geq 3k large enough in a manner depending on the local representations ρf|G\rho_{f}|_{G_{\mathbb{Q}_{\ell}}} for |Np\ell|Np and for 𝔉\mathfrak{F}-many nn, there exists a representation

τn:GGL2(𝒪)\tau_{n}:G_{\mathbb{Q}}\to GL_{2}(\mathcal{O})

such that:

  • τnρf(modπj)\tau_{n}\equiv\rho_{f}\pmod{\pi^{j}} (for some choice of basis of TfT_{f});

  • detτn=χ\det\tau_{n}=\chi;

  • the local representations τn|G\tau_{n}|_{G_{\mathbb{Q}_{\ell}}} lie on the same irreducible component of the framed local deformation ring as ρf|G\rho_{f}|_{G_{\mathbb{Q}_{\ell}}} if Qn\ell\nmid Q_{n}, and are Steinberg representations twisted by the unramified character FrobqnϵQn(qn)\operatorname{Frob}_{q_{n}}\mapsto\epsilon_{Q_{n}}(q_{n}) for all qn|Qnq_{n}|Q_{n}.

It remains to apply a modularity lifting theorem to conclude that τn\tau_{n} arises from a suitable modular form gng_{n}. We claim that T¯f|G(ζp)\overline{T}_{f}|_{G_{\mathbb{Q}(\zeta_{p})}} is absolutely irreducible: suppose otherwise for contradiction. Then the image GG of the action of GG_{\mathbb{Q}} on T¯f\overline{T}_{f} fixes a pair of lines. Since the image of the inertia group IpI_{p} contains a matrix whose square has distinct eigenvalues by (TW), the pair of lines must be the eigenspaces of IpI_{p}. But IpI_{p} surjects onto Gal((ζp)/)\operatorname{Gal}(\mathbb{Q}(\zeta_{p})/\mathbb{Q}), so by assumption no element of GG_{\mathbb{Q}} interchanges the two eigenspaces of IpI_{p}, and T¯f\overline{T}_{f} is reducible – a contradiction. Thus the Taylor-Wiles hypothesis in Kisin’s result [29] is satisfied, and τn\tau_{n} indeed arises from a newform gng_{n}. ∎

7.2. Annihilating two Selmer groups

7.2.1.

In order to apply Theorem 7.1.5, we must make a suitable choice of {𝖰,ϵ𝖰}𝖭\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}. In this subsection, we show that such a choice as possible.

Proposition 7.2.2.

There exists a finite set of places Σ\Sigma, containing all v|Npv|Np\infty, such that

=Σ10.{}_{\Sigma}^{1}=0.
Proof.

It suffices to show that

(46) H1((L¯)/,L¯)=0,H^{1}(\mathbb{Q}(\overline{L}^{\ast})/\mathbb{Q},\overline{L}^{\ast})=0,

for if so, for any 0cH1(,L¯)0\neq c\in H^{1}(\mathbb{Q},\overline{L}^{\ast}), cc restricts to a nonzero homomorphism G(L¯)L¯,G_{\mathbb{Q}(\overline{L}^{\ast})}\to\overline{L}^{\ast}, and there exist primes which are totally split in (L¯)\mathbb{Q}(\overline{L}^{\ast}) but not in the extension cut out by the restriction of cc.

We now show (46). If μp(L¯)\mu_{p}\not\subset\mathbb{Q}(\overline{L}), then the center of Gal((L¯)/)\operatorname{Gal}(\mathbb{Q}(\overline{L}^{\ast})/\mathbb{Q}) contains elements that act by nontrivial scalars on L¯,\overline{L}^{\ast}, and (46) follows from inflation-restriction. So suppose that μp(L¯)\mu_{p}\subset\mathbb{Q}(\overline{L}); then the projective image G¯=Gal((L¯)/)\overline{G}=\operatorname{Gal}(\mathbb{Q}(\overline{L})/\mathbb{Q}) of the (irreducible) Galois action on T¯f\overline{T}_{f} has a cyclic quotient of order p1,p-1, and a classical result of Dickson implies that p=3p=3 and G¯\overline{G} is either a dihedral group, or S4.S_{4}. In the former case, the order of Gal((L¯)/)\operatorname{Gal}(\mathbb{Q}(\overline{L}^{\ast})/\mathbb{Q}) is prime to pp, so (46) still holds. We are left to consider the case G¯=S4\overline{G}=S_{4} and p=3.p=3. Let G=Gal((T¯f)/)G=\operatorname{Gal}(\mathbb{Q}(\overline{T}_{f})/\mathbb{Q}) be the image of the Galois action; since we have assumed that det:G𝔽3×\det:G\to\mathbb{F}_{3}^{\times} factors through G¯\overline{G}, a complex conjugation cc in GG projects to a transposition in G¯\overline{G}. Let H¯G¯\overline{H}\subset\overline{G} be a copy of S3S_{3} containing the image of cc, and HH the normalizer of its preimage in GG, which is contained in a unique Borel subgroup BB. Let NN be the unipotent radical of BGB\cap G. To prove (46), it suffices to check that

H1(H,L¯)=H1(N,L¯)H/N=0.H^{1}(H,\overline{L}^{\ast})=H^{1}(N,\overline{L}^{\ast})^{H/N}=0.

This holds because im(N1)\operatorname{im}(N-1) is isomorphic to the subgroup (0)L¯\begin{pmatrix}\ast&\ast\\ 0&\ast\end{pmatrix}\subset\overline{L}^{\ast}, while cHc\in H acts on NN by 1-1 and on L¯/im(N1)\overline{L}^{\ast}/\operatorname{im}(N-1) by 11. ∎

We now require a more elaborate version of Theorem 3.3.4; the proof is inspired by [8], and begins with a series of lemmas.

Lemma 7.2.3.

There exists an integer jj that, for all n0,n\geq 0,

πjH1(K(Tf)/,L/πn)=πjH1(K(Tf)/,L/πn)=0.\pi^{j}H^{1}(K(T_{f})/\mathbb{Q},L/\pi^{n})=\pi^{j}H^{1}(K(T_{f})/\mathbb{Q},L^{\dagger}/\pi^{n})=0.
Proof.

Let E=(μp)K(Tf)E=\mathbb{Q}(\mu_{p^{\infty}})\subset K(T_{f}), and note that LL and LL^{\dagger} are isomorphic GEG_{E}-modules. Since ff is non-CM, (Lp)GE=0(L\otimes\mathbb{Q}_{p})^{G_{E}}=0, and so (L/πn)GE(L/\pi^{n})^{G_{E}} is uniformly bounded in nn.

The pro-pp-Sylow subgroup of Gal(K(Tf)/E)\operatorname{Gal}(K(T_{f})/E) is a compact pp-adic Lie group with semisimple Lie algebra; hence, by [17, Lemma B.1], H1(K(Tf)/E,L/πn)H^{1}(K(T_{f})/E,L/\pi^{n}) is uniformly bounded in nn.

Now, by inflation-restriction, we have exact sequences

(47) 0H1(E/,(L/πn)GE)H1(K(Tf)/,L/πn)H1(K(Tf)/E,L/πn),0H1(E/,(L/πn)GE)H1(K(Tf)/,L/πn)H1(K(Tf)/E,L/πn),\begin{split}&0\to H^{1}(E/\mathbb{Q},(L/\pi^{n})^{G_{E}})\to H^{1}(K(T_{f})/\mathbb{Q},L/\pi^{n})\to H^{1}(K(T_{f})/E,L/\pi^{n}),\\ &0\to H^{1}(E/\mathbb{Q},(L^{\dagger}/\pi^{n})^{G_{E}})\to H^{1}(K(T_{f})/\mathbb{Q},L^{\dagger}/\pi^{n})\to H^{1}(K(T_{f})/E,L^{\dagger}/\pi^{n}),\end{split}

where the outer terms are isomorphic and uniformly bounded in nn; the lemma follows. ∎

For the next lemma, we abbreviate LmL/πmL_{m}\coloneqq L/\pi^{m}, LmL/πmL^{\dagger}_{m}\coloneqq L^{\dagger}/\pi^{m}, and TmTf/πmT_{m}\coloneqq T_{f}/\pi^{m}. Moreover, if yMy\in M for any torsion 𝒪\mathcal{O}-module MM, let ord(y)\operatorname{ord}(y) be the smallest integer t0t\geq 0 such that πty=0\pi^{t}y=0.

Lemma 7.2.4.

There is a global constant CC, depending on TfT_{f}, with the following property. Given cocycles ϕH1(,Lm)\phi\in H^{1}(\mathbb{Q},L_{m}), ψH1(,Lm),\psi\in H^{1}(\mathbb{Q},L^{\dagger}_{m}), and c1,c2H1(K,Tm)δc_{1},c_{2}\in H^{1}(K,T_{m})^{\delta} for some δ=±1,\delta=\pm 1, there exist infinitely many primes qNpq\nmid Np such that all the cocycles are unramified at qq and:

  • The Frobenius of qq in Gal(K(Tm)/)\operatorname{Gal}(K(T_{m})/\mathbb{Q}) is a complex conjugation; in particular, qq is mm-admissible with sign δ\delta.

  • ordlocqϕordϕC.\operatorname{ord}\operatorname{loc}_{q}\phi\geq\operatorname{ord}\phi-C.

  • ordlocqψordψC\operatorname{ord}\operatorname{loc}_{q}^{\dagger}\psi\geq\operatorname{ord}\psi-C, or locqψ=0\operatorname{loc}_{q}^{\dagger}\psi=0, as desired.

  • ordlocqciordciC\operatorname{ord}\operatorname{loc}_{q}c_{i}\geq\operatorname{ord}c_{i}-C for i=1,2.i=1,2.

Proof.

Let us first fix a complex conjugation cGc\in G_{\mathbb{Q}} and choose a basis for TmT_{m} in which cc acts as (δ00δ)\begin{pmatrix}-\delta&0\\ 0&\delta\end{pmatrix}.

The restriction of the cocycles ϕ,ψ,ci\phi,\psi,c_{i} to GK(Tm)G_{K(T_{m})} may be considered as a homomorphism

h:GK(Tm)LmLm(Tm)2h:G_{K(T_{m})}\to L_{m}\oplus L^{\dagger}_{m}\oplus(T_{m})^{2}

compatible with the action of GKG_{K}; let HH be the image of this homomorphism. Since there exists an element of gzGKg_{z}\in G_{K} that acts by a scalar z±1z\neq\pm 1 on TfT_{f}, we have:

H\displaystyle H\supset (gzz)(gzz2)H+(gzz)(gz1)H+(gzz2)(gz1)H\displaystyle(g_{z}-z)(g_{z}-z^{2})H+(g_{z}-z)(g_{z}-1)H+(g_{z}-z^{2})(g_{z}-1)H
(z1)(z21)(z2z)(πLm(H)πLm(H)πTm2(H),)\displaystyle\supset(z-1)(z^{2}-1)(z^{2}-z)\left(\pi_{L_{m}}(H)\oplus\pi_{L_{m}^{\dagger}}(H)\oplus\pi_{T_{m}^{2}}(H),\right)

where π\pi_{\bullet} are the projection operators. Now, since LL and LL^{\dagger} are absolutely irreducible, the natural maps p[GK]End(Lp)\mathbb{Q}_{p}[G_{K}]\to\operatorname{End}(L\otimes\mathbb{Q}_{p}) and p[GK]End(Lp)\mathbb{Q}_{p}[G_{K}]\to\operatorname{End}(L^{\dagger}\otimes\mathbb{Q}_{p}) are surjective. Combining these observations with Lemma 7.2.3, we see that, for some constant CC depending only on TfT_{f}, there exists γGK(Tf)\gamma\in G_{K(T_{f})} satisfying:

  • The (1001)\begin{pmatrix}1&0\\ 0&-1\end{pmatrix} component of ϕ(γ)\phi(\gamma) has order at least ordϕC{\operatorname{ord}\phi-C}.

  • The (0010)\begin{pmatrix}0&0\\ 1&0\end{pmatrix} component of ψ(γψ)\psi(\gamma_{\psi}) has order at least ordψC{\operatorname{ord}\psi-C}, or is 0, as desired.

  • The components of ci(γ)c_{i}(\gamma) and c2(γ)c_{2}(\gamma) in the δ\delta eigenspace have order at least ordciC\operatorname{ord}c_{i}-C, where i=1,2i=1,2.

For the final item, we are using the elementary fact that a group cannot be the union of two trivial subgroups, as well as the irreducibility of TfT_{f}.

Since ϕ(c2)=cϕ(c)+ϕ(c)=0,\phi(c^{2})=c\phi(c)+\phi(c)=0, ϕ(c)\phi(c) lies in the 1-1 eigenspace for complex conjugation, whereas (1001)\begin{pmatrix}1&0\\ 0&-1\end{pmatrix} has eigenvalue 1; hence the (1001)\begin{pmatrix}1&0\\ 0&-1\end{pmatrix} component of ϕ(cγ)\phi(c\gamma) has order at least ordϕC{\operatorname{ord}\phi-C}. Similarly, the (0010)\begin{pmatrix}0&0\\ 1&0\end{pmatrix} component of ψ(cγψ)\psi(c\gamma_{\psi}) has order at least ordψC{\operatorname{ord}\psi-C}, or is 0, as desired.

Any prime with Frobenius cγc\gamma in kerh\ker h satisfies the conclusion of the lemma; cf. the proof of Lemma 3.3.6 for the assertions about cic_{i}. ∎

Corollary 7.2.5.

Suppose given a finite set of ultraprimes 𝖳\mathsf{T} and non-torsion cocycles:

  • ϕ𝖧1(𝖳/,L)\phi\in\mathsf{H}^{1}(\mathbb{Q}^{\mathsf{T}}/\mathbb{Q},L);

  • ψ𝖧1(𝖳/,L)\psi\in\mathsf{H}^{1}(\mathbb{Q}^{\mathsf{T}}/\mathbb{Q},L^{\dagger});

  • c1,c2𝖧1(K𝖳/K,Tf)δc_{1},c_{2}\in\mathsf{H}^{1}(K^{\mathsf{T}}/K,T_{f})^{\delta} for δ=±1.\delta=\pm 1.

Then there exist infinitely many admissible ultraprimes 𝗊𝖳\mathsf{q}\not\in\mathsf{T} with sign δ\delta such that:

  • loc𝗊ϕ0\operatorname{loc}_{\mathsf{q}}\phi\neq 0.

  • Either loc𝗊ψ0\operatorname{loc}^{\dagger}_{\mathsf{q}}\psi\neq 0 or loc𝗊ψ=0\operatorname{loc}_{\mathsf{q}}^{\dagger}\psi=0, as desired.

  • loc𝗊ci0.\operatorname{loc}_{\mathsf{q}}c_{i}\neq 0.

