1 Introduction

Let A be an elliptic curve defined over \(\mathbf{Q} \), having ordinary reduction at a rational prime \(p>3\). Let \(\varrho _{1}\) and \(\varrho _{2}\) be odd, irreducible, two-dimensional Artin representations of the absolute Galois group of \(\mathbf{Q} \), which are unramified at p and satisfy the self-duality condition

$$\begin{aligned} \det (\varrho _{1})=\det (\varrho _{2})^{-1}. \end{aligned}$$

By modularity, the triple \((A,\varrho _{1},\varrho _{2})\) arises from a triple (fgh) of cuspidal p-ordinary newforms of weights \(w_{o}=(2,1,1)\). Let \(f_{\alpha }\) be the ordinary p-stabilisation of f, and fix p-stabilisations \(g_{\alpha }\) and \(h_{\alpha }\) of g and h respectively. Set \(\varrho =\varrho _{1}\otimes {}\varrho _{2}\). In the recent paper [6] we proposed a p-adic analogue of the Birch and Swinnerton-Dyer conjecture for the leading term at \(w_{o}\) of the 3-variable Garrett–Hida p-adic L-function \(L_{p}^{\alpha \alpha }(A,\varrho )=L_{p}(\varvec{f},\varvec{g}_{\alpha },\varvec{h}_{\alpha })\) associated with the triple \((\varvec{f},\varvec{g}_{\alpha },\varvec{h}_{\alpha })\) of Hida families specialising to \((f_{\alpha },g_{\alpha },h_{\alpha })\) at \(w_{o}\). In this article we verify our conjecture in the analytic rank-zero exceptional cases, viz. when the complex Garrett L-function \(L(A,\varrho ,s)=L(f\otimes {}g\otimes {}h,s)\) does not vanish at \(s=1\) and \(L_{p}^{\alpha \alpha }(A,\varrho )\) has an exceptional zero at \(w_{o}\) in the sense of Mazur–Tate–Teitelbaum (cf. Theorem 2.1 and Sect. 2.1 below). Moreover, when \(L(A,\varrho ,1)=0\) and \(L_{p}^{\alpha \alpha }(A,\varrho )\) has an exceptional zero, we propose a conjecture relating the value at \(w_{o}\) of the fourth partial derivative of \(L_{p}^{\alpha \alpha }(A,\varrho )\) along the \(\varvec{f}\)-direction to the p-adic logarithms of two global points on A rational over the number field cut out by \(\varrho \) (cf. Conjecture 2.3).

2 Setting and notations

Fix algebraic closures \({\bar{\mathbf{Q }}}\) and \({\bar{\mathbf{Q }}}_{p}\) of \(\mathbf{Q} \) and \(\mathbf{Q} _{p}\) respectively, and field embeddings and . With the notations of the Introduction, let

$$\begin{aligned} \xi =\sum _{n\geqslant {}1}a_{n}(\xi )\cdot {}q^{n}\in {}S_{u}(N_{\xi },\chi _{\xi })_{{\bar{\mathbf{Q }}}} \end{aligned}$$

denote one of the cuspidal newforms f, g and h. Here u and \(N_{\xi }\) are the weight and the conductor of \(\xi \) respectively, and \(S_{u}(N_{\xi },\chi _{\xi })_{F}\) is the space of cuspidal modular forms of level \(\Gamma _{1}(N_{\xi })\), weight u, character \(\chi _{\xi }\) and Fourier coefficients in the subfield F of \({\bar{\mathbf{Q }}}_{p}\). Fix a number field \(\mathbf{Q} (\varrho )\) containing for any \(\xi \) the Fourier coefficients \(a_{n}(\xi )\), as well as the roots \(\alpha _{\xi }\) and \(\beta _{\xi }\) of the pth Hecke polynomials \(P_{\xi ,p}=X^{2}-a_{p}(\xi )\cdot {}X+\chi _{\xi }(p)\cdot {}p\). Let \(V_{\varrho _{i}}\) be a two-dimensional \(\mathbf{Q} (\varrho )\)-vector space affording the representation \(\varrho _{i}\), and let \(K_{\varrho }\) be a Galois number field such that \(\varrho _{i}\) factors through \(\text {Gal}(K_{\varrho }/\mathbf{Q} )\). Set

$$\begin{aligned} V_{\varrho }=V_{\varrho _{1}}\otimes _\mathbf{Q (\varrho )}V_{\varrho _{2}}\ \ \text {and}\ \ V_{p}(A,\varrho )=V_{p}(A)\otimes _\mathbf{Q }V_{\varrho }, \end{aligned}$$

where is the p-adic Tate module of A with \(\mathbf{Q} _{p}\)-coefficients. Throughout this note we make the following

Assumption 1.1

  1. 1.

    (Self-duality) The characters \(\chi _{g}\) and \(\chi _{h}\) are inverse to each other.

  2. 2.

    (Local signs) The conductors \(N_{g}\) and \(N_{h}\) are coprime to \(p\cdot {}N_{f}\).

  3. 3.

    (Étaleness) The forms g and h are cuspidal, p-regular and do not have RM by a real quadratic field in which p splits.

