1 Introduction

Fix a prime \(p>3\) and an integer \(N>4\) such that \(p\not \mid N\phi (N)\). Let

$$\begin{aligned} f_o=\sum _{n>0}a_nq^n\in S_k(X_0(N)) \end{aligned}$$

be a \(p\)-ordinary newform of even weight \(k=2r\ge 2\) and trivial nebentypus. Thus \(f_o\) is an eigenvector for all the Hecke operators \(T_n\) with associated eigenvalues \(a_n\), and \(a_p\) is a \(p\)-adic unit for a choice of embeddings \(\iota _\infty :\overline{{\mathbf{{Q}}}}\hookrightarrow {{\mathbf{{C}}}}\) and \(\iota _p:\overline{{\mathbf{{Q}}}}\hookrightarrow {\overline{{\mathbf{{Q}}}}}_p\) that will remain fixed throughout this paper. Also let \(\mathcal{O }\) denote the ring of integers of a (sufficiently large) finite extension \(L/{\mathbf{{Q}}}_p\) containing all the \(a_n\).

For \(s>0\), let \(X_s\) be the compactified modular curve of level

$$\begin{aligned} \Gamma _s:=\Gamma _0(N)\cap \Gamma _1(p^s), \end{aligned}$$

and consider the tower

$$\begin{aligned} \cdots \rightarrow X_s\xrightarrow {\alpha }X_{s-1}\rightarrow \cdots \end{aligned}$$

with respect to the degeneracy maps described on the non-cuspidal moduli by

$$\begin{aligned} (E,\alpha _E,\pi _E)\mapsto (E,\alpha _E,p\cdot \pi _E), \end{aligned}$$

where \(\alpha _E\) denotes a cyclic \(N\)-isogeny on the elliptic curve \(E\), and \(\pi _E\) a point of \(E\) of exact order \(p^s\). The group \(({\mathbf{{Z}}}/p^s{\mathbf{{Z}}})^\times \) acts on \(X_s\) via the diamond operators

$$\begin{aligned} \langle d\rangle :(E,\alpha _E,\pi _E)\mapsto (E,\alpha _E,d\cdot \pi _E) \end{aligned}$$

compatibly with \(\alpha \) under the reduction \(({\mathbf{{Z}}}/p^s{\mathbf{{Z}}})^\times \rightarrow ({\mathbf{{Z}}}/p^{s-1}{\mathbf{{Z}}})^\times \). Set \(\Gamma :=1+p{\mathbf{{Z}}}_p\). Letting \(J_s\) be the Jacobian variety of \(X_s\), the inverse limit of the system induced by Albanese functoriality,

$$\begin{aligned} \cdots \rightarrow \mathrm{Ta}_p(J_s)\otimes _{{\mathbf{{Z}}}_p}\mathcal{O }\rightarrow \mathrm{Ta}_p(J_{s-1})\otimes _{{\mathbf{{Z}}}_p}\mathcal{O }\rightarrow \cdots , \end{aligned}$$
(1.1)

is equipped with an action of the Iwasawa algebras \(\widetilde{\Lambda }_{\mathcal{O }}:=\mathcal{O }[[{\mathbf{{Z}}}_p^\times ]]\) and

$$\begin{aligned} \Lambda _\mathcal{O }:=\mathcal{O }[[\Gamma ]]. \end{aligned}$$

Let \(\mathfrak{h }_s\) be the \(\mathcal{O }\)-algebra generated by the Hecke operators \(T_\ell \) (\(\ell \not \mid Np\)), \(U_\ell :=T_\ell \) (\(\ell \vert Np\)), and the diamond operators \(\langle d\rangle \) (\(d\in ({\mathbf{{Z}}}/p^s{\mathbf{{Z}}})^\times \)) acting on the space \(S_k(X_s)\) of cusp forms of weight \(k\) and level \(\Gamma _s\). Hida’s ordinary projector

$$\begin{aligned} e^\mathrm{ord}:=\lim _{n\rightarrow \infty }U_p^{n!} \end{aligned}$$

defines an idempotent of \(\mathfrak{h }_s\), projecting to the maximal subspace of \(\mathfrak{h }_s\) where \(U_p\) acts invertibly. We make each \(\mathfrak{h }_s\) into a \(\widetilde{\Lambda }_\mathcal{O }\)-algebra by letting the group-like element attached to \(z\in {\mathbf{{Z}}}_p^\times \) act as \(z^{k-2}\langle z\rangle \).

Taking the projective limit with respect to the restriction maps induced by the natural inclusion \(S_k(X_{s-1})\hookrightarrow S_k(X_s)\), we obtain a \(\widetilde{\Lambda }_\mathcal{O }\)-algebra

$$\begin{aligned} \mathfrak{h }^\mathrm{ord}:={\mathop {\leftarrow } \limits ^{\text{ lim}}_{s}}e^\mathrm{ord}\mathfrak{h }_s \end{aligned}$$
(1.2)

which can be seen to be independent of the weight \(k\ge 2\) used in its construction.

After a highly influential work [16] of Hida, one can associate with \(f_o\) a certain local domain \(\mathbb{I }\) quotient of \(\mathfrak{h }^\mathrm{ord}\), finite flat over \(\Lambda _\mathcal{O }\), with the following properties. For each \(n\), let \({\mathbf{{a}}}_n\in \mathbb{I }\) be the image of \(T_n\) under the projection \(\mathfrak{h }^\mathrm{ord}\rightarrow \mathbb{I }\), and consider the formal \(q\)-expansion

$$\begin{aligned} {\mathbf{{f}}}=\sum _{n>0}{\mathbf{{a}}}_nq^n\in \mathbb{I }[[q]]. \end{aligned}$$

We say that a continuous \(\mathcal{O }\)-algebra homomorphism \(\nu :\mathbb{I }\rightarrow \overline{{\mathbf{{Q}}}}_p\) is an arithmetic prime if there is an integer \(k_\nu \ge 2\), called the weight of \(\nu \), such that the composition \(\Gamma \rightarrow \mathbb{I }^\times \rightarrow \overline{{\mathbf{{Q}}}}_p^\times \) agrees with \(\gamma \mapsto \gamma ^{k_\nu -2}\) on an open subgroup of \(\Gamma \) of index \(p^{s_\nu -1}\ge 1\). Denote by \(\mathcal{X }_\mathrm{arith}(\mathbb{I })\) the set of arithmetic primes of \(\mathbb{I }\), which will often be seen as sitting inside \(\mathrm{Spf}(\mathbb{I })(\overline{{\mathbf{{Q}}}}_p)\). If \(\nu \in \mathcal{X }_\mathrm{arith}(\mathbb{I })\), \(F_\nu \) will denote its residue field. Then:

  • for every \(\nu \in \mathcal{X }_\mathrm{arith}(\mathbb{I })\), there exists an ordinary \(p\)-stabilized newformFootnote 1

    $$\begin{aligned} {\mathbf{{f}}}_\nu \in S_{k_\nu }(X_{s_\nu }) \end{aligned}$$

    such that \(\nu ({\mathbf{{f}}})\in F_\nu [[q]]\) gives the \(q\)-expansion of \({\mathbf{{f}}}_\nu \);

  • if \(s_\nu =1\) and \(k_\nu \equiv k\pmod {2(p-1)}\), there exists a normalized newform \({\mathbf{{f}}}_\nu ^\sharp \in S_{k_\nu }(X_0(N))\) such that

    $$\begin{aligned} {\mathbf{{f}}}_\nu (q)={\mathbf{{f}}}_\nu ^\sharp (q)-\frac{p^{k_\nu -1}}{\nu ({\mathbf{{a}}}_p)} {\mathbf{{f}}}_\nu ^\sharp (q^p); \end{aligned}$$
    (1.3)
  • there exists a unique \(\nu _o\in \mathcal{X }_\mathrm{arith}(\mathbb{I })\) such that \(f_o={\mathbf{{f}}}_{\nu _o}^\sharp \).

In particular, after “\(p\)-stabilization” (1.3), the form \(f_o\) fits in the \(p\)-adic family \({\mathbf{{f}}}\).

Similarly for the associated Galois representation \(V_{f_o}\): the continuous \(\mathfrak{h }^\mathrm{ord}\)-linear action of the absolute Galois group \(G_{\mathbf{{Q}}}\) on the module

$$\begin{aligned} \mathbb{T }:=\mathbb{T }^\mathrm{ord}\otimes _{\mathfrak{h }^\mathrm{ord}}\mathbb{I },\quad \quad \mathrm{where} \ \mathbb{T }^\mathrm{ord}:={\mathop {\leftarrow } \limits ^{\text{ lim}}_{s}{{e}^\mathrm{ord}}}(\mathrm{Ta}_p(J_s)\otimes _{{\mathbf{{Z}}}_p}\mathcal{O }), \end{aligned}$$
(1.4)

gives rise to a “big” Galois representation \(\rho _{{\mathbf{{f}}}}:G_{{\mathbf{{Q}}}}\rightarrow \mathrm{Aut}(\mathbb{T })\) such that

$$\begin{aligned} \nu (\rho _{{\mathbf{{f}}}})\cong \rho _{{\mathbf{{f}}}_\nu }^* \quad \quad \mathrm{for \ every}\, \nu \in \mathcal{X }_\mathrm{arith}(\mathbb{I }), \end{aligned}$$

where \(\rho _{{\mathbf{{f}}}_\nu }^*\) is the contragredient of the (cohomological) \(p\)-adic Galois representation \(\rho _{{\mathbf{{f}}}_\nu }:G_{\mathbf{{Q}}}\rightarrow \mathrm{Aut}(V_{{\mathbf{{f}}}_\nu })\) attached to \({\mathbf{{f}}}_\nu \) by Deligne; in particular, one recovers \(\rho _{f_o}^*\) from \(\rho _{{\mathbf{{f}}}}\) by specialization at \(\nu _o\).

Assume from now on that the residual representation \(\bar{\rho }_{f_o}\) is irreducible; then \(\mathbb T \) can be shown to be free of rank 2 over \(\mathbb{I }\). (See [23, Théorème 7].) Let \(K\) be an imaginary quadratic field with ring of integers \(\mathcal{O }_K\) containing an ideal \(\mathfrak{N }\subset \mathcal{O }_K\) with

$$\begin{aligned} \mathcal{O }_K/\mathfrak{N }\cong {\mathbf{{Z}}}/N{\mathbf{{Z}}}, \end{aligned}$$
(1.5)

and denote by \(H\) the Hilbert class field of \(K\). Under this Heegner hypothesis relative to \(N\) (but with no extra assumptions on the prime \(p\)), the work [19] of Howard produces a compatible sequence \(U_p^{-s}\cdot \mathfrak{X }_{s}\) of cohomology classes with values in a certain twist of the ordinary part of (1.1), giving rise to a canonical “big” cohomology class \(\mathfrak{X }\), the big Heegner point (of conductor 1), in the cohomology of a self-dual twist \(\mathbb{T }^\dagger \) of \(\mathbb{T }\). Moreover, if every prime factor of \(N\) splits in \(K\), it follows from his results that the class

$$\begin{aligned} \mathfrak{Z }:=\mathrm{Cor}_{H/K}(\mathfrak{X }) \end{aligned}$$

lies in Nekovář’s extended Selmer group \(\widetilde{H}_f^1(K,\mathbb{T }^\dagger )\). In particular, for every \(\nu \in \mathcal{X }_\mathrm{arith}(\mathbb{I })\) with \(s_\nu =1\) and \(k_\nu \equiv k\pmod {2(p-1)}\) as above, the specialization \(\mathfrak{Z }_\nu \) belongs to the Bloch-Kato Selmer group \(H_{f}^1(K,V_{{\mathbf{{f}}}_\nu ^\sharp }(k_\nu /2))\) of the self-dual representation \(\mathbb{T }^\dagger \otimes _{\mathbb{I }}F_\nu \cong V_{{\mathbf{{f}}}_\nu ^\sharp }(k_\nu /2)\). The classes \(\mathfrak{Z }_\nu \) may thus be regarded as a natural higher weight analogue of the Kummer images of Heegner points on modular Abelian varieties (associated with weight \(2\) eigenforms).

But for any of the above \({\mathbf{{f}}}_\nu ^\sharp \), one has an alternate (and completely different!) method of producing such a higher weight analogue. Briefly, if \(k_\nu =2r_\nu > 2\), associated to any elliptic curve \(A\) with CM by \(\mathcal{O }_K\), there is a null-homologous cycle \(\Delta ^\mathrm{heeg}_{A,r_\nu }\), a so-called Heegner cycle, on the \((2r_\nu -1)\)-dimensional Kuga–Sato variety \(W_{r_\nu }\), giving rise to an \(H\)-rational class in the Chow group \(\mathrm{CH}^{r_\nu +1}(W_{r_\nu })_0\) with \({\mathbf{{Q}}}\)-coefficients. Since the representation \(V_{{\mathbf{{f}}}_\nu ^\sharp }(r_\nu )\) appears in the étale cohomology of \(W_{r_\nu }\):

$$\begin{aligned} H_{{{\acute{\mathrm{e}}}}\mathrm{t}}^{2r_\nu -1}(\overline{W}_{r_\nu }, {\mathbf{{Q}}}_{p})(r_\nu )\xrightarrow {\pi _{{\mathbf{{f}}}_\nu ^\sharp }} V_{{\mathbf{{f}}}_\nu ^\sharp }(r_\nu ), \end{aligned}$$

by taking the images of the cycles \(\Delta ^\mathrm{heeg}_{A,r_\nu }\) under the \(p\)-adic étale Abel-Jacobi map

$$\begin{aligned} \Phi _{H}^{{{{\acute{\mathrm{e}}}}\mathrm{t}}}:\mathrm{CH}^{r_\nu +1}(W_{r_\nu })_0(H) \rightarrow H^1(H,H_{{{{\acute{\mathrm{e}}}}\mathrm{t}}}^{2r_\nu -1}(\overline{W}_{r_\nu },{\mathbf{{Q}}}_p)(r_\nu )) \end{aligned}$$

and composing with the map induced by \(\pi _{{\mathbf{{f}}}_\nu ^\sharp }\) on \(H^1\)’s, we may consider the classes

$$\begin{aligned} \Phi _\mathbf{f _\nu ^\sharp ,K}^{{\acute{\mathrm{e}}}\mathrm{t}}(\Delta ^\mathrm{heeg}_{r_\nu }) :=\mathrm{Cor}_{H/K}(\pi _{{\mathbf{{f}}}_\nu ^\sharp }\Phi _{H}^{{{{\acute{\mathrm{e}}}}\mathrm{t}}}(\Delta ^\mathrm{heeg}_{A,r_\nu })). \end{aligned}$$

By the work [28] of Nekovář, these classes are known to lie in the same Selmer group as \(\mathfrak{Z }_\nu \), and the question of their comparison thus naturally arises.

Main Theorem

(Thm. 5.11) Assume that \(p\) splits in \(K\) and that \(\mathfrak{Z }\) is not \(\mathbb{I }\)-torsion. Then for any \(\nu \in \mathcal{X }_\mathrm{arith}(\mathbb{I })\) of weight \(k_\nu =2r_\nu > 2\) with \(k_\nu \equiv k\pmod {2(p-1)}\) and trivial character, we have

$$\begin{aligned} \langle \mathfrak{Z }_{\nu },\mathfrak{Z }_\nu \rangle _K =\left(1-\frac{p^{r_\nu -1}}{\nu ({\mathbf{{a}}}_p)}\right)^4 \frac{\langle \Phi _{{\mathbf{{f}}}_\nu ^\sharp ,K}^{{{{\acute{\mathrm{e}}}}\mathrm{t}}}(\Delta ^\mathrm{heeg}_{r_\nu }), \Phi _{{\mathbf{{f}}}_\nu ^\sharp ,K}^{{{{\acute{\mathrm{e}}}}\mathrm{t}}}(\Delta ^\mathrm{heeg}_{r_\nu })\rangle _K}{{u^2(4D)^{r_\nu -1}}}, \end{aligned}$$

where \(\langle ,\rangle _K\) is the cyclotomic \(p\)-adic height pairing on \(H_f^1(K,V_{{\mathbf{{f}}}_\nu ^\sharp }(r_\nu ))\), \(u:=\vert \mathcal{O }_K^\times \vert /2\), and \(-D<0\) is the discriminant of \(K\).

Thus assuming the non-degeneracy of the \(p\)-adic height pairing, it follows that the étale Abel-Jacobi images of Heegner cycles are \(p\)-adically interpolated by \(\mathfrak{Z }\). We also note that \(\mathfrak{Z }\) is conjecturally always not \(\mathbb{I }\)-torsion ([19, Conj. 3.4.1]), and that by [18, Cor. 5] this conjecture can be verified in any given case by exhibiting the non-vanishing of an appropriate \(L\)-value (a derivative, in fact).

This paper is organized as follows. Section 2 is aimed at proving an expression for the formal group logarithms of ordinary CM points on \(X_s\) using Coleman’s theory of \(p\)-adic integration. Our methods here are drawn from [1], Sect. 4], which we extend in weight 2 to the case of level divisible by an arbitrary power of \(p\), but with ramification restricted to a potentially crystalline setting. Not quite surprisingly, this restriction turns out to make our computations essentially the same as theirs, and will suffice for our purposes.

In Sect. 3 we recall the generalised Heegner cycles and the formula for their \(p\)-adic Abel-Jacobi images from loc.cit., and discuss the relation between these and the more classical Heegner cycles.

In Sect. 4 we deduce from the work [30] of Ochiai a “big” logarithm map that will allow as to move between different weights in the Hida family.

Finally, in Sect. 5 we prove our main results. The key observation is that, when \(p\) splits in \(K\), the combination of CM points on \(X_s\) taken in Howard’s construction appears naturally in the evaluation of the critical twist of a \(p\)-adic modular form at a canonical trivialized elliptic curve. The expression from Sect. 2 thus yields, for infinitely many \(\nu \) of weight 2, a formula for the \(p\)-adic logarithm of the localization of \(\mathfrak{Z }_\nu \) in terms of certain values of a \(p\)-adic modular form of weight 0 associated with \({\mathbf{{f}}}_\nu \) (Thm. 5.8). When extended by \(p\)-adic continuity to an arithmetic prime \(\nu \) of higher even weight, this expression is seen to agree with the formula from Sect. 3, and by the interpolation properties of the big logarithm map it corresponds to the \(p\)-adic logarithm of the localization of \(\mathfrak{Z }_\nu \). Our main results follow easily from this.

Finally, we note that an extension of the results in this paper, and in particular of the Main Theorem above, has a number of arithmetic applications arising from the connection with the theory of \(p\)-adic \(L\)-functions. (See [5].)

2 Preliminaries

2.1 \(p\)-Adic modular forms

To avoid some issues related to the representability of certain moduli problems, in this section we change notations from the Introduction, letting \(X_s\) be the compactified modular curve of level \(\Gamma _s:=\Gamma _1(Np^s)\), viewed as a scheme over \(\mathrm{Spec}(\mathbf Q _p)\). Let \(\pi :\mathcal{E }_s\rightarrow X_s\) be the universal generalized elliptic curve over \(X_s\), and let

$$\begin{aligned} \underline{\omega }_{X_s}:=\pi _{*}\Omega ^1_{\mathcal{E }_s/X_s}(\log Z_s) \end{aligned}$$

be the pushforward of the invertible sheaf of relative differentials on \(\mathcal{E }_s/X_s\) with possible log-poles along the inverse image of the cuspidal subscheme \(Z_s\subset X_s\).