Proof.

Choose a sequence TnT_{n} representing 𝖳\mathsf{T} and sequences ϕn,ψn,cn1,cn2\phi_{n},\psi_{n},c_{n}^{1},c_{n}^{2} representing the respective cocycles in H1(Tn/,L/πn)H^{1}(\mathbb{Q}^{T_{n}}/\mathbb{Q},L/\pi^{n}), etc. For each nn, apply Lemma 7.2.4 with m=nm=n and the appropriate desideratum for ψn\psi_{n}; by definition, any resulting admissible ultraprime 𝗊\mathsf{q}, represented by a sequence qnq_{n}, satisfies the desired conclusion. ∎

Proposition 7.2.6.

Suppose given a self-dual Selmer structure (,𝖳)(\mathcal{F},\mathsf{T}) for TfT_{f}. Then there exists {𝖰,ϵ𝖰}𝖭𝖳\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}_{\mathsf{T}} such that

r𝖰=d𝖰=0.r_{\mathsf{Q}}=d_{\mathsf{Q}}=0.

(Recall that r𝖰=rk𝒪Sel(𝖰)(Tf).r_{\mathsf{Q}}=\operatorname{rk}_{\mathcal{O}}\operatorname{Sel}_{\mathcal{F}(\mathsf{Q})}(T_{f}).)

Proof.

Without loss of generality, by Corollary 3.3.7 we may assume that r1=0r_{1}=0; for if not, choose any {𝖰,ϵ𝖰}𝖭𝖳\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}_{\mathsf{T}} with r𝖰=0,r_{\mathsf{Q}}=0, and then relabel (𝖰)\mathcal{F}(\mathsf{Q}) as \mathcal{F}.

We will show that, if d1>0d_{1}>0, we may find {𝖰,ϵ𝖰}\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\} such that r𝖰=0r_{\mathsf{Q}}=0 and d𝖰<d1d_{\mathsf{Q}}<d_{1}; this clearly suffices by induction. By Proposition 7.1.4, there exist non-torsion elements ϕSel𝒮(L),ψSel𝒮(L).\phi\in\operatorname{Sel}_{\mathcal{S}}(L),\;\psi\in\operatorname{Sel}_{\mathcal{S}^{\dagger}}(L^{\dagger}). We choose any admissible 𝗊𝖳\mathsf{q}\not\in\mathsf{T} with sign ϵ𝗊\epsilon_{\mathsf{q}} such that loc𝗊ϕ0,\operatorname{loc}_{\mathsf{q}}\phi\neq 0, loc𝗊ψ0.\operatorname{loc}_{\mathsf{q}}^{\dagger}\psi\neq 0. Then by Proposition 2.7.4,

rk𝒪Sel𝒮,𝗊(L)+rk𝒪Sel𝒮𝗊(L)=2+rk𝒪Sel𝒮𝗊(L)+rk𝒪Sel𝒮𝗊(L)\operatorname{rk}_{\mathcal{O}}\operatorname{Sel}_{\mathcal{S}^{\dagger,\mathsf{q}}}(L^{\dagger})+\operatorname{rk}_{\mathcal{O}}\operatorname{Sel}_{\mathcal{S}^{\mathsf{q}}}(L)=2+\operatorname{rk}_{\mathcal{O}}\operatorname{Sel}_{\mathcal{S}^{\dagger}_{\mathsf{q}}}(L^{\dagger})+\operatorname{rk}_{\mathcal{O}}\operatorname{Sel}_{\mathcal{S}_{\mathsf{q}}}(L)

(in the notation of (43)). The images of the localization maps

loc𝗊:Sel𝒮𝗊(L)Sel𝒮𝗊(L)𝖧ord1(𝗊,L)𝖧ordunr1(𝗊,L)𝖧unr1(𝗊,L)𝖧ordunr1(𝗊,L)\operatorname{loc}_{\mathsf{q}}:\frac{\operatorname{Sel}_{\mathcal{S}^{\mathsf{q}}}(L)}{\operatorname{Sel}_{\mathcal{S}_{\mathsf{q}}}(L)}\hookrightarrow\frac{\mathsf{H}^{1}_{\operatorname{ord}}(\mathbb{Q}_{\mathsf{q}},L)}{\mathsf{H}^{1}_{\operatorname{ord}\cap\operatorname{unr}}(\mathbb{Q}_{\mathsf{q}},L)}\oplus\frac{\mathsf{H}^{1}_{\operatorname{unr}}(\mathbb{Q}_{\mathsf{q}},L)}{\mathsf{H}^{1}_{\operatorname{ord}\cap\operatorname{unr}}(\mathbb{Q}_{\mathsf{q}},L)}

and

loc𝗊:Sel𝒮,𝗊(L)Sel𝒮𝗊(L)𝖧ord1(𝗊,L)𝖧ordunr1(𝗊,L)𝖧unr1(𝗊,L)𝖧ordunr1(𝗊,L)\operatorname{loc}^{\dagger}_{\mathsf{q}}:\frac{\operatorname{Sel}_{\mathcal{S}^{{}^{\dagger},\mathsf{q}}}(L^{\dagger})}{\operatorname{Sel}_{\mathcal{S}^{\dagger}_{\mathsf{q}}}(L^{\dagger})}\hookrightarrow\frac{\mathsf{H}^{1}_{\operatorname{ord}}(\mathbb{Q}_{\mathsf{q}},L^{\dagger})}{\mathsf{H}^{1}_{\operatorname{ord}\cap\operatorname{unr}}(\mathbb{Q}_{\mathsf{q}},L^{\dagger})}\oplus\frac{\mathsf{H}^{1}_{\operatorname{unr}}(\mathbb{Q}_{\mathsf{q}},L^{\dagger})}{\mathsf{H}^{1}_{\operatorname{ord}\cap\operatorname{unr}}(\mathbb{Q}_{\mathsf{q}},L^{\dagger})}

have total rank two and annihilate each other under the induced Tate pairing by Proposition 2.7.2. Hence the image in the ordinary part is zero for both maps, and d𝗊<d1.d_{\mathsf{q}}<d_{1}. However, by adding 𝗊\mathsf{q}, we have made r𝗊=1.r_{\mathsf{q}}=1. Let cSel(𝗊)(Tf)c\in\operatorname{Sel}_{\mathcal{F}(\mathsf{q})}(T_{f}) be a generator; since 𝗊c0\partial_{\mathsf{q}}c\neq 0 by Proposition 3.3.3, cc has nonzero component in the ϵ𝗊\epsilon_{\mathsf{q}} eigenspace for τ\tau.

Now consider the set 𝖯\mathsf{P} of admissible ultraprimes 𝗌\mathsf{s} with sign ϵ𝗌=ϵ𝗊\epsilon_{\mathsf{s}}=\epsilon_{\mathsf{q}} such that loc𝗌c0.\operatorname{loc}_{\mathsf{s}}c\neq 0. If, for any 𝗌𝖯\mathsf{s}\in\mathsf{P}, d𝗊𝗌d𝗊d_{\mathsf{q}\mathsf{s}}\leq d_{\mathsf{q}}, then we may take 𝖰=𝗊𝗌\mathsf{Q}=\mathsf{q}\mathsf{s} and complete our induction step. For example, this will occur provided d𝗊>0,d_{\mathsf{q}}>0, by the argument above; so without loss of generality, d𝗊=0d_{\mathsf{q}}=0 and d𝗊𝗌=1d_{\mathsf{q}\mathsf{s}}=1 for all 𝗌𝖯\mathsf{s}\in\mathsf{P}. By definition, we therefore have non-torsion elements ϕ(𝗌)Sel𝒮(𝗊𝗌)(L)\phi(\mathsf{s})\in\operatorname{Sel}_{\mathcal{S}(\mathsf{q}\mathsf{s})}(L) and ψ(𝗌)Sel𝒮(𝗊𝗌)(L)\psi(\mathsf{s})\in\operatorname{Sel}_{\mathcal{S}^{\dagger}(\mathsf{q}\mathsf{s})}(L^{\dagger}) such that loc𝗌ϕ(s)\operatorname{loc}_{\mathsf{s}}\phi(s) and loc𝗌ψ(s)\operatorname{loc}_{\mathsf{s}}\psi(s) do not lie in the unramified subspace of the ordinary cohomology.

Choose any 𝗌1𝖯\mathsf{s}_{1}\in\mathsf{P}, and then choose 𝗌2𝖯\mathsf{s}_{2}\in\mathsf{P} such that loc𝗌2ϕ(𝗌1)0\operatorname{loc}_{\mathsf{s}_{2}}\phi(\mathsf{s}_{1})\neq 0 but loc𝗌2ψ(𝗌1)=0.\operatorname{loc}_{\mathsf{s}_{2}}\psi(\mathsf{s}_{1})=0. By another application of Proposition 3.3.3, r𝗊𝗌1𝗌2=1,r_{\mathsf{q}\mathsf{s}_{1}\mathsf{s}_{2}}=1, and a generator cc^{\prime} of Sel(𝗊𝗌1𝗌2)(Tf)\operatorname{Sel}_{\mathcal{F}(\mathsf{q}\mathsf{s}_{1}\mathsf{s}_{2})}(T_{f}) again has nonzero component in the ϵ𝗊\epsilon_{\mathsf{q}} eigenspace. We now choose 𝗌3𝖯\mathsf{s}_{3}\in\mathsf{P} such that loc𝗌3c0\operatorname{loc}_{\mathsf{s}_{3}}c^{\prime}\neq 0, loc𝗌3ϕ(𝗌2)0,\operatorname{loc}_{\mathsf{s}_{3}}\phi(\mathsf{s}_{2})\neq 0, and loc𝗌3ψ(𝗌1)0.\operatorname{loc}_{\mathsf{s}_{3}}\psi(\mathsf{s}_{1})\neq 0. Clearly r𝗊𝗌1𝗌2𝗌3=0.r_{\mathsf{q}\mathsf{s}_{1}\mathsf{s}_{2}\mathsf{s}_{3}}=0. Note that rk𝒪Sel𝒮𝗌1𝗌2𝗌3(𝗊)=3;\operatorname{rk}_{\mathcal{O}}\operatorname{Sel}_{\mathcal{S}^{\mathsf{s}_{1}\mathsf{s}_{2}\mathsf{s}_{3}}(\mathsf{q})}=3; up to torsion, ϕ(𝗌i)\phi(\mathsf{s}_{i}) are generators. So to show that d𝗊𝗌1𝗌2𝗌3=d𝗊d_{\mathsf{q}\mathsf{s}_{1}\mathsf{s}_{2}\mathsf{s}_{3}}=d_{\mathsf{q}}, it suffices to show that the images of ϕ(𝗌i)\phi(\mathsf{s}_{i}) form a rank-three subspace of

Si=13𝖧unr+ord1(𝗌i,L)𝖧ord1(𝗌i,L)S\coloneqq\bigoplus_{i=1}^{3}\frac{\mathsf{H}^{1}_{\operatorname{unr}+\operatorname{ord}}(\mathbb{Q}_{\mathsf{s}_{i}},L)}{\mathsf{H}^{1}_{\operatorname{ord}}(\mathbb{Q}_{\mathsf{s}_{i}},L)}

under the localization

loc:Sel𝒮𝗌1𝗌2𝗌3(𝗊)(L)Sel𝒮𝗌1𝗌2𝗌3(𝗊)(L)S.\operatorname{loc}:\frac{\operatorname{Sel}_{\mathcal{S}^{\mathsf{s}_{1}\mathsf{s}_{2}\mathsf{s}_{3}}}(\mathsf{q})(L)}{\operatorname{Sel}_{\mathcal{S}_{\mathsf{s}_{1}\mathsf{s}_{2}\mathsf{s}_{3}}(\mathsf{q})}(L)}\hookrightarrow S.

By pairing ϕ(𝗌i)\phi(\mathsf{s}_{i}) and ψ(𝗌j)\psi(\mathsf{s}_{j}) for iji\neq j and applying Proposition 2.7.2 once more, we see that loc𝗌iϕ(𝗌j)0\operatorname{loc}_{\mathsf{s}_{i}}\phi(\mathsf{s}_{j})\neq 0 if and only if loc𝗌jψ(𝗌i)0.\operatorname{loc}_{\mathsf{s}_{j}}\psi(\mathsf{s}_{i})\neq 0. Hence, the images of ϕ(𝗌i)\phi(\mathsf{s}_{i}) in SS are of the form:

loc(ϕ(𝗌1))\displaystyle\operatorname{loc}(\phi(\mathsf{s}_{1})) =(0,,)\displaystyle=(0,\ast,\cdot)
loc(ψ(𝗌2))\displaystyle\operatorname{loc}(\psi(\mathsf{s}_{2})) =(0,0,),\displaystyle=(0,0,\ast),
loc(ψ(𝗌3))\displaystyle\operatorname{loc}(\psi(\mathsf{s}_{3})) =(,0,0),\displaystyle=(\ast,0,0),

where \ast is nonzero and \cdot may or may not be zero. This completes the inductive step since d𝗊𝗌1𝗌2𝗌3=d𝗊<d1.d_{\mathsf{q}\mathsf{s}_{1}\mathsf{s}_{2}\mathsf{s}_{3}}=d_{\mathsf{q}}<d_{1}.

8. Proof of main results

For this section, let ff be a modular form of weight two, level NN, and trivial character, with ring of integers 𝒪f\mathcal{O}_{f}, and let 𝒪f\wp\subset\mathcal{O}_{f} be an ordinary prime lying over p2N.p\nmid 2N. Denote by 𝒪\mathcal{O} the completion.

8.1. A result of Skinner-Urban

The following result is a corollary to the proof of the main conjecture [47].

Theorem 8.1.1 (Skinner-Urban).

Let KK be an imaginary quadratic field of discriminant prime to NpNp in which pp splits. Assume that \wp is ordinary for ff and that:

  • the mod \wp representation T¯f\overline{T}_{f} is absolutely irreducible;

  • N=N1N2,N=N_{1}N_{2}, where every factor of N1N_{1} is split in KK and N2N_{2} is the squarefree product of an odd number primes inert in KK.

If SelΛ(𝐖f)\operatorname{Sel}_{\mathcal{F}_{\Lambda}}(\mathbf{W}_{f}) is Λ\Lambda-cotorsion, then

charΛSelΛ(𝐖f)(𝝀(1))2\operatorname{char}_{\Lambda}\operatorname{Sel}_{\mathcal{F}_{\Lambda}}(\mathbf{W}_{f})^{\vee}\subset(\bm{\lambda}(1))^{2}

as ideals of Λ\Lambda, where 𝛌(1)Λ\bm{\lambda}(1)\in\Lambda is the element constructed in (6.1).

Proof.