The first condition is a reformulation of the self-duality condition mentioned in the Introduction, namely \(\det (\varrho _{1})=\text {det}(\varrho _{2})^{-1}\). Recall that the form \(\xi \) is p-regular if \(P_{\xi ,p}\) has distinct roots. Moreover, one says that a weight-one eigenform has RM (real multiplication) if it is the theta series associated with a ray class character of a real quadratic field. Assumption 1.1.3 is equivalent to require that \(V_{\varrho _{i}}\) is irreducible, not isomorphic to \(\text {Ind}^\mathbf{Q }_{K}\chi \) for a finite order character \(\chi : G_{K}{\mathop {\longrightarrow }\limits ^{}}\mathbf{Q} (\varrho )^{*}\) of a real quadratic field K in which p splits, and that an arithmetic Frobenius at p acts on \(V_{\varrho _{i}}\) with distinct eigenvalues. For \(\xi =g,h\), this assumption guarantees that the p-adic Coleman–Mazur–Buzzard eigencurve of tame level \(N_{\xi }\) is étale over the weight space at the points corresponding to the p-stabilisations of \(\xi \) (cf. [2]). It is used in [6] to construct the Garrett–Nekovář height which appears in the main result of this note. To explain the relevance of Assumptions 1.1.1 and 1.1.2, let \(\alpha _{f}\) be the unit root of \(P_{f,p}\) and fix roots \(\alpha _{g}\) and \(\alpha _{h}\) of \(P_{g,p}\) and \(P_{h,p}\) respectively. Fix a finite extension L of \(\mathbf{Q} _{p}\) containing \(\mathbf{Q} (\varrho )\) and the roots of unity of order \(\text {lcm}(N_{f},N_{g},N_{h})\). Let \(\xi \) be one of f, g and h, and let \(u_{o}\) be the weight of \(\xi \). According to the results of [2, 10, 18], there exists a unique Hida family

$$\begin{aligned} \varvec{\xi }_{\alpha }=\sum _{n\geqslant {}1}a_{n}(\varvec{\xi }_{\alpha })\cdot {}q^{n}\in {}\mathscr {O}_{\varvec{\xi }}[\![q]\!] \end{aligned}$$

which specialises at \(u_{o}\) to the p-stabilised newform

$$\begin{aligned} \xi _{\alpha }=\xi (q)-\frac{\chi _{\xi }(p)p^{u-1}}{\alpha _{\xi }}\cdot {}\xi (q^{p})\in {}S_{u_{o}}(p\cdot {}M_{\xi },\chi _{\xi })_{L}. \end{aligned}$$

Here \(M_{\xi }=N_{\xi }/p^{\text {ord}_{p}(N_{\xi })}\) is the tame level of \(\xi \) (so that \(M_{\xi }=N_{\xi }\) if \(\xi =g,h\)), and \(\mathscr {O}_{\varvec{\xi }}\) is the ring of bounded analytic functions on a (sufficiently small) connected open disc \(U_{\varvec{\xi }}\) in the p-adic weight space over L. For each classical weight u in \(U_{\varvec{\xi }}\cap {}\mathbf{Z} _{\geqslant {}3}\), the weight-u specialisation \(\varvec{\xi }_{\alpha ,u}=\sum _{n\geqslant {}1}a_{n}(\varvec{\xi }_{\alpha })(u)\cdot {}q^{n}\in {}L[\![q]\!]\) of \(\varvec{\xi }_{\alpha }\) is the q-expansion of the ordinary p-stabilisation of a newform \(\xi _{u}\) in \(S_{u}(M_{\xi },\chi _{\xi })_{L}\). Since f has a unique p-ordinary p-stabilisation \(f_{\alpha }\), we simply write \(\varvec{f}\) for \(\varvec{f}_{\!\alpha }\).

Assumption 1.1.1 guarantees that for each classical triple \(w=(k,l,m)\) in the set

$$\begin{aligned} \Sigma =U_{\!\varvec{f}}\times {}U_{\varvec{g}}\times {}U_{\varvec{h}}\cap {}\mathbf{Z} _{\geqslant {}1}^{3} \end{aligned}$$

the complex Garrett L-function \(L(f_{k}\otimes {}g_{l}\otimes {}h_{m},s)\) admits an analytic continuation to all of \(\mathbf{C} \) and satisfies a functional equation relating its values at s and \(k+l+m-2-s\), with root number \(\varepsilon (w)=\prod _{\ell \leqslant \infty }\varepsilon _{\ell }(w)\) equal to \(+1\) or to \(-1\). Assumption 1.1.2 implies that all the local signs \(\varepsilon _{\ell }(w)\) are equal to \(+1\) for every w in the f-unbalanced region \(\Sigma _{f}=\{w=(k,l,m)\in {}\Sigma : k\geqslant {}l+m\}\) (cf. [11]). Under these assumptions, [12] associates with \((\varvec{f},\varvec{g}_{\alpha },\varvec{h}_{\alpha })\) an analytic function

$$\begin{aligned} \mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )=\mathscr {L}_{p}(\varvec{f},\varvec{g}_{\alpha },\varvec{h}_{\alpha }) \end{aligned}$$