Algebraically, \(H^0(X_s,\underline{\omega }_{X_s}^{\otimes {2}})\) gives the space of modular forms of weight 2 and level \(\Gamma _s\) (defined over \({\mathbf{{Q}}}_p\)). Consider the complex

$$\begin{aligned} \Omega _{X_s/{\mathbf{{Q}}}_p}^\bullet (\log {Z}_s):0\longrightarrow \mathcal{O }_{X_s}\xrightarrow {\;\;d\;\;} \Omega _{X_s/{\mathbf{{Q}}}_p}^1(\log {Z}_s)\longrightarrow 0 \end{aligned}$$
(2.1)

of sheaves on \(X_s\). The algebraic de Rham cohomology of \(X_s\)

$$\begin{aligned} H_\mathrm{dR}^1(X_s/{\mathbf{{Q}}}_p):=\mathbb{H }^1(X_s, \Omega _{X_s/{\mathbf{{Q}}}_p}^\bullet (\log {Z_s})) \end{aligned}$$

is a finite-dimensional \({\mathbf{{Q}}}_p\)-vector space equipped with a Hodge filtration

$$\begin{aligned} 0\;\subset \; H^0(X_s,\Omega _{X_s/{\mathbf{{Q}}}_p}^1(\log {Z_s}))\;\subset \; H_\mathrm{dR}^1(X_s/{\mathbf{{Q}}}_p), \end{aligned}$$

and by the Kodaira–Spencer isomorphism \(\underline{\omega }_{X_s}^{\otimes 2}\cong \Omega ^1_{X_s/{\mathbf{{Q}}}_p}(\log {Z_s})\), every cusp form \(f\in S_2(X_s)\) (in particular) defines a cohomology class \(\omega _f\in H_\mathrm{dR}^1(X_s/{\mathbf{{Q}}}_p)\).

Let \(X\) be the complete modular curve of level \(\Gamma _1(N)\), also viewed over \(\mathrm{Spec}({\mathbf{{Q}}}_p)\), and consider the subspaces of the associated rigid analytic space \(X^\mathrm{an}\):

$$\begin{aligned} X^\mathrm{ord}\;\subset \;X_{<1/(p+1)}\;\subset \; X_{<p/(p+1)}\;\subset \;X^\mathrm{an}. \end{aligned}$$

To define these, let \(\mathcal{X }_{/{\mathbf{{Z}}}_p}\) be the canonical integral model of \(X\) over \(\mathrm{Spec}({\mathbf{{Z}}}_p)\), and let \(X_{{\mathbf{F}}_p}:=\mathcal{X }\times _{{\mathbf{{Z}}}_p}\mathbf{F}_p\) denote its special fiber. The supersingular points \(SS\subset X_{{\mathbf{F}}_p}({\overline{\mathbf{F}}}_{p})\) is the finite set of points corresponding to the moduli of supersingular elliptic curves (with \(\Gamma _1(N)\)-level structure) in characteristic \(p\). Let \(E_{p-1}\) be the Eisenstein series of weight \(p-1\) and level 1, seen as a global section of the sheaf \(\underline{\omega }_{X}^{\otimes (p-1)}\). (Recall that we are assuming \(p\ge 5\).) The reduction of \(E_{p-1}\) to \(X_{\mathbf{F}_p}\) is the Hasse invariant, which defines a section of the reduction of \(\underline{\omega }_{X}^{\otimes (p-1)}\) with \(SS\) as its locus of (simple) zeroes. If \(x\in X(\overline{\mathbf{Q }}_p)\), let \(\bar{x}\in X_\mathbf{F _p}(\overline{\mathbf{F }}_p)\) denote its reduction. Each point \(\bar{x}\in SS\) is smooth in \(X_{\mathbf{F}_p}\), and the ordinary locus of \(X\).

$$\begin{aligned} X^\mathrm{ord}:=X^\mathrm{an}\backslash \bigcup _{\bar{x}\in SS}D_{\bar{x}} \end{aligned}$$

is defined to be the complement of their residue discs \(D_{\bar{x}}\subset X^\mathrm{an}\). The function \(\vert E_{p-1}(x)\vert _p\) defines a local parameter on \(D_{\bar{x}}\), and with the normalization \(\vert {p}\vert _p=p^{-1}\), \(X_{<1/(p+1)}\) (resp. \(X_{<p/(p+1)}\)) is defined to be complement in \(X^\mathrm{an}\) of the subdiscs of \(D_{\bar{x}}\) where \(\vert E_{p-1}(x)\vert _p\le p^{-1/(p+1)}\) (resp. \(\vert E_{p-1}(x)\vert _p\le p^{-p/(p+1)}\)), for all \(\bar{x}\in SS\).

Using the canonical subgroup \(H_E\) (of order \(p\)) attached to every elliptic curve \(E\) corresponding to a closed point in \(X_{<p/(p+1)}\), the Deligne–Tate map

$$\begin{aligned} \phi _0:X_{<1/(p+1)}\rightarrow X_{<p/(p+1)} \end{aligned}$$

is defined by sending \(E\mapsto E/H_E\) (with the induced action on the level structure) under the moduli interpretation. This map is a finite morphism which by definition lifts to characteristic zero the absolute Frobenius on \(X_{{\mathbf{F}}_p}\). (See [21, Thm. 3.1].)

For every \(s>0\), the Deligne–Tate map \(\phi _0\) can be iterated \(s-1\) times on the open rigid subspace \(X_{<p^{2-s}/(p+1)}\) of \(X^\mathrm{an}\) where \(\vert E_{p-1}(x)\vert _p>p^{-p^{2-s}/(p+1)}\). Letting \(\alpha _s:X_s\rightarrow X\) be the map forgetting the “\(\Gamma _1(p^s)\)-part” of the level structure, define

$$\begin{aligned} \mathcal{W }_1(p^s)\subset X_s^\mathrm{an} \end{aligned}$$

to be the open rigid subspace of \(X_s\) whose closed points correspond to triples \((E,\alpha _E,\pi _E)\) whose image under \(\alpha _s\) lands inside \(X_{<p^{2-s}(p+1)}\) and are such that \(\pi _E\) generates the canonical subgroup of \(E\) of order  \(p^s\) (as in [4, Def. 3.4]).

Define \(\mathcal W _2(p^s)\subset X_s^\mathrm{an}\) is the same manner, replacing \(p^{2-s}/(p+1)\) by \(p^{1-s}/(p+1)\) in the definition of \(\mathcal W _1(p^s)\). Then we obtain a lifting of Frobenius \(\phi =\phi _s\) on \(X_s\) making the diagram

figure a1

commutative by sending a point \(x\!=\!(E,\alpha _E,\imath _E)\!\in \!\mathcal{W }_2(p^s)\), where \(\imath _E\!:\!\varvec{\mu }_{p^s}\!\hookrightarrow \! E[p^s]\) is an embedding giving the \(\Gamma _1(p^s)\)-level structure on \(E\), to \(x^{\prime }=(\phi _0E,\phi _0\alpha _E,\imath _E^{\prime })\), where \(\imath _E^{\prime }\) is determined by requiring that \(\alpha _s(x^{\prime })\) lands in \(X_{<p^{2-s}/(p+1)}\) and for each \(\zeta \in \varvec{\mu }_{p^s}-\{1\}\), \(\imath _E^{\prime }(\zeta )=\phi _0Q\) if \(\imath _E(\zeta )=pQ\). (Cf. [11, Sect. B.2].)

Let \(k\in {\mathbf{{Z}}}\), and denote by \(\underline{\omega }_{X_s^\mathrm{an}}\) the rigid analytic sheaf on \(X_s^\mathrm{an}\) deduced from \(\underline{\omega }_{X_s}\). Let \(I_s:=\{v\in \mathbf Q :0\le v< p^{2-s}/(p+1)\}\), and for \(p^{-v}\in I_s\) define the affinoid subdomain \(X_s(v)\) of \(X_s^\mathrm{an}\) inside \(\mathcal W _1(p^s)\) whose closed points \(x\) satisfy \(\vert E_{p- 1}(x)\vert _p\ge p^{-v}\). Then \(X_s(0)\) is the connected component of the ordinary locus of \(X_s\) containing the cusp \(\infty \). The space of \(p\) -adic modular forms of weight \(k\) and level \(\Gamma _s\) (defined over \({\mathbf{{Q}}}_p\)) is the \(p\)-adic Banach space

$$\begin{aligned} M_k^\mathrm{ord}(X_s):=H^0(X_s(0),\underline{\omega }_{X_s^\mathrm{an}}^{\otimes k}), \end{aligned}$$

and the space of overconvergent  \(p\) -adic modular forms of weight \(k\) and level \(\Gamma _s\) is the \(p\)-adic Fréchet space

$$\begin{aligned} M_k^\mathrm{rig}(X_s):={\mathop {\leftarrow } \limits ^{\text{ lim}}_{v}} H^0(X_s(v),\underline{\omega }_{X_s^\mathrm{an}}^{\otimes k}), \end{aligned}$$

where the limit is with respect to the natural restriction maps as \(v\in I_s\) increasingly approaches \(p^{2-s}/(p+1)\). By restriction, a classical modular form in \(H^0(X_s,\underline{\omega }_{X_s}^{\otimes k})\) defines an (obviously) overconvergent \(p\)-adic modular form of the same weight an level. Moreover, the action of the diamond operators on \(X_s\) gives rise to an action of \(({\mathbf{{Z}}}/p^s{\mathbf{{Z}}})^\times \) on the spaces of \(p\)-adic modular forms which agrees with the action on \(H^0(X_s,\underline{\omega }_{X_s}^{\otimes k})\) under restriction.

We say that a ring \(R\) is a \(p\) -adic ring if the natural map \(R\rightarrow {\mathop {\leftarrow } \limits ^{\text{ lim}}} R/p^nR\) is an isomorphism. For varying \(s>0\), the data of a compatible sequence of embeddings \(\varvec{\mu }_{p^s}\hookrightarrow E\) as \(R\)-group schemes, amounts to the data of an embedding \(\varvec{\mu }_{p^\infty }\hookrightarrow E[p^\infty ]\) of \(p\)-divisible groups, and also to the given of a trivialization of \(E\) over \(R\), i.e. an isomorphism

$$\begin{aligned} \imath _E:\hat{E}\rightarrow \hat{\mathbf{G}}_m \end{aligned}$$

of the associated formal groups. The space \(\mathbf{M}(N)\) of Katz  \(p\)-adic modular functions of tame level \(N\) (over \({\mathbf{{Z}}}_p\)) is the space of functions \(f\) on trivialized elliptic curves with \(\Gamma _1(N)\)-level structure over arbitrary \(p\)-adic rings, assigning to the isomorphism class of a triple \((E,\alpha _E,\imath _E)\) over \(R\) a value \(f(E,\alpha _E,\imath _E)\in R\) whose formation is compatible under base change. If \(R\) is a fixed \(p\)-adic ring, by only considering \(p\)-adic rings which are \(R\)-algebras, we obtain the notion of Katz \(p\)-adic modular functions defined over \(R\), forming the space \(\mathbf M (N)\widehat{\otimes }_\mathbf{Z _p}R\) which will also be denoted by \(\mathbf M (N)\) with an abuse of notation.

The action of \(z\in {\mathbf{{Z}}}_p^\times \) on a trivialization gives rise to an action of \({\mathbf{{Z}}}_p^\times \) on \(\mathbf{M}(N)\):

$$\begin{aligned} \langle z\rangle f(E,\alpha _E,\imath _E):=f(E,\alpha _E,z\cdot \imath _E), \end{aligned}$$

and given a character \(\chi \in \mathrm{Hom}_\mathrm{cont}({\mathbf{{Z}}}_p^\times ,R^\times )\), we say that \(f\in \mathbf{M}(N)\) has weight-nebentypus \(\chi \) if \(\langle z\rangle f=\chi (z)f\) for all \(z\in {\mathbf{{Z}}}_p^\times \). If \(k\) is an integer, denoting by \(z^k\) the \(k\)-th power character on \({\mathbf{{Z}}}_p^\times \), the subspace \(M_k^\mathrm{ord}(Np^s,\varepsilon )\) of \(M_k^\mathrm{ord}(X_s)\) consisting of \(p\)-adic modular forms with nebentypus \(\varepsilon :({\mathbf{{Z}}}/p^s{\mathbf{{Z}}})^\times \rightarrow R^\times \) can be recovered as

$$\begin{aligned} M_k^\mathrm{ord}(Np^s,\varepsilon )\cong \{f\in \mathbf{M}(N):\langle z\rangle f=z^k\varepsilon (z)f,\quad \mathrm{for \ all}\,\, z\in {\mathbf{{Z}}}_p^\times \}. \end{aligned}$$
(2.2)

Since it will play an important role later, we next recall from [14, Sect. III.6.2] the definition in terms of moduli of the twist of \(p\)-adic modular forms by characters of not necessarily finite order. Let \(R\) be a \(p\)-adic ring, and let \((E,\alpha _E,\imath _E)\) be a trivialized elliptic curve with \(\Gamma _1(N)\)-level structure over \(R\). For each \(s>0\), consider the quotient \(E_0:=E/\imath _E^{-1}(\varvec{\mu }_{p^s})\), and let \(\varphi _0:E\rightarrow E_0\) denote the projection. Since \(p\not \mid N\), \(\varphi _0\) induces a \(\Gamma _1(N)\)-level structure \(\alpha _{E_0}\) on \(E_0\), and since \(\mathrm{ker}(\varphi _0)\cong \varvec{\mu }_{p^s}\), the dual \(\check{\varphi }_0:E_0\rightarrow E\) is étale, inducing an isomorphism of the associated formal groups. Thus (with a slight abuse of notation) \(\imath _{E_0}:=\imath _E\circ \check{\varphi }_0: \hat{E}_0\xrightarrow {\sim }\hat{\mathbf{G}}_m\) is a trivialization of \(E_0\), and since we have an embedding \(\jmath :{\mathbf{{Z}}}/p^s{\mathbf{{Z}}}\cong \mathrm{ker}(\check{\varphi }_0)\hookrightarrow E_0[p^s]\), we deduce an isomorphism

$$\begin{aligned} E_0[p^s]\cong \varvec{\mu }_{p^s}\oplus {\mathbf{{Z}}}/p^s{\mathbf{{Z}}}\end{aligned}$$

which we use to bijectively attach a \(p^s\)-th root of unity \(\zeta _{C}\) to every étale subgroup \(C\subset E_0[p^s]\) of order \(p^s\), in such a way that 1 is attached to \(\mathrm{ker}(\check{\varphi }_0)\).

Now for \(f\in \mathbf{M}(N)\) and \(a\in {\mathbf{{Z}}}_p\), define \(f\otimes \mathbb 1 _{a+p^s{\mathbf{{Z}}}_p}\) to be the rule on trivialized elliptic curves given by

$$\begin{aligned} f\otimes \mathbb 1 _{a+p^s{\mathbf{{Z}}}_p}(E,\alpha _E,\imath _E) =\frac{1}{p^s}\sum _{C}\zeta _C^{-a}\cdot f(E_0/C,\alpha _C,\imath _C) \end{aligned}$$
(2.3)

where the sum is over the étale subgroups \(C\subset E_0[p^s]\) of order \(p^s\), and where \(\alpha _C\) (resp. \(\imath _C\)) denotes the \(\Gamma _1(N)\)-level structure (resp. trivialization) on the quotient \(E_0/C\) naturally induced by \(\alpha _{E_0}\) (resp. \(\imath _{E_0}\)).

Lemma 2.1

The assignment \(a+p^s{\mathbf{{Z}}}_p\rightsquigarrow \left(f\mapsto f\otimes \mathbb 1 _{a+p^s{\mathbf{{Z}}}_p}\right)\) gives rise to an \(\mathrm{End}_R\mathbf{M}(N)\)-valued measure \(\mu _\mathrm{Gou}\) on \({\mathbf{{Z}}}_p\).

Proof

Let \(\sum _na_nq^n\) be the \(q\)-expansion of \(f\), i.e. the value that it takes at the triple \((\mathrm{Tate}(q),\alpha _\mathrm{can},\imath _\mathrm{can})=(\mathbf{G}_m/q^{\mathbf{{Z}}},\zeta _N,\varvec{\mu }_{p^\infty } \hookrightarrow \mathbf{G}_m/q^{\mathbf{{Z}}})\) over the \(p\)-adic completion of \({R((q))}\). By the \(q\)-expansion principle, the claim follows immediately from the equality

$$\begin{aligned} f\otimes \mathbb 1 _{a+p^s{\mathbf{{Z}}}_p}(q)=\sum _{n\equiv a\;\mathrm{mod}\;{p^s}}a_nq^n, \end{aligned}$$

which is shown by adapting the arguments in [14, p. 102]. \(\square \)

Definition 2.2

(Gouvêa) Let \(f\in \mathbf{M}(N)\) and \(\chi :{\mathbf{{Z}}}_p\rightarrow R\) be any continuous multiplicative function. The twist of \(f\) by \(\chi \) is

$$\begin{aligned} f\otimes \chi :=\left(\int _{{\mathbf{{Z}}}_p}\chi (x)d\mu _\mathrm{Gou}(x)\right) (f)\in \mathbf{M}(N). \end{aligned}$$

This operation is compatible with the usual character twist of Hecke eigenforms:

Lemma 2.3

Let \(\chi :{\mathbf{{Z}}}_p^\times \rightarrow R^\times \) be a continuous character extended by zero on \(p{\mathbf{{Z}}}_p\). If \(f\in \mathbf{M}(N)\) has \(q\)-expansion \(\sum _na_nq^n\), then \(f\otimes \chi \) has \(q\)-expansion \(\sum _n\chi (n)a_nq^n\), and if \(f\) has weight-nebentypus \(\kappa \in \mathrm{Hom}_\mathrm{cts}({\mathbf{{Z}}}_p^\times ,R^\times )\), then \(f\otimes \chi \) has weight-nebentypus \(\chi ^2\kappa \).

Proof

See [14], Cor. III.6.8.\(i\)] and [14, Cor. III.6.9]). \(\square \)

In particular, twisting by the identity function of \({\mathbf{{Z}}}_p\) we obtain an operator \(d:\mathbf{M}(N)\rightarrow \mathbf{M}(N)\) whose effect on \(q\)-expansions is \(q\frac{d}{dq}\). For every \(k\in {\mathbf{{Z}}}\), we see from (2.2) and Lemma 2.3, that this restricts to a map

$$\begin{aligned} d:M_k^\mathrm{ord}(X_s)\rightarrow M_{k+2}^\mathrm{ord}(X_s) \end{aligned}$$

which increases the weight by 2 and preserves the nebentypus. Moreover, for \(k=0\), the arguments in [9, Prop. 4.3] can be adapted to show that \(d\) restricts to a linear map \(M_0^\mathrm{rig}(X_s)\rightarrow M_{2}^\mathrm{rig}(X_s)\), viewing \(M_k^\mathrm{rig}(X_s)\hookrightarrow M_k^\mathrm{ord}(X_s)\) by restriction.

2.2 Comparison isomorphisms

Let \(\zeta _s\) be a primitive \(p^s\)-th root of unity, and let \({F}\) be a finite extension of \({\mathbf{{Q}}}_p(\zeta _s)\) over which \(X_s\) acquires stable reduction, i.e. such that the base extension \(X_s\times _{{\mathbf{{Q}}}_p}F\) admits a stable model over the ring of integers \(\mathcal{O }_{F}\) of \({F}\). For the ease of notation, from now on we will denote \(X_s\times _{{\mathbf{{Q}}}_p}{F}\) (as well as the associated rigid analytic space) simply by \(X_s\).

Let \({\fancyscript{X}}_s\) be the minimal regular model of \(X_{s}\) over \(\mathcal{O }_{F}\), and denote by \(F_0\) the maximal unramified subfield of \({F}\). The work [17] of Hyodo–Kato endows the \({F}\)-vector space \(H_\mathrm{dR}^1(X_s/{F})\) with a canonical \(F_0\)-structure

$$\begin{aligned} H_\mathrm{log-cris}^1({\fancyscript{X}}_s) \hookrightarrow H_\mathrm{dR}^1(X_s/{F}) \end{aligned}$$
(2.4)

equipped with a semi-linear Frobenius operator \(\varphi \).

After the proof [33] of the Semistable conjecture of Fontaine–Jannsen, these structures are known to agree with those attached by Fontaine’s theory to the \(p\)-adic \(G_{F}\)-representation

$$\begin{aligned} V_s:=H_{{{{\acute{\mathrm{e}}}}\mathrm{t}}}^1(\overline{X}_s,{\mathbf{{Q}}}_p). \end{aligned}$$
(2.5)

More precisely, since \(X_s\) has semistable reduction, \(V_s\) is semistable in the sense of Fontaine, and there is a canonical isomorphism \(D_\mathrm{st}(V_s)\xrightarrow {} H_\mathrm{log-cris}^1({\fancyscript{X}}_s)\), inducing an isomorphism

$$\begin{aligned} D_\mathrm{dR}(V_s)\xrightarrow {\sim } H_\mathrm{dR}^1(X_s/{F}) \end{aligned}$$
(2.6)

as filtered \(\varphi \)-modules after extension of scalars to \(F\).