We must explain some details and notations of [47], in which it is assumed that T¯f\overline{T}_{f} is ramified at every |N2.\ell|N_{2}. As in [47], we let OLO_{L} be the ring of integers of a suitable finite extension of p\mathbb{Q}_{p} and consider ff as a specialization of a suitable Hida family 𝐟\mathbf{f}. This family is parametrized by 𝕀,\mathbb{I}, which is a normal domain and a finite integral extension of OLWO_{L}\llbracket W\rrbracket. We write ΓK=ΓK+×ΓK\Gamma_{{K}}=\Gamma_{{K}}^{+}\times\Gamma_{{K}}^{-} for the Galois group of the maximal p\mathbb{Z}_{p}-extension of K{K} and its decomposition into cyclotomic/anticyclotomic components. For a sufficiently large finite set of primes Σ\Sigma, there is [47, Theorem 12.3.1] a three-variable pp-adic LL-function 𝐟,KΣ𝕀ΓK.\mathcal{L}_{\mathbf{f},{K}}^{\Sigma}\in\mathbb{I}\llbracket\Gamma_{{K}}\rrbracket. (Here the superscript Σ\Sigma refers to removing Euler factors at primes in Σ\Sigma, or relaxing local conditions for a Selmer group.) Letting γ\gamma^{-} be a topological generator of ΓK\Gamma_{{K}}^{-}, we may expand:

(48) 𝐟,KΣ=a0+a1(γ1)+a2(γ1)2+\mathcal{L}_{\mathbf{f},{K}}^{\Sigma}=a_{0}+a_{1}(\gamma^{-}-1)+a_{2}(\gamma^{-}-1)^{2}+\ldots

where ai𝕀ΓK+a_{i}\in\mathbb{I}\llbracket\Gamma_{{K}}^{+}\rrbracket. Let ChKΣ(𝐟)𝕀ΓKCh_{{K}_{\infty}}^{\Sigma}(\mathbf{f})\subset\mathbb{I}\llbracket\Gamma_{{K}}\rrbracket be the characteristic ideal of the three-variable Selmer group as considered in [47]. Skinner and Urban deduce

(49) ChKΣ(𝐟)(𝐟,KΣ)Ch_{K_{\infty}}^{\Sigma}(\mathbf{f})\subset(\mathcal{L}_{\mathbf{f},K}^{\Sigma})

by proving (see [47, Theorem 6.5.4, Proposition 12.3.6, Proposition 13.4.1]):

  1. (1)

    If P𝕀ΓKP\subset\mathbb{I}\llbracket\Gamma_{K}\rrbracket is a height one prime which is not of the form P+𝕀ΓKP_{+}\mathbb{I}\llbracket\Gamma_{K}\rrbracket for some P+𝕀ΓK+P_{+}\subset\mathbb{I}\llbracket\Gamma_{K}^{+}\rrbracket, then

    ordPChKΣ(𝐟)ordP(𝐟,KΣ).\operatorname{ord}_{P}Ch_{K_{\infty}}^{\Sigma}(\mathbf{f})\geq\operatorname{ord}_{P}(\mathcal{L}_{\mathbf{f},K}^{\Sigma}).
  2. (2)

    If T¯f\overline{T}_{f} is ramified at every |N2,\ell|N_{2}, then ordP(𝐟,KΣ)=0\operatorname{ord}_{P}(\mathcal{L}_{\mathbf{f},K}^{\Sigma})=0 for all height one primes PP of the form P+𝕀ΓKP_{+}\mathbb{I}\llbracket\Gamma_{K}\rrbracket for some P+𝕀ΓK+P_{+}\subset\mathbb{I}\llbracket\Gamma_{K}^{+}\rrbracket.

Although (2) does not apply, we claim that we may replace (49) by the weaker inclusion:

(50) ChKΣ(𝐟)(ai)(𝐟,KΣ)Ch_{K_{\infty}}^{\Sigma}(\mathbf{f})\cdot(a_{i})\supset(\mathcal{L}_{\mathbf{f},K}^{\Sigma})

where aia_{i} is any of the terms in (48). Indeed, because both sides of (50) are divisorial, it suffices to check that ordP(𝐟,KΣ)ordP(ai)\operatorname{ord}_{P}(\mathcal{L}_{\mathbf{f},K}^{\Sigma})\leq\operatorname{ord}_{P}(a_{i}) for all PP as in (2). But this is clear: if (𝐟,KΣ)(\mathcal{L}_{\mathbf{f},K}^{\Sigma}) is zero modulo PkP^{k} for such a prime PP, then aia_{i} is as well. By [50] aia_{i} may be chosen so that its image under the specialization map 𝟙:𝕀ΓK+OL\mathbbm{1}:\mathbb{I}\llbracket\Gamma_{K}^{+}\rrbracket\to O_{L} is nonzero. Fix such a choice α~\widetilde{\alpha}.

The divisibility (50) also (trivially) implies a divisibility for the Fitting ideal of the 3-variable Selmer group:

(51) (α~)FittKΣ(𝐠)(𝐠,KΣ).(\widetilde{\alpha})Fitt_{K_{\infty}}^{\Sigma}(\mathbf{g})\subset(\mathcal{L}_{\mathbf{g},K}^{\Sigma}).

Specializing (51) to the anticyclotomic variable, we obtain

charΛSelΣ(K,f)LpΣ(K,f) in Λp,\operatorname{char}_{\Lambda}\operatorname{Sel}^{\Sigma}(K_{\infty},f)\subset L_{p}^{\Sigma}(K_{\infty},f)\text{ in }\Lambda\otimes\mathbb{Q}_{p},

where LpΣ(K,f)L_{p}^{\Sigma}(K_{\infty},f) is a certain Σ\Sigma-primitive anticyclotomic LL-function, and SelΣ(K,f)\operatorname{Sel}^{\Sigma}(K_{\infty},f) is the Σ\Sigma-primitive Selmer group. Replacing [47, Proposition 3.3.19] by [40, Proposition A.2] (and using the hypothesis that the Selmer group is Λ\Lambda-cotorsion), we may convert this to an imprimitive divisibility

(52) charΛSelΛ(𝐖f)(𝝀(1))2 in Λp.\operatorname{char}_{\Lambda}\operatorname{Sel}_{\mathcal{F}_{\Lambda}}(\mathbf{W}_{f})^{\vee}\subset(\bm{\lambda}(1))^{2}\text{ in }\Lambda\otimes\mathbb{Q}_{p}.

NB: The anticyclotomic pp-adic LL-function appearing in [47, §12.3.5], which emerges naturally from the specialization of the three-variable pp-adic LL-function, is normalized using Hida’s canonical period, whereas 𝝀(1)2\bm{\lambda}(1)^{2} is the LL-function constructed in [3], normalized using Gross’s period. However, these LL-functions differ only by a power of pp. Similarly, the local cohomology groups H1(K,𝐖f)H^{1}(K_{\ell},\mathbf{W}_{f}) for |N2\ell|N_{2} have characteristic ideal a power of ()(\wp), so the choice of local condition at primes |N2\ell|N_{2} does not change the characteristic ideal in Λp\Lambda\otimes\mathbb{Q}_{p}. See the appendix, and [40] for a detailed discussion.

To upgrade (52) to a divisibility in Λ\Lambda, we simply note that the μ\mu-invariant of 𝝀(1)\bm{\lambda}(1) is 0 by [50]. ∎

8.2. The Heegner point main conjecture

In this subsection, we prove the following main theorem.

Theorem 8.2.1.

Let ff be a modular form of weight two, level NN, and trivial character, with an ordinary prime \wp of its ring of integers 𝒪f\mathcal{O}_{f}, and let KK be an imaginary quadratic field. Assume:

  • N=N+NN=N^{+}N^{-}, where every factor of N+N^{+} is split in KK, and NN^{-} is a squarefree product of primes inert in KK.

  • The residue characteristic pp of \wp does not divide 2DKN2D_{K}N, and pp splits in KK.

  • The modulo \wp representation T¯f\overline{T}_{f} associated to ff is absolutely irreducible; if p=3,p=3, assume that T¯f\overline{T}_{f} is not induced from a character of G3G_{\mathbb{Q}\sqrt{-3}}.

Then, for all {𝖰,ϵ𝖰}𝖭ν(N)\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}^{\nu(N^{-})} such that (𝛋(𝖰,ϵ𝖰))0,(\bm{\kappa}(\mathsf{Q},\epsilon_{\mathsf{Q}}))\neq 0, we have

rkΛSelΛ(𝖰)(𝐓)=crkΛSelΛ(𝖰)(𝐖f)=1\operatorname{rk}_{\Lambda}\operatorname{Sel}_{\mathcal{F}_{\Lambda}(\mathsf{Q})}(\mathbf{T})=\operatorname{crk}_{\Lambda}\operatorname{Sel}_{\mathcal{F}_{\Lambda}(\mathsf{Q})}(\mathbf{W}_{f})=1

and

charΛ((SelΛ(𝖰)(𝐖f))tors)=charΛ(SelΛ(𝖰)(𝐓)(𝜿(𝖰)))2 in Λp.\operatorname{char}_{\Lambda}\left(\left(\operatorname{Sel}_{\mathcal{F}_{\Lambda}(\mathsf{Q})}(\mathbf{W}_{f})^{\vee}\right)_{\operatorname{tors}}\right)=\operatorname{char}_{\Lambda}\left(\frac{\operatorname{Sel}_{\mathcal{F}_{\Lambda}(\mathsf{Q})}(\mathbf{T})}{(\bm{\kappa}(\mathsf{Q}))}\right)^{2}\text{ in }\Lambda\otimes\mathbb{Q}_{p}.

For all {𝖰,ϵ𝖰}𝖭ν(N)+1\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}^{\nu(N^{-})+1} such that 𝛌(𝖰)0,\bm{\lambda}(\mathsf{Q})\neq 0,

rkΛSelΛ(𝖰)(𝐓)=crkΛSelΛ(𝖰)(𝐖f)=0\operatorname{rk}_{\Lambda}\operatorname{Sel}_{\mathcal{F}_{\Lambda}(\mathsf{Q})}(\mathbf{T})=\operatorname{crk}_{\Lambda}\operatorname{Sel}_{\mathcal{F}_{\Lambda}(\mathsf{Q})}(\mathbf{W}_{f})=0

and

charΛ(SelΛ(𝖰)(𝐖f))=(𝝀(𝖰))2 in Λp.\operatorname{char}_{\Lambda}\left(\operatorname{Sel}_{\mathcal{F}_{\Lambda}(\mathsf{Q})}(\mathbf{W}_{f})^{\vee}\right)=(\bm{\lambda}(\mathsf{Q}))^{2}\text{ in }\Lambda\otimes\mathbb{Q}_{p}.

If moreover the image of the GG_{\mathbb{Q}} action on T¯f\overline{T}_{f} contains a nontrivial scalar, then the equalities hold in Λ\Lambda.

Proof.

Given ff, apply Proposition 7.2.6 to the standard Selmer structure (,𝖲)(\mathcal{F},\mathsf{S}) on TfT_{f} (with local conditions the image of the Kummer map at all v¯\underline{v} such that v|Npv|Np). Let {Qn,ϵQn}\left\{Q_{n},\epsilon_{Q_{n}}\right\} be a sequence of weakly admissible pairs representing the resulting pair {𝖰,ϵ𝖰}𝖭\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}. Let gng_{n} be the resulting sequence of newforms of level NQnNQ_{n} obtained from Theorem 7.1.5 (and Proposition 7.2.2); gng_{n} may only be defined for 𝔉\mathfrak{F}-many nn.

Step 1.

{𝖰,ϵ𝖰}𝖭ν(N)+1\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}^{\nu(N^{-})+1}.

Proof.

The prime (T)(T), corresponding to the trivial character, does not lie in the exceptional set Σ\Sigma for Λ\mathcal{F}_{\Lambda} (see the proofs of [33, Lemma 5.3.13] and [24, Lemma 2.2.7]). Hence SelΛ(𝖰)(𝐓)=0,\operatorname{Sel}_{\mathcal{F}_{\Lambda}(\mathsf{Q})}(\mathbf{T})=0, which by Theorem 3.4.6 and the nontriviality of (𝜿,𝝀)(\bm{\kappa},\bm{\lambda}) implies the claim. ∎

Step 2.

For any fixed jj,

(𝝀(𝖰))(𝝀gn(1))(modj,Tj)(\bm{\lambda}(\mathsf{Q}))\equiv(\bm{\lambda}_{g_{n}}(1))\pmod{\wp^{j},T^{j}}

for 𝔉\mathfrak{F}-many nn.

Proof.

Recall notations of (5.1) and (6.1). By definition, the image of 𝝀(𝖰)\bm{\lambda}(\mathsf{Q}) modulo (j,Tj)(\wp^{j},T^{j}) is a map Gal(Kj/K)𝒪\operatorname{Gal}(K_{j}/K)\to\mathcal{O} obtained, for 𝔉\mathfrak{F}-many nn, by evaluating a surjective map Fn:MQn𝕋Qn𝒪(f)𝒪(f)/jF_{n}:M_{Q_{n}}\otimes_{\mathbb{T}^{Q_{n}}}\mathcal{O}(f)\to\mathcal{O}(f)/\wp^{j} of 𝕋Qn\mathbb{T}^{Q_{n}}-modules at certain CM points, where 𝒪(f)\mathcal{O}(f) is 𝒪\mathcal{O} with 𝕋Qn\mathbb{T}^{Q_{n}}-action by ff. Recall that the map is chosen to factor through multiplicatation by 𝒪(f)/πj+C\mathcal{O}(f)/\pi^{j+C} for the constant CC of Lemma 5.1.4, and that πC(MQn𝕋Qn𝒪(f))\pi^{C}(M_{Q_{n}}\otimes_{\mathbb{T}^{Q_{n}}}\mathcal{O}(f)) is principal. When gng_{n} has a sufficiently deep congruence to ff, 𝒪(gn)/πj+C=𝒪(f)/πj+C\mathcal{O}(g_{n})/\pi^{j+C}=\mathcal{O}(f)/\pi^{j+C} as 𝕋Qn\mathbb{T}^{Q_{n}}-modules, and the composite Gn:MQn𝒪(gn)𝒪(gn)/πjG_{n}:M_{Q_{n}}\to\mathcal{O}(g_{n})\to\mathcal{O}(g_{n})/\pi^{j} induces a unit multiple of FnF_{n}, where MQn𝒪(gn)M_{Q_{n}}\to\mathcal{O}(g_{n}) is the quaternionic modular form associated to gng_{n}. But GnG_{n} is the very map whose evaluation at CM points is used to define 𝝀gn(1)\bm{\lambda}_{g_{n}}(1), and the claim follows. ∎

Step 3.