in the ring \(\mathscr {O}_{\!\varvec{f}\varvec{g}\varvec{h}}=\mathscr {O}_{\!\varvec{f}}{\hat{\otimes }}_{L}\mathscr {O}_{\varvec{g}}{\hat{\otimes }}_{L}\mathscr {O}_{\varvec{h}}\), whose square

$$\begin{aligned} L_{p}^{\alpha \alpha }(A,\varrho )=L_{p}(\varvec{f},\varvec{g}_{\alpha },\varvec{h}_{\alpha })=\mathscr {L}_{p}(\varvec{f},\varvec{g}_{\alpha },\varvec{h}_{\alpha })^{2} \end{aligned}$$

satisfies the following interpolation property. For each \(w=(k,l,m)\) in \(\Sigma _{f}\), the value of \(L_{p}^{\alpha \alpha }(A,\varrho )\) at w is an explicit non-zero complex multiple of

$$\begin{aligned} \left( 1-\frac{\beta _{k}\alpha _{l}\alpha _{m}}{p^{c_{w}}}\right) ^{2} \left( 1-\frac{\beta _{k}\beta _{l}\alpha _{m}}{p^{c_{w}}}\right) ^{2} \left( 1-\frac{\beta _{k}\alpha _{l}\beta _{m}}{p^{c_{w}}}\right) ^{2} \left( 1-\frac{\beta _{k}\beta _{l}\beta _{m}}{p^{c_{w}}}\right) ^{2}\cdot {}L(f_{k}\otimes {}g_{l}\otimes {}h_{m},c_{w}). \end{aligned}$$
(1)

Here \(c_{w}=\frac{k+l+m-2}{2}\), and for \(\varvec{\xi }=\varvec{f},\varvec{g}_{\alpha },\varvec{h}_{\alpha }\) one denotes by \(\alpha _{u}\) the unit root of \(P_{\xi _{u},p}\) and sets \(\beta _{u}\cdot {}\alpha _{u}=\chi _{\xi }^{\prime }(p)\cdot {}p^{u-1}\), where \(\chi _{\xi }^{\prime }\) is the prime-to-p part of \(\chi _{\xi }\) (so that \(\chi _{\xi }^{\prime }=\chi _{\xi }\) for \(\xi =g,h\), and \(\chi _{f}^{\prime }\) is the trivial character modulo \(M_{f}\)). We refer to Theorem A of loc. cit. for the precise interpolation formula. We call \(L_{p}^{\alpha \alpha }(A,\varrho )=L_{p}(\varvec{f},\varvec{g}_{\alpha },\varvec{h}_{\alpha })\) the Garrett–Hida p-adic L-function associated with \((A,\varrho )\) (or with \((\varvec{f},\varvec{g}_{\alpha },\varvec{h}_{\alpha })\)).

3 Exceptional zero formulae

The p-adic variant of the Birch and Swinnerton-Dyer conjecture formulated in [6] predicts that the leading term of \(L_{p}^{\alpha \alpha }(A,\varrho )\) at \(w_{o}=(2,1,1)\) is encoded by the discriminant of the Garrett– Nekovář height pairing

(2)

constructed in Section 2 of loco citato, where \(\mathscr {I}\) is the ideal of functions in \(\mathscr {O}_{\!\varvec{f}\varvec{g}\varvec{h}}\) which vanish at \(w_{o}\) and the p-extended Mordell–Weil group \(A^{\dag }(K_{\varrho })^{\varrho }\) is defined as follows. When A has good reduction at p, one sets \(A^{\dag }(K_{\varrho })^{\varrho }=A(K_{\varrho })^{\varrho }\), where \(A(K_{\varrho })^{\varrho }\) is a shorthand for the \(\text {Gal}(K_{\varrho }/\mathbf{Q} )\)-invariants of \(A(K_{\varrho })\otimes _\mathbf{Z }V_{\varrho }\). If A has multiplicative reduction at p, then \(\alpha _{f}=a_{p}(f)=\pm 1\) and the maximal p-unramified quotient \(V_{p}(A)^{-}\) of \(V_{p}(A)\) is a 1-dimensional \(\mathbf{Q} _{p}\)-vector space on which an arithmetic Frobenius acts as multiplication by \(\alpha _{f}\). Let \(q_{A}\) in \(p\mathbf{Z} _{p}\) be the p-adic Tate period of the base change \(A_\mathbf{Q _{p}}\) of A to \(\mathbf{Q} _{p}\) (cf. Chapter V of [15]), and let \(\mathbf{Q} _{p^{2}}\) be the quadratic unramified extension of \(\mathbf{Q} _{p}\). The Tate uniformisation yields a rigid analytic morphism

$$\begin{aligned} \wp _{\text {Tate}} : \mathbf{G} _{m,\mathbf{Q} _{p^{2}}}^{rig}{\mathop {\longrightarrow }\limits ^{}}A_\mathbf{Q _{p^{2}}} \end{aligned}$$

with kernel \(q_{A}^\mathbf{Z }\) and unique up to sign. Set

$$\begin{aligned} q(A)=p^{-}\big ((\wp _{\text {Tate}}(\!\!\root p^{n} \of {q_{A}}\,))_{n\geqslant {}1}\big )\in {}V_{p}(A)^{-}, \end{aligned}$$