Consider the étale Abel-Jacobi map \(\mathrm{CH}^1(X_s)_0({F})\rightarrow H^1({F},V_s(1))\) constructed in [28], which in this case agrees with the usual Kummer map

$$\begin{aligned} {\delta }_{{F}}:J_s({F})\rightarrow H^1({F},{\mathbf{{Q}}}_p\otimes \mathrm{Ta}_p(J_s)), \end{aligned}$$

where \(J_s=\mathrm{Pic}^0(X_s)\) is the connected Picard variety of \(X_s\). (See [loc.cit., Ex. (2.3)]).

Let \(g\in S_2(X_s)\) be a newform with primitive nebentypus of \(p\)-power conductor, let \(V_g\) the \(p\)-adic Galois representation attached to \(g\), which is equipped with a Galois-equivariant projection \(V_s\rightarrow V_g\), and let \(V_g^*\) be the representation contragredient to \(V_g\), so that \(V_g(1)\) and \(V_g^*\) are in Kummer duality. Also let \(L_g\) be a finite extension of \({\mathbf{{Q}}}_p\) over which the Hecke eigenvalues of \(g\) are defined. By [3, Ex. 3.11], the image of the induced composite map:

$$\begin{aligned} \delta _{g,F}:J_s({F})\xrightarrow {\delta _{{F}}}H^1({F},V_s(1))\rightarrow H^1({F},V_g(1)) \end{aligned}$$
(2.7)

lies in the Bloch-Kato “finite” subspace \(H_f^1({F},V_g(1))\), and by our assumption on the nebentypus of \(g\), the Bloch-Kato exponential map gives an isomorphism

$$\begin{aligned} \mathrm{exp}_{{F},V_g(1)}:\frac{D_\mathrm{dR}(V_g(1))}{\mathrm{Fil}^0D_\mathrm{dR}(V_g(1))}\rightarrow H^1_f({F},V_g(1)) \end{aligned}$$
(2.8)

whose inverse will be denoted by \(\mathrm{log}_{{F},V_g(1)}\).

Our aim in this section is to compute the images of certain degree 0 divisors on \(X_s\) under the \(p\)-adic Abel-Jacobi map \(\delta _{g,F}^{(p)}\), defined as the composition

$$\begin{aligned} \delta _{g,{F}}^{(p)}:J_s({F})\xrightarrow {{\delta }_{g,{F}}} H_f^1({F},V_g(1))\xrightarrow {\mathrm{log}_{{F},V_g(1)}}\frac{D_\mathrm{dR}(V_g(1))}{\mathrm{Fil}^0D_\mathrm{dR}(V_g(1))}\xrightarrow {\sim }(\mathrm{Fil}^0D_\mathrm{dR}(V_g^*))^\vee ,\nonumber \\ \end{aligned}$$
(2.9)

where the last identification arises from the de Rham pairing

$$\begin{aligned} \langle \;,\;\rangle :D_\mathrm{dR}(V_g(1))\times D_\mathrm{dR}(V_g^*)\rightarrow D_\mathrm{dR}({\mathbf{{Q}}}_p(1))\otimes _{{\mathbf{{Q}}}_p}L_g\cong L_g \end{aligned}$$
(2.10)

with respect to which \(\mathrm{Fil}^0D_\mathrm{dR}(V_g(1))\) and \(\mathrm{Fil}^0D_\mathrm{dR}(V_g^*)\) are exact annihilators of each other. A basic ingredient for this computation will be the following alternate description of the logarithm map \(\mathrm{log}_{{F},V_g(1)}\).

Recall the interpretation of \(H^1({F},V_g(1))\) as the space \(\mathrm{Ext}_{\underline{\mathrm{Rep}}(G_{{F}})}^1({\mathbf{{Q}}}_p,V_g(1))\) of extensions of \(V_g(1)\) by \({\mathbf{{Q}}}_p\) in the category of \(p\)-adic \(G_{F}\)-representations. Since \({F}\) contains \({\mathbf{{Q}}}_p(\zeta _s)\), \(V_g\) is a crystalline \(G_{{F}}\)-representation in the sense of Fontaine, and under that interpretation the Bloch-Kato “finite” subspace corresponds to those extensions which are crystalline (see [26, Prop. 1.26], for example):

$$\begin{aligned} H_f^1({F},V_g(1))\cong \mathrm{Ext}_{\underline{\mathrm{Rep}}_\mathrm{cris}(G_{{F}})}^1({\mathbf{{Q}}}_p,V_g(1)). \end{aligned}$$
(2.11)

Now consider a crystalline extension

$$\begin{aligned} 0\rightarrow V_g(1)\rightarrow W\rightarrow {\mathbf{{Q}}}_p\rightarrow 0. \end{aligned}$$
(2.12)

Since \(D_\mathrm{cris}(V_g(1))^{\varphi =1}=0\) by our assumptions, the resulting extension of \(\varphi \)-modules

$$\begin{aligned} 0\rightarrow D_\mathrm{cris}(V_g(1))\rightarrow D_\mathrm{cris}(W)\rightarrow F_0\rightarrow 0 \end{aligned}$$
(2.13)

admits a unique section \(s_W^\mathrm{frob}:F_0\rightarrow D_\mathrm{cris}(W)\) with \(s_W^\mathrm{frob}(1)\in D_\mathrm{cris}(W)^{\varphi =1}\). Extending scalars from \(F_0\) to \({F}\) in (2.13) and taking \(\mathrm{Fil}^0\)-parts, we take an arbitrary section \(s_W^\mathrm{fil}:{F}\rightarrow \mathrm{Fil}^0D_\mathrm{dR}(W)\) of the resulting exact sequence of \({F}\)-vector spaces

$$\begin{aligned} 0\rightarrow \mathrm{Fil}^0D_\mathrm{dR}(V_g(1))\rightarrow \mathrm{Fil}^0D_\mathrm{dR}(W)\rightarrow {F}\rightarrow 0 \end{aligned}$$
(2.14)

and form the difference

$$\begin{aligned} t_W:=s_W^\mathrm{fil}(1)-s_W^\mathrm{frob}(1), \end{aligned}$$

which can be seen in \(D_\mathrm{dR}(V_g(1))\), and whose image modulo \(\mathrm{Fil}^0D_\mathrm{dR}(V_g(1))\) is well-defined.

Lemma 2.4

Under the identification (2.11), the above assignment

$$\begin{aligned} 0\rightarrow V_g(1)\rightarrow W\rightarrow {\mathbf{{Q}}}_p\rightarrow 0\quad \rightsquigarrow \quad t_W\;\mathrm{mod}\;\mathrm{Fil}^0D_\mathrm{dR}(V_g(1)) \end{aligned}$$

defines an isomorphism which agrees with the Bloch-Kato logarithm map

$$\begin{aligned} \mathrm{log}_{{F},V_g(1)}:H_f^1({F},V_g(1))\xrightarrow {\sim }\frac{D_\mathrm{dR}(V_g(1))}{\mathrm{Fil}^0D_\mathrm{dR}(V_g(1))}. \end{aligned}$$

Proof

See [26, Lem. 2.7], for example. \(\square \)

Let \(\Delta \in J_s({F})\) be the class of a degree 0 divisor on \(X_s\) with support contained in the finite set of points \(S\subset X_s({F})\). The extension class \(W=W_\Delta \) (2.12) corresponding to \({\delta }_{g,{F}}(\Delta )\) can then be constructed from the étale cohomology of the open curve \(Y_s:=X_s\!\backslash \! S\), as explained in [1, Sect. 4.1]. We describe the associated \(s_{W_\Delta }^\mathrm{fil}\) and \(s_{W_\Delta }^\mathrm{frob}\).

By [33] (or also [13]), denoting \(g\)-isotypical components by the superscript \(g\), there is a canonical isomorphism of \(F_0\otimes _{{\mathbf{{Q}}}_p}L_g\)-modules

$$\begin{aligned} D_\mathrm{cris}(V_g)\cong H_\mathrm{log-cris}^1({\fancyscript{X}}_s)^g \end{aligned}$$
(2.15)

compatible with \(\varphi \)-actions and inducing an \({F}\otimes _{{\mathbf{{Q}}}_p}L_g\)-module isomorphism

$$\begin{aligned} D_\mathrm{dR}(V_g)\cong H_\mathrm{dR}^1(X_s/{F})^g \end{aligned}$$
(2.16)

after extension of scalars.

Writing \(\Delta =\sum _{Q\in S}n_Q.Q\) for some \(n_Q\in {\mathbf{{Z}}}\), we assume from now on that the reductions of the points \(Q\in S\) are smooth and pair-wise distinct. Assume from now on that the reduction of \(S\) in the special fiber is stable under the absolute Frobenius. Like \(H^1_\mathrm{dR}(X_s/{F})\), the \({F}\)-vector space \(H^1_\mathrm{dR}(Y_s/{F})\) is equipped with a canonical \(F_0\)-structure

$$\begin{aligned} H_\mathrm{log-cris}^1({\fancyscript{Y}}_s)\hookrightarrow H_\mathrm{dR}^1(Y_s/{F}), \end{aligned}$$
(2.17)

a Frobenius operator still denoted by \(\varphi \), and a Hecke action compatible with that in (2.4). Thus for \(W=W_\Delta \) the exact sequence (2.13) is obtained as the pullback

(2.18)

of the bottom extension of \(\varphi \)-modules with respect to the \(F_0\otimes _{{\mathbf{{Q}}}_p}L_g\)-linear map sending \(1\mapsto (n_Q)_{Q\in S}\), where the subscript 0 indicates taking the degree 0 subspace.

On the other hand, after extending scalars from \(F_0\) to \({F}\) and taking \(\mathrm{Fil}^0\)-parts, (2.14) is given by the pullbackFootnote 2

(2.19)

of the bottom exact sequence of free \({F}\otimes _{{\mathbf{{Q}}}_p}L_g\)-modules with respect to the \({F}\otimes _{{\mathbf{{Q}}}_p} L_g\)-linear map sending \(1\mapsto (n_Q)_{Q\in S}\).

Let \(g^*\in S_2(X_s)\) be the form dual to \(g\), defined as the newform associated with the twist \(g\otimes \varepsilon _g^{-1}\), and let \(\omega _{g^*}\in H^0(X_s,\Omega _{X_s/{F}}^1)\) be its associated differential, so that \(\mathrm{Fil}^0D_\mathrm{dR}(V_{g}^*)=\mathrm{Fil}^1D_\mathrm{dR}(V_{g^*})=({F}\otimes _{{\mathbf{{Q}}}_p}L_g).\omega _{g^*}\). Thus \(\delta _{g,{F}}^{(p)}(\Delta )\) is determined by the value

$$\begin{aligned} {\delta }_{g,{F}}^{(p)}(\Delta )(\omega _{g^*})=\langle t_{W_{\Delta }},\omega _{g^*}\rangle \end{aligned}$$
(2.20)

of the pairing (2.10), which corresponds to the Poincaré pairing on \(H_\mathrm{dR}^1(X_s/{F})\) under the identification (2.16). Using rigid analysis, we now give an expression for the latter pairing that will make (2.20) amenable to computations.

Let \(\mathcal{X }_{s}\) be the canonical balanced model of \(X_s\) over \({\mathbf{{Z}}}_p[\zeta _s]\) constructed by Katz and Mazur (see [22, Ch. 13]). The special fiber \(\mathcal{X }_s\times _{{\mathbf{{Z}}}_p[\zeta _s]}\mathbf{F}_p\) is a reduced disjoint union of Igusa curves over \(\mathbf{F}_p\) intersecting at the supersingular points. Exactly two of these components are isomorphic to the Igusa curve \(\mathrm{Ig}(\Gamma _s)\) representing the moduli problem \(([\Gamma _1(N)],[\mathrm{bal}.\Gamma _1(p^s)^\mathrm{can}])\) over \(\mathbf{F}_p\), and we let \(I_\infty \) be the one that contains the reduction of \(\mathcal{W }_1(p^s)\times _{{\mathbf{{Q}}}_p}{{\mathbf{{Q}}}_p(\zeta _s)}\), and \(I_0\) be the other. (These two are the two “good” components in the terminology of [24]).

By the universal property of the regular minimal model, there exists a morphism

$$\begin{aligned} {\fancyscript{X}}_s\rightarrow \mathcal{X }_s\times _{{\mathbf{{Z}}}_p[\zeta _s]}\mathcal{O }_{F}\end{aligned}$$
(2.21)

which reduces to a sequence of blow-ups on the special fiber. Letting \(\kappa \) be the residue field of \({F}\), define \(\mathcal{W }_{\infty }\subset X_s\) (resp. \(\mathcal{W }_{0}\subset X_s\)) to be the inverse image under the reduction map via \({\fancyscript{X}}_s\) of the unique irreducible component of \({\fancyscript{X}}_s\times _{\mathcal{O }_{F}}\kappa \) mapping bijectively onto \(I_\infty \times _{\mathbf{F}_p}\kappa \) (resp. \(I_0\times _{\mathbf{F}_p}\kappa \)) in \(\mathcal{X }_s\times _{{\mathbf{{Z}}}_p[\zeta _s]}\kappa \) via the reduction of (2.21). Similarly, define \(\mathcal{U }\subset X_s\) by considering the irreducible components of \(\mathcal{X }_s\times _{{\mathbf{{Z}}}_p[\zeta _s]}\kappa \) different from \(I_\infty \times _{\mathbf{F}_p}\kappa \) and \(I_0\times _{\mathbf{F}_p}\kappa \). Letting \(SS\) denote (the degree of) the supersingular divisor of \(\mathrm{Ig}(\Gamma _s)\), it follows that \(\mathcal{U }\) intersects \(\mathcal{W }_{\infty }\) (resp. \(\mathcal{W }_{0}\)) in a union of \(SS\) supersingular annuli.

Since they reduce to smooth points, the residue class \(D_Q\) of each \(Q\in S\) is conformal to the open unit disc \(D\subset {\mathbf{{C}}}_p\). Fix an isomorphism \(h_Q:D_Q\xrightarrow {\sim }D\) that sends \(Q\) to 0, and for a real number \(r_Q<1\) in \(p^{\mathbf{{Q}}}\), denote by \(\mathcal{V }_Q\subset D_Q\) the annulus consisting of the points \(x\in D_Q\) with \(r_Q<\vert h_Q(x)\vert _p<1\). In the same manner, we define annuli \(\mathcal{V }_z\) for each \(z\) in the cuspidal subscheme \(Z_s\subset X_s\).

Attached to any (oriented) annulus \(\mathcal{V }\), there is a \(p\) -adic annular residue map

$$\begin{aligned} \mathrm{Res}_{\mathcal{V }}:\Omega _\mathcal{V }^1\rightarrow {\mathbf{{C}}}_p \end{aligned}$$

defined by expanding \(\omega \in \Omega ^1_\mathcal{V }\) as \(\omega =\sum _{n\in {\mathbf{{Z}}}}a_nT^n\frac{dT}{T}\) for a fixed uniformizing parameter \(T\) on \(\mathcal{V }\) (compatible with the orientation), and setting \(\mathrm{Res}_\mathcal{V }(\omega )=a_0\). This descends to a linear functional on \(\Omega _\mathcal{V }^1/d\mathcal{O }_\mathcal{V }\). (Cf. [6, Lem. 2.1]).

For any basic wide-open  \(\mathcal{W }\) (as in [4, p. 34]), define

$$\begin{aligned} H_\mathrm{rig}^1(\mathcal{W }):=\mathbb{H }^1(\mathcal{W },\Omega _{}^\bullet (\log {Z})) \cong {\Omega _{\mathcal{W }}^1}/{d\mathcal{O }_{\mathcal{W }}}, \end{aligned}$$
(2.22)

where \(\Omega _{}^\bullet (\log {Z})\) denotes the complex of rigid analytic sheaves on \(\mathcal{W }\) deduced from (2.1) by analytification and pullback, and consider the basic wide-opens

$$\begin{aligned} \widetilde{\mathcal{W }}_\infty :=\mathcal{W }_\infty \backslash \bigcup _{Q\in S}(D_Q\backslash \mathcal{V }_Q)\quad \mathrm{and}\quad \widetilde{\mathcal{W }}_0:=\mathcal{W }_0\backslash \bigcup _{Q\in S}(D_Q\backslash \mathcal{V }_Q). \end{aligned}$$

As follows from the arguments in [2, Lem. 4.4.1], the spaces \(H_\mathrm{rig}^1(\widetilde{\mathcal{W }}_\infty )\) and \(H_\mathrm{rig}^1(\widetilde{\mathcal{W }}_0)\) are each equipped with a natural action of the Hecke operators \(T_\ell \) (\(\ell \not \mid Np\)) compatible with the Hecke action on \(H^1_\mathrm{dR}(Y_s/{F})\) under restriction.

Lemma 2.5

  • The natural restriction maps induce an isomorphism

    $$\begin{aligned} H_\mathrm{dR}^1(Y_s/{F})^g\;\xrightarrow {}\;H_\mathrm{rig}^1(\widetilde{\mathcal{W }}_{\infty })^g\oplus H_\mathrm{rig}^1(\widetilde{\mathcal{W }}_0)^g. \end{aligned}$$
  • A class \(\omega \in H^1_\mathrm{dR}(Y_s/{F})^g\) belongs to the natural image of \(H^1_\mathrm{dR}(X_s/{F})^g\) if an only if it can be represented by a pair of differentials \((\omega _\infty ,\omega _0)\in \Omega ^1_{\widetilde{\mathcal{W }}_\infty } \times \Omega ^1_{\widetilde{\mathcal{W }}_0}\) with vanishing \(p\)-adic annular residues.

  • If \(\eta \) and \(\omega \) are any two classes in \(H_\mathrm{dR}^1(X_s/{F})^g\), their Poincaré pairing can be computed as

    $$\begin{aligned} \langle \eta ,\omega \rangle = \sum _{\mathcal{V }\subset \widetilde{\mathcal{W }}_\infty }\mathrm{Res}_{\mathcal{V }}(F_{\omega _\infty \vert _\mathcal{V }}\cdot \eta _{\infty }\vert _\mathcal{V })+ \sum _{\mathcal{V }\subset \widetilde{\mathcal{W }}_0}\mathrm{Res}_{\mathcal{V }}(F_{\omega _0\vert _\mathcal{V }}\cdot \eta _0\vert _\mathcal{V }), \end{aligned}$$
    (2.23)

    where for each annulus \(\mathcal{V }\), \(F_{\omega _\mathcal{V }}\) denotes any solution to \(dF_{\omega _\mathcal{V }}=\omega _\mathcal{V }\) on \(\mathcal{V }\).

Proof

By an excision argument, the first assertion is easily deduced from [10, Thm. 2.1] as in [2, Lem. 4.4.2]; the second and third are shown by adapting the arguments in [9, §5] for each of the two components, as they are proven in [7, Prop. 1.3] for \(s=1\). (See also [10, §\({3}\)].) \(\square \)

2.3 Coleman \(p\)-adic integration

Coleman’s theory provides a coherent choice of local primitives that will allow us to compute (2.20) using the formula (2.23).

Recall the lift of Frobenius \(\phi :\mathcal{W }_2(p^s)\rightarrow \mathcal{W }_1(p^s)\) described in Sect. 2.1, where \(\mathcal{W }_i(p^s)\) are the strict neighborhoods of the connected component \(X_s(0)\) of the ordinary locus of \(X_s\) containing the cusp \(\infty \) described there. Recall also the wide open space \(\mathcal W _\infty \) described in the preceding section, which also contains \(X_s(0)\) by construction.