For any fixed jj,

FittΛSelΛ(𝖰)(𝐖f)FittΛSelgn,Λ(𝐖gn)(modj,Tj)\operatorname{Fitt}_{\Lambda}\operatorname{Sel}_{\mathcal{F}_{\Lambda}(\mathsf{Q})}(\mathbf{W}_{f})^{\vee}\equiv\operatorname{Fitt}_{\Lambda}\operatorname{Sel}_{\mathcal{F}_{g_{n},\Lambda}}(\mathbf{W}_{g_{n}})^{\vee}\pmod{\wp^{j},T^{j}}

for 𝔉\mathfrak{F}-many nn.

Proof.

Since fitting ideals are stable under base change and T¯f\overline{T}_{f} has no GKG_{K}-fixed points, it suffices to show that

(53) FittΛ(SelΛ(𝖰)(𝐖f[πj,Tj]))=FittΛ(Selgn,Λ(𝐖gn[πj,Tj]))\operatorname{Fitt}_{\Lambda}\left(\operatorname{Sel}_{\mathcal{F}_{\Lambda}(\mathsf{Q})}(\mathbf{W}_{f}[\pi^{j},T^{j}])\right)=\operatorname{Fitt}_{\Lambda}\left(\operatorname{Sel}_{\mathcal{F}_{g_{n},\Lambda}}(\mathbf{W}_{g_{n}}[\pi^{j},T^{j}])\right)

for 𝔉\mathfrak{F}-many nn. Note that, for 𝔉\mathfrak{F}-many nn, 𝐖f[πj,Tj]=𝐖gn[πj,Tj]\mathbf{W}_{f}[\pi^{j},T^{j}]=\mathbf{W}_{g_{n}}[\pi^{j},T^{j}] as finite Galois modules, and

SelΛ(𝖰)(𝐖f[πj,Tj])\operatorname{Sel}_{\mathcal{F}_{\Lambda}(\mathsf{Q})}(\mathbf{W}_{f}[\pi^{j},T^{j}])

is isomorphic to a submodule of H1(KΣQn/K,𝐖f[πj,Tj])H^{1}(K^{\Sigma\cup Q_{n}}/K,\mathbf{W}_{f}[\pi^{j},T^{j}]) defined by certain local conditions. We will show that these local conditions coincide with the ones defining Selgn,Λ(𝐖gn[πj,Tj])\operatorname{Sel}_{\mathcal{F}_{g_{n},\Lambda}}(\mathbf{W}_{g_{n}}[\pi^{j},T^{j}]). At a jj-admissible prime qn|Qnq_{n}|Q_{n}, this is clear (cf. the proof of Construction 5.2.2(1)). At v|Nv|N, the local conditions are simply the kernels

ker(H1(Kv,𝐖f[πj,Tj])H1(Kv,𝐖f)),ker(H1(Kv,𝐖gn[πj,Tj])H1(Kv,𝐖gn)).\ker\left(H^{1}(K_{v},\mathbf{W}_{f}[\pi^{j},T^{j}])\to H^{1}(K_{v},\mathbf{W}_{f})\right),\;\;\ker\left(H^{1}(K_{v},\mathbf{W}_{g_{n}}[\pi^{j},T^{j}])\to H^{1}(K_{v},\mathbf{W}_{g_{n}})\right).

If v|Nv|N is a prime of multiplicative reduction for ff, then 𝐖f=𝐖gn\mathbf{W}_{f}=\mathbf{W}_{g_{n}} as GKvG_{K_{v}} modules for 𝔉\mathfrak{F}-many nn, so the local conditions clearly coincide. For other places of bad reduction, the inertia co-invariants Tf,IvT_{f,I_{v}} and Tgn,IvT_{g_{n},I_{v}} are finite, and we may assume they are isomorphic, say with exponent bounded by πMj\pi^{M-j} for some M0M\geq 0. It follows that the local conditions are also given by the kernels of

ker(H1(Kv,𝐖f[πj,Tj])H1(Kv,𝐖f[πM])),ker(H1(Kv,𝐖gn[πj,Tj])H1(Kv,𝐖gn[πM])),\ker\left(H^{1}(K_{v},\mathbf{W}_{f}[\pi^{j},T^{j}])\to H^{1}(K_{v},\mathbf{W}_{f}[\pi^{M}])\right),\;\;\ker\left(H^{1}(K_{v},\mathbf{W}_{g_{n}}[\pi^{j},T^{j}])\to H^{1}(K_{v},\mathbf{W}_{g_{n}}[\pi^{M}])\right),

which also agree for 𝔉\mathfrak{F}-many nn.

For primes v|pv|p, it suffices to compare the kernels

(H1(Kv,gr𝐖f[πj,Tj])H1(Kv,gr𝐖f)),ker(H1(Kv,gr𝐖gn[πj,Tj])H1(Kv,gr𝐖gn)).\left(H^{1}(K_{v},\operatorname{gr}\mathbf{W}_{f}[\pi^{j},T^{j}])\to H^{1}(K_{v},\operatorname{gr}\mathbf{W}_{f})\right),\;\;\ker\left(H^{1}(K_{v},\operatorname{gr}\mathbf{W}_{g_{n}}[\pi^{j},T^{j}])\to H^{1}(K_{v},\operatorname{gr}\mathbf{W}_{g_{n}})\right).

A similar argument as above applies provided that H0(Kv,gr𝐖f)=w|vH0(K,w,grWf)H^{0}(K_{v},\operatorname{gr}\mathbf{W}_{f})=\prod_{w|v}H^{0}(K_{\infty,w},\operatorname{gr}W_{f}) is finite, which it is because apa_{p} cannot be a root of unity. ∎

Step 4.

Conclusion of the proof.

Step 1 shows that NQnN^{-}Q_{n} is the squarefree product of an odd number primes inert in KK for 𝔉\mathfrak{F}-many nn, and by Theorem 3.4.6 applied to the Euler system (𝜿gn,𝝀gn)(\bm{\kappa}_{g_{n}},\bm{\lambda}_{g_{n}}) for TgnT_{g_{n}}, Selgn,Λ(𝐖gn)\operatorname{Sel}_{\mathcal{F}_{g_{n},\Lambda}}(\mathbf{W}_{g_{n}}) is then Λ\Lambda-cotorsion. By Theorem 8.1.1, for such nn we have:

(54) FittΛSelgn,Λ(𝐖gn)(𝝀gn(1))2Λ.\operatorname{Fitt}_{\Lambda}\operatorname{Sel}_{\mathcal{F}_{g_{n},\Lambda}}(\mathbf{W}_{g_{n}})^{\vee}\subset(\bm{\lambda}_{g_{n}}(1))^{2}\subset\Lambda.

On the other hand, by Theorem 3.4.6, the theorem would follow from:

(55) FittΛSelΛ(𝖰)(𝐖f)(𝝀(𝖰))2Λ.\operatorname{Fitt}_{\Lambda}\operatorname{Sel}_{\mathcal{F}_{\Lambda}(\mathsf{Q})}(\mathbf{W}_{f})^{\vee}\subset(\bm{\lambda}(\mathsf{Q}))^{2}\subset\Lambda.

For the passage between characteristic ideal and fitting ideals, recall that the characteristic ideal of any Λ\Lambda-module is the smallest divisorial ideal containing the Fitting ideal; cf. [47, Corollary 3.2.9]. Steps 2 and 3 allow us to pass from (54) to (55). ∎

Corollary 8.2.2.

Under the hypotheses of Theorem 8.2.1, if additionally ν(N)\nu(N^{-}) is even, then the Heegner point main conjecture holds for ff in Λp\Lambda\otimes\mathbb{Q}_{p}; that is, there is a pseudo-isomorphism of Λ\Lambda-modules:

Sel(K,Af[])ΛMM\operatorname{Sel}(K_{\infty},A_{f}[\wp^{\infty}])^{\vee}\approx\Lambda\oplus M\oplus M

for some torsion Λ\Lambda-module MM, and

charΛ(Sel(K,TfΛ)Λ𝜿(1))=charΛ(M)\operatorname{char}_{\Lambda}\left(\frac{\operatorname{Sel}(K,T_{f}\otimes\Lambda)}{\Lambda\bm{\kappa}(1)}\right)=\operatorname{char}_{\Lambda}(M)

as ideals of Λp\Lambda\otimes\mathbb{Q}_{p}. If additionally the image of the Galois action on T¯f\overline{T}_{f} contains a nontrivial scalar, then the equality is true in Λ.\Lambda.

Corollary 8.2.3.

Under the hypotheses of Theorem 8.2.1, the bipartite Euler system (κ(1,),λ(1,))(\kappa(1,\cdot),\lambda(1,\cdot)) of (6.2.10) is nontrivial.

Proof.

Let {𝖰,ϵ𝖰}𝖭\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N} be such that Sel(𝖰)(Tf)=0,\operatorname{Sel}_{\mathcal{F}(\mathsf{Q})}(T_{f})=0, where again \mathcal{F} is the standard Selmer structure for TfT_{f}. As noted in Step 1 of the proof of Theorem 8.2.1, we have 𝟙(SelΛ(𝖰)(𝐖f))0\mathbbm{1}(\operatorname{Sel}_{\mathcal{F}_{\Lambda}(\mathsf{Q})}(\mathbf{W}_{f}))\neq 0, so 𝟙(𝝀(𝖰))0\mathbbm{1}(\bm{\lambda}(\mathsf{Q}))\neq 0; this implies λ(1,𝖰)0\lambda(1,\mathsf{Q})\neq 0 by Remark 6.2.12.∎

Corollary 8.2.3 is generalized by Theorem A.1.1 of the appendix.

8.3. Kolyvagin’s conjecture

Let ff, \wp, KK, and N+NN^{+}N^{-} be as in (3.1.1) and (5.1.1).

8.3.1.

For any 𝗆𝖪\mathsf{m}\in\mathsf{K}, define the 𝗆\mathsf{m}-transverse Selmer ranks

(56) r𝗆±=rk𝒪Sel(𝗆)(Tf)±,r_{\mathsf{m}}^{\pm}=\operatorname{rk}_{\mathcal{O}}\operatorname{Sel}_{\mathcal{F}(\mathsf{m})}(T_{f})^{\pm},

where ±\pm refers to the τ\tau eigenvalue; note that this is well-defined because the local conditions defining (𝗆)\mathcal{F}(\mathsf{m}) are all τ\tau-stable. When 𝗆=1,\mathsf{m}=1, the r1±r_{1}^{\pm} are the classical Selmer ranks of ff.

Proposition 8.3.2.

For all 𝗆𝗅𝖪\mathsf{m}\mathsf{l}\in\mathsf{K}, and for each δ{±}\delta\in\left\{\pm\right\}, either:

  • r𝗆𝗅δ=r𝗆δ1,r_{\mathsf{m}\mathsf{l}}^{\delta}=r_{\mathsf{m}}^{\delta}-1, loc𝗅δ(Sel(𝗆)(Tf))δ0,\operatorname{loc}_{\mathsf{l}}^{\delta}(\operatorname{Sel}_{\mathcal{F}(\mathsf{m})}(T_{f}))^{\delta}\neq 0, and 𝗅δ(Sel(𝗆𝗅)(Tf))δ=0.\partial_{\mathsf{l}}^{\delta}(\operatorname{Sel}_{\mathcal{F}(\mathsf{m}\mathsf{l})}(T_{f}))^{\delta}=0.

  • r𝗆𝗅δ=r𝗆δ+1,r_{\mathsf{m}\mathsf{l}}^{\delta}=r_{\mathsf{m}}^{\delta}+1, loc𝗅δ(Sel(𝗆)(Tf))δ=0,\operatorname{loc}_{\mathsf{l}}^{\delta}(\operatorname{Sel}_{\mathcal{F}(\mathsf{m})}(T_{f}))^{\delta}=0, and 𝗅δ(Sel(𝗆𝗅)(Tf))δ0.\partial_{\mathsf{l}}^{\delta}(\operatorname{Sel}_{\mathcal{F}(\mathsf{m}\mathsf{l})}(T_{f}))^{\delta}\neq 0.

Proof.

If 𝗅(𝗆,𝖰)=(𝗆𝗅,𝖰)+(𝗆,𝖰)\mathcal{F}^{\mathsf{l}}(\mathsf{m},\mathsf{Q})=\mathcal{F}(\mathsf{m}\mathsf{l},\mathsf{Q})+\mathcal{F}(\mathsf{m},\mathsf{Q}) and 𝗅(𝗆,𝖰)=(𝗆𝗅,𝖰)(𝗆,𝖰)\mathcal{F}_{\mathsf{l}}(\mathsf{m},\mathsf{Q})=\mathcal{F}(\mathsf{m}\mathsf{l},\mathsf{Q})\cap\mathcal{F}(\mathsf{m},\mathsf{Q}), then we have a τ\tau-equivariant exact sequence

0Sel𝗅(𝗆,𝖰)(Tf)Sel𝗅(𝗆,𝖰)(Tf)𝖧1(K𝗅,Tf),0\to\operatorname{Sel}_{\mathcal{F}_{\mathsf{l}}(\mathsf{m},\mathsf{Q})}(T_{f})\to\operatorname{Sel}_{\mathcal{F}^{\mathsf{l}}(\mathsf{m},\mathsf{Q})}(T_{f})\to\mathsf{H}^{1}(K_{\mathsf{l}},T_{f}),

where the image of the final arrow has rank two and is self-annihilating under the local Tate pairing by Propositions 2.7.2 and 2.7.4. Since the Tate pairing of two classes with opposite τ\tau eigenvalues is necessarily zero, the proposition follows. ∎

Lemma 8.3.3.

Suppose given elements c±𝖧1(K,Tf)±c^{\pm}\in\mathsf{H}^{1}(K,T_{f})^{\pm}, where ±\pm is the τ\tau eigenvalue. Then there exists a Kolyvagin-admissible ultraprime 𝗅\mathsf{l} such that

c±0loc𝗅±c±.c^{\pm}\neq 0\implies\operatorname{loc}_{\mathsf{l}}^{\pm}c^{\pm}.

If (sclr) holds for TfT_{f}, then the same is true for elements c±𝖧1(K,Tf/πj)c^{\pm}\in\mathsf{H}^{1}(K,T_{f}/\pi^{j}).

Proof.

The proof of Theorem 3.3.4 applies almost verbatim, except that in the proof of Lemma 3.3.6 we will have two homomorphisms ϕ±HomGK(GL,T¯f)±\phi^{\pm}\in\operatorname{Hom}_{G_{K}}(G_{L},\overline{T}_{f})^{\pm}, and we must choose gGLg\in G_{L} so that ϕϵ(g)\phi^{\epsilon}(g) has nonzero component in the ϵ\epsilon eigenspace of τ\tau for both signs ϵ\epsilon (unless ϕϵ\phi^{\epsilon} is itself 0); for each ϵ\epsilon, this condition is satisfied outside a proper subgroup of GLG_{L}, so indeed there exists gGLg\in G_{L} such that both conditions are satisfied. With this modification, the rest of the proof applies unchanged. ∎

Lemma 8.3.4.