where \(p^{-}\) denotes the projection \(V_{p}(A){\mathop {\longrightarrow }\limits ^{}}V_{p}(A)^{-}\) and \((\!\!\root p^{n} \of {q_{A}}\,)_{n\geqslant {}1}\) is any compatible system of \(p^{n}\)-th roots of \(q_{A}\), and define

$$\begin{aligned} A^{\dag }(K_{\varrho })^{\varrho }=A(K_{\varrho })^{\varrho }\oplus {} {\mathcal {Q}}_{p}(A,\varrho ) \end{aligned}$$

to be the direct sum of \(A(K_{\varrho })^{\varrho }\) and the \(\mathbf{Q} (\varrho )\)-submodule

$$\begin{aligned} {\mathcal {Q}}_{p}(A,\varrho )=H^{0}(\mathbf{Q} _{p},\mathbf{Q} (\varrho )\cdot {}q(A)\otimes _\mathbf{Q (\varrho )}V_{\varrho }) \end{aligned}$$

of \(H^{0}(\mathbf{Q} _{p},V_{p}(A)^{-}\otimes _\mathbf{Q }V_{\varrho })\). The Garrett–Nekovář height depends on the choice of suitably normalised \(G_\mathbf{Q }\)-equivariant embeddings

(3)

where \(V(\xi )=V(\varvec{\xi }_{\alpha })\otimes _{1}L\) (for \(\xi =g,h\)) is the weight-one specialisation of the big Galois representation \(V(\varvec{\xi }_{\alpha })\) associated with \(\varvec{\xi }_{\alpha }\). (We refer to Sect. 3.1 below for precise definitions.) More precisely, denote by V(f) the \(f_{\alpha }\)-isotypic component of the cohomology group , where \(X_{1}(N_{f},p)_{{\bar{\mathbf{Q }}}}\) is the base change to \({\bar{\mathbf{Q }}}\) of the compact modular curve \(X_{1}(N_{f},p)\) of level \(\Gamma _{1}(N_{f})\cap {}\Gamma _{0}(p)\) over \(\mathbf{Q} \), and set

$$\begin{aligned} V(f,g,h)=V(f)\otimes _\mathbf{Q _{p}}V(g)\otimes _{L}V(h). \end{aligned}$$

Section 2 of [6] constructs a canonical Garrett–Nekovář p-adic height pairing

(4)

on the naive extended Selmer group of V(fgh) over \(\mathbf{Q} \), defined as the direct sum of the Bloch–Kato Selmer group \(\text {Sel}(\mathbf{Q} ,V(f,g,h))\) of V(fgh) over \(\mathbf{Q} \) and the module \(H^{0}(\mathbf{Q} _{p},V(f,g,h)^{-})\) of \(G_\mathbf{Q _{p}}\)-invariants of the maximal p-unramified quotient \(V(f,g,h)^{-}\) of V(fgh). (The definition of is briefly recalled in Sect. 3.2.3 below.) Fix a modular parametrisation \(\wp _{\infty } : X_{1}(N_{f},p){\mathop {\longrightarrow }\limits ^{}}A\), under which one identifies V(f) and \(V_{p}(A)\). The embeddings \(\gamma _{g}\) and \(\gamma _{h}\) and the global Kummer map on \(A(K_{\varrho })\) then induce an embedding . The pairing (2) is defined to be composition of the canonical Garrett–Nekovář height and \(\gamma _{gh}^{\otimes {}2}\). The pairings (2) and (4) are skew-symmetric, and the discriminant of (2) in \((\mathscr {I}^{r^{\dag }(A,\varrho )}/\mathscr {I}^{r^{\dag }(A,\varrho )+1})/\mathbf{Q} (\varrho )^{*2}\), where \(r^{\dag }(A,\varrho )=\dim _\mathbf{Q (\varrho )}A^{\dag }(K_{\varrho })^{\varrho }\), is independent of the choice of \(\wp _{\infty }\), \(\gamma _{g}\) and \(\gamma _{h}\). We refer to [6] for more details.

If \(\xi \) denotes either g or h, then the restriction to \(G_\mathbf{Q _{p}}\) of the Artin representation \(V(\xi )\) is the direct sum of the submodules \(V(\xi )_{\alpha }\) and \(V(\xi )_{\beta }\) on which an arithmetic Frobenius acts as multiplication by \(\alpha _{\xi }\) and \(\beta _{\xi }\) respectively (cf. Assumption 1.1.3). The \(G_\mathbf{Q _{p}}\)-representation \(V(f,g,h)^{-}\) then decomposes as the direct sum of the subspaces

$$\begin{aligned} V(f)^{-}_{ij}=V(f)^{-}\otimes _\mathbf{Q _{p}}V(g)_{i}\otimes _{L}V(h)_{j}, \end{aligned}$$

where (ij) is a pair of elements of \(\{\alpha ,\beta \}\). If \(\xi \) denotes either g or h, Sect. 3.1.1 below recalls the definition of canonical weight-one differentials