Proposition 2.6

(Coleman) Let \(g=\sum _{n>0}b_nq^n\in S_2(X_s)\) be a normalized newform with primitive nebentypus of \(p\)-power conductor, so that \(b_p\) is such that \(U_pg=b_pg\). Then there exists a locally analytic function \(F_{\omega _g}\) on \(\mathcal{W }_{\infty }\) which is unique up to a constant on \(\mathcal{W }_\infty \) and such that

  • \(dF_{\omega _g}=\omega _{g}\) on \({\mathcal{W }_\infty }\), and

  • \(F_{\omega _g}-\frac{b_p}{p}\phi ^*F_{\omega _g} \in M_0^\mathrm{rig}(X_s).\)

Proof

This follows from the general result of Coleman [8, Thm. 10.1]. Indeed, a computation on \(q\)-expansions shows that the action of the Frobenius lift \(\phi \) on differentials agrees with that of \(pV\), with \(V\) the map sending \(q\mapsto q^p\), in the sense that \(\phi ^*\omega _g=p\omega _{Vg}\) on \(\mathcal{W }_\infty ^{\prime }:=\phi ^{-1}(\mathcal{W }_\infty \cap \mathcal{W }_1(p^s))\). Since the differential \(\omega _{g^{[p]}}=\omega _{g}-b_p\omega _{Vg}\) attached to

$$\begin{aligned} g^{[p]}=\sum _{(n,p)=1}b_nq^n \end{aligned}$$

becomes exact upon restriction to \(\mathcal{W }_\infty ^{\prime }\), this shows that the polynomial \(L(T)=1-\frac{b_p}{p}T\) is such that

$$\begin{aligned} L(\phi ^*)\omega _g=0. \end{aligned}$$

Finally, since \(g\) has primitive nebentypus, \(b_p\) has complex absolute value \(p^{1/2}\), and hence [8, Thm. 10.1] can be applied with \(L(T)\) as above. \(\square \)

Attached to a primitive \(p^s\)-th root of unity \(\zeta \), there is an automorphism \(w_\zeta \) of \(X_s\) which interchanges the components \(\mathcal{W }_\infty \) and \(\mathcal{W }_0\) (see [2, Lem. 4.4.3]).

Corollary 2.7

Define \(\phi ^{\prime }:=w_\zeta \circ \phi \circ w_\zeta \). With hypotheses as in Proposition 2.6, there exists a unique locally analytic function \(F_{\omega _g}^{\prime }\) on \(\mathcal{W }_{0}\) which vanishes at 0, satisfies \(dF_{\omega _g}^{\prime }=\omega _{g}\) on \(\mathcal{W }_0\), and \(F_{\omega _g}^{\prime }-\frac{b_p}{p}(\phi ^{\prime })^*F_{\omega _g}^{\prime }\) is rigid analytic on a wide-open neighborhood \(\mathcal{W }_0^{\prime }\) of \(w_\zeta X_s(0)\) in \(\mathcal{W }_0\).

Proof

Proposition 2.6 applied to the differential \(\omega _g^{\prime }:=w_\zeta ^*\omega _g\) gives the existence of a locally analytic function \(F_{\omega _g^{\prime }}\) with \(F_{\omega _g}^{\prime }:=w_\zeta ^*F_{\omega _g^{\prime }}\) having the desired properties. The uniqueness of \(F_{\omega _g}^{\prime }\) follows immediately from that of \(F_{\omega _g^{\prime }}\). \(\square \)

We refer to the locally analytic function \(F_{\omega _g}\) (resp. \(F_{\omega _g}^{\prime }\)) appearing in Proposition 2.6 as the Coleman primitive of \(g\) on \(\mathcal{W }_\infty \) (resp. \(\mathcal{W }_0\)). Let \(g=\sum _{n>0}b_nq^n\) be as in Proposition 2.6. The \(q\)-expansion \(\sum _{(n,p)=1}\frac{b_n}{n}q^n\) corresponds to a \(p\)-adic modular form \(g^{\prime }\) vanishing at \(\infty \) and satisfying \(dg^{\prime }=g^{[p]}\), where \(d\) is the operator described at the end of Section 2.1, which here corresponds to the differential operator \(\mathcal O _\mathcal{W }\rightarrow \Omega _\mathcal{W }^1\) for any subspace \(\mathcal W \subset X_s\). Set \(d^{-1}g^{[p]}:=g^{\prime }\).

Corollary 2.8

If \(F_{\omega _g}\) is the Coleman primitive of \(g\) on \(\mathcal{W }_\infty \) which vanishes at \(\infty \), then

$$\begin{aligned} F_{\omega _g}-\frac{b_p}{p}\phi ^*F_{\omega _g}=d^{-1}g^{[p]}. \end{aligned}$$

Proof

Since \(d^{-1}g^{[p]}\) is an overconvergent rigid analytic primitive of \(\omega _{g^{[p]}}\), and the operator \(L(\phi ^*)=1-\frac{b_p}{p}\phi ^*\) acting on the space of locally analytic functions on \(\mathcal{W }^{\prime }_\infty \) is invertible, we see that \(L(\phi ^*)^{-1}(d^{-1}g^{[p]})\) satisfies the defining properties of \(F_{\omega _g}\). Since \(d^{-1}g^{[p]}\) vanishes at \(\infty \), the result follows. \(\square \)

Now we can give a closed formula for the \(p\)-adic Abel-Jacobi images of certain degree 0 divisors on \(X_s\).

Proposition 2.9

Assume \(s>1\). Let \(g\in S_2(X_s)\) be a normalized newform with primitive nebentypus of \(p\)-power conductor, let \(P\) be an \({F}\)-rational point of \(X_s\) factoring through \(X_s(0)\subset X_s\), and let \(\Delta \in J_s({F})\) be the divisor class of \((P)-(\infty )\). Then

$$\begin{aligned} \delta ^{(p)}_{g,{F}}(\Delta )(\omega _{g^*})=F_{\omega _{g^*}}(P), \end{aligned}$$
(2.24)

where \(F_{\omega _{g^*}}\) is the Coleman primitive of \(\omega _{g^*}\) on \(\mathcal{W }_\infty \) which vanishes at \(\infty \).

Proof

By (2.20), we must compute \(\langle t_{W_\Delta },\omega _{g^*}\rangle = \langle s_{W_\Delta }^\mathrm{fil},\omega _{g^*}\rangle - \langle s_{W_\Delta }^\mathrm{frob},\omega _{g^*}\rangle \), where

  • \(s_{W_\Delta }^\mathrm{fil}\in \mathrm{Fil}^1D_\mathrm{dR}(W_\Delta )\) is such that \(\rho (s_{W_\Delta }^\mathrm{fil})=1\) in (2.19), and

  • \(s_{W_\Delta }^\mathrm{frob}\in D_\mathrm{cris}(W_\Delta )^{\varphi =1}\) is such that \(\rho (s_{W_\Delta }^\mathrm{frob})=1\) in (2.18).

By Lemma 2.5, we see that these can be represented, respectively, by

  • \(\eta _{\Delta }^\mathrm{fil}\) a section of \(\Omega _{X_s/{F}}^1\) over \(Y_s\) with simple poles at \(P\) and \(\infty \) and with

    • \(\mathrm{Res}_{P}(\eta _{\Delta }^\mathrm{fil})=1\), while \(\mathrm{Res}_{\infty }(\eta _{\Delta }^\mathrm{fil})=0\) for all \(Q\in S-\{P\}\);

    • \(\mathrm{Res}_{\infty }(\eta _{\Delta }^\mathrm{fil})=-1\), while \(\mathrm{Res}_{z}(\eta _{\Delta }^\mathrm{fil})=0\) for all \(z\in Z_s-\{\infty \}\),

  • \(\eta _{\Delta }^\mathrm{frob}=(\eta _{\infty }^\mathrm{frob},\eta _{0}^\mathrm{frob}) \in \Omega _{\widetilde{\mathcal{W }}_\infty }^1\times \Omega _{\widetilde{\mathcal{W }}_0}^1\) with

    • \((\phi ^*\eta _{\infty }^\mathrm{frob},(\phi ^{\prime })^*\eta _{\Delta }^\mathrm{frob})=(p\cdot \eta _{\infty }^\mathrm{frob}+dG_\infty ,p\cdot \eta _{0}^\mathrm{frob}+dG_0)\) with \(G_\infty \) and \(G_0\) rigid analytic on \(\phi ^{-1}\widetilde{\mathcal{W }}_\infty \) and \((\phi ^{\prime })^{-1}\widetilde{\mathcal{W }}_0\), respectively;

    • \(\mathrm{Res}_\mathcal{V }(\eta _\Delta ^\mathrm{frob})=0\) for all supersingular annuli \(\mathcal{V }\); and

    • \(\mathrm{Res}_{\mathcal{V }_Q}(\eta _\Delta ^\mathrm{frob})=\mathrm{Res}_Q(\eta _\Delta ^\mathrm{fil})\) (\(Q\in S\)), \(\mathrm{Res}_{\mathcal{V }_z}(\eta _\Delta ^\mathrm{frob})=\mathrm{Res}_z(\eta _\Delta ^\mathrm{fil})\) (\(z\in Z_s\)).

The arguments in [1, Prop. 3.21] can now be straightforwardly adapted to deduce the result. Indeed, using the defining properties of the Coleman primitives \(F_{\omega _{g^*}}\) and \(F_{\omega _{g^*}}^{\prime }\) of \(\omega _{g^*}\) on \(\mathcal{W }_\infty \) and \(\mathcal{W }_0\), respectively, one first shows that

$$\begin{aligned} \sum _{\mathcal{V }\subset \widetilde{\mathcal{W }}_\infty }\mathrm{Res}_\mathcal{V }(F_{\omega _{g^*}}\cdot \eta _\infty ^\mathrm{frob})=0\quad \mathrm{and}\quad \sum _{\mathcal{V }\subset \widetilde{\mathcal{W }}_0}\mathrm{Res}_\mathcal{V }(F_{\omega _{g^*}}^{\prime }\cdot \eta _0^\mathrm{frob})=0 \end{aligned}$$
(2.25)

as in [loc.cit., Lemma 3.20]. On the other hand, using the same primitives, one shows as in [loc.cit., Lemma 3.19] that

$$\begin{aligned} \sum _{\mathcal{V }\subset \widetilde{\mathcal{W }}_\infty }\mathrm{Res}_\mathcal{V }(F_{\omega _{g^*}}\cdot \eta _\Delta ^\mathrm{fil}) \!=\!F_{\omega _{g^*}}(P) \quad \mathrm{and}\quad \sum _{\mathcal{V }\subset \widetilde{\mathcal{W }}_0}\mathrm{Res}_\mathcal{V }(F_{\omega _{g^*}}^{\prime }\cdot \eta _\Delta ^\mathrm{fil})=0.\qquad \end{aligned}$$
(2.26)

Substituting (2.26) and (2.25) into the formula (2.23) for the Poincaré pairing (and using that \(s>1\), so that there is no overlap between the supersingular annuli in \(\widetilde{\mathcal{W }}_\infty \) and the supersingular annuli in \(\widetilde{\mathcal{W }}_0\)), the result follows. \(\square \)

3 Generalised Heegner cycles

Let \(X_1(N)\) be the compactified modular curve of level \(\Gamma _1(N)\) defined over \({\mathbf{{Q}}}\), and let \(\mathcal{E }\) be the universal generalized elliptic curve over \(X_1(N)\). (Recall that \(N>4\)). For \(r>1\), denote by \(W_{r}\) the \((2r-1)\)-dimensional Kuga-Sato varietyFootnote 3, defined as the canonical desingularization of the \((2r-2)\)-nd fiber product of \(\mathcal{E }\) with itself over \(X_1(N)\). By construction, the variety \(W_r\) is equipped with a proper morphism

$$\begin{aligned} \pi _r:W_r\rightarrow X_1(N) \end{aligned}$$

whose fibers over a noncuspidal closed point of \(X_1(N)\) corresponding to an elliptic curve \(E\) with \(\Gamma _1(N)\)-level structure is identified with \(2r-2\) copies of \(E\). (For a more detailed description, see [1, Sect. 3.1].)

Let \(K\) be an imaginary quadratic field of odd discriminant \(-D<0\). It will be assumed throughout that \(K\) satisfies the following hypothesis:

Assumption 3.1

All the prime factors of \(N\) split in \(K\).

Denote by \(\mathcal{O }_K\) the ring of integers of \(K\), and note that by this assumption we may choose an ideal \(\mathfrak{N }\subset \mathcal{O }_K\) with \(\mathcal{O }_K/\mathfrak{N }\cong {\mathbf{{Z}}}/N{\mathbf{{Z}}}\) that we fix once and for all.

Let \(A\) be a fixed elliptic curve with CM by \(\mathcal{O }_K\). The pair \((A,A[\mathfrak{N }])\) defines a point \(P_A\) on \(X_0(N)_{}\) rational over \(H\), the Hilbert class field of \(K\). Choose one of the square-roots \(\sqrt{-D}\in \mathcal{O }_K\), let \(\Gamma _{\sqrt{-D}}\subset A\times A\) be the graph of \(\sqrt{-D}\), and define

$$\begin{aligned} \Upsilon _{A,r}^\mathrm{heeg}:=\Gamma _{\sqrt{-D}}\times \overset{(r-1)}{\cdots } \times \Gamma _{\sqrt{-D}} \end{aligned}$$

viewed inside \(W_r\) by the natural inclusion \((A\times A)^{r-1}\rightarrow W_r\) as the fiber of \(\pi _r\) over a point on \(X_1(N)\) lifting \(P_A\). Let \(\epsilon _W\) be the projector from [1, (2.1.2)], and set

$$\begin{aligned} \Delta ^\mathrm{heeg}_{A,r}:=\epsilon _W\Upsilon _{A,r}^\mathrm{heeg}, \end{aligned}$$
(3.1)

which is an \((r-1)\)-dimensional null-homologous cycle on \(W_r\) defining an \(H\)-rational class in the Chow group \(\mathrm{CH}^r(W_r)_0\) (taken with \({\mathbf{{Q}}}\)-coefficients, as always here).

These cycles (3.1) are usually referred to as Heegner cycles (of conductor one, weight \(2r\)), and they share with classical Heegner points (as in [15]) many of their arithmetic properties (see [25, 27, 35]).

We next recall a variation of the previous construction introduced in the recent work [1] of Bertolini–Darmon–Prasanna. Let \(A\) be the CM elliptic curve fixed above, and consider the varietyFootnote 4

$$\begin{aligned} X_r:=W_r\times A^{2r-2}. \end{aligned}$$

For each class \([\mathfrak a ]\in \mathrm{Pic}(\mathcal{O }_K)\), represented by an ideal \(\mathfrak a \subset \mathcal{O }_K\) prime to \(N\), let \(A_\mathfrak a :=A/A[\mathfrak a ]\) and denote by \(\varphi _\mathfrak a \) the degree \(\mathrm{N}\mathfrak a \)-isogeny

$$\begin{aligned} \varphi _\mathfrak a :A\rightarrow A_\mathfrak a . \end{aligned}$$

The pair \(\mathfrak a *(A,A[\mathfrak{N }]):=(A_\mathfrak a ,A_\mathfrak a [\mathfrak{N }])\) defines a point \(P_{A_\mathfrak a }\) in \(X_0(N)\) rational over \(H\). Let \(\Gamma _{\varphi _\mathfrak a }^t\subset A_\mathfrak a \times A\) be the transpose of the graph of \(\varphi _\mathfrak a \), and set

$$\begin{aligned} \Upsilon ^\mathrm{bdp}_{\varphi _\mathfrak a ,r}:=\Gamma _{\varphi _\mathfrak a }^t \times \overset{(2r-2)}{\cdots }\times \Gamma _{\varphi _\mathfrak a }^t\; \subset \;(A_\mathfrak a \times A)^{2r-2}=A_\mathfrak{a }^{2r-2}\times A^{2r-2}\xrightarrow {(\imath _\mathfrak a ,\mathrm{id}_A)} X_r, \end{aligned}$$

where \(\imath _\mathfrak a \) is the natural inclusion \(A_\mathfrak{a }^{2r-2}\rightarrow W_r\) as the fiber of \(\pi _r\) over a point on \(X_1(N)\) lifting \(P_{A_\mathfrak a }\). Letting \(\epsilon _A\) be the projector from [1, (1.4.4)], the cycles

$$\begin{aligned} \Delta _{\varphi _\mathfrak a ,r}^\mathrm{bdp}:=\epsilon _A\epsilon _W\Upsilon _{\varphi _\mathfrak a ,r}^\mathrm{bdp} \end{aligned}$$
(3.2)

define classes in \(\mathrm{CH}^{2r-1}(X_r)_0(H)\) and are referred to as generalised Heegner cycles.

We will assume for the rest of this paper that \(K\) also satisfies the following:

Assumption 3.2

The prime \(p\) splits in \(K\).

Let \(g\in S_{2r}(X_0(N))\) be a normalized newform, and let \(V_g\) be the \(p\)-adic Galois representation associated to \(g\) by Deligne. By the Künneth formula, there is a map

$$\begin{aligned} H_{{{{\acute{\mathrm{e}}}}\mathrm{t}}}^{4r-3}(\overline{X}_r,{\mathbf{{Q}}}_p(2r-1))\longrightarrow H_{{{{\acute{\mathrm{e}}}}\mathrm{t}}}^{2r-1}(\overline{W}_r,{\mathbf{{Q}}}_p(1)) \otimes \mathrm{Sym}^{2r-2}H^1_{{{{\acute{\mathrm{e}}}}\mathrm{t}}}(\overline{A},{\mathbf{{Q}}}_p(1)), \end{aligned}$$

which composed with the natural Galois-equivariant projection

$$\begin{aligned} H_{{{{\acute{\mathrm{e}}}}\mathrm{t}}}^{2r-1}(\overline{W}_r,{\mathbf{{Q}}}_p(1)) \otimes \mathrm{Sym}^{2r-2}H^1_{{{{\acute{\mathrm{e}}}}\mathrm{t}}}(\overline{A},{\mathbf{{Q}}}_p(1))\xrightarrow {\pi _g\otimes \pi _{\mathrm{N}^{r-1}}} V_g(r) \end{aligned}$$

induces a map

$$\begin{aligned} \pi _{g,\mathrm{N}^{r-1}}:H^1(F,H_{{{{\acute{\mathrm{e}}}}\mathrm{t}}}^{4r-3}(\overline{X}_r,{\mathbf{{Q}}}_p(2r-1))) \longrightarrow H^1(F,V_g(r))\nonumber \end{aligned}$$

over any number field \(F\). In the following we fix a number field \(F\) containing \(H\).

Now consider the étale Abel-Jacobi map

$$\begin{aligned} \Phi _F^{{{{\acute{\mathrm{e}}}}\mathrm{t}}}:\mathrm{CH}^{2r-1}(X_r)_0(F)\rightarrow H^1(F,H_{{{{\acute{\mathrm{e}}}}\mathrm{t}}}^{4r-3}(\overline{X}_r,{\mathbf{{Q}}}_p)(2r-1)) \end{aligned}$$

constructed in [28]. Let \(F_p\) be the completion of \(\imath _p(F)\), and denote by \(\mathrm{loc}_p\) the induced localization map from \(G_F\) to \(\mathrm{Gal}(\overline{{\mathbf{{Q}}}}_p/F_p)\). Then we may define the \(p\) -adic Abel-Jacobi map \(\mathrm{AJ}_{F_p}\) by the commutativity of the diagram

(3.3)

where the existence of the dotted arrow follows from [28, Thm.(3.1)(i)], and the vertical map is given by the logarithm map of Bloch-Kato, as it appeared in (2.9) for \(r=1\). Using the comparison isomorphism of Faltings [12], the map \(\mathrm{AJ}_{F_p}\) may be evaluated at the class \(\omega _g\otimes e_\zeta ^{\otimes r-1}\), with \(e_\zeta \) an \(F_p\)-basis of \(D_\mathrm{dR}({\mathbf{{Q}}}_p(1))\cong F_p\).

The main result of [1] yields the following formula for the \(p\)-adic Abel-Jacobi images of the generalised Heegner cycles (3.2) which we will need.

Theorem 3.3

(Bertolini–Darmon–Prasanna) Let \(g=\sum _nb_nq^n\in S_{2r}(X_0(N))\) be a normalized newform of weight \(2r\ge 2\) and level \(N\) prime to \(p\). Then

$$\begin{aligned} (1-b_pp^{-r}+p^{-1})&\sum _{[\mathfrak a ]\in \mathrm{Pic}(\mathcal{O }_K)}\mathrm{N}\mathfrak a ^{1-r}\cdot \mathrm{AJ}_{F_p}(\Delta _{\varphi _\mathfrak a ,r}^\mathrm{bdp}) (\omega _g\otimes e_\zeta ^{\otimes r-1})\\&\quad \quad =(-1)^{r-1}(r-1)!\sum _{[\mathfrak a ]\in \mathrm{Pic}(\mathcal{O }_K)}d^{-r}g^{[p]}(\mathfrak a *(A,A[\mathfrak{N }])), \end{aligned}$$

where \(g^{[p]}=\sum _{(n,p)=1}b_nq^n\) is the \(p\)-depletion of \(g\).