Suppose that the bipartite Euler system (κ(1,),λ(1,))(\kappa(1,\cdot),\lambda(1,\cdot)) of (6.2.10) is nontrivial. Then, for all 𝗆𝖪\mathsf{m}\in\mathsf{K}, (κ(𝗆,),λ(𝗆,))(\kappa(\mathsf{m},\cdot),\lambda(\mathsf{m},\cdot)) is nontrivial.

In particular, for all 𝗆𝖪\mathsf{m}\in\mathsf{K} and {𝖰,ϵ𝖰}𝖭𝗆\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}_{\mathsf{m}}:

rk𝒪Sel(𝗆,𝖰)(Tf)1{κ(𝗆,𝖰)0,ν(N)+|𝖰| evenλ(𝗆,𝖰)0,ν(N)+|𝖰| odd.\operatorname{rk}_{\mathcal{O}}\operatorname{Sel}_{\mathcal{F}(\mathsf{m},\mathsf{Q})}(T_{f})\leq 1\iff\begin{cases}\kappa(\mathsf{m},\mathsf{Q})\neq 0,&\nu(N^{-})+|\mathsf{Q}|\text{ even}\\ \lambda(\mathsf{m},\mathsf{Q})\neq 0,&\nu(N^{-})+|\mathsf{Q}|\text{ odd}.\end{cases}
Proof.

Proposition 6.2.11 implies that, for fixed 𝗆\mathsf{m}, the pair (κ(𝗆,),λ(𝗆,))(\kappa(\mathsf{m},\cdot),\lambda(\mathsf{m},\cdot)) forms a bipartite Euler system with sign ν(N)\nu(N^{-}) for the self-dual Selmer structure ((𝗆),𝖲𝗆)(\mathcal{F}(\mathsf{m}),\mathsf{S}\cup\mathsf{m}) on TfT_{f}. We will prove that, for any 𝗆𝗅𝖪\mathsf{m}\mathsf{l}\in\mathsf{K}, if (κ(𝗆,),λ(𝗆,))(\kappa(\mathsf{m},\cdot),\lambda(\mathsf{m},\cdot)) is nontrivial then so is (κ(𝗆𝗅,),λ(𝗆𝗅,)).(\kappa(\mathsf{m}\mathsf{l},\cdot),\lambda(\mathsf{m}\mathsf{l},\cdot)).

Choose {𝖰,ϵ𝖰}𝖭𝗆ν(N)+1\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}_{\mathsf{m}}^{\nu(N^{-})+1} such that Sel(𝗆,𝖰)(Tf)=0\operatorname{Sel}_{\mathcal{F}(\mathsf{m},\mathsf{Q})}(T_{f})=0 and 𝗅𝖰\mathsf{l}\not\in\mathsf{Q}; this is possible by Corollary 3.3.7. By Proposition 8.3.2, we may choose a nonzero

dSel(𝗆𝗅,𝖰)(Tf).d\in\operatorname{Sel}_{\mathcal{F}(\mathsf{m}\mathsf{l},\mathsf{Q})}(T_{f}).

Applying Theorem 3.3.4 to dd, let 𝗊\mathsf{q} be admissible with sign ϵ𝗊\epsilon_{\mathsf{q}} such that 𝗊𝖰𝗆𝗅\mathsf{q}\not\in\mathsf{Q}\mathsf{m}\mathsf{l} and loc𝗊d0.\operatorname{loc}_{\mathsf{q}}d\neq 0. By Proposition 3.3.3 for the Selmer structures (𝗆,𝖰𝗊)\mathcal{F}(\mathsf{m},\mathsf{Q}\mathsf{q}) and (𝗆,𝖰)\mathcal{F}(\mathsf{m},\mathsf{Q}),

(57) rk𝒪Sel(𝗆,𝖰𝗊)(Tf)=1.\operatorname{rk}_{\mathcal{O}}\operatorname{Sel}_{\mathcal{F}(\mathsf{m},\mathsf{Q}\mathsf{q})}(T_{f})=1.

Hence, by hypothesis, κ(𝗆,𝖰𝗊)\kappa(\mathsf{m},\mathsf{Q}\mathsf{q}) generates Sel(𝗆,𝖰𝗊)(Tf)\operatorname{Sel}_{\mathcal{F}(\mathsf{m},\mathsf{Q}\mathsf{q})}(T_{f}) up to finite index, and in particular 𝗊κ(𝗆,𝖰𝗊)0\partial_{\mathsf{q}}\kappa(\mathsf{m},\mathsf{Q}\mathsf{q})\neq 0. Now, taking the sum of local pairings and using Proposition 2.7.2,

(58) 0=𝗏d,κ(𝗆,𝖰𝗊)𝗏=d,κ(𝗆,𝖰𝗊)𝗅+d,κ(𝗆,𝖰𝗊)𝗊.0=\sum_{\mathsf{v}}\langle d,\kappa(\mathsf{m},\mathsf{Q}\mathsf{q})\rangle_{\mathsf{v}}=\langle d,\kappa(\mathsf{m},\mathsf{Q}\mathsf{q})_{\mathsf{l}}+\langle d,\kappa(\mathsf{m},\mathsf{Q}\mathsf{q})\rangle_{\mathsf{q}}.

Since the latter pairing is nonzero by construction, the former is as well, and so, by Proposition 6.2.11(1),

Res𝗅κ(𝗆,𝖰𝗊))0κ(𝗆𝗅,𝖰𝗊)0.\operatorname{Res}_{\mathsf{l}}\kappa(\mathsf{m},\mathsf{Q}\mathsf{q}))\neq 0\implies\kappa(\mathsf{m}\mathsf{l},\mathsf{Q}\mathsf{q})\neq 0.

8.3.5.

For any 𝗆𝖪\mathsf{m}\in\mathsf{K}, define the vanishing order of the Kolyvagin system at 𝗆\mathsf{m}:

(59) ν𝗆={min{|𝗇|:𝗇𝖪,λ(𝗇𝗆,1)0},ν(N) odd,min{|𝗇|:𝗇𝖪,κ(𝗇𝗆,1)0},ν(N) even.\nu_{\mathsf{m}}=\begin{cases}\min\left\{|\mathsf{n}|\,:\,\mathsf{n}\in\mathsf{K},\,\lambda(\mathsf{n}\cup\mathsf{m},1)\neq 0\right\},&\nu(N^{-})\text{ odd,}\\ \min\left\{|\mathsf{n}|\,:\,\mathsf{n}\in\mathsf{K},\,\kappa(\mathsf{n}\cup\mathsf{m},1)\neq 0\right\},&\nu(N^{-})\text{ even}.\end{cases}
Corollary 8.3.6.

If (κ(1,),λ(1,))(\kappa(1,\cdot),\lambda(1,\cdot)) is nontrivial, and in particular under the hypotheses of Theorem A.1.1, we have for all 𝗆𝖪\mathsf{m}\in\mathsf{K}:

  • If ν(N)\nu(N^{-}) is odd, then ν𝗆=max{r𝗆+,r𝗆}\nu_{\mathsf{m}}=\max\left\{r_{\mathsf{m}}^{+},r_{\mathsf{m}}^{-}\right\} and r𝗆±ϵf12(mod2).r_{\mathsf{m}}^{\pm}\equiv\frac{\epsilon_{f}-1}{2}\pmod{2}.

  • If ν(N)\nu(N^{-}) is even, then ν𝗆=max{r𝗆+,r𝗆}1\nu_{\mathsf{m}}=\max\left\{r_{\mathsf{m}}^{+},r_{\mathsf{m}}^{-}\right\}-1 and ϵf(1)|𝗆|+ν𝗆+1\epsilon_{f}\cdot(-1)^{|\mathsf{m}|+\nu_{\mathsf{m}}+1} is the larger τ\tau eigenspace.

In particular, if rk𝒪Sel(K,Tf)=1,\operatorname{rk}_{\mathcal{O}}\operatorname{Sel}(K,T_{f})=1, then L(f/K,1)0.L^{\prime}(f/K,1)\neq 0.

Proof.

Note that the parity statement in Theorem 3.3.8 implies r𝗆++r𝗆ν(N)+1(mod2)r_{\mathsf{m}}^{+}+r_{\mathsf{m}}^{-}\equiv\nu(N^{-})+1\pmod{2} for all 𝗆𝖪\mathsf{m}\in\mathsf{K} by Proposition 8.3.2. So if we have rk𝒪Sel(𝗆𝗇)(Tf)1,\operatorname{rk}_{\mathcal{O}}\operatorname{Sel}_{\mathcal{F}(\mathsf{m}\mathsf{n})}(T_{f})\leq 1, for some 𝗇𝖪\mathsf{n}\in\mathsf{K} such that 𝗆𝗇=\mathsf{m}\cap\mathsf{n}=\emptyset, then the τ\tau-equivariant localization map

Sel(𝗆)(Tf)𝗅𝗇Hunr1(K𝗅,Tf)\operatorname{Sel}_{\mathcal{F}(\mathsf{m})}(T_{f})\to\oplus_{\mathsf{l}\in\mathsf{n}}H^{1}_{\operatorname{unr}}(K_{\mathsf{l}},T_{f})

has kernel of rank either zero (if ν(N)\nu(N^{-}) is odd), or at most one (if ν(N)\nu(N^{-}) is even). It follows that ν𝗆max{r𝗆+,r𝗆}\nu_{\mathsf{m}}\geq\max\left\{r_{\mathsf{m}}^{+},r_{\mathsf{m}}^{-}\right\} in the former case and ν𝗆max{r𝗆+,r𝗆}1\nu_{\mathsf{m}}\geq\max\left\{r_{\mathsf{m}}^{+},r_{\mathsf{m}}^{-}\right\}-1 in the latter. The opposite inequality follows readily from repeated applications of Proposition 8.3.2 and Lemma 8.3.3. The additional claims about parity and τ\tau eigenvalues follow from Proposition 6.2.7.

For the final statement, take 𝗆=1.\mathsf{m}=1. The corollary implies ν1=0,\nu_{1}=0, so κ(1,1)0.\kappa(1,1)\neq 0. Since κ(1,1)\kappa(1,1) is the Kummer image of the classical Heegner point yKE(K),y_{K}\in E(K), the result follows from the Gross-Zagier theorem of [52]. ∎

It remains to relate the vanishing of the patched Kolyvagin classes to the classical vanishing order

(60) νclassical{min{ν(m):λ¯(m,1)0},ν(N) odd,min{ν(m):c¯(m,1)0},ν(N) even.\nu_{\text{classical}}\coloneqq\begin{cases}\min\left\{\nu(m)\,:\,\overline{\lambda}(m,1)\neq 0\right\},&\nu(N^{-})\text{ odd,}\\ \min\left\{\nu(m)\,:\,\overline{c}(m,1)\neq 0\right\},&\nu(N^{-})\text{ even.}\end{cases}
Corollary 8.3.7.

If (κ(1,),λ(1,))(\kappa(1,\cdot),\lambda(1,\cdot)) is nontrivial, and in particular under the hypotheses of Theorem A.1.1, νclassical\nu_{\text{classical}} is finite. If (sclr) holds for ff, then νclassical=ν1,\nu_{\text{classical}}=\nu_{1}, and in particular:

  • If ν(N)\nu(N^{-}) is odd, then νclassical=max{r1+,r1}\nu_{\text{classical}}=\max\left\{r_{1}^{+},r_{1}^{-}\right\} and r1±ϵf12(mod2).r_{1}^{\pm}\equiv\frac{\epsilon_{f}-1}{2}\pmod{2}.

  • If ν(N)\nu(N^{-}) is even, then νclassical=max{r1+,r1}1\nu_{\text{classical}}=\max\left\{r_{1}^{+},r_{1}^{-}\right\}-1 and ϵf(1)1+νclassical\epsilon_{f}\cdot(-1)^{1+\nu_{\text{classical}}} is the larger τ\tau eigenspace.

Proof.

The finiteness of the classical vanishing order is clear by construction: if a patched Kolyvagin class or element is nontrivial, then infinitely many of the classical Kolyvagin classes or elements defining it are nontrivial. This also shows νclassicalν1.\nu_{\text{classical}}\leq\nu_{1}. We will check that equality holds under the condition (sclr). Suppose first ν(N)\nu(N^{-}) is even. We abbreviate by cj(m,1)H1(K,Tj)c_{j}(m,1)\in H^{1}(K,T_{j}) the image of c¯(m,1)\overline{c}(m,1) when v(Im)j.v_{\wp}(I_{m})\geq j. Given some nonzero cj(m,1)c_{j}(m,1), one may show as in [34, p. 309] that there exist classes cj(mn,1)0c_{j}(m_{n},1)\neq 0 with v(Imn)v_{\wp}(I_{m_{n}})\to\infty and ν(mn)=ν(m).\nu(m_{n})=\nu(m). (In [34], additional hypotheses are put on the image of the Galois action, but the argument goes through by invoking Lemma 8.3.3.) In particular, the sequence mnm_{n} defines a nonzero κ(𝗆,1)\kappa(\mathsf{m},1) witnessing ν1νclassical\nu_{1}\leq\nu_{\text{classical}}.

Now suppose that ν(N)\nu(N^{-}) is odd, and that λj(m,1)0\lambda_{j}(m,1)\neq 0 where ν(m)=νclassical\nu(m)=\nu_{\text{classical}}. We choose an auxiliary {𝗊,ϵ𝗊}𝖭\left\{\mathsf{q},\epsilon_{\mathsf{q}}\right\}\in\mathsf{N} with the following properties:

  • ϵ𝗊\epsilon_{\mathsf{q}} is the sign of the larger τ\tau eigenspace in Sel(Tf).\operatorname{Sel}_{\mathcal{F}}(T_{f}).

  • The localization map loc𝗊\operatorname{loc}_{\mathsf{q}} is trivial on Sel(Tf).\operatorname{Sel}_{\mathcal{F}}(T_{f}).