$$\begin{aligned} \omega _{\xi _{\alpha }}\in {}(V(\xi )_{\alpha }\otimes _\mathbf{Q _{p}}\mathbf{Q} _{p}^{\text {nr}})^{G_\mathbf{Q _{p}}}\ \ \ \text {and}\ \ \ \eta _{\xi _{\alpha }}\in {}(V(\xi )_{\beta }\otimes _\mathbf{Q _{p}}\mathbf{Q} _{p}^{\text {nr}})^{G_\mathbf{Q _{p}}}, \end{aligned}$$
(5)

where \(\mathbf{Q} _{p}^{\text {nr}}\) is the maximal unramified extension of \(\mathbf{Q} _{p}\). If A is multiplicative at p, set

$$\begin{aligned} q(f)=\wp _{\infty }^{-1}(q(A))\in {}V(f)^{-}, \end{aligned}$$

where one denotes again by \(\wp _{\infty } : V(f)^{-}\simeq {}V_{p}(A)^{-}\) the isomorphism arising form the fixed modular parametrisation \(\wp _{\infty } : X_{1}(N_{f},p){\mathop {\longrightarrow }\limits ^{}}A\).

Under the running assumptions, the \(\mathbf{Q} (\varrho )\)-module \({\mathcal {Q}}_{p}(A,\varrho )\) (resp., the L-module \(H^{0}(\mathbf{Q} _{p},V(f,g,h)^{-})\)) is non-zero precisely A is multiplicative at p and

$$\begin{aligned} \alpha _{f}=\alpha _{g}\cdot {}\alpha _{h}\ \ \ \text {or}\ \ \ \alpha _{f}=\beta _{g}\cdot {}\alpha _{h}, \end{aligned}$$

in which case it has dimension 2 and one says that \((A,\varrho )\) is exceptional at p. More precisely, note that \(\alpha _{g}\not =\beta _{g}\) by Assumptions 1.1.3, hence only one of the previous identities can be satisfied. Moreover \(\alpha _{f}=\alpha _{g}\cdot {}\alpha _{h}\) (resp., \(\alpha _{f}=\beta _{g}\cdot {}\alpha _{h}\)) if and only if \(\alpha _{f}=\beta _{g}\cdot {}\beta _{h}\) (resp., \(\alpha _{f}=\alpha _{g}\cdot {}\beta _{h}\)) by Assumption 1.1.1. Fix an auxiliary integer \(m_{p}\) such that p splits (resp., is inert) in \(\mathbf{Q} \left[ \sqrt{m_{p}}\right] \) if \(\alpha _{f}=+1\) (resp., \(\alpha _{f}=-1\)), so that \(G_\mathbf{Q _{p}}\) acts trivially on \(\sqrt{m_{p}}\cdot {}q(f)\) in \(V(f)^{-}\otimes _\mathbf{Q _{p}}\mathbf{Q} _{p}^{\text {nr}}\). If \(\alpha _{f}=\alpha _{g}\cdot {}\alpha _{h}\), then \(G_\mathbf{Q _{p}}\) acts trivially on \(V(f)^{-}_{\alpha \alpha }\) and \(V(f)^{-}_{\beta \beta }\), hence the p-adic periods

$$\begin{aligned} q_{\alpha \alpha }=\sqrt{m_{p}}\cdot {}q(f)\otimes \omega _{g_{\alpha }}\otimes \omega _{h_{\alpha }}\ \ \ \text {and}\ \ \ q_{\beta \beta }=\sqrt{m_{p}}\cdot {}q(f) \otimes \eta _{g_{\alpha }}\otimes \eta _{h_{\alpha }} \end{aligned}$$

can naturally be viewed as elements of \(V(f)^{-}_{\alpha \alpha }\) and \(V(f)^{-}_{\beta \beta }\) respectively, which generate \(H^{0}(\mathbf{Q} _{p},V(f,g,h)^{-})\). Similarly, if \(\alpha _{f}=\beta _{g}\cdot {}\alpha _{h}\), then the periods

$$\begin{aligned} q_{\alpha \beta }=\sqrt{m_{p}}\cdot {}q(f)\otimes \omega _{g_{\alpha }}\otimes \eta _{g_{h}}\ \ \ \text {and}\ \ \ q_{\beta \alpha }=\sqrt{m_{p}}\cdot {}q(f)\otimes \eta _{g_{\alpha }}\otimes \omega _{h_{\alpha }} \end{aligned}$$

can naturally be viewed as generators of \(H^{0}(\mathbf{Q} _{p},V(f,g,h)^{-})\).