Proof

See the proof of [1, Thm. 5.13]. \(\square \)

We end this section by relating the images of Heegner cycles and of generalised Heegner cycles under the \(p\)-adic height pairing. (Cf. [1, Sect. 3.4]).

Consider \(\Pi _r:=W_r\times A^{r-1}\) seen as a subvariety of \(W_r\times X_r=W_r\times W_r\times (A^2)^{r-1}\) via the map

$$\begin{aligned} (\mathrm{id}_{W_r},\mathrm{id}_{W_r},(\mathrm{id}_A,\sqrt{-D})^{r-1}). \end{aligned}$$

Denoting by \(\pi _W\) and \(\pi _X\) the projections onto the first and second factors of \(W_r\times X_r\), the rational equivalence class of the cycle \(\Pi _r\) gives rise to a map on Chow groups

$$\begin{aligned} \Pi _r:\mathrm{CH}^{2r-1}(X_r)\rightarrow \mathrm{CH}^{r+1}(W_{r}) \end{aligned}$$

induced by \(\Pi _r(\Delta )=\pi _{W,*}(\Pi _r\cdot \pi _X^*\Delta )\).

Lemma 3.4

We have

$$\begin{aligned} \langle \Delta _{A,r}^\mathrm{heeg},\Delta _{A,r}^\mathrm{heeg}\rangle _{W_r}= (4D)^{r-1}\cdot \langle \Delta _{\mathrm{id}_A,r}^\mathrm{bdp},\Delta _{\mathrm{id}_A,r}^\mathrm{bdp}\rangle _{X_r}, \end{aligned}$$

where \(\langle ,\rangle _{W_r}\) and \(\langle ,\rangle _{X_r}\) are the \(p\)-adic height pairings of [26] on \(\mathrm{CH}^{r+1}(W_r)_0\) and \(\mathrm{CH}^{2r-1}(X_r)_0\), respectively.

Proof

The image \(\Phi _{F}^{{{{\acute{\mathrm{e}}}}\mathrm{t}}}(\Delta _{A,r}^\mathrm{{heeg}})\) remains unchanged if we replace \(\Gamma _{\sqrt{-D}}\) by \(Z_A:=\Gamma _{\sqrt{-D}}-(A\times \{0\})-D(\{0\}\times A)\) (see [27, §II(3.6)]). Since \(Z_A\cdot Z_A=-2D\), we easily see from the construction of \(\Pi _r\) that

$$\begin{aligned} \Phi _F^{{{{\acute{\mathrm{e}}}}\mathrm{t}}}(\Delta _{A,r}^\mathrm{heeg})=(-2D)^{r-1}\cdot \Phi _F^{{{{\acute{\mathrm{e}}}}\mathrm{t}}}(\Pi _r(\Delta _{\mathrm{id}_A,r}^\mathrm{bdp})). \end{aligned}$$
(3.4)

On the other hand, if \(\langle ,\rangle _A\) denotes the Poincaré pairing on \(H_\mathrm{dR}^1(A/F)\), we have

$$\begin{aligned} \langle (\sqrt{-D})^*\omega ,(\sqrt{-D})^*\omega ^{\prime }\rangle _A =D\cdot \langle \omega ,\omega ^{\prime }\rangle _A, \end{aligned}$$

for all \(\omega , \omega ^{\prime }\in H_\mathrm{dR}^1(A/F)\). By the definition of the \(p\)-adic height pairings \(\langle ,\rangle _{W_r}\) and \(\langle ,\rangle _{X_{r}}\) (factoring through \(\Phi _{F}^{{{{\acute{\mathrm{e}}}}\mathrm{t}}}\)), we thus see that

$$\begin{aligned} \langle \Delta _{\mathrm{id}_A,r}^\mathrm{bdp},\Delta _{\mathrm{id}_A,r}^\mathrm{bdp}\rangle _{X_r}=D^{r-1}\cdot \langle \Pi _r(\Delta _{\mathrm{id}_A,r}^\mathrm{bdp}),\Pi _r(\Delta _{\mathrm{id}_A,r}^\mathrm{bdp})\rangle _{W_r}. \end{aligned}$$
(3.5)

Combining (3.4) and (3.5), the result follows. \(\square \)

4 Big logarithm map

Let \({\mathbf{{f}}}=\sum _{n>0}{\mathbf{{a}}}_nq^n\in \mathbb{I }[[q]]\) be a Hida family passing through (the ordinary \(p\)-stabilization of) a \(p\)-ordinary newform \(f_o\in S_k(X_0(N))\) as described in the Introduction. We begin this section by recalling the definition of a certain twist of \({\mathbf{{f}}}\) such that all of its specializations at arithmetic primes of even weight correspond to \(p\)-adic modular forms with trivial weight-nebentypus.

Decompose the \(p\)-adic cyclotomic character \(\varepsilon _\mathrm{cyc}\) as the product

$$\begin{aligned} \varepsilon _\mathrm{cyc}=\omega \cdot \epsilon :G_{{\mathbf{{Q}}}}\rightarrow {\mathbf{{Z}}}_p^\times =\varvec{\mu }_{p-1}\times \Gamma . \end{aligned}$$

Since \(k\) is even, the character \(\omega ^{k-2}\) admits a square root \(\omega ^{\frac{k-2}{2}}:G_{\mathbf{{Q}}}\rightarrow \varvec{\mu }_{p-1}\), and in fact two different square roots, corresponding to the two different lifts of \(k-2\in {\mathbf{{Z}}}/(p-1){\mathbf{{Z}}}\) to \({\mathbf{{Z}}}/2(p-1){\mathbf{{Z}}}\). Fix for now a choice of \(\omega ^{\frac{k-2}{2}}\), and define the critical character to be

$$\begin{aligned} \Theta :=\omega ^{\frac{k-2}{2}}\cdot [\epsilon ^{1/2}]:G_{\mathbf{{Q}}}\rightarrow \Lambda _\mathcal{O }^\times , \end{aligned}$$
(4.1)

where \(\epsilon ^{1/2}:G_{\mathbf{{Q}}}\rightarrow \Gamma \) denotes the unique square root of \(\epsilon \) taking values in \(\Gamma \).

Remark 4.1

As noted in [19, Rem. 2.1.4], the above choice of \(\Theta \) is for most purposes largely indistinguishable from the other choice, namely \(\omega ^{\frac{p-1}{2}}\Theta \), where

$$\begin{aligned} \omega ^{\frac{p-1}{2}}:\mathrm{Gal}({\mathbf{{Q}}}(\sqrt{p^*})/{\mathbf{{Q}}})\xrightarrow {\sim }\{\pm {1}\}\quad \quad (p^*=(-1)^{\frac{p-1}{2}}p). \end{aligned}$$

Nonetheless, for a given \(f_o\) as above, our main result (Theorem 5.11) will specifically apply to only one of the two possible choices for the critical character.

The critical twist of \(\mathbb{T }\) is then defined to be the module

$$\begin{aligned} \mathbb{T }^\dagger :=\mathbb{T }\otimes _\mathbb{I }\mathbb{I }^\dagger \end{aligned}$$
(4.2)

equipped with the diagonal \(G_{\mathbf{{Q}}}\)-action, where \(\mathbb{I }^\dagger =\mathbb{I }(\Theta ^{-1})\) is \(\mathbb{I }\) as a module over itself with \(G_{\mathbf{{Q}}}\) acting via the character \(G_{\mathbf{{Q}}}\xrightarrow {\Theta ^{-1}}\Lambda _\mathcal{O }^\times \rightarrow \mathbb{I }^\times .\)

Lemma 4.2

Let \(\rho _{\mathbb{T }^\dagger }:G_{\mathbf{{Q}}}\rightarrow \mathrm{Aut}(\mathbb{T }^\dagger )\) be the Galois representation carried by \(\mathbb{T }^\dagger \). Then for every \(\nu \in \mathcal{X }_\mathrm{arith}(\mathbb{I })\) of even weight \(k_\nu =2r_\nu \ge 2\) we have

$$\begin{aligned} \nu (\rho _{\mathbb{T }^\dagger })\cong \rho _{{\mathbf{{f}}}_\nu ^{\prime }}\otimes \varepsilon _\mathrm{cyc}^{r_\nu }, \end{aligned}$$

where \({\mathbf{{f}}}_\nu ^{\prime }\) is a character twist of \({\mathbf{{f}}}_\nu \) of the same weight with trivial nebentypus. In other words, defining \(\mathbb V _\nu ^\dagger :=\mathbb{T }^\dagger \otimes _{\mathbb{I }}F_\nu \) and letting \(V_{{\mathbf{{f}}}_\nu ^{\prime }}\) be the representation space of \(\rho _{{\mathbf{{f}}}_\nu ^{\prime }}\), we have

$$\begin{aligned} \mathbb V _\nu ^\dagger \cong V_{{\mathbf{{f}}}_\nu ^{\prime }}(r_\nu ), \end{aligned}$$
(4.3)

and in particular \(\mathbb V _\nu ^\dagger \) is isomorphic to its Kummer dual.

Proof

This follows from a straightforward computation explained in [29, (3.5.2)] for example (where \(\mathbb{T }^\dagger \) is denoted by \(T\)). \(\square \)

Let \(\theta :{\mathbf{{Z}}}_p^\times \rightarrow \Lambda _\mathcal{O }^\times \) be such that \(\Theta =\theta \circ \varepsilon _\mathrm{cyc}\). It follows from the preceding lemma that the formal \(q\)-expansion

$$\begin{aligned} {\mathbf{{f}}}^\dagger ={\mathbf{{f}}}\otimes \theta ^{-1} :=\sum _{n>0}\theta ^{-1}(n){\mathbf{{a}}}_nq^n\in \mathbb{I }[[q]] \end{aligned}$$

(where we put \(\theta ^{-1}(n)=0\) whenever \(p\vert n\)) is such that, for every \(\nu \in \mathcal X _\mathrm{arith}(\mathbb I )\) of even weight, \(\mathbb V _\nu ^\dagger \) is the Galois representation attached to the specialization \({\mathbf{{f}}}_\nu \otimes \theta _\nu ^{-1}\) of \({\mathbf{{f}}}^\dagger \), which by Lemma 2.3 is a \(p\)-adic modular form of weight 0 and trivial nebentypus.

We next recall some of the local properties of the big Galois representation \(\mathbb{T }\). Let \(I_w\subset D_w\subset G_{\mathbf{{Q}}}\) be the inertia and decomposition groups at the place \(w\vert p\) induced by our fixed embedding \(\imath _p:\overline{{\mathbf{{Q}}}}\hookrightarrow \overline{{\mathbf{{Q}}}}_p\). In the following we will identify \(D_w\) with the absolute Galois group \(G_{{\mathbf{{Q}}}_p}\). Then by a result of Mazur and Wiles (see [34, Thm. 2.2.2]) there exists a filtration of \(\mathbb{I }[D_w]\)-modules

$$\begin{aligned} 0\rightarrow {\fancyscript{F}}_w^+\mathbb{T }\rightarrow \mathbb{T }\rightarrow {\fancyscript{F}}_w^-\mathbb{T }\rightarrow 0 \end{aligned}$$
(4.4)

with \({\fancyscript{F}}_w^{\pm }\mathbb{T }\) free of rank one over \(\mathbb{I }\) and with the Galois action on \({\fancyscript{F}}_w^-\mathbb{T }\) unramified, given by the character \(\alpha :D_w/I_w\rightarrow \mathbb{I }^\times \) sending an arithmetic Frobenius \(\sigma _p\) to \({\mathbf{{a}}}_p\). Twisting (4.4) by \(\Theta ^{-1}\) we define \({\fancyscript{F}}_w^\pm \mathbb{T }^\dagger \) in the natural manner.

Let \(\mathbb{T }^*:=\mathrm{Hom}_{\mathbb{I }}(\mathbb{T },\mathbb{I })\) be the contragredientFootnote 5 of \(\mathbb{T }\), and consider the \(\mathbb{I }\)-module

$$\begin{aligned} \mathbb{D }:=({\fancyscript{F}}_w^+\mathbb{T }^*\widehat{\otimes }_{{\mathbf{{Z}}}_p} \widehat{{\mathbf{{Z}}}}_p^\mathrm{nr})^{G_{{\mathbf{{Q}}}_p}}, \end{aligned}$$
(4.5)

where \({\fancyscript{F}}_w^+\mathbb{T }^*:=\mathrm{Hom}_\mathbb{I }({\fancyscript{F}}^-\mathbb{T },\mathbb{I })\subset \mathbb{T }^*\), and \(\widehat{{\mathbf{{Z}}}}_p^\mathrm{nr}\) is the completion of the ring of integers of the maximal unramified extension of \({\mathbf{{Q}}}_p\) in \(\overline{{\mathbf{{Q}}}}_p\).

Fix once and for all a compatible system \(\zeta =\{\zeta _s\}\) of primitive \(p^s\)-th roots of unity, and denote by \(e_\zeta \) the basis of \(D_\mathrm{dR}({\mathbf{{Q}}}_p(1))\) corresponding to \(1\in {\mathbf{{Q}}}_p\) under the resulting identification \(D_\mathrm{dR}({\mathbf{{Q}}}_p(1))\cong {\mathbf{{Q}}}_p\).

Lemma 4.3

The module \(\mathbb{D }\) is free of rank one over \(\mathbb{I }\), and for every \(\nu \in \mathcal{X }_\mathrm{arith}(\mathbb{I })\) of even weight \(k_\nu =2r_\nu \ge 2\) there is a canonical isomorphism

$$\begin{aligned} \mathbb{D }_\nu \otimes D_\mathrm{dR}({\mathbf{{Q}}}_p(r_\nu )) \cong \frac{D_\mathrm{dR}(V_{{\mathbf{{f}}}_\nu }(r_\nu ))}{\mathrm{Fil}^0D_\mathrm{dR}(V_{{\mathbf{{f}}}_\nu }(r_\nu ))}. \end{aligned}$$
(4.6)

Proof

Since the action on \({\fancyscript{F}}_w^+\mathbb{T }^*\) is unramified, the first claim follows from [30, Lemma 3.3] in light of the definition (4.5) of \(\mathbb{D }\). The second can be deduced from [30, Lemma 3.2] as in the proof of [30, Lemma 3.6]. \(\square \)

With the same notations as in Lemma 4.3, we denote by \(\langle ,\rangle _\mathrm{dR}\) the pairing

$$\begin{aligned} \langle ,\rangle _\mathrm{dR}:\mathbb{D }_\nu \otimes D_\mathrm{dR}({\mathbf{{Q}}}_p(r_\nu )) \times \mathrm{Fil}^1D_\mathrm{dR}(V_{{\mathbf{{f}}}_\nu ^*}(r_\nu -1))\rightarrow F_\nu \end{aligned}$$
(4.7)

deduced from the usual de Rham pairing

$$\begin{aligned} \frac{D_\mathrm{dR}(V_{{\mathbf{{f}}}_\nu }(r_\nu ))}{\mathrm{Fil}^0D_\mathrm{dR}(V_{{\mathbf{{f}}}_\nu }(r_\nu ))} \times \mathrm{Fil}^0D_\mathrm{dR}(V_{{\mathbf{{f}}}_\nu }^*(1-r_\nu ))\rightarrow F_\nu \end{aligned}$$

via the identification (4.6) and the isomorphism \(V_{{\mathbf{{f}}}_\nu }^*\cong V_{{\mathbf{{f}}}_\nu ^*}(k_\nu -1)\).

Theorem 4.4

(Ochiai) Assume that the residual representation \(\bar{\rho }_{f_o}\) is irreducible, fix an \(\mathbb{I }\)-basis \(\mathbf{\eta }\) of \(\mathbb{D }\), and set \(\lambda :={\mathbf{{a}}}_p-1\). There exists an \(\mathbb{I }\)-linear map

$$\begin{aligned} \mathrm{Log}_{\mathbb{T }^\dagger }^{(\eta )}:H^1({\mathbf{{Q}}}_p,{\fancyscript{F}}_w^+\mathbb{T }^\dagger ) \rightarrow \mathbb{I }[\lambda ^{-1}] \end{aligned}$$

such that if \(\mathfrak{Y }\in H^1({\mathbf{{Q}}}_p,{\fancyscript{F}}_w^+\mathbb{T }^\dagger )\) and \(\nu \in \mathcal{X }_\mathrm{arith}(\mathbb{I })\) has weight \(k_\nu =2r_\nu \ge 2\), then

$$\begin{aligned} \nu (\mathrm{Log}^{(\eta )}_{\mathbb{T }^\dagger }(\mathfrak{Y }))&= \frac{(-1)^{r_\nu -1}}{(r_\nu -1)!}\nonumber \\&\!\times \!\left\{ \begin{array}{l@{\quad }l} \left(1-\frac{p^{r_\nu -1}}{\nu ({\mathbf{{a}}}_p)}\right)^{-1} \left(1-\frac{\nu ({\mathbf{{a}}}_p)}{p^{r_\nu }}\right) \langle \mathrm{log}_{V_{{\mathbf{{f}}}_\nu }(r_\nu )} (\mathfrak{Y }_\nu ),\eta _\nu ^{\prime }\rangle _\mathrm{dR}&\mathrm{if}\,\vartheta _\nu =\mathbb 1 ;\\ \frac{1}{G(\vartheta _\nu ^{-1})} \left(\frac{\nu ({\mathbf{{a}}}_p)}{p^{r_\nu -1}}\right)^{s_\nu } \langle \mathrm{log}_{s,V_{{\mathbf{{f}}}_\nu }(r_\nu )}(\mathfrak{Y }_\nu ), \eta _\nu ^{\prime }\rangle _\mathrm{dR}&\mathrm{if} \,\vartheta _\nu \ne \mathbb 1 , \end{array}\right.\nonumber \\ \end{aligned}$$
(4.8)

where

  • \(\mathrm{log}_{V_{{\mathbf{{f}}}_\nu }(r_\nu )}\) (resp. \(\mathrm{log}_{s,V_{{\mathbf{{f}}}_\nu }(r_\nu )}\)) is the Bloch-Kato logarithm map for \(V_{{\mathbf{{f}}}_\nu }(r_\nu )\) over \({\mathbf{{Q}}}_p\) (resp. \({\mathbf{{Q}}}_{p,s}:={\mathbf{{Q}}}_p(\varvec{\mu }_{p^s})\)),

  • \(\eta _\nu ^{\prime }\in \mathrm{Fil}^1D_\mathrm{dR}(V_{{\mathbf{{f}}}_\nu ^*}(r_\nu -1))\) is such that \(\langle \eta _\nu \otimes e_\zeta ^{\otimes r_\nu },\eta _\nu ^{\prime }\rangle _\mathrm{dR}=1\),

  • \(\vartheta _\nu :{\mathbf{{Z}}}_p^\times \rightarrow F_\nu ^\times \) is the finite order character \(z\mapsto \theta _\nu (z)z^{1-r_\nu }\),

  • \(s>0\) is such that the conductor of \(\vartheta _\nu \) is \(p^{s}\), and

  • \(G(\vartheta _\nu ^{-1})\) is the Gauss sum \(\sum _{x\;\mathrm{mod}\;p^{s}} \vartheta _\nu ^{-1}(x)\zeta _{s}^{x}\).