To ensure the second condition, we may choose Frob𝗊G\operatorname{Frob}_{\mathsf{q}}\in G_{\mathbb{Q}} to be a complex conjugation. Let {qn,ϵqn}\left\{q_{n},\epsilon_{q_{n}}\right\} represent {𝗊,ϵ𝗊}\left\{\mathsf{q},\epsilon_{\mathsf{q}}\right\} as in Remark 5.1.6. Once again, the argument of [34, p. 309] implies that, for each nn, there exists mnm_{n} with cj(mn,qn)0c_{j}(m_{n},q_{n})\neq 0 and v(Imn).v_{\wp}(I_{m_{n}})\to\infty. We therefore obtain a nonzero patched class κ(𝗆,𝗊)\kappa(\mathsf{m},\mathsf{q}) with |𝗆|=νclassical.|\mathsf{m}|=\nu_{\text{classical}}. Repeating the argument of Lemma 8.3.4, it follows that ν1νclassical+1.\nu_{1}\leq\nu_{\text{classical}}+1. For contradiction, we assume that

νclassical=ν11=rϵ𝗊1.\nu_{\text{classical}}=\nu_{1}-1=r^{\epsilon_{\mathsf{q}}}-1.

This implies 𝗊κ(𝗆,𝗊)=0,\partial_{\mathsf{q}}\kappa(\mathsf{m},\mathsf{q})=0, so by Lemma 8.3.4 and Proposition 3.3.3, we conclude

rk𝒪Sel(𝗆,1)(Tf)=2.\operatorname{rk}_{\mathcal{O}}\operatorname{Sel}_{\mathcal{F}(\mathsf{m},1)}(T_{f})=2.

However, by Proposition 8.3.2 and the assumption |𝗆|=ν11=rϵ𝗊1,|\mathsf{m}|=\nu_{1}-1=r^{\epsilon_{\mathsf{q}}}-1, rk𝒪Sel(𝗆,1)(Tf)ϵ𝗊\operatorname{rk}_{\mathcal{O}}\operatorname{Sel}_{\mathcal{F}(\mathsf{m},1)}(T_{f})^{\epsilon_{\mathsf{q}}} is odd, hence equal to one. Proposition 8.3.2 then implies

Sel(𝗆,1)(Tf)ϵ𝗊Sel(Tf)ϵ𝗊,\operatorname{Sel}_{\mathcal{F}(\mathsf{m},1)}(T_{f})^{\epsilon_{\mathsf{q}}}\subset\operatorname{Sel}_{\mathcal{F}}(T_{f})^{\epsilon_{\mathsf{q}}},

so

loc𝗊(Sel(𝗆,1)(Tf))=0\operatorname{loc}_{\mathsf{q}}(\operatorname{Sel}_{\mathcal{F}(\mathsf{m},1)}(T_{f}))=0

by the choice of 𝗊\mathsf{q}. However, this contradicts Proposition 3.3.3, so we must have ν1=νclassical.\nu_{1}=\nu_{classical}.

Appendix A Kolyvagin’s conjecture for inert or non-ordinary pp

A.1. The main result

In this appendix, we shall prove the following:

Theorem A.1.1.

Let ff be a non-CM modular form of weight two, level NN, and trivial character, with a prime \wp of its ring of integers 𝒪f\mathcal{O}_{f}, and let KK be an imaginary quadratic field. Assume:

  • N=N+NN=N^{+}N^{-}, where every factor of N+N^{+} is split in KK, and NN^{-} is a squarefree product of an even number of primes inert in KK.

  • The residue characteristic pp of \wp does not divide 2DKN2D_{K}N.

  • The modulo \wp representation T¯f\overline{T}_{f} associated to ff is absolutely irreducible; if p=3,p=3, assume that T¯f\overline{T}_{f} is not induced from a character of G3G_{\mathbb{Q}\sqrt{-3}}.

  • If pp is inert in KK or or apa_{p} is not a \wp-adic unit, then there exists some prime 0||N\ell_{0}||N.

  • If apa_{p} is not a \wp-adic unit, then either 0\ell_{0} may be chosen above so that AfA_{f} has non-split toric reduction at 0\ell_{0}, or the image of the Galois action on TfT_{f} contains a conjugate of SL2(p)SL_{2}(\mathbb{Z}_{p}).

Then (κ(1,),λ(1,))(\kappa(1,\cdot),\lambda(1,\cdot)) is nontrivial.

A.1.2.

If pp is split in KK, then this is simply Corollary 8.2.3. If pp is non-ordinary or inert in KK, then the anticyclotomic main conjecture is currently not known in full generality; however, since all we are interested in specialization the trivial character, we will show that the result may instead be obtained, more circuitously, by combining main conjectures for quadratic twists of ff. The proof applies equally well to the split ordinary case.

A.2. Comparing periods

A.2.1.

Let ff be a modular form of weight two, level NN, and trivial character, with ring of integers 𝒪f\mathcal{O}_{f} of its coefficient field, and let 𝒪f\wp\subset\mathcal{O}_{f} be an ordinary prime lying over p2Np\nmid 2N, with associated completion 𝒪\mathcal{O}; we assume that T¯f\overline{T}_{f} is absolutely irreducible. There are two ways to normalize the anticyclotomic pp-adic LL-function, as explained in [50, 40]. For any factorization N=N1N2,N=N_{1}N_{2}, where N1N_{1} and N2N_{2} are coprime, the congruence ideal ηf(N1,N2)𝒪\eta_{f}(N_{1},N_{2})\subset\mathcal{O} is defined as

(61) πf(Ann𝕋N1,N2(kerπf)),\pi_{f}(\operatorname{Ann}_{\mathbb{T}_{N_{1},N_{2}}}(\ker\pi_{f}))\cdot,

where πf:𝕋N1,N2𝒪\pi_{f}:\mathbb{T}_{N_{1},N_{2}}\to\mathcal{O} is the projection giving the Hecke eigenvalues of ff. Hida’s canonical period [23] is defined (up to pp-adic units) by:

(62) Ωfcan=(f,f)ηf(N,1),\Omega_{f}^{can}=\frac{(f,f)}{\eta_{f}(N,1)},

where (f,f)(f,f) is the Peterson inner product.

On the other hand, if N=N1N2N=N_{1}N_{2} where N2N_{2} is squarefree with an odd number of prime factors, then the ff-isotypic part of the Hecke module p[XN1,N2]\mathbb{Z}_{p}[X_{N_{1},N_{2}}] is free of rank one, generated by an element φf,N2\varphi_{f,N_{2}}. Gross’s period is defined as:

(63) Ωf,N2=(f,f)φf,N2,φf,N2,\Omega_{f,N_{2}}=\frac{(f,f)}{\langle\varphi_{f,N_{2}},\varphi_{f,N_{2}}\rangle},

where ,\langle\cdot,\cdot\rangle is the canonical intersection pairing on p[XN1,N2]\mathbb{Z}_{p}[X_{N_{1},N_{2}}]. This period occurs naturally in anticyclotomic Iwasawa theory due to the well-known special value formula of Gross.

Proposition A.2.2.

Let KK be an imaginary quadratic field of discriminant prime to NpNp, and suppose that N=N+NN=N^{+}N^{-} where all factors of N+N^{+} are split in KK, and NN^{-} is a squarefree product of an odd number primes inert in KK. Then L(f/K,1)¯Ωf,NL(f/K,1)\in\overline{\mathbb{Q}}\cdot\Omega_{f,N^{-}} and the element λ(1)Λ\lambda(1)\in\Lambda constructed in (6.2.10) satisfies:

(64) L(f/K,1)Ωf,N=λ(1)2\frac{L(f/K,1)}{\Omega_{f,N^{-}}}=\lambda(1)^{2}

up to pp-adic units.

A.2.3.

Consider the Néron model 𝒜f\mathcal{A}_{f} of AfA_{f} over \mathbb{Z}_{\ell}. The Tamagawa numbers over KK are:

(65) tf()=lg𝒪Φ(𝒜f)(𝒪K/)t_{f}(\ell)=\lg_{\mathcal{O}}\Phi(\mathcal{A}_{f})(\mathcal{O}_{K}/\ell)_{\wp}

We also require a variant:

(66) cf()=lg𝒪Φ(𝒜f)(𝔽¯)c_{f}(\ell)=\lg_{\mathcal{O}}\Phi(\mathcal{A}_{f})(\overline{\mathbb{F}}_{\ell})_{\wp}

The number cf()c_{f}(\ell) is the maximal exponent ee such that Af[fe]A_{f}[\wp^{e}_{f}] is unramified at \ell. If \ell is inert in KK, then tf()=cf().t_{f}(\ell)=c_{f}(\ell). The following theorem generalizes [44, 28] and [40, Theorem 6.8].

Proposition A.2.4.

Suppose N=N1N2N=N_{1}N_{2} where N2N_{2} is squarefree with an odd number of prime factors and coprime to N1N_{1}. For any 0||N\ell_{0}||N, we have:

vηf(N,1)vφf,N2,φf,N2|N2cf()σ(N2)cf(0).v_{\wp}\eta_{f}(N,1)-v_{\wp}\langle\varphi_{f,N_{2}},\varphi_{f,N_{2}}\rangle\geq\sum_{\ell|N_{2}}c_{f}(\ell)-\sigma(N_{2})c_{f}(\ell_{0}).
Proof.

For a decomposition N=N1N2N=N_{1}^{\prime}N_{2}^{\prime} with N2N_{2}^{\prime} the squarefree product of an even number of primes that do not divide N1N_{1}^{\prime}, one defines δ(N1,N2)𝒪\delta(N_{1}^{\prime},N_{2}^{\prime})\subset\mathcal{O} to be the “degree” of an optimal modular parametrization JN1,N2AfJ_{N_{1}^{\prime},N_{2}^{\prime}}\to A_{f} as explained in [40, 28]. By [40, Proposition 6.6], we have:

(67) vδ(N,1)=cf(0)+vφf,0,φf,0.v_{\wp}\delta(N,1)=c_{f}(\ell_{0})+v_{\wp}\langle\varphi_{f,\ell_{0}},\varphi_{f,\ell_{0}}\rangle.

On the other hand, [15, Lemma 4.17] implies that:

(68) vηf(N/0,0)vφf,0,φf,0.v_{\wp}\eta_{f}(N/\ell_{0},\ell_{0})\geq v_{\wp}\langle\varphi_{f,\ell_{0}},\varphi_{f,\ell_{0}}\rangle.

Because 𝕋N/0,0\mathbb{T}_{N/\ell_{0},\ell_{0}} is a quotient of 𝕋N,1\mathbb{T}_{N,1}, we conclude that:

(69) vηf(N,1)vδ(N,1)cf(0).v_{\wp}\eta_{f}(N,1)\geq v_{\wp}\delta(N,1)-c_{f}(\ell_{0}).

We apply [40, Proposition 6.6] again, this time to the decomposition N=N1N2N=N_{1}N_{2} and any r|N2.r|N_{2}. This yields:

(70) vδ(N1r,N2/r)=cf(r)+vφf,N2,φf,N2.v_{\wp}\delta(N_{1}r,N_{2}/r)=c_{f}(r)+v_{\wp}\langle\varphi_{f,N_{2}},\varphi_{f,N_{2}}\rangle.

If N2N_{2} is prime, this is sufficient to conclude. If not, we may choose r0.r\neq\ell_{0}.

The results of Ribet-Takahashi and Khare [44, 28] imply that:

(71) vδ(N,1)vδ(N1r,N2/r)+|N2/rcf()σ(N2/r)cf(0).v_{\wp}\delta(N,1)\geq v_{\wp}\delta(N_{1}r,N_{2}/r)+\sum_{\ell|N_{2}/r}c_{f}(\ell)-\sigma(N_{2}/r)\cdot c_{f}(\ell_{0}).

(If 0|N1,\ell_{0}|N_{1}, we are using the fact that both rr and 0\ell_{0} exactly divide N1r.N_{1}r.) Combining (69), (70), and (71) completes the proof.

Remark A.2.5.

If 0\ell_{0} is residually ramified, the inequality is an equality. In [53, Theorem 6.4], more restrictive conditions are given under which this result holds.

A.2.6.

We will also require the following related result:

Proposition A.2.7.

Let ff and \wp be as above and suppose 0||N\ell_{0}||N. Then, in the notation of the proof of Theorem A.2.4,

vηf(N,1)vφf,0,φf,0cf(0).v_{\wp}\eta_{f}(N,1)-v_{\wp}\langle\varphi_{f,\ell_{0}},\varphi_{f,\ell_{0}}\rangle\leq c_{f}(\ell_{0}).
Proof.

Let J=J0(N)J=J_{0}(N) be the modular Jacobian and let 𝕋\mathbb{T} be the full Hecke algebra of level NN. Write π:𝕋𝔪𝒪f,\pi:\mathbb{T}_{\mathfrak{m}}\to\mathcal{O}_{f,\wp} for the projection associated to gg and let II be its kernel. The claim will follow from (67) once we establish

(72) vηf(N,1)vδ(N,1).v_{\wp}\eta_{f}(N,1)\leq v_{\wp}\delta(N,1).

Indeed, if JAJ\to A is an optimal parametrization, then the dual map AJA^{\vee}\to J^{\vee} is an inclusion. The composition

ϕ:JAAJwNJ\phi:J\to A\to A^{\vee}\to J^{\vee}\xrightarrow{w_{N}}J

is a Hecke-equivariant endomorphism; by (22), its image in End(J)𝔪\operatorname{End}(J)_{\mathfrak{m}} may be identified with some y𝕋𝔪y\in\mathbb{T}_{\mathfrak{m}}. Because imϕJ[I],\operatorname{im}\phi\subset J[I], we have yAnn(I)y\in\operatorname{Ann}(I). By the definition of δ(N,1)\delta(N,1),

(π(y))=δ(N,1)𝒪.(\pi(y))=\delta(N,1)\subset\mathcal{O}.

This implies (72). ∎

A.3. Cyclotomic Iwasawa theory: ordinary case

A.3.1.

Let Λ=𝒪Gal(/)\Lambda_{\mathbb{Q}_{\infty}}=\mathcal{O}\llbracket\operatorname{Gal}(\mathbb{Q}_{\infty}/\mathbb{Q})\rrbracket be the cyclotomic Iwasawa algebra. We denote by 𝟙:Λ𝒪\mathbbm{1}:\Lambda_{\mathbb{Q}_{\infty}}\to\mathcal{O} and 𝟙:Λ𝒪\mathbbm{1}:\Lambda\to\mathcal{O} the specializations at the trivial character. If \wp is ordinary and Σ\Sigma is a finite set of rational primes, we consider the Σ\Sigma-ordinary cyclotomic Selmer group

Sel(,Wf)=ker(H1(,𝐖f)vΣ{p}H1(Iv,𝐖f)×H1(p,𝐖f)Hord1(p,𝐖f)),\operatorname{Sel}(\mathbb{Q}_{\infty},W_{f})=\ker\left(H^{1}(\mathbb{Q},\mathbf{W}_{f})\to\prod_{v\not\in\Sigma\cup\left\{p\right\}}H^{1}(I_{v},\mathbf{W}_{f})\times\frac{H^{1}(\mathbb{Q}_{p},\mathbf{W}_{f})}{H^{1}_{\operatorname{ord}}(\mathbb{Q}_{p},\mathbf{W}_{f})}\right),

and denote by Ch,fΣΛCh_{\mathbb{Q}_{\infty},f}^{\Sigma}\subset\Lambda_{\mathbb{Q}_{\infty}} the characteristic ideal of its Pontryagin dual.