Equation (1) shows that the value of the square-root Garrett–Hida L-function \(\mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )\) at \(w_{o}\) is a non-zero multiple of

$$\begin{aligned} \left( 1-\frac{\alpha _{g}\alpha _{h}}{\alpha _{f}}\right) \left( 1-\frac{\beta _{g}\alpha _{h}}{\alpha _{f}}\right) \left( 1-\frac{\alpha _{g}\beta _{h}}{\alpha _{f}}\right) \left( 1-\frac{\beta _{g}\beta _{h}}{\alpha _{f}}\right) \cdot {}\sqrt{L(A,\varrho ,1)}, \end{aligned}$$

where \(L(A,\varrho ,s)=L(f\otimes {}g\otimes {}h,s)\). The previous discussion then shows that \((A,\varrho )\) is exceptional at p precisely if one of the Euler factors which appear in the previous expression is zero, id est if \(\mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )\) (or \(L_{p}^{\alpha \alpha }(A,\varrho )\)) has an exceptional zero in the sense of Mazur–Tate–Teitelbaum [13]. In this case Lemma 9.8 of [7] proves that the restriction \(\mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )|_{\mathsf {L}}\) of \(\mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )\) to the improving line \(\mathsf {L}\) defined by the equations \(\varvec{m}=1\) and \(\varvec{k}=l+1\) admits the factorisation

$$\begin{aligned} \mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )|_{\mathsf {L}}=\mathscr {E}_{f}\cdot {}\mathscr {E}_{g}\cdot {}\mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )^{\star } \end{aligned}$$

in the ring \({\mathcal {O}}(\mathsf {L})\) of analytic functions on \(\mathsf {L}\), where

$$\begin{aligned} \mathscr {E}_{f}=\left. 1-\frac{a_{p}(\varvec{f})}{a_{p}(\varvec{g}_{\alpha })\cdot {}a_{p}(\varvec{h}_{\alpha })}\right| _{\mathsf {L}}\ \ \text {and}\ \ \mathscr {E}_{g}=\left. 1-\chi _{h}(p)\cdot {}\frac{a_{p}(\varvec{g}_{\alpha })}{a_{p}(\varvec{f})\cdot {}a_{p}(\varvec{h}_{\alpha })}\right| _{\mathsf {L}}. \end{aligned}$$

Moreover, the value at \(w_{o}\) of the improved p-adic L-function \(\mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )^{\star }\) is an explicit algebraic number in \(\mathbf{Q} (\varrho )\), equal to zero precisely if \(L(A,\varrho ,s)\) vanishes at \(s=1\). We refer to the proof of Proposition 8.3 of [12] for details.

The following is the main result of this note.

Theorem 2.1

Assume that \((A,\varrho )\) is exceptional at p. Let \((q_{\flat },q_{\natural })\) denote either the pair \((q_{\alpha \alpha },q_{\beta \beta })\) or \((q_{\alpha \beta },q_{\beta \alpha })\), depending on whether \(\alpha _{f}=\alpha _{g}\cdot {}\alpha _{h}\) or \(\alpha _{f}=\beta _{g}\cdot {}\alpha _{h}\) respectively. Then the following equality holds in \(\mathscr {I}/\mathscr {I}^{2}\) up to sign.

Theorem 2.1 is proved in Sect. 4 below. More precisely, Sects. 3.3 and 3.4 below prove that the following equality holds in \(\mathscr {I}/\mathscr {I}^{2}\) up to sign:

(6)

where \(\varepsilon =+1\) if \(\alpha _{f}=\alpha _{g}\cdot {}\alpha _{h}\) and \(\varepsilon =-1\) if \(\alpha _{f}=\beta _{g}\cdot {}\beta _{h}\), and where

$$\begin{aligned} -\frac{1}{2}\cdot {}{\mathfrak {L}}_{\varvec{\xi }}^{\text {an}}=d\log a_{p}(\varvec{\xi })_{\varvec{u}=u_{o}} \end{aligned}$$
(7)

is the value at the centre \(u_{o}\) of \(U_{\varvec{\xi }}\) of the logarithmic derivative of the p-th Fourier coefficient of the Hida family \(\varvec{\xi }=\varvec{f},\varvec{g}_{\alpha },\varvec{h}_{\alpha }\). In Sect. 4 we then deduce Theorem  2.1 from Eq. (6) and the study carried out in [7, Section 9] of the linear term of \(\mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )\) at \(w_{o}\) in the exceptional case.

It should be possible to extend Theorem 2.1 (and Conjecture 2.3 below) to the case of p-new eigenforms of even weight \(k\geqslant {}2\) and trivial character (cf. Section 1.1 of [6]). We have not checked the details.

3.1 The rank-zero exceptional case of [6, Conjecture 1.1]

Assume in this section that \((A,\varrho )\) is exceptional at p, and that the Garrett complex L-function \(L(A,\varrho ,s)=L(f\otimes {}g\otimes {}h,s)\) does not vanish at \(s=1\):

$$\begin{aligned} L(A,\varrho ,1)\not =0. \end{aligned}$$

According to the main result of [8] (see also Theorem B of [3]), one has

$$\begin{aligned} A(K_{\varrho })^{\varrho }=0, \end{aligned}$$

hence \(A^{\dag }(K_{\varrho })^{\varrho }={\mathcal {Q}}_{p}(A,\varrho )\). The Garrett– Nekovář p-adic regulator \(R_{p}^{\alpha \alpha }(A,\varrho )\), viz. the discriminant of the p-adic height on \(A^{\dag }(K_{\varrho })^{\varrho }\), is then given by

in \((\mathscr {I}^{2}/\mathscr {I}^{3})/\mathbf{Q} (\varrho )^{*2}\), where \((q_{1},q_{2})\) is a \(\mathbf{Q} (\varrho )\)-basis of \({\mathcal {Q}}_{p}(A,\varrho )\).