Proof

Let \(\Lambda _\mathrm{cyc}={\mathbf{{Z}}}_p[[\Gamma _\mathrm{cyc}]]\) be the cyclotomic Iwasawa algebra, where

$$\begin{aligned} \Gamma _\mathrm{cyc}:=\mathrm{Gal}(\mathbf Q _{p,\infty }/\mathbf Q _p)\cong \mathbf Z _p, \end{aligned}$$

and consider the \(\mathbb{I }\widehat{\otimes }_{{\mathbf{{Z}}}_p}\Lambda _\mathrm{cyc}\)-modules \(\mathcal{D }:=\mathbb{D }\widehat{\otimes }_{{\mathbf{{Z}}}_p}\Lambda _\mathrm{cyc}\) and \({\fancyscript{F}}_w^+\mathcal{T }^*:={\fancyscript{F}}_w^+\mathbb{T }^*\widehat{\otimes }_{{\mathbf{{Z}}}_p} \Lambda _\mathrm{cyc}\otimes \omega ^{\frac{k-2}{2}}\), the latter being equipped with the diagonal action of \(G_{{\mathbf{{Q}}}_p}\). Also let \(\gamma _o\) be a topological generator of \(\Gamma _\mathrm{cyc}\) and \(\mathcal I :=(\lambda ,\gamma _o)\subset \mathbb I \widehat{\otimes }_\mathbf{Z _p}\Lambda _\mathrm{cyc}\cong \mathbb I [[\Gamma _\mathrm{cyc}]]\). Consider the \(\mathbb I \)-algebra isomorphism

$$\begin{aligned} \mathrm{Tw}_{\theta _1}:\mathbb I [[\Gamma _\mathrm{cyc}]]\rightarrow \mathbb I [[\Gamma _\mathrm{cyc}]] \end{aligned}$$
(4.9)

given by \(\mathrm{Tw}_{\theta _1}([\sigma ])=\epsilon ^{1/2}(\sigma )[\sigma ]\) for \(\sigma \in \Gamma _\mathrm{cyc}\), where \(\epsilon ^{1/2}\) is the unique square-root of the wild component of the cyclotomic character. By [30, Prop. 5.3] there exists an injective \(\mathbb{I }\widehat{\otimes }_{{\mathbf{{Z}}}_p}\Lambda _\mathrm{cyc}\)-linear map

$$\begin{aligned} \mathrm{Exp}_{{\fancyscript{F}}_w^+{\mathcal{T }^*}} :\mathcal ID \rightarrow H^1({\mathbf{{Q}}}_p,{\fancyscript{F}}_w^+{\mathcal{T }^*}) \end{aligned}$$

with cokernel killed by \(\mathcal{I }\) which interpolates the Bloch-Kato exponential over the arithmetic primes of \(\mathbb I \) and of \(\Lambda _\mathrm{cyc}\). Notice that letting \(\fancyscript{F}_w^+\mathcal T ^\dagger \) be the module \(\fancyscript{F}_w^+\mathcal T ^*\) with the \(\mathbb I [[\Gamma _\mathrm{cyc}]]\)-action twisted by \(\theta _1\), there is a Galois equivariant projection \(\fancyscript{F}_w^+\mathcal T ^\dagger \rightarrow \fancyscript{F}_w^+\mathbb T ^\dagger \). The composition

$$\begin{aligned} \mathcal{ID }\xrightarrow {\mathrm{Exp}_{{\fancyscript{F}}_{w}^{+}\mathcal{T }^{*}}}H^{1}(\mathbf{Q }_{p},{\fancyscript{F}}_{w}^{+}\mathcal{T }^{*})\xrightarrow {\mathrm{Tw}_{}} H^{1}(\mathbf{Q }_{p},{\fancyscript{F}}_{w}^{+}\mathcal{T }^{\dagger })\xrightarrow {\mathrm{Cor}} H^{1}(\mathbf{Q }_{p},{\fancyscript{F}}_{w}^{+}\mathbb{T }^{\dagger })\nonumber \\ \end{aligned}$$
(4.10)

is an \(\mathbb I \)-linear map making for every \(\nu \in \mathcal{X }_\mathrm{arith}(\mathbb{I })\) as in the statement the diagram

figure a2

commutative, where \(\mathrm{Sp}_{\nu ,\zeta }\) is given by the composition of (4.6) with the map

$$\begin{aligned} \mathcal{D }=\mathbb{D }\widehat{\otimes }_{{\mathbf{{Z}}}_p}\Lambda _\mathrm{cyc} \rightarrow \mathbb{D }_\nu \otimes D_\mathrm{dR}({\mathbf{{Q}}}_p(r_\nu ))\otimes {\mathbf{{Q}}}_{p,s} \end{aligned}$$

induced by specialization at \(\nu \) on \(\mathbb{D }\) and \(\sigma \mapsto e_{\zeta }^{\otimes r_\nu }\otimes \zeta _s^\sigma \) (\(\sigma \in \Gamma _\mathrm{cyc}\)) on \(\Lambda _\mathrm{cyc}\), and where the bottom horizontal arrow is given by:

$$\begin{aligned} (-1)^{r_\nu -1}(r_\nu -1)!\times \left\{ \begin{array}{l@{\quad }l} \left(1-\frac{p^{r_\nu -1}}{\nu ({\mathbf{{a}}}_p)}\right) \left(1-\frac{\nu ({\mathbf{{a}}}_p)}{p^{r_\nu }}\right)^{-1}\mathrm{exp}_{V_{{\mathbf{{f}}}_\nu }(r_\nu )}&\mathrm{if}\, \vartheta _\nu =\mathbb 1 ;\\ G(\vartheta _\nu ^{-1})\left(\frac{p^{r_\nu -1}}{\nu ({\mathbf{{a}}}_p)}\right)^{s_\nu } \mathrm{exp}_{s,V_{{\mathbf{{f}}}_\nu }(r_\nu )}&\mathrm{if}\, \vartheta _\nu \ne \mathbb 1 \end{array}\right. \end{aligned}$$

with \(\mathrm{exp}_{V_{{\mathbf{{f}}}_\nu }(r_\nu )}\) (resp. \(\mathrm{exp}_{s,V_{{\mathbf{{f}}}_\nu }(r_\nu )}\)) the Bloch-Kato exponential map for \(V_{{\mathbf{{f}}}_\nu }(r_\nu )\) over \({\mathbf{{Q}}}_p\) (resp. \({\mathbf{{Q}}}_{p,s}\)). The map \(\mathrm{Exp}_{\fancyscript{F}_w^+\mathbb T ^\dagger }\) factors through an injective \(\mathbb I \)-linear map

$$\begin{aligned} \mathrm{Exp}_{\fancyscript{F}_w^+\mathbb T ^\dagger }:\mathbb D ^\dagger \rightarrow H^1(\mathbf Q _p,\fancyscript{F}_w^+\mathbb T ^\dagger ), \end{aligned}$$

where \(\mathbb D ^\dagger :=\mathcal ID \otimes _\mathbf{Z _p}\mathbb I [[\Gamma _\mathrm{cyc}]]/(\gamma _o^2-\gamma _o^{\prime })\) with \(\gamma _o^{\prime }\) a topological generator of \(\Gamma \). (Recall for the Introduction that \(\Gamma \) acts on \(\mathbb I \) via the diamond operators.)

Now if \(\mathfrak{Y }\in H^1({\mathbf{{Q}}}_p,{\fancyscript{F}}_w^+\mathbb{T }^\dagger )\), then \(\lambda \cdot \mathfrak{Y }\) lands in the image \(\mathrm{Exp}_{{\fancyscript{F}}_w^+\mathbb{T }^\dagger }\) and so

$$\begin{aligned} \mathrm{Log}_{\mathbb{T }^\dagger }(\mathfrak{Y }):=\lambda ^{-1} \cdot (\mathrm{Exp}_{{\fancyscript{F}}_w^+\mathbb{T }^\dagger })^{-1}(\lambda \cdot \mathfrak{Y }) \in \mathbb{I }[\lambda ^{-1}]\otimes _\mathbb{I }\mathbb{D }^\dagger \end{aligned}$$

is well-defined. Thus defining \(\mathrm{Log}^{(\eta )}_{\mathbb{T }^\dagger }(\mathfrak{Y })\in \mathbb{I }[\lambda ^{-1}]\) by the relation

$$\begin{aligned} \mathrm{Log}_{\mathbb{T }^\dagger }(\mathfrak{Y })=\mathrm{Log}^{(\eta )}_{\mathbb{T }^\dagger }(\mathfrak{Y })\cdot \eta \otimes 1 \end{aligned}$$

the result follows. \(\square \)

5 The big Heegner point

In this chapter we prove the main results of this paper, relating the étale Abel-Jacobi images of Heegner cycles to the specializations at higher even weights of the big Heegner point \(\mathfrak{Z }\) (whose definition is recalled below), from where a deformation of the \(p\)-adic Gross-Zagier formula of Nekovář over a Hida family follows at once. There are two key points to the proof: the properties of the big logarithm map deduced from the work of Ochiai as explained in the preceding section, and the local study of (almost all) the weight 2 specializations of \(\mathfrak{Z }\) taken up in the following.

5.1 Weight two specializations

Recall form Sect. 3 that \(K\) is a fixed imaginary quadratic field in which all prime factors of \(N\) split, and that \(\mathfrak{N }\subset \mathcal{O }_K\) is a fixed cyclic \(N\)-ideal, i.e. such that \(\mathcal{O }_K/\mathfrak{N }\cong {\mathbf{{Z}}}/N{\mathbf{{Z}}}\). We also assume that \(p\) splits in \(K\), and let \(\mathfrak p \) be the prime of \(K\) above \(p\) induced by our fixed embedding \(\imath _p\), and by \(\bar{\mathfrak{p }}\) the other. Finally, \(A\) is a fixed elliptic curve with CM by \(\mathcal{O }_K\) defined over the Hilbert class field \(H\) of \(K\).

Let \(R_0=\widehat{{\mathbf{{Z}}}}_p^\mathrm{nr}\) be the completion of the ring of integers of the maximal unramified extension of \(\mathbf Q _p\), which we view as an overfield of \(H\) via \(\imath _p\). Since \(p\) splits in \(K\), \(A\) admits a trivialization

$$\begin{aligned} \imath _A:\hat{A}\rightarrow \hat{\mathbf{G}}_m \end{aligned}$$

over \(R_0\) with \(\imath _A^{-1}(\varvec{\mu }_{p^s})=A[\mathfrak p ^s]\) for every \(s>0\). Letting \(\alpha _A\) be the cyclic \(N\)-isogeny on \(A\) with kernel \(A[\mathfrak{N }]\), the triple \((A,\alpha _A,\imath _A)\) thus defines a trivialized elliptic curve with \(\Gamma _0(N)\)-level structure defined over \(R_0\).

Set \(A_0:=A/A[\mathfrak p ^s]\) and let \((A_0,\alpha _{A_0},\imath _{A_0})\) be the trivialized elliptic curve deduced from \((A,\alpha _A,\imath _A)\) via the projection \(A\rightarrow A_0\). Let \(C\subset A_0[p^s]\) be any étale subgroup of order \(p^s\), and set \(A_s:=A_0/C\). Finally, let \((A_s,\alpha _{A_s},\imath _{A_s})\) be the trivialized elliptic curve with \(\Gamma _0(N)\)-level structure deduced from \((A_0,\alpha _{A_0},\imath _{A_0})\) via the projection \(A_0\rightarrow A_s\), and consider the triple

$$\begin{aligned} h_s=(A_s,\alpha _{A_s},\imath _{A_s}(\zeta _s)), \end{aligned}$$
(5.1)

which defines an algebraic point on the modular curve \(X_s\).

Write \(p^*=(-1)^{\frac{p-1}{2}}p\), and let \(\vartheta \) be the unique continuous character

$$\begin{aligned} \vartheta :G_{{\mathbf{{Q}}}(\sqrt{p^*})}\rightarrow {\mathbf{{Z}}}_p^\times /\{\pm {1}\} \end{aligned}$$
(5.2)

such that \(\vartheta ^2=\varepsilon _\mathrm{cyc}\). Notice the inclusion \(G_{H_{p^s}}\subset G_{{\mathbf{{Q}}}(\sqrt{p^*})}\) for any \(s>0\), where \(H_{p^s}\) denotes the ring class field of \(K\) of conductor \(p^s\).

Lemma 5.1

The curve \(A_s\) has CM by the order \(\mathcal{O }_{p^s}\) of \(K\) of conductor \(p^s\), and the point \(h_s\) is rational over \(L_{p^s}:=H_{p^s}(\varvec{\mu }_{p^s})\). In fact we have

$$\begin{aligned} h_s^\sigma =\langle \vartheta (\sigma )\rangle \cdot h_s\end{aligned}$$
(5.3)

for all \(\sigma \in \mathrm{Gal}(L_{p^s}/H_{p^s})\).

Proof

The first assertion is clear, and immediately from the construction we also see that \(\alpha _{A_s}\) is the cyclic \(N\)-isogeny on \(A_s\) with kernel \(A_s[\mathfrak{N }\cap \mathcal{O }_{p^s}]\). It follows that the point (5.1) gives rise to precisely the point \(h_s\in X_s({\mathbf{{C}}})\) in [19, Eq. (4)]. The result thus follows from [loc.cit., Cor. 2.2.2]. \(\square \)

If \(\nu \) is an arithmetic prime of \(\mathbb{I }\), we let \(\psi _\nu \) denote its wild character, defined as the composition of \(\nu :\mathbb{I }\rightarrow \overline{{\mathbf{{Q}}}}_p\) with the structure map \(\Gamma =1+p{\mathbf{{Z}}}_p\rightarrow \mathbb{I }^\times \). The nebentypus of \({\mathbf{{f}}}_\nu \) is then given by \(\varepsilon _{{\mathbf{{f}}}_\nu }=\psi _\nu \omega ^{k-k_\nu }\), where \(\omega :({\mathbf{{Z}}}/p{\mathbf{{Z}}})^\times \rightarrow \varvec{\mu }_{p-1} \subset {\mathbf{{Z}}}_p^\times \) is the Teichmüller character.

Recall the critical characters \(\Theta \) and \(\theta \) from Sect. 4, and for every \(\nu \in \mathcal{X }_\mathrm{arith}(\mathbb{I })\) of weight 2, consider the \(F_\nu ^\times \)-valued Hecke character of \(K\) given by

$$\begin{aligned} \chi _\nu (x)=\Theta _\nu (\mathrm{art}_{\mathbf{{Q}}}(\mathrm{N}_{K/{\mathbf{{Q}}}}(x))) \end{aligned}$$
(5.4)

for all \(x\in \mathbb{A }_K^\times \). Notice that since \(\chi _\nu \) has finite order, it may alternately be seen as character on \(G_K\) via the Artin reciprocity map \(\mathrm{art}_K:\mathbb{A }_K^\times \rightarrow G_K^\mathrm{ab}\).

Let \(\mathcal{O }_{{\mathbf{{C}}}_p}\) be the ring of integers of the completion of \(\overline{{\mathbf{{Q}}}}_p\). For every \(\nu \in \mathcal{X }_\mathrm{arith}(\mathbb{I })\), after fixing an embedding \(F_\nu \rightarrow \overline{{\mathbf{{Q}}}}_p\), the form \({\mathbf{{f}}}_\nu \in S_{k_\nu }(X_{s_\nu })\) defines a \(p\)-adic modular form \({\mathbf{{f}}}_\nu \in \mathbf{M}(N)\). Finally, recall the dual form \({\mathbf{{f}}}_\nu ^*\) defined as in the paragraph before (2.20).

Lemma 5.2

Let \(\nu \in \mathcal{X }_\mathrm{arith}(\mathbb{I })\) have weight 2 and non-trivial wild character, and let \(s>1\) be the \(p\)-power of the conductor of \(\psi _\nu \). Then

$$\begin{aligned} d^{-1}{\mathbf{{f}}}_\nu ^{*[p]}\otimes \theta _\nu (A,\alpha _A,\imath _A) \!=\!\frac{u}{G(\theta _\nu ^{-1})}\sum _{\sigma \in \mathrm{Gal}(H_{p^s}/H)}\chi _\nu ^{-1}(\tilde{\sigma }) \cdot d^{-1}{\mathbf{{f}}}_\nu ^{*[p]}(h_s^{\tilde{\sigma }}),\qquad \end{aligned}$$
(5.5)

where \(u=\vert \mathcal{O }_K^\times \vert /2\), \(G(\theta _\nu ^{-1})\) is the Gauss sum \(\sum _{x\;\mathrm{mod}\;p^s}\theta _\nu ^{-1}(x)\zeta _s^{x}\), and for every \(\sigma \in \mathrm{Gal}(H_{p^s}/H)\), \(\tilde{\sigma }\) is any lift of \(\sigma \) to \(\mathrm{Gal}(L_{p^s}/H)\).

Proof

Notice that the expression in the right hand side of (5.5) does not depend on the choice of lifts \(\tilde{\sigma }\). Indeed, as explained in [18, p. 808] the character \(\chi _{0,\nu }:=\chi _\nu \vert _{\mathbb{A }_{\mathbf{{Q}}}^\times }\), seen as a Dirichlet character in the usual manner, is such that \(\chi _{0,\nu }^{-1}=\theta _\nu ^2\). But since the weight of \(\nu \) is 2, we have \(\theta _\nu ^2=\varepsilon _{{\mathbf{{f}}}_\nu }=\varepsilon _{{\mathbf{{f}}}_\nu ^*}^{-1}\) (see [18, p. 806]), and our claim thus follows immediately from (5.3).

To compute the above value of the twist \(d^{-1}{\mathbf{{f}}}_\nu ^{*[p]}\otimes \theta _\nu ^{}\) we follow Definition 2.2. The integer \(s>1\) in the statement is such that \(\theta _\nu \) factors through \(({\mathbf{{Z}}}/p^s{\mathbf{{Z}}})^\times \), therefore

$$\begin{aligned} d^{-1}{\mathbf{{f}}}_\nu ^{*[p]}\!\otimes \!\theta _\nu (A,\alpha _A,\imath _A)\;=\;\sum _{a\;\mathrm{mod}\;p^s}\theta _\nu ^{}(a)\left(\int _{a+p^s{\mathbf{{Z}}}_p}d\mu _\mathrm{Gou}(x)\!\right)(d^{-1}{\mathbf{{f}}}_\nu ^{*[p]})(\!A,\alpha _A,\imath _A)\nonumber \\ \;=\;\frac{1}{p^s}\!\!\sum _{a\;\mathrm{mod}\;p^s}\theta _\nu ^{}(a)\!\!\sum _C\!\zeta _C^{-a}\!\cdot d^{-1}{\mathbf{{f}}}_\nu ^{*[p]}(A_0/C,\alpha _C,\imath _C),\qquad \quad \end{aligned}$$
(5.6)

where as before \(A_0:=A/\imath _A^{-1}(\varvec{\mu }_{p^s})=A/A[\mathfrak p ^s]\) and the sum is over the étale subgroups \(C\subset A_0[p^s]\) of order \(p^s\). Letting \(\gamma _s\) be a generator of \({\mathbf{{Z}}}/p^s{\mathbf{{Z}}}\), these subgroups correspond bijectively with the cyclic subgroups \(C_u=\langle \zeta _s^u.\gamma _s\rangle \subset \varvec{\mu }_{p^s}\times {\mathbf{{Z}}}/p^s{\mathbf{{Z}}}\), with \(u\) running over the integers modulo \(p^s\), and we set \(\zeta _{C_u}=\zeta _s^u\).

Since \(\theta _\nu \) does not factor through \(({\mathbf{{Z}}}/p^{s-1}{\mathbf{{Z}}})^\times \), we have \(\sum _{a\;\mathrm{mod}\;p^s}\theta _\nu (a)\zeta _s^{-ua}=0\) whenever \(u\notin ({\mathbf{{Z}}}/p^s{\mathbf{{Z}}})^\times \). Continuing from (5.6), we thus obtain

$$\begin{aligned} d^{-1}{\mathbf{{f}}}_\nu ^{*[p]}\!\otimes \!\theta _\nu ^{}(\!A,\alpha _A,\imath _A\!)\;=\;\frac{1}{p^s}\!\sum _{a\;\mathrm{mod}\;p^s}\theta _\nu ^{}(a)\!\sum _{u\;\mathrm{mod}\;p^s}\zeta _s^{-ua}\cdot d^{-1}{\mathbf{{f}}}_\nu ^{*[p]}(A_{C_u},\alpha _{C_u},\imath _{C_u})\nonumber \\ \;=\;\frac{1}{p^s}\!\sum _{u\in {({\mathbf{{Z}}}/p^s{\mathbf{{Z}}})^\times }} d^{-1}{\mathbf{{f}}}_\nu ^{*[p]}(A_{C_u},\alpha _{C_u},\imath _{C_u}) \!\sum _{a\;\mathrm{mod}\;p^s}\theta _\nu ^{}(a)\zeta _s^{-ua}\nonumber \\ \;=\;\frac{1}{G(\theta _\nu ^{-1})}\sum _{u\in {({\mathbf{{Z}}}/p^s{\mathbf{{Z}}})^\times }} \theta _\nu ^{-1}(u)\cdot d^{-1}{\mathbf{{f}}}_\nu ^{*[p]}(A_{C_u}, \alpha _{C_u},\imath _{C_u}), \end{aligned}$$

with the last equality obtained by a change of variables. The result thus follows from the relation

$$\begin{aligned} \!\sum _{u\in ({\mathbf{{Z}}}/p^s{\mathbf{{Z}}})^\times }\theta _\nu ^{-1}(u)\cdot d^{-1}{\mathbf{{f}}}_\nu ^{*[p]}(A_{C_u},\alpha _{C_u},\imath _{C_u}) \!=\!u\!\sum _{\sigma \in \mathrm{Gal}(H_{p^s}/H)}\chi _\nu ^{-1}(\!\tilde{\sigma }\!) \cdot d^{-1}{\mathbf{{f}}}_\nu ^{*[p]}(h_s^{\tilde{\sigma }}),\nonumber \end{aligned}$$

where \(u=\vert \mathcal{O }_K^\times \vert /2\), and for each \(\sigma \in \mathrm{Gal}(H_{p^s}/H)\), \(\tilde{\sigma }\in \mathrm{Gal}(L_{p^s}/H)\) lifts \(\sigma \). \(\square \)

Keeping the above notations, let \(\Delta _s\in J_s(L_{p^s})\) be the divisor class of \((h_s)-(\infty )\), and consider the element in \(J_s(L_{p^s})\otimes _{{\mathbf{{Z}}}}F_\nu \) given by

$$\begin{aligned} \widetilde{Q}_{\chi _\nu }:=\sum _{\sigma \in \mathrm{Gal}(H_{p^s}/H)}\Delta _s^{\tilde{\sigma }} \otimes \chi _\nu ^{-1}(\tilde{\sigma }), \end{aligned}$$
(5.7)

where for every \(\sigma \in \mathrm{Gal}(H_{p^s}/H)\), \(\tilde{\sigma }\) is any lift to \(\mathrm{Gal}(L_{p^s}/H)\).