From the work of Skinner and Urban, we deduce the following result.

Theorem A.3.2 (Skinner-Urban).

Let KK be an imaginary quadratic field of discriminant prime to NpNp in which pp splits. Assume that \wp is good ordinary for ff and that:

  • the mod \wp representation T¯f\overline{T}_{f} is absolutely irreducible;

  • N=N1N2,N=N_{1}N_{2}, where every factor of N1N_{1} is split in KK and N2N_{2} is the squarefree product of an odd number primes inert in KK.

Then there exists an element αΛ\alpha\in\Lambda_{\mathbb{Q}_{\infty}} such that 𝟙(α)\mathbbm{1}(\alpha) divides

Ωf,N2Ωfcanηf(N,1)φf,N2,φf,N2\frac{\Omega_{f,N_{2}}}{\Omega_{f}^{can}}\sim\frac{\eta_{f}(N,1)}{\langle\varphi_{f,N_{2}},\varphi_{f,N_{2}}\rangle}

in 𝒪\mathcal{O} and

(α)Ch,fCh,fχK(Lp(,f))(Lp(,fχK)).(\alpha)Ch_{\mathbb{Q}_{\infty},f}Ch_{\mathbb{Q}_{\infty},f\otimes\chi_{K}}\subset(L_{p}(\mathbb{Q}_{\infty},f))(L_{p}(\mathbb{Q}_{\infty},f\otimes\chi_{K})).
Proof.

Recall the divisibility established in the course of the proof of Theorem 8.1.1 for the Fitting ideal of the 3-variable Selmer group:

(73) (α~)FittKΣ(𝐠)(𝐟,KΣ),(\widetilde{\alpha})Fitt_{K_{\infty}}^{\Sigma}(\mathbf{g})\subset(\mathcal{L}_{\mathbf{f},K}^{\Sigma}),

where α~𝕀[ΓK+]\widetilde{\alpha}\in\mathbb{I}[\Gamma_{K}^{+}] may be chosen such that α~\widetilde{\alpha} specializes to a unit multiple of Ωf,N2/Ωfcan\Omega_{f,N_{2}}/\Omega_{f}^{can} at the trivial character (by [50]). By Lemma 3.2.5, Corollary 3.2.9(i), and Corollary 3.2.20(iii) of [47], specializing to the cyclotomic variable yields a divisibility

(74) (α)Ch,fΣCh,fχKΣ(LpΣ(,f))(LpΣ(,fχK)),(\alpha)Ch^{\Sigma}_{\mathbb{Q}_{\infty},f}Ch^{\Sigma}_{\mathbb{Q}_{\infty},f\otimes\chi_{K}}\subset(L^{\Sigma}_{p}(\mathbb{Q}_{\infty},f))(L^{\Sigma}_{p}(\mathbb{Q}_{\infty},f\otimes\chi_{K})),

where α\alpha is the image of α~\widetilde{\alpha}. The desired divisibility for the imprimitive LL-functions and Selmer groups follows by [47, Proposition 3.2.18]. ∎

A.4. Cyclotomic Iwasawa theory: general case

A.4.1.

Kato gave a formulation [27] of the cyclotomic main conjectures which also applies to non-ordinary primes. For any finite set of rational primes Σ\Sigma, consider the strict Selmer group:

Selstr(,Wf)=ker(H1(,𝐖f)vΣ{p}H1(Iv,𝐖f)×H1(p,𝐖f)),\operatorname{Sel}_{\operatorname{str}}(\mathbb{Q}_{\infty},W_{f})=\ker\left(H^{1}(\mathbb{Q},\mathbf{W}_{f})\to\prod_{v\not\in\Sigma\cup\left\{p\right\}}H^{1}(I_{v},\mathbf{W}_{f})\times H^{1}(\mathbb{Q}_{p},\mathbf{W}_{f})\right),

and denote by ChKato,fΣΛCh_{\operatorname{Kato},f}^{\Sigma}\subset\Lambda_{\mathbb{Q}_{\infty}} the characteristic ideal of its Pontryagin dual.

The Iwasawa cohomology H1(Σ/,𝐓)H^{1}(\mathbb{Q}^{\Sigma}/\mathbb{Q},\mathbf{T}) (for Σ\Sigma the set of primes dividing NpNp) is free of rank one when T¯f\overline{T}_{f} is absolutely irreducible, and under that hypothesis Kato defined an element zKatoH1(Σ/,𝐓)z_{\operatorname{Kato}}\in H^{1}(\mathbb{Q}^{\Sigma}/\mathbb{Q},\mathbf{T}) which is closely related to LL-values of ff.

Remark A.4.2.

In [27, Theorem 12.4], it is only asserted that zKatoz_{\operatorname{Kato}} is integral when the image of the Galois action on TfT_{f} contains a conjugate of SL2(p)SL_{2}(\mathbb{Z}_{p}). However, the proof (13.14 of loc. cit.) only requires the fact that any two 𝒪\mathcal{O}-lattices in TfpT_{f}\otimes\mathbb{Q}_{p} are homothetic, which holds whenever T¯f\overline{T}_{f} is absolutely irreducible. Note that zKato=𝐳γ(p)z_{\operatorname{Kato}}=\mathbf{z}_{\gamma}^{(p)} in the notation of [27], where γTf\gamma\in T_{f} is the sum of any generators of the two eigenspaces for complex conjugation.

A.4.3.

Kato’s main conjecture [27, Conjecture 12.10] then asserts that:

(75) charΛ(H1(Σ/,𝐓)ΛzKato)=ChKato,f.\operatorname{char}_{\Lambda_{\mathbb{Q}_{\infty}}}\left(\frac{H^{1}(\mathbb{Q}^{\Sigma}/\mathbb{Q},\mathbf{T})}{\Lambda_{\mathbb{Q}_{\infty}}\cdot z_{\operatorname{Kato}}}\right)=Ch_{\operatorname{Kato},f}.

If \wp is ordinary, then this conjecture is equivalent to the usual cyclotomic main conjecture for ff by [27, 17.13]. We abbreviate by ZKato,fΛZ_{\operatorname{Kato},f}\subset\Lambda_{\mathbb{Q}_{\infty}} the left side of (75).

The following is the analogue of Theorem A.3.2 in the non-ordinary case, due to Wan.

Theorem A.4.4 (Wan).

Let KK be an imaginary quadratic field in which pp splits. Assume that \wp is a good, non-ordinary prime for ff and that:

  • For all primes |N\ell|N ramified in KK, that Tf|GT_{f}|_{G_{\mathbb{Q}_{\ell}}} is Steinberg with sign 1-1. At least one such prime exists.

  • For all |N\ell|N, \ell is either split or ramified in KK.

Then

ChKato,fChKato,fχKZKato,fZKato,fχK in Λ.Ch_{\operatorname{Kato},f}\cdot Ch_{\operatorname{Kato},f\otimes\chi_{K}}\subset Z_{\operatorname{Kato},f}\cdot Z_{\operatorname{Kato},f\otimes\chi_{K}}\text{ in }\Lambda_{\mathbb{Q}_{\infty}}.
Proof.

This is proven in [51, p. 29]; compare to the proof of Corollary 3.32, where it is assumed that ZKato,fχKCharKato,fχKZ_{\operatorname{Kato},f\otimes\chi_{K}}\subset\operatorname{Char}_{\operatorname{Kato},f\otimes\chi_{K}}. ∎

Additionally, Kato [27] has proven one direction of his conjecture in our setting:

Theorem A.4.5 (Kato).

Let ff be a modular forms of weight two, level NN, and trivial character, and 𝒪f\wp\subset\mathcal{O}_{f} a prime of good reduction with odd residue characteristic. Then ChKato,f0Ch_{\operatorname{Kato},f}\neq 0 and

ZKato,fChKato,fZ_{\operatorname{Kato},f}\subset Ch_{\operatorname{Kato},f}

in Λp.\Lambda_{\mathbb{Q}_{\infty}}\otimes\mathbb{Q}_{p}. In particular, if apa_{p} is a \wp-adic unit, then Sel(,Wf)\operatorname{Sel}(\mathbb{Q}_{\infty},W_{f}) is Λ\Lambda_{\mathbb{Q}_{\infty}}-cotorsion and

Lp(,f)Ch,fL_{p}(\mathbb{Q}_{\infty},f)\subset Ch_{\mathbb{Q}_{\infty},f}

in Λp.\Lambda_{\mathbb{Q}_{\infty}}\otimes\mathbb{Q}_{p}. If the image of the Galois action on TfT_{f} contains SL2(p)SL_{2}(\mathbb{Z}_{p}), then all of the inclusions hold in Λ\Lambda_{\mathbb{Q}_{\infty}}.

A.4.6.

Denote by μ(f)\mu(f) the μ\mu-invariant of Ch,fCh_{\mathbb{Q}_{\infty},f} or ChKato,fCh_{\operatorname{Kato},f} in the ordinary or non-ordinary case, respectively. To control the powers of \wp in Theorem A.4.5, we will use the following.

Lemma A.4.7.

Let ff and gg be modular forms of weight two and trivial character such that T¯f\overline{T}_{f} is absolutely irreducible. Suppose that ff and gg have a congruence modulo j\wp^{j}, i.e. there is a common completion 𝒪\mathcal{O} of 𝒪f\mathcal{O}_{f} and 𝒪g\mathcal{O}_{g} and, in some basis, a congruence of 𝒪\mathcal{O}-valued associated Galois representations

TfTg(modj).T_{f}\equiv T_{g}\pmod{\wp^{j}}.

If μ(f)<j\mu(f)<j, then μ(g)=μ(f).\mu(g)=\mu(f).

Proof.

For the sake of notation, assume \wp is ordinary; this makes no difference to the proof. By [18], μ(f)\mu(f) is also the μ\mu-invariant of Ch,fΣCh_{\mathbb{Q}_{\infty},f}^{\Sigma} for any finite set of primes Σ\Sigma, and likewise for gg. If Σ\Sigma contains all primes dividing the level of either ff or gg, then we have:

(76) SelΣ(,Wf)[j]SelΣ(,Wg)[j]\operatorname{Sel}^{\Sigma}(\mathbb{Q}_{\infty},W_{f})[\wp^{j}]\simeq\operatorname{Sel}^{\Sigma}(\mathbb{Q}_{\infty},W_{g})[\wp^{j}]

as Λ\Lambda_{\mathbb{Q}_{\infty}}-modules. Let MfM_{f} and MgM_{g} be the Pontryagin duals of SelΣ(,Wf)\operatorname{Sel}^{\Sigma}(\mathbb{Q}_{\infty},W_{f}) and SelΣ(,Wg)[j]\operatorname{Sel}^{\Sigma}(\mathbb{Q}_{\infty},W_{g})[\wp^{j}], respectively, and let 𝔓=()Λ\mathfrak{P}=(\wp)\subset\Lambda_{\mathbb{Q}_{\infty}}. Then we have a congruence

MfΛ/𝔓jMgΛ/𝔓j.M_{f}\otimes\Lambda_{\mathbb{Q}_{\infty}}/\mathfrak{P}^{j}\simeq M_{g}\otimes\Lambda_{\mathbb{Q}_{\infty}}/\mathfrak{P}^{j}.

Since μ(f)=lgMf,(𝔓)<j,\mu(f)=\lg M_{f,(\mathfrak{P})}<j, where (𝔓)(\mathfrak{P}) denotes the localization,

(77) Mf,(𝔓)Λ/𝔓j=Mf,(𝔓)Λ/𝔓j1,M_{f,(\mathfrak{P})}\otimes\Lambda_{\mathbb{Q}_{\infty}}/\mathfrak{P}^{j}=M_{f,(\mathfrak{P})}\otimes\Lambda_{\mathbb{Q}_{\infty}}/\mathfrak{P}^{j-1},

which implies the same for gg. Therefore Mg,(𝔓)Λ/𝔓j=Mg,(𝔓)M_{g,(\mathfrak{P})}\otimes\Lambda_{\mathbb{Q}_{\infty}}/\mathfrak{P}^{j}=M_{g,(\mathfrak{P})} and the result follows. ∎

Proof of Theorem A.1.1.

Let us suppose first that AfA_{f} has non-split toric reduction at 0\ell_{0} if \wp is non-ordinary. Fix once and for all an auxiliary quadratic imaginary field 𝒦\mathcal{K}, not contained in the fixed field K(Tf)K(T_{f}), such that:

  • If \wp is ordinary, then 0\ell_{0} is inert in 𝒦\mathcal{K} and every other factor of NpNp is split in 𝒦\mathcal{K}.

  • If \wp is non-ordinary, then 0\ell_{0} is ramified in 𝒦\mathcal{K} and every other factor of NpNp is split in 𝒦\mathcal{K}.

As in the proof of Theorem 8.2.1, begin by applying Proposition 7.2.6 and Theorem 7.1.5 to obtain some {𝖰,ϵ𝖰}𝖭\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N}, represented by QnQ_{n}, and a resulting sequence of newforms gng_{n} of NQnNQ_{n}; we make sure to choose each 𝗊𝖰\mathsf{q}\in\mathsf{Q} such that Frob𝗊\operatorname{Frob}_{\mathsf{q}} has trivial image in Gal(𝒦/)\operatorname{Gal}(\mathcal{K}/\mathbb{Q}), which is clearly possible.

Claim.

There exists a constant CC, depending only on ff, such that

(78) v(L(gn/K,1)Ωgncan)lg𝒪Sel(K,Wgn)+|NQntgn()+Cv_{\wp}\left(\frac{L(g_{n}/K,1)}{\Omega_{g_{n}}^{can}}\right)\leq\lg_{\mathcal{O}}\operatorname{Sel}(K,W_{g_{n}})+\sum_{\ell|NQ_{n}}t_{g_{n}}(\ell)+C

for 𝔉\mathfrak{F}-many nn.

Proof of claim.