Let be the \(G_\mathbf{Q }\)-equivariant embedding defined by the tensor product of the isomorphism \(V_{p}(A)^{-}\simeq {}V(f)^{-}\) induced by \(\wp _{\infty }\), \(\gamma _{g}\) and \(\gamma _{h}\) (cf. Eq. (3)). The normalisation imposed on the embeddings \(\gamma _{g}\) and \(\gamma _{h}\) (and described in Sect. 3.1.1 below) implies that the matrix M in \(\text {GL}_{2}(L)\) defined by the identity \((q_{\flat } \ q_{\natural })\cdot {}M=(\gamma _{gh}(q_{1})\ \gamma _{gh}(q_{2}))\) has determinant in \(\mathbf{Q} (\varrho )^{*}\). In light of the above discussion, Theorem  2.1 then proves the following corollary, which together with Eq. (6) establishes [6, Conjecture 1.1] in the present setting.

Corollary 2.2

If \(L(A,\varrho ,s)\) does not vanish at \(s=1\), then \(A^{\dag }(K_{\varrho })^{\varrho }={\mathcal {Q}}_{p}(A,\varrho )\) and the following equality holds in the quotient of \(\mathscr {I}^{2}/\mathscr {I}^{3}\) by the action of \(\mathbf{Q} (\varrho )^{*2}\).

$$\begin{aligned} L_{p}^{\alpha \alpha }(A,\varrho )\!\!\!\pmod {\mathscr {I}^{3}}=R_{p}^{\alpha \alpha }(A,\varrho ) \end{aligned}$$

3.2 Exceptional zeros and rational points (cf. [14])

Assume in this section that \((A,\varrho )\) is exceptional at p, and that the Garrett complex L-function \(L(A,\varrho ,s)\) vanishes at the central critical point \(s=1\):

$$\begin{aligned} L(A,\varrho ,1)=0. \end{aligned}$$

Set \(\{\flat ,\natural \}=\{\alpha \alpha ,\beta \beta \}\) of \(\{\flat ,\natural \}=\{\alpha \beta ,\beta \alpha \}\), depending on whether

$$\begin{aligned} \alpha _{f}=\alpha _{g}\cdot {}\alpha _{h} \hbox {or} \alpha _{f}=\beta _{g}\cdot {}\alpha _{h}. \end{aligned}$$

The p-adic L-function \(\mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )\) belongs to \(\mathscr {I}^{2}\) (cf. Theorem 2.1) and Conjecture 2.3 of [6] predicts that its image in \((\mathscr {I}^{2}/\mathscr {I}^{3})/\mathbf{Q} (\varrho )^{*}\) equals

for two rational points P and Q in \(A(K_{\varrho })^{\varrho }\). (Recall that the p-adic height is skew-symmetric, hence the previous expression is a square root of its discriminant on the \(\mathbf{Q} (\varrho )\)-submodule of \(A^{\dag }(K_{\varrho })^{\varrho }\) generated by \(q_{\flat }, q_{\natural }, P\) and Q.) One has

by Eq. (6). Moreover, Sect. 3.5 below proves that

(8)

for each Selmer class x in \(\text {Sel}(\mathbf{Q} ,V(f,g,h))\), where

$$\begin{aligned} \log _{\flat }=\langle \log _{p}(\cdot ),q_{\natural }\rangle _{fgh} : H^{1}_{\text {fin}}(\mathbf{Q} _{p},V(f,g,h)){\mathop {\longrightarrow }\limits ^{}}L. \end{aligned}$$

Here \(\log _{p} : H^{1}_{\text {fin}}(\mathbf{Q} _{p},V(f,g,h))\simeq {}D_{\text {dR}}(V(f,g,h))/\text {Fil}^{0}\) is the Bloch–Kato p-adic logarithm (cf. Lemma 9.1 of [7]), and \(\left<{}\cdot ,\cdot \right>_{fgh} : D_{\text {dR}}(V(f,g,h))^{\otimes {}2}{\mathop {\longrightarrow }\limits ^{}}L\) is the pairing induced by the natural Kummer duality \(\pi _{fgh} : V(f,g,h)^{\otimes 2}{\mathop {\longrightarrow }\limits ^{}}L(1)\) defined in Sect. 3.1.1 below (cf. Eq. (11)). We are then led to the following

Conjecture 2.3

Assume that \(A(K_{\varrho })^{\varrho }\) is a 2-dimensional \(\mathbf{Q} (\varrho )\)-vector space. Then for any \(\mathbf{Q} (\varrho )\)-basis (PQ) of \(A(K_{\varrho })^{\varrho }\), the equality

$$\begin{aligned} \frac{\partial ^{2}\mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )}{\partial \varvec{k}^{2}}(w_{o})=\log _{\flat }(P)\cdot {}\log _{\natural }(Q)- \log _{\natural }(P)\cdot {}\log _{\flat }(Q) \end{aligned}$$

holds in L up to multiplication by a non-zero scalar in \(\mathbf{Q} (\varrho )^{*}\).