Let \(F_{s}\) be the completion of \(\imath _p(L_{p^s})\), and consider the \(p\)-adic Abel-Jacobi map \(\delta _{\mathbf{f}_{\nu}, F_{s}}^{(p)}\) defined in (2.9) which we extend by \(F_\nu \)-linearity to a map

$$\begin{aligned} \delta _{{\mathbf{{f}}}_\nu ,F_s}^{(p)}:J_s(L_{p^s})\otimes _{{\mathbf{{Z}}}}F_\nu \longrightarrow (\mathrm{Fil}^1D_\mathrm{dR}(V_{{\mathbf{{f}}}_\nu ^*}))^\vee . \end{aligned}$$

Proposition 5.3

Let \(\nu \in \mathcal{X }_\mathrm{arith}(\mathbb{I })\) and \(s>1\) be as in Lemma 5.2. Then

$$\begin{aligned} \sum _{\sigma \in \mathrm{Gal}(H_{p^s}/H)}\chi _\nu ^{-1}(\tilde{\sigma }) \cdot d^{-1}{\mathbf{{f}}}_\nu ^{*[p]}(h_s^{\tilde{\sigma }})= \delta _{{\mathbf{{f}}}_\nu ,F_s}^{(p)}(\widetilde{Q}_{\chi _\nu })(\omega _{{\mathbf{{f}}}_\nu ^*}). \end{aligned}$$
(5.8)

Proof

The integer \(s>1\) in the statement is so that the nebentypus \(\varepsilon _{{\mathbf{{f}}}_\nu }\) of \({\mathbf{{f}}}_\nu \) is primitive modulo \(p^s\). Moreover, since \(p\) splits in \(K\), we see from the construction that the point \(h_s\) lies in the connected component \(X_s(0)\) of the ordinary locus of \(X_s\) containing the cusp \(\infty \). Thus Proposition 2.9 applies, giving

$$\begin{aligned} \delta ^{(p)}_{{\mathbf{{f}}}_\nu ,F_s}(\Delta _s)(\omega _{{\mathbf{{f}}}_\nu ^{*}}) =F_{\omega _{{\mathbf{{f}}}_\nu ^*}}(h_s), \end{aligned}$$

where \(F_{\omega _{{\mathbf{{f}}}_\nu ^*}}\) is the Coleman primitive of \(\omega _{{\mathbf{{f}}}_\nu ^*}\) from Proposition 2.6 vanishing at \(\infty \), and by linearity

$$\begin{aligned} \sum _{\sigma \in \mathrm{Gal}(H_{p^s}/H)}\chi _\nu ^{-1}(\tilde{\sigma }) \cdot F_{\omega _{{\mathbf{{f}}}_\nu }}(h_s^{\tilde{\sigma }})= \delta _{{\mathbf{{f}}}_\nu ,F_s}^{(p)}(\widetilde{Q}_{\chi _\nu })(\omega _{{\mathbf{{f}}}_\nu ^*}). \end{aligned}$$
(5.9)

Since \(\phi \) lifts the Deligne–Tate map to \(X_s\), we see that \(\phi h_s\) is defined over the subfield \(H_{p^{s-1}}(\zeta _s)\subset L_{p^s}\). If \(b_p\) denotes the \(U_p\)-eigenvalue of \({\mathbf{{f}}}_\nu ^*\), by Corollary 2.8 we obtain

$$\begin{aligned} \sum _{\sigma }\chi _\nu ^{-1}(\tilde{\sigma })\cdot d^{-1}{\mathbf{{f}}}_\nu ^{*[p]}(h_s^{\tilde{\sigma }})&= \sum _{\sigma }\chi _\nu ^{-1}(\tilde{\sigma })\cdot F_{\omega _{{\mathbf{{f}}}_\nu ^*}}(h_s^{\tilde{\sigma }}) -\frac{b_p}{p}\sum _{\sigma }\chi _\nu ^{-1}(\tilde{\sigma })\cdot F_{\omega _{{\mathbf{{f}}}_\nu ^*}}(\phi h_s^{\tilde{\sigma }})\nonumber \\&= \sum _{\sigma }\chi _\nu ^{-1}(\tilde{\sigma })\cdot F_{\omega _{{\mathbf{{f}}}_\nu ^*}}(h_s^{\tilde{\sigma }}), \end{aligned}$$

where all the sums are over \(\sigma \in \mathrm{Gal}(H_{p^s}/H)\), and the second equality follows immediately from the fact \(\theta _\nu \) is primitive modulo \(p^s\). The result thus follows from (5.9). \(\square \)

Still with the same notations, recall Hida’s ordinary projector (1.2) and set \(y_s:=e^\mathrm{ord}h_s\), which naturally lies in \(e^\mathrm{ord}J_s(L_{p^s})\) (see [19, p.100]). Equation (5.3) then amounts to the fact that

$$\begin{aligned} y_s^\sigma =\Theta (\sigma )\cdot y_s\end{aligned}$$
(5.10)

for all \(\sigma \in \mathrm{Gal}(L_{p^s}/H_{p^s})\), where \(\Theta \) is the critical character (4.1). Denoting by \(J_s^\mathrm{ord}(L_{p^s})^\dagger \) the module \(e^\mathrm{ord}J_s(L_{p^s})\) with the Galois action twisted by \(\Theta ^{-1}\), and by \(y_s^\dagger \) the point \(y_s\) seen in this new module, (5.10) translates into the statement that

$$\begin{aligned} y_s^\dagger \in H^0(H_{p^s},J_s^\mathrm{ord}(L_{p^s})^\dagger ). \end{aligned}$$

Lemma 5.4

(Howard) The classes

$$\begin{aligned} x_s:=\mathrm{Cor}_{H_{p^s}/H}(y_s^\dagger )\in H^0(H,J_s^\mathrm{ord}(L_{p^s})^\dagger ) \end{aligned}$$
(5.11)

are such that

$$\begin{aligned} \alpha _*x_{s+1}=U_p\cdot x_s,\quad \quad \mathrm{for \, all}\, s>0 \end{aligned}$$

under the Albanese maps induced from the degeneracy maps \(\alpha :X_{s+1}\rightarrow X_s\).

Proof

This is shown in the course of the proof of [19, Lemma 2.2.4]. \(\square \)

Abbreviate by \(\mathrm{Ta}_p^\mathrm{ord}(J_s)\) the module \(e^\mathrm{ord}(\mathrm{Ta}_p(J_s)\otimes _{{\mathbf{{Z}}}_p}\mathcal{O })\) from the Introduction, and denote by \(\mathrm{Ta}_p^\mathrm{ord}(J_s)^\dagger \) this same module with the Galois action twisted by \(\Theta ^{-1}\). By the Galois and Hecke-equivariance of the twisted Kummer map

$$\begin{aligned} \mathrm{Kum}_s:H^0(H,J_s^\mathrm{ord}(L_{p^s})^\dagger )\rightarrow H^1(H,\mathrm{Ta}_p^\mathrm{ord}(J_s)^\dagger ) \end{aligned}$$

constructed in [19, p. 101], Lemma 5.4 implies that the cohomology classes \(\mathfrak{X }_s:=\mathrm{Kum}_s(x_s)\) are such that \(\alpha _*\mathfrak{X }_{s+1}=U_p\cdot \mathfrak{X }_s\), for all \(s>0\).

Definition 5.5

(Howard) The big Heegner point of conductor one is the cohomology class \(\mathfrak X \) given by the image of \({\mathop {\leftarrow s} \nolimits ^{\text{ lim}}}U_p^{-s}\cdot \mathfrak{X }_{s}\) under the natural map induced by the \(\mathfrak{h }^\mathrm{ord}[G_{{\mathbf{{Q}}}}]\)-linear projection \({\mathop {\leftarrow s} \nolimits ^{\text{ lim}}}(\mathrm{Ta}_p^\mathrm{ord}(J_s)^\dagger )\rightarrow \mathbb{T }^\dagger \).

Our object of study is in fact

$$\begin{aligned} \mathfrak{Z }:=\mathrm{Cor}_{H/K}(\mathfrak{X }), \end{aligned}$$
(5.12)

which [19, Conj. 3.4.1] predicts to be not \(\mathbb{I }\)-torsion. For \(\nu \in \mathcal{X }_\mathrm{arith}(\mathbb{I })\) of weight 2, let \(L(s,{\mathbf{{f}}}_\nu ,\chi _\nu )\) be the Rankin-Selberg convolution \(L\)-function of [20, §1]. In the spirit of the classical Gross-Zagier theorem, one has the following criterion.

Theorem 5.6

(Howard) If \(\nu \in \mathcal{X }_\mathrm{arith}(\mathbb{I })\) has weight 2 and non-trivial nebentypus, then

$$\begin{aligned} \mathfrak{Z }_\nu \ne 0\quad \Longleftrightarrow \quad L^{\prime }(1,{\mathbf{{f}}}_{\nu },\chi _{\nu })\ne 0, \end{aligned}$$
(5.13)

and if the non-vanishing holds for at least one such \(\nu \), then \(\mathfrak{Z }\) is not \(\mathbb{I }\)-torsion.

Proof

See [18, Prop. 3] for the equivalence (5.13), and [loc.cit, Cor. 5] for the last implication. We outline the proof for future reference. For every \(\nu \in \mathcal{X }_\mathrm{arith}(\mathbb{I })\) of weight 2 and non-trivial nebentypus, consider (with the same notations as above)

$$\begin{aligned} Q_{\chi _\nu }:=\sum _{\tau \in \mathrm{Gal}(L_{p^s}/K)}\Delta _s^{\tau }\otimes \chi _\nu ^{-1}(\tau ) \in J_s(L_{p^s})\otimes _{{\mathbf{{Z}}}}\overline{{\mathbf{{Q}}}}. \end{aligned}$$
(5.14)

If \(e_{{\mathbf{{f}}}_\nu }\) denotes the idempotent of the Hecke algebra (tensored with \(\overline{{\mathbf{{Q}}}}\)) defined by the eigenform \({\mathbf{{f}}}_\nu \), the arguments in [18, pp. 809–810] show that

$$\begin{aligned} \mathfrak{Z }_\nu \ne 0\quad \Longleftrightarrow \quad e_{{\mathbf{{f}}}_\nu }Q_{\chi _\nu }\ne 0, \end{aligned}$$
(5.15)

and by the “twisted Gross-Zagier theorem” [20, Thm. 4.6.2], one has

$$\begin{aligned} e_{{\mathbf{{f}}}_\nu }Q_{\chi _\nu }\ne 0\quad \Longleftrightarrow \quad L^{\prime }(1,{\mathbf{{f}}}_\nu ,\chi _\nu )\ne 0. \end{aligned}$$

\(\square \)

Corollary 5.7

Assume that there is a \(\nu ^{\prime }\in \mathcal{X }_\mathrm{arith}(\mathbb{I })\) of weight 2 and non-trivial nebentypus such that \(L^{\prime }(1,{\mathbf{{f}}}_{\nu ^{\prime }},\chi _{\nu ^{\prime }})\ne 0\). Then the localization map

$$\begin{aligned} \mathrm{loc}_\mathfrak p :H_{f}^1(K,\mathbb V _\nu ^\dagger ) \rightarrow H^1_f({\mathbf{{Q}}}_p,\mathbb V _\nu ^\dagger ) \end{aligned}$$

is injective at all but finitely many \(\nu \in \mathcal{X }_\mathrm{arith}(\mathbb{I })\).

Proof

By [18, Cor. 5], the assumption implies that \(\mathfrak{Z }\) is nontorsion, and by [19, Cor. 3.4.3] that \(\widetilde{H}_f^1(K,\mathbb{T }^\dagger )\) has rank 1 over \(\mathbb{I }\). By [19, Lemma 2.1.7], it follows that

$$\begin{aligned} H_f^1(K,\mathbb V _\nu ^\dagger )=\mathfrak{Z }_\nu .F_\nu , \end{aligned}$$

for all but finitely many \(\nu \) of weight 2 and non-trival nebentypus. On the other hand, since \(\mathrm{dim}_{F_\nu }H_f^1({\mathbf{{Q}}}_p,\mathbb V _\nu ^\dagger )\) for every \(\nu \) of weight 2 with non-trivial nebentypus, we see that it suffices to show that one has the implication

$$\begin{aligned} \mathfrak{Z }_\nu \ne 0\quad \Longrightarrow \quad \mathrm{loc}_\mathfrak p (\mathfrak{Z }_\nu )\ne 0 \end{aligned}$$
(5.16)

for every \(\nu \in \mathcal{X }_\mathrm{arith}(\mathbb{I })\) of weight 2 and non-trivial nebentypus. (Indeed, (5.16) will show that \(\mathrm{loc}_\mathfrak p \) is injective at infinitely many \(\nu \), and by [19, Lemma 2.1.7] it will follow that the kernel of the localization map

$$\begin{aligned} \mathrm{loc}_\mathfrak p :\widetilde{H}_f^1(K,\mathbb{T }^\dagger )\rightarrow H^1({\mathbf{{Q}}}_p,\mathbb{T }^\dagger ) \end{aligned}$$

must be \(\mathbb{I }\)-torsion, hence contained in only finitely arithmetic primes).

The point \(Q_{\chi _\nu }\) (5.14) defines a \(K\)-rational point on a twist \(J_{\chi _\nu }\) of \(J_s\) by the character \(\chi _\nu ^{-1}\). Since the localization map

$$\begin{aligned} J_{\chi _\nu }(K)\otimes _{\mathbf{{Z}}}{\mathbf{{Q}}}\rightarrow J_{\chi _\nu }(K_\mathfrak p ) \end{aligned}$$

is injective, we thus see that (5.16) follows from (5.15), hence the result. \(\square \)

For any class \([\mathfrak a ]\in \mathrm{Pic}(\mathcal{O }_K)\), taking a representative \(\mathfrak a \subset \mathcal{O }_K\) prime to \(Np\), define

$$\begin{aligned} \mathfrak a *(A,\alpha _A,\imath _A):=(A_\mathfrak a ,\alpha _{A_\mathfrak a },\imath _{A_\mathfrak a }), \end{aligned}$$

where \(A_\mathfrak a =A/A[\mathfrak{N }]\), \(\alpha _{A_\mathfrak a }=A_\mathfrak a [\mathfrak{N }]\), and \(\imath _{A_\mathfrak a }\) is the trivialization \(\hat{A}_\mathfrak a \xrightarrow {\hat{\varphi }_\mathfrak a ^{-1}}\hat{A} \xrightarrow {\imath _{A}}\hat{\mathbf{G}}_m\) induced by the projection \(\varphi _\mathfrak a :A\rightarrow A_\mathfrak a \).

Theorem 5.8

Let \(\nu \in \mathcal{X }_\mathrm{arith}(\mathbb{I })\) have weight \(2\) and non-trivial wild character \(\psi _\nu \), and let \(s>1\) be the \(p\)-power of the conductor of \(\psi _\nu \). Then

$$\begin{aligned} \sum _{[\mathfrak a ]\in \mathrm{Pic}(\mathcal{O }_K)}d^{-1}{\mathbf{{f}}}_\nu ^{[p]} \otimes \theta _\nu ^{-1}(\mathfrak a *(A,\alpha _A,\imath _A)) \!=\!u\frac{\nu ({\mathbf{{a}}}_p)^s}{G(\theta _\nu ^{-1})}\mathrm{log}_{s,V_{{\mathbf{{f}}}_\nu }(1)}(\mathrm{loc}_\mathfrak p (\mathfrak{Z }_\nu ))(\omega _{{\mathbf{{f}}}_\nu ^*}),\nonumber \\ \end{aligned}$$
(5.17)

where \(u=\vert \mathcal{O }_K^\times \vert /2\), and \(G(\theta _\nu ^{-1})\) is the Gauss sum \(\sum _{x\;\mathrm{mod}\;p^s}\theta _\nu ^{-1}(x)\zeta _s^{x}\).

Proof

Since clearly \(d^{-1}{\mathbf{{f}}}_\nu ^{[p]}\otimes \theta _\nu ^{-1}=d^{-1}{\mathbf{{f}}}_\nu ^{*[p]} \otimes \theta _\nu \), letting \(F_s\) be the completion of \(\imath _p(L_{p^s})\) it suffices to establish the equality

$$\begin{aligned} d^{-1}{\mathbf{{f}}}_\nu ^{*[p]}\otimes \theta _\nu (A,\alpha _A,\imath _A) =u\frac{\nu ({\mathbf{{a}}}_p)^s}{G(\theta _\nu ^{-1})}\mathrm{log}_{F_s,V_{{\mathbf{{f}}}_\nu }(1)}(\mathrm{loc}_\mathfrak p (\mathfrak{X }_\nu ))(\omega _{{\mathbf{{f}}}_\nu ^*}). \end{aligned}$$
(5.18)

Combining the formulas from Lemma 5.2 and Proposition 5.3, we have

$$\begin{aligned} d^{-1}{\mathbf{{f}}}_\nu ^{*[p]}\otimes \theta _\nu (A,\alpha _A,\imath _A) =\frac{u}{G(\theta _\nu ^{-1})} \delta _{{\mathbf{{f}}}_\nu ,F_s}^{(p)}(\widetilde{Q}_{\chi _\nu }). \end{aligned}$$
(5.19)

Now the integer \(s>1\) is such that the natural map \(\mathbb{T }\rightarrow \mathbb V _\nu \) can be factored as

$$\begin{aligned} \mathbb{T }\rightarrow \mathrm{Ta}_p^\mathrm{ord}(J_s)\rightarrow \mathbb V _\nu , \end{aligned}$$
(5.20)

and we have \(\mathbb V _\nu ^\dagger \cong \mathbb V _\nu \) as \(G_{L_{p^s}}\)-modules. Tracing through the construction of \(\mathfrak{X }\), we see that the image of \(U_p^s\cdot \mathfrak{X }_\nu \) in \(H^1(L_{p^s},\mathbb V _\nu ^\dagger )\) agrees with the image of \(\widetilde{Q}_{\chi _\nu }\) under the composite map (where the unlabelled arrow is induced by (5.20))

$$\begin{aligned}&J_s(L_{p^s})\otimes _{{\mathbf{{Z}}}}F_\nu \xrightarrow {\mathrm{Kum}_s} H^1(L_{p^s},\mathrm{Ta}_p(J_s)\otimes _{{\mathbf{{Z}}}}F_\nu )\xrightarrow {e^\mathrm{ord}}H^1(L_{p^s},\mathrm{Ta}_p^\mathrm{ord}(J_s)\otimes _{{\mathbf{{Z}}}}F_\nu )\nonumber \\&\quad \longrightarrow H^1(L_{p^s},\mathbb V _\nu )\cong H^1(L_{p^s},\mathbb V _\nu ^\dagger ). \end{aligned}$$
(5.21)

Since \(U_p\) acts on \(\mathbb V _\nu ^\dagger \) as multiplication by \(\nu ({\mathbf{{a}}}_p)\), we thus arrive at the equality

$$\begin{aligned} \mathrm{Kum}_s(e^\mathrm{ord}\widetilde{Q}_{\chi _\nu })=\nu ({\mathbf{{a}}}_p)^s\cdot \mathrm{res}_{L_{p^s}/H}(\mathfrak{X }_\nu )\in H^1(L_{p^s},\mathbb V _\nu ). \end{aligned}$$
(5.22)

By [32, Prop. 1.6.8], this shows that the restriction to \(\mathrm{loc}_\mathfrak p (\mathfrak{X }_\nu )\) to \(G_{F_s}\) is contained in the Bloch-Kato finite subspace \(H_f^1(F_s,\mathbb V _\nu )\cong H_f^1(F_s,\mathbb V _\nu ^\dagger )\). Since the map \(\delta _{{\mathbf{{f}}}_\nu ,F_s}^{(p)}\) is defined by the commutativity of the diagram

figure a3

we thus see that (5.18) follows from (5.19) and (5.22). \(\square \)

Remark 5.9

The expression in the left hand side of (5.17) can be interpreted as the value of a certain \(p\)-adic Rankin \(L\)-series at a point outside the range of classical interpolation, and hence Theorem 5.8 may be seen as a \(p\)-adic analogue of the Gross-Zagier formula for the classes \(\mathfrak{Z }_\nu \), in the same spirit as the main result of [1]. This interpretation, which does not play a direct role in this paper, is studied further in the companion paper [5].