Consider first the ordinary case. By Lemma A.4.7 and Theorem A.4.5,

(79) μ(fχ𝒦)(Lp(,gnχ𝒦))Ch,gnχ𝒦 in Λ\wp^{\mu(f\otimes\chi_{\mathcal{K}})}\cdot(L_{p}(\mathbb{Q}_{\infty},g_{n}\otimes\chi_{\mathcal{K}}))\subset Ch_{\mathbb{Q}_{\infty},g_{n}\otimes\chi_{\mathcal{K}}}\text{ in }\Lambda_{\mathbb{Q}_{\infty}}

for 𝔉\mathfrak{F}-many nn. By Theorem A.3.2 for gng_{n}, for 𝔉\mathfrak{F}-many nn we have

(80) (α)μ(fχK)Ch,gnCh,gnχ𝒦(Lp(,gn))Ch,gnχ𝒦,(\alpha)\cdot\wp^{\mu(f\otimes\chi_{K})}\cdot Ch_{\mathbb{Q}_{\infty},g_{n}}\cdot Ch_{\mathbb{Q}_{\infty},g_{n}\otimes\chi_{\mathcal{K}}}\subset(L_{p}(\mathbb{Q}_{\infty},g_{n}))\cdot Ch_{\mathbb{Q}_{\infty},g_{n}\otimes\chi_{\mathcal{K}}},

where, by Proposition A.2.7, 𝟙(α)\mathbbm{1}(\alpha) divides cf(0)\wp^{c_{f}(\ell_{0})} in 𝒪\mathcal{O}. Since Ch,gnχ𝒦0Ch_{\mathbb{Q}_{\infty},g_{n}\otimes\chi_{\mathcal{K}}}\neq 0, and since characteristic ideals are divisorial, we conclude that

(81) (α)μ(fχK)Ch,gn(Lp(,gn)).(\alpha)\cdot\wp^{\mu(f\otimes\chi_{K})}\cdot Ch_{\mathbb{Q}_{\infty},g_{n}}\subset(L_{p}(\mathbb{Q}_{\infty},g_{n})).

Applying the same argument to gnχKg_{n}\otimes\chi_{K}, we have:

(82) (α)2μ(fχKχ𝒦)+μ(fχ𝒦)Ch,gnCh,gnχK(Lp(,gn)(Lp(,gnχK)).(\alpha)^{2}\cdot\wp^{{\mu(f\otimes\chi_{K}\otimes\chi_{\mathcal{K}})}+\mu(f\otimes\chi_{\mathcal{K}})}\cdot Ch_{\mathbb{Q}_{\infty},g_{n}}\cdot Ch_{\mathbb{Q}_{\infty},g_{n}\otimes\chi_{K}}\subset(L_{p}(\mathbb{Q}_{\infty},g_{n})\cdot(L_{p}(\mathbb{Q}_{\infty},g_{n}\otimes\chi_{K})).

The result now follows from standard interpolation properties of both sides of (82), cf. e.g. [47, Theorem 3.6.11].

The non-ordinary case is similar: combining Theorem A.4.4 and Theorem A.4.5, we have for 𝔉\mathfrak{F}-many gng_{n}

μ(fχ𝒦)ChKato,gnZKato,gn,\wp^{\mu(f\otimes\chi_{\mathcal{K}})}\cdot Ch_{\operatorname{Kato},g_{n}}\subset Z_{\operatorname{Kato},g_{n}},

and likewise for the twist gnχKg_{n}\otimes\chi_{K}. Since ZKato,gnZ_{\operatorname{Kato},g_{n}} is principal, the result follows as in [51, Corollary 3.35]. Note that the pp-adic Tamagawa factor appearing there is trivial by [2, Proposition II.2]. ∎

As in Step 3 of the proof of Theorem 8.2.1, #Sel(K,Wgn)=#Sel(𝖰)(Wf)<\#\operatorname{Sel}(K,W_{g_{n}})=\#\operatorname{Sel}_{\mathcal{F}(\mathsf{Q})}(W_{f})<\infty for 𝔉\mathfrak{F}-many nn (the local conditions at v|Nv|N may be compared in the same way, and at v|pv|p we use [20, Lemma 7]).

Now, by combining the claim above with Proposition A.2.4, we have for 𝔉\mathfrak{F}-many nn:

(83) v(L(gn/K,1)Ωgn,NQn)lg𝒪Sel(K,Wgn)+|N+tgn()+Cv_{\wp}\left(\frac{L(g_{n}/K,1)}{\Omega_{g_{n},N^{-}Q_{n}}}\right)\leq\lg_{\mathcal{O}}\operatorname{Sel}(K,W_{g_{n}})+\sum_{\ell|N^{+}}t_{g_{n}}(\ell)+C^{\prime}

for a constant CC^{\prime} that does not depend on nn. In particular, for 𝔉\mathfrak{F}-many nn, L(gn/K,1)0L(g_{n}/K,1)\neq 0, which by parity considerations implies that ν(N)+|𝖰|\nu(N^{-})+|\mathsf{Q}| is odd. Exactly as in Step 2 of the proof of Theorem 8.2.1, we then conclude from (83) that λ(1,𝖰)0.\lambda(1,\mathsf{Q})\neq 0.

If we assume instead that \wp is non-ordinary but the image of the Galois action on TfT_{f} contains a conjugate of SL2(p)SL_{2}(\mathbb{Z}_{p}), then, rather than choosing 𝒦\mathcal{K} at the beginning, we choose {𝖰,ϵ𝖰}𝖭\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\}\in\mathsf{N} as in the proof of Theorem 8.2.1, but we also ensure that ϵ𝖰(𝗊)=1\epsilon_{\mathsf{Q}}(\mathsf{q})=-1 for at least one 𝗊\mathsf{q}. If {Qn,ϵQn}\left\{Q_{n},\epsilon_{Q_{n}}\right\} represents {𝖰,ϵ𝖰}\left\{\mathsf{Q},\epsilon_{\mathsf{Q}}\right\} and qn|Qnq_{n}|Q_{n} represents 𝗊\mathsf{q}, then for each nn, we choose an auxiliary imaginary quadratic field 𝒦n\mathcal{K}_{n} such that qnq_{n} is ramified in 𝒦n\mathcal{K}_{n} and all other factors of NQnpNQ_{n}p are split. Note that, by the proof of [17, Lemma 6.15], the image of the Galois action on TgnT_{g_{n}} contains a conjugate of SL2(p)SL_{2}(\mathbb{Z}_{p}) for 𝔉\mathfrak{F}-many nn. The argument is then the same as above; μ(fχ𝒦n)\mu(f\otimes\chi_{\mathcal{K}_{n}}) and μ(fχ𝒦nχK)\mu(f\otimes\chi_{\mathcal{K}_{n}}\otimes\chi_{K}) may not be uniformly bounded in nn, but by Kato’s result these error terms are not needed under the large-image hypothesis. ∎

Remark A.4.8.

In all cases, we crucially use the prime 0\ell_{0} to make the error terms uniform in nn.

References

  • [1] Patrick B Allen. Deformations of polarized automorphic galois representations and adjoint selmer groups. Duke Mathematical Journal, 165(13):2407–2460, 2016.
  • [2] Laurent Berger. Tamagawa numbers of some crystalline representations. arXiv preprint math/0209233, 2002.
  • [3] Massimo Bertolini and Henri Darmon. Iwasawa’s main conjecture for elliptic curves over anticyclotomic zp-extensions. Annals of mathematics, pages 1–64, 2005.
  • [4] Spencer Bloch and Kazuya Kato. L-functions and tamagawa numbers of motives. In The Grothendieck Festschrift, pages 333–400. Springer, 2007.
  • [5] Siegfried Bosch and Xavier Xarles. Component groups of néron models via rigid uniformization. Mathematische Annalen, 306(3):459–486, 1996.
  • [6] Ashay Burungale, Francesc Castella, and Chan-Ho Kim. A proof of perrin-riou’s heegner point main conjecture. arXiv preprint arXiv:1908.09512, 2019.
  • [7] Ashay A Burungale and Ye Tian. p-converse to a theorem of gross–zagier, kolyvagin and rubin. Inventiones mathematicae, 220(1):211–253, 2020.
  • [8] Francesc Castella, Giada Grossi, Jaehoon Lee, and Christopher Skinner. On the anticyclotomic iwasawa theory of rational elliptic curves at eisenstein primes. arXiv preprint arXiv:2008.02571, 2020.
  • [9] Francesc Castella and Xin Wan. Iwasawa main conjecture for heegner points: Supersingular case. arXiv preprint arXiv:1506.02538, 2015.
  • [10] Francesc Castella and Xin Wan. The iwasawa main conjectures for gl2\rm gl_{2} and derivatives of pp-adic ll-functions. arXiv preprint arXiv:2001.03878, 2020.
  • [11] Masataka Chida and Ming-Lun Hsieh. On the anticyclotomic iwasawa main conjecture for modular forms. Compositio Mathematica, 151(5):863–897, 2015.
  • [12] Christophe Cornut and Vinayak Vatsal. Nontriviality of rankin-selberg l-functions and cm points. London Mathematical Society Lecture Note Series, 320:121, 2007.
  • [13] Henri Darmon. Rational points on modular elliptic curves. Number 101. American Mathematical Soc., 2004.
  • [14] Henri Darmon, Fred Diamond, and Richard Taylor. Fermat’s last theorem. Current Developments in Mathematics, 2000.
  • [15] Henri Darmon, Fred Diamond, and Richard Taylor. Fermat’s last theorem. Current Developments in Mathematics, 2000.
  • [16] Fred Diamond, Richard Taylor, et al. Lifting modular mod \ell representations. Duke Mathematical Journal, 74(2):253–269, 1994.
  • [17] Najmuddin Fakhruddin, Chandrashekhar Khare, and Stefan Patrikis. Relative deformation theory, relative selmer groups, and lifting irreducible galois representations. arXiv preprint arXiv:1904.02374, 2019.
  • [18] Ralph Greenberg and Vinayak Vatsal. On the iwasawa invariants of elliptic curves. Inventiones mathematicae, 142(1):17–63, 2000.
  • [19] Benedict H Gross. Kolyvagin’s work on modular elliptic curves. L-functions and arithmetic (Durham, 1989), 153:235–256, 1991.
  • [20] Benedict H Gross and James A Parson. On the local divisibility of heegner points. In Number theory, analysis and geometry, pages 215–241. Springer, 2012.
  • [21] Alexandre Grothendieck and Michel Raynaud. Modeles de Néron et monodromie. In Groupes de Monodromie en Géométrie Algébrique, pages 313–523. Springer, 1972.
  • [22] David Helm. On maps between modular jacobians and jacobians of shimura curves. Israel Journal of Mathematics, 160(1):61–117, 2007.
  • [23] Haruzo Hida. Modules of congruence of hecke algebras and l-functions associated with cusp forms. American Journal of Mathematics, 110(2):323–382, 1988.
  • [24] Benjamin Howard. The heegner point kolyvagin system. Compositio Mathematica, 140(6):1439–1472, 2004.
  • [25] Benjamin Howard. Bipartite euler systems. 2006.
  • [26] Benjamin Howard. Variation of heegner points in hida families. Inventiones mathematicae, 167(1):91–128, 2007.
  • [27] Kazuya Kato et al. p-adic hodge theory and values of zeta functions of modular forms. Astérisque, 295:117–290, 2004.
  • [28] C. Khare. On isomorphisms between deformation rings and hecke rings. Inventiones mathematicae, 154:199–222, 2003.
  • [29] Mark Kisin. Moduli of finite flat group schemes, and modularity. Annals of Mathematics, pages 1085–1180, 2009.
  • [30] VA Kolyvagin. On the structure of selmer groups. Mathematische Annalen, 291(1):253–259, 1991.
  • [31] VA Kolyvagin. On the structure of shafarevich-tate groups. In Algebraic geometry, pages 94–121. Springer, 1991.
  • [32] Jeffrey Manning. Ultrapatching. unpublished notes, 2019.
  • [33] Barry Mazur and Karl Rubin. Kolyvagin systems. American Mathematical Society, 2004.
  • [34] William G McCallum. Kolyvagin’s work on shafarevich-tate groups. L-functions and arithmetic (Durham, 1989), 153:295–316, 1991.
  • [35] James S Milne. Arithmetic duality theorems. Citeseer, 2006.
  • [36] Jan Nekovár. Selmer complexes, volume 310. Société mathématique de France, 2006.
  • [37] Jan Nekovár. The euler system method for cm points on shimura curves. London Mathematical Society Lecture Note Series, 320:471, 2007.
  • [38] Jürgen Neukirch, Alexander Schmidt, and Kay Wingberg. Cohomology of number fields, volume 323. Springer Science & Business Media, 2013.
  • [39] Bernadette Perrin-Riou. Fonctions ll pp-adiques, théorie d’iwasawa et points de heegner. Bulletin de la Société Mathématique de France, 115:399–456, 1987.
  • [40] Robert Pollack and Tom Weston. On anticyclotomic μ\mu-invariants of modular forms. Compositio Mathematica, 147(5):1353–1381, 2011.
  • [41] Ravi Ramakrishna. Deforming galois representations and the conjectures of serre and fontaine-mazur. Annals of mathematics, 156(1):115–154, 2002.
  • [42] KA Ribet. On modular representations of gal (q/q) arising from. Invent. math, 100:476, 1990.
  • [43] Kenneth A Ribet. On l-adic representations attached to modular forms ii. Glasgow Mathematical Journal, 27:185–194, 1985.
  • [44] Kenneth A Ribet and Shuzo Takahashi. Parametrizations of elliptic curves by shimura curves and by classical modular curves. Proceedings of the National Academy of Sciences, 94(21):11110–11114, 1997.
  • [45] Peter Scholze. On the p-adic cohomology of the lubin-tate tower. arXiv preprint arXiv:1506.04022, 2015.
  • [46] Christopher Skinner. A converse to a theorem of gross, zagier, and kolyvagin. Annals of Mathematics, 191(2):329–354, 2020.
  • [47] Christopher Skinner and Eric Urban. The iwasawa main conjectures for gl 2. Inventiones mathematicae, 195(1):1–277, 2014.
  • [48] Christopher Skinner and Wei Zhang. Indivisibility of heegner points in the multiplicative case. arXiv preprint arXiv:1407.1099, 2014.
  • [49] Jacques Tilouine. Hecke algebras and the gorenstein property. In Modular forms and Fermat’s last theorem, pages 327–342. Springer, 1997.
  • [50] Vinayak Vatsal. Special values of anticyclotomic ll-functions. Duke Mathematical Journal, 116(2):219–261, 2003.
  • [51] Xin Wan. Iwasawa main conjecture for non-ordinary modular forms. arXiv preprint arXiv:1607.07729, 2016.
  • [52] Shou-Wu Zhang. Gross–zagier formula for {GL}_2\mathrm{\{}GL\}\_2. Asian Journal of Mathematics, 5(2):183–290, 2001.
  • [53] Wei Zhang. Selmer groups and the indivisibility of heegner points. Cambridge Journal of Mathematics, 2(2):191–253, 2014.

Declarations

Funding. This research was funded by NSF grant #DGE1745303.

Conflicts of interest. None.

Data transparency. Not applicable.

Code availability. Not applicable.