As explained in [5], the main result of [1] can be used to prove cases of Conjecture 2.3 when g and h are theta series associated with certain ray class characters of the same imaginary quadratic field in which p is inert (and P and Q are Heegner points). By combining this with an extension of the height computations carried out in [16, 17], the article [4] proves instances of Conjecture 1.1 of [6] in this setting.

Remark 2.4

In light of the aforementioned results of [5], Rivero proposes in [14, Conjecture 4.5] a variant of Conjecture 2.3. He also asks (cf. Question 5.3 of [14]) if one can expect a similar description of \(\frac{\partial ^{2}\mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )}{\partial \varvec{k}^{2}}(w_{o})\) when A has good reduction at p. The previous discussion places Rivero’s conjecture within a conceptual framework and sheds some light on this question.

4 Height computations

Throughout the rest of this note we assume that \((A,\varrho )\) is exceptional at p. In particular A has multiplicative reduction at p, id est p divides exactly \(N_{f}\).

4.1 Setting and notations

This subsection briefly recalls the needed definitions and notations from our previous articles [6, 7].

4.1.1 Galois representations

Set \(N=\text {lcm}(N_{f},N_{g},N_{h})\) and let \(G_\mathbf{Q ,N}\) be the Galois group of the maximal extension of \(\mathbf{Q} \) contained in \({\bar{\mathbf{Q }}}\) and unramified outside \(N\infty \). If \(\varvec{\xi }\) denotes one of \(\varvec{f},\varvec{g}_{\alpha }\) and \(\varvec{h}_{\alpha }\), let \(V(\varvec{\xi })\) be the big Galois representation associated with \(\varvec{\xi }\) (cf. Section 5 of [7]). It is a free \(\mathscr {O}_{\varvec{\xi }}\)-module of rank two, equipped with a continuous linear action \(G_\mathbf{Q ,N}\). For each u in \(U_{\varvec{\xi }}\cap {}\mathbf{Z} _{\geqslant {}2}\) the base change \(V(\varvec{\xi })\otimes _{u}L\) of \(V(\varvec{\xi })\) along evaluation at u on \(\mathscr {O}_{\varvec{\xi }}\) is canonically isomorphic to the homological p-adic Deligne representation of \(\varvec{\xi }_{u}\) with coefficients in L (cf. loco citato for more details). In particular if \(\varvec{\xi }=\varvec{f}\) and \(u=2\) there is a natural specialisation isomorphism \(\rho _{2} : V(\varvec{f})\otimes _{2}L\simeq {}V(f)\). If \(\varvec{\xi }=\varvec{g}_{\alpha },\varvec{h}_{\alpha }\) and \(u=1\) set \(V(\xi )=V(\varvec{\xi })\otimes _{1}L\) (cf. Sect. 1). It is a two-dimensional L-vector space affording the dual of the p-adic Deligne–Serre representation of \(\xi =g,h\) with coefficients in L. In order to have a uniform notation, in this case one defines \(\rho _{1} : V(\varvec{\xi })\otimes _{1}L{\mathop {\longrightarrow }\limits ^{}}V(\xi )\) to be the identity.

The restriction of \(V(\varvec{\xi })\) to \(G_\mathbf{Q _{p}}\) (via the embedding \(i_{p}\) fixed at the outset) fits into a short exact sequence of \(\mathscr {O}_{\varvec{\xi }}[G_\mathbf{Q _{p}}]\)-modules with \(V(\varvec{\xi })^{\pm }\) free of rank one over \(\mathscr {O}_{\varvec{\xi }}\). More precisely, let \(\chi _{\text {cyc}} : G_\mathbf{Q }{\mathop {\longrightarrow }\limits ^{}}\mathbf{Z} _{p}^{*}\) be the p-adic cyclotomic character, and let \({\check{a}}_{p}(\varvec{\xi }) : G_\mathbf{Q _{p}}{\mathop {\longrightarrow }\limits ^{}}\mathscr {O}_{\varvec{\xi }}^{*}\) be the unramified character sending an arithmetic Frobenius to the p-th Fourier coefficients \(a_{p}(\varvec{\xi })\) of \(\varvec{\xi }\). Then

$$\begin{aligned} V(\varvec{\xi })^{+}\simeq {}\mathscr {O}_{\varvec{\xi }}\big (\chi _{\text {cyc}}^{\varvec{u}-1}\cdot {}\chi _{\xi }{\check{a}}_{p}(\varvec{\xi })^{-1}\big )\ \ \ \text {and}\ \ \ V(\varvec{\xi })^{-}\simeq {}\mathscr {O}_{\varvec{\xi }}({\check{a}}_{p}(\varvec{\xi })), \end{aligned}$$
(9)

where \(\chi _{\text {cyc}}^{\varvec{u}-1} : G_\mathbf{Q }{\mathop {\longrightarrow }\limits ^{}}\mathscr {O}_{\varvec{\xi }}^{*}\) satisfies \(\chi _{\text {cyc}}^{\varvec{u}-1}(\sigma )(u)=\chi _{\text {cyc}}(\sigma )^{u-1}\) for each u in \(U_{\varvec{\xi }}\cap {}\mathbf{Z} \). (The freeness of