5.2 Higher weight specializations

Now we can prove our main result. Recall from the Introduction that \(f_o\) is a \(p\)-ordinary newform of level \(N\) prime to \(p\), even weight \(k\ge 2\) and trivial nebentypus, that

$$\begin{aligned} {\mathbf{{f}}}=\sum _{n>0}{\mathbf{{a}}}_nq^n\in \mathbb{I }[[q]] \end{aligned}$$

is the Hida family passing through the ordinary \(p\)-stabilization of \(f_o\), and that \(K\) is an imaginary quadratic field such that every prime factor of \(pN\) is split in \(K\).

If \({\mathbf{{f}}}_\nu \) is the ordinary \(p\)-stabilization of a \(p\)-ordinary newform \({\mathbf{{f}}}_\nu ^\sharp \) of even weight \(2r_\nu >2\) and trivial nebentypus, the Heegner cycle \(\Delta _{A,r_\nu }^\mathrm{heeg}\) has been defined in Sect. 3, and by [28, Thm. (3.1)(i)] the class

$$\begin{aligned} \Phi ^{{{{\acute{\mathrm{e}}}}\mathrm{t}}}_{{\mathbf{{f}}}_\nu ^\sharp ,K}(\Delta _{r_\nu }^\mathrm{heeg}) :=\mathrm{Cor}_{H/K}(\Phi ^{{{{\acute{\mathrm{e}}}}\mathrm{t}}}_{{\mathbf{{f}}}_\nu ^\sharp ,H}(\Delta _{A,r_\nu }^\mathrm{heeg})) \end{aligned}$$
(5.23)

lies in the Bloch-Kato Selmer group \(H_f^1(K,V_{{\mathbf{{f}}}_\nu ^\sharp }(r_\nu ))\).

On the other hand, by [19, Prop. 2.4.5], the big Heegner point \(\mathfrak{X }\) lies in the strict Greenberg Selmer group \(\mathrm{Sel}_\mathrm{Gr}(H,\mathbb{T }^\dagger )\) (defined in [loc.cit., Def. 2.4.2]), and since \(\mathrm{Sel}_\mathrm{Gr}(K,\mathbb V _\nu ^\dagger )\cong H_f^1(K,\mathbb V _\nu ^\dagger )\) as explained in [19, p. 114]) and \(\mathbb V _\nu ^\dagger \cong V_{{\mathbf{{f}}}_\nu ^\sharp }(r_\nu )\) by Lemma 4.2, the class

$$\begin{aligned} \mathfrak{Z }_\nu =\mathrm{Cor}_{H/K}(\mathfrak{X }_\nu ) \end{aligned}$$

naturally lies in \(H_f^1(K,V_{{\mathbf{{f}}}_\nu ^\sharp }(r_\nu ))\) as well. Our main result relates these two classes.

Recall that the following hypotheses are being assumed throughout this paper.

Assumption 5.10

The residual representation \(\bar{\rho }_{f_o}\) is irreducible, and the semi-simplification of \(\overline{\rho }_{f_o}\vert _{G_{{\mathbf{{Q}}}_p}}\) is non-scalar.

Theorem 5.11

Let \(\nu _o\) be the arithmetic prime of \(\mathbb{I }\) such that \({\mathbf{{f}}}_{\nu _o}\) is the ordinary \(p\)-stabilization of \(f_o\), and let \(\mathbb{T }^\dagger =\mathbb{T }\otimes \Theta ^{-1}\) be the critical twist of \(\mathbb{T }\) such that \(\vartheta _{\nu _o}\) is the trivial characterFootnote 6. Assume that there is a \(\nu ^{\prime }\in \mathcal{X }_\mathrm{arith}(\mathbb{I })\) of weight 2 and non-trivial nebentypus such that

$$\begin{aligned} L^{\prime }(1,{\mathbf{{f}}}_{\nu ^{\prime }},\chi _{\nu ^{\prime }})\ne 0. \end{aligned}$$
(5.24)

Then for all but finitely many arithmetic primes \(\nu \in \mathcal{X }_\mathrm{arith}(\mathbb{I })\) of weight \(2r_\nu >2\) with \(2r_{\nu }\equiv k\pmod {2(p-1)}\), we have

$$\begin{aligned} \langle \mathfrak{Z }_{\nu },\mathfrak{Z }_\nu \rangle _K =\left(1-\frac{p^{r_\nu -1}}{\nu ({\mathbf{{a}}}_p)}\right)^4 \frac{\langle \Phi _{{\mathbf{{f}}}_\nu ^\sharp ,K}^{{{{\acute{\mathrm{e}}}}\mathrm{t}}}(\Delta ^\mathrm{heeg}_{r_\nu }), \Phi _{{\mathbf{{f}}}_\nu ^\sharp ,K}^{{{{\acute{\mathrm{e}}}}\mathrm{t}}}(\Delta ^\mathrm{heeg}_{r_\nu })\rangle _K}{{u^2(4D)^{r_\nu -1}}}, \end{aligned}$$
(5.25)

where \(\langle ,\rangle _K\) is the cyclotomic \(p\)-adic height pairing on \(H_f^1(K,V_{{\mathbf{{f}}}_\nu ^\sharp }(r_\nu ))\), \(u=\vert \mathcal{O }_K^\times \vert /2\), and \(-D<0\) is the discriminant of \(K\).

Proof

Since \(\mathfrak{Z }\in \mathrm{Sel}_\mathrm{Gr}(K,\mathbb{T }^\dagger )\), the localization \(\mathrm{loc}_\mathfrak p (\mathfrak{Z })\) lies in the kernel of the natural map

$$\begin{aligned} H^1({\mathbf{{Q}}}_p,\mathbb{T }^\dagger )\rightarrow H^1({\mathbf{{Q}}}_p,{\fancyscript{F}}_w^-\mathbb{T }^\dagger ), \end{aligned}$$

and since \(H^0({\mathbf{{Q}}}_p,{\fancyscript{F}}_w^-\mathbb{T }^\dagger )=0\) by [19, Lemma 2.4.4], the class \(\mathrm{loc}_\mathfrak p (\mathfrak{Z })\) can be seen as sitting inside \(H^1({\mathbf{{Q}}}_p,{\fancyscript{F}}_w^+\mathbb{T }^\dagger )\). Thus upon taking an \(\mathbb{I }\)-basis \({\eta }\) of \(\mathbb{D }\), we can form

$$\begin{aligned} \mathcal{L }_\mathfrak p ^\mathrm{arith}({\mathbf{{f}}}^\dagger ):=u\cdot \mathrm{Log}_{\mathbb{T }^\dagger }^{({\eta })}(\mathrm{loc}_\mathfrak p (\mathfrak{Z }))\in \mathbb{I }[\lambda ^{-1}]\quad \quad (\lambda :={\mathbf{{a}}}_p-1). \end{aligned}$$

On the other hand, consider the continuous function on \(\mathrm{Spf}(\mathbb{I })(\overline{{\mathbf{{Q}}}}_p)\) given by

$$\begin{aligned} \mathcal{L }_\mathfrak p ^\mathrm{analy}({\mathbf{{f}}}^\dagger ):\nu \mapsto \sum _{[\mathfrak a ]\in \mathrm{Pic}(\mathcal{O }_K)}d^{-1}{\mathbf{{f}}}_\nu ^{[p]}\otimes \theta _\nu ^{-1} (\mathfrak a *(A,\alpha _A,\imath _A)). \end{aligned}$$

(Its continuity can be checked by staring at the \(q\)-expansion of \(d^{-1}{\mathbf{{f}}}_\nu ^{[p]}\otimes \theta _\nu ^{-1}\) and appealing to the results in [14, § I.3.5], for example.)

By the specialization property (4.8) of the map \(\mathrm{Log}_{\mathbb{T }^\dagger }^{({\eta })}\), we see that Theorem 5.8 can be reformulated as follows: For every \(\nu \in \mathcal{X }_\mathrm{arith}(\mathbb{I })\) of weight 2 and non-trivial wild character, there exists a unit \(\Omega _{\nu }^{(\eta )}\in \mathcal{O }_\nu ^\times \) such that

$$\begin{aligned} \nu \left(\mathcal{L }_\mathfrak p ^\mathrm{analy}({\mathbf{{f}}}^\dagger )\right)= \Omega _{\nu }^{(\eta )}\cdot \nu \left(\mathcal{L }_\mathfrak p ^\mathrm{arith}({\mathbf{{f}}}^\dagger )\right). \end{aligned}$$
(5.26)

In fact,

$$\begin{aligned} \Omega _{\nu }^{(\eta )}=\langle \eta _\nu \otimes {e_\zeta ^{\otimes r_\nu }},\omega _{{\mathbf{{f}}}_\nu ^*}\rangle _\mathrm{dR} \end{aligned}$$
(5.27)

under the pairing (4.7), so that \(\omega _{{\mathbf{{f}}}_\nu ^*}=\Omega _{\nu }^{(\eta )}\cdot \eta _\nu ^{\prime }\) with \(\eta _\nu ^{\prime }\) as defined in Theorem 4.4.Footnote 7 Since both \(\mathfrak{L }_\mathfrak p ^\mathrm{arith}({\mathbf{{f}}}^\dagger )\) and \(\mathfrak{L }_\mathfrak p ^\mathrm{analy}({\mathbf{{f}}}^\dagger )\) are continuous functions of \(\nu \), (5.26) shows that the map \(\nu \mapsto \Omega _{\nu }^{(\eta )}\) is continuous, and hence (5.27) is valid for all \(\nu \in \mathcal{X }_\mathrm{arith}(\mathbb{I })\).

Now let \(\nu \in \mathcal{X }_\mathrm{arith}(\mathbb{I })\) be as in the statement. Then \(\theta _{\nu }(z)=z^{r_\nu -1}\vartheta _{\nu }(z)=z^{r_\nu -1}\) as characters on \({\mathbf{{Z}}}_p^\times \), from where if follows that

$$\begin{aligned} \nu \left(\mathcal{L }_\mathfrak p ^\mathrm{analy}({\mathbf{{f}}}^\dagger )\right)&= \sum _{[\mathfrak a ]\in \mathrm{Pic}(\mathcal{O }_K)}d^{-1}{\mathbf{{f}}}_{\nu }^{[p]}\otimes \theta _{\nu }^{-1}(\mathfrak a *(A,\alpha _A,\imath _A))\nonumber \\&= \sum _{[\mathfrak a ]\in \mathrm{Pic}(\mathcal{O }_K)}d^{-r_\nu }{\mathbf{{f}}}_{\nu }^{[p]}(\mathfrak a *(A,A[\mathfrak{N }])). \end{aligned}$$

By Theorem 3.3, setting

$$\begin{aligned} \Delta _{r_\nu }^\mathrm{bdp}\!:=\!\mathrm{Norm}_{H/K}(\Delta _{\varphi _{(1)},r_\nu }^\mathrm{bdp}) \!=\!\sum _{[\mathfrak a ]\in \mathrm{Pic}(\mathcal{O }_K)}\mathrm{N}\mathfrak a ^{1-r}\cdot \Delta _{\varphi _\mathfrak a ,r_\nu }^\mathrm{bdp}\in \mathrm{CH}^{2r_\nu -1}(X_{r_\nu })_0(K),\nonumber \\ \end{aligned}$$
(5.28)

this shows that

$$\begin{aligned} \nu \left(\mathcal{L }_\mathfrak p ^\mathrm{analy}({\mathbf{{f}}}^\dagger )\right)&= \mathcal{E }_\nu (r_\nu )\mathcal{E }_\nu ^*(r_\nu ) \frac{(-1)^{r_\nu -1}}{(r_\nu -1)!} \mathrm{AJ}_{{\mathbf{{Q}}}_p}(\Delta _{r_\nu }^\mathrm{bdp})(\omega _{{\mathbf{{f}}}_\nu ^\sharp }\otimes e_\zeta ^{\otimes r_\nu -1})\nonumber \\&= \mathcal{E }_\nu (r_\nu )\mathcal{E }_\nu ^*(r_\nu ) \frac{(-1)^{r_\nu -1}}{(r_\nu -1)!}\mathrm{log}_\mathbb{V _\nu ^\dagger }(\mathrm{loc}_\mathfrak p (\Phi ^{{{{\acute{\mathrm{e}}}}\mathrm{t}}}_{{\mathbf{{f}}}_\nu ^\sharp ,K} (\Delta _{r_\nu }^\mathrm{bdp})))(\omega _{{\mathbf{{f}}}_\nu ^\sharp }\otimes e_\zeta ^{\otimes r_\nu -1}),\nonumber \\ \end{aligned}$$
(5.29)

where

$$\begin{aligned} \mathcal{E }_{\nu }(r_\nu ):=\left(1-\frac{p^{r_\nu -1}}{\nu ({\mathbf{{a}}}_p)}\right),\quad \quad \mathcal{E }_{\nu }^*(r_\nu ):=\left(1-\frac{\nu ({\mathbf{{a}}}_p)}{p^{r_\nu }}\right), \end{aligned}$$

and \(\Phi ^{{{{\acute{\mathrm{e}}}}\mathrm{t}}}_{{\mathbf{{f}}}_\nu ^\sharp ,K} :=\pi _{{\mathbf{{f}}}_\nu ^\sharp ,\mathrm{N}^{r_\nu -1}}\circ \Phi _{K}^{{{{\acute{\mathrm{e}}}}\mathrm{t}}}\) with notations as in the diagram (3.3) defining \(\mathrm{AJ}_{{\mathbf{{Q}}}_p}\).

On the other hand, by the specialization property of the map \(\mathrm{Log}_{\mathbb{T }^\dagger }^{({\eta })}\) we have

$$\begin{aligned} \nu \left(\mathcal{L }_\mathfrak p ^\mathrm{arith}({\mathbf{{f}}}^\dagger )\right)=u\frac{(-1)^{r_{\nu }-1}}{(r_{\nu }-1)!}\mathcal{E }_{\nu }(r_\nu )^{-1}\mathcal{E }_{\nu }^*(r_\nu ) \mathrm{log}_\mathbb{V _{\nu }^\dagger }(\mathrm{loc}_\mathfrak p (\mathfrak{Z }_{\nu }))(\eta _\nu ^{\prime }). \end{aligned}$$
(5.30)

Comparing (5.30) and (5.29), we thus conclude form (5.26) that

$$\begin{aligned} \mathrm{log}_\mathbb{V _{\nu }^\dagger }(\mathrm{loc}_\mathfrak p (\mathfrak{Z }_{\nu }))(\omega _{{\mathbf{{f}}}_{\nu }^\sharp } \!\otimes \! e_\zeta ^{\otimes r_\nu -1}) \!=\! \frac{1}{u}\mathcal{E }_{\nu }(r_\nu )^2\mathrm{log}_\mathbb{V _{\nu }^\dagger }(\mathrm{loc}_\mathfrak p (\Phi ^{{\acute{\mathrm{e}}}\mathrm{t}}_{{\mathbf{{f}}}_{\nu }^\sharp ,K}(\Delta _{r_\nu }^\mathrm{bdp})))(\omega _{{\mathbf{{f}}}_{\nu }^\sharp }\!\otimes \! e_\zeta ^{\otimes r_\nu -1}). \end{aligned}$$

Since \(\mathrm{Fil}^{1}D_\mathrm{dR}(V_{{\mathbf{{f}}}_\nu ^\sharp }(r_\nu -1))\) is spanned by \(\omega _{{\mathbf{{f}}}_{\nu }^\sharp }\otimes e_\zeta ^{\otimes r_\nu -1}\), it follows that

$$\begin{aligned} \mathrm{log}_\mathbb{V _{\nu }^\dagger }(\mathrm{loc}_\mathfrak p (\mathfrak{Z }_{\nu }))=\frac{1}{u}\mathcal{E }_{\nu }(r_\nu )^2\mathrm{log}_\mathbb{V _{\nu }^\dagger }(\mathrm{loc}_\mathfrak p (\Phi ^{{\acute{\mathrm{e}}}\mathrm{t}}_{{\mathbf{{f}}}_{\nu }^\sharp ,K}(\Delta _{r_\nu }^\mathrm{bdp}))), \end{aligned}$$

and since \(\mathrm{log}_\mathbb{V _{\nu }^\dagger }\) is an isomorphism, that

$$\begin{aligned} \text{ loc}_\mathfrak p (\mathfrak{Z }_{\nu })=\frac{1}{u}\mathcal{E }_{\nu }(r_\nu )^2\mathrm{loc}_\mathfrak p (\Phi ^{{\acute{\mathrm{e}}}\mathrm{t}}_{{\mathbf{{f}}}_{\nu }^\sharp ,K}(\Delta _{r_\nu }^\mathrm{bdp})). \end{aligned}$$
(5.31)

Our nonvanishing assumption (5.24) implies on the one hand, by Theorem 5.6, that \(\mathfrak{Z }_\nu \) is non-zero for all but finitely many \(\nu \in \mathcal{X }_\mathrm{arith}(\mathbb{I })\), and on the other hand, by Corollary 5.7, that the localization map \(\mathrm{loc}_\mathfrak p \) is injective for all but finitely many \(\nu \in \mathcal{X }_\mathrm{arith}(\mathbb{I })\). In particular, we thus see from (5.31) that we have

$$\begin{aligned} \langle \mathfrak{Z }_{\nu },\mathfrak{Z }_\nu \rangle _K&= \frac{1}{u^2}\mathcal{E }_{r_\nu }(r_\nu )^4 \langle \Phi _{{\mathbf{{f}}}_\nu ^\sharp ,K}^{{{{\acute{\mathrm{e}}}}\mathrm{t}}}(\Delta ^\mathrm{bdp}_{r_\nu }), \Phi _{{\mathbf{{f}}}_\nu ^\sharp ,K}^{{{{\acute{\mathrm{e}}}}\mathrm{t}}}(\Delta ^\mathrm{bdp}_{r_\nu })\rangle _K\\&= \frac{1}{u^2}\mathcal{E }_{r_\nu }(r_\nu )^4 \frac{\langle \Phi _{{\mathbf{{f}}}_\nu ^\sharp ,K}^{{{{\acute{\mathrm{e}}}}\mathrm{t}}}(\Delta ^\mathrm{heeg}_{r_\nu }), \Phi _{{\mathbf{{f}}}_\nu ^\sharp ,K}^{{{{\acute{\mathrm{e}}}}\mathrm{t}}}(\Delta ^\mathrm{heeg}_{r_\nu })\rangle _K}{{(4D)^{r_\nu -1}}} \end{aligned}$$

for all but finitely many \(\nu \) as in the statement, where the last equality follows from Lemma 3.4 in light of the definitions (5.23) and (5.28). The result follows. \(\square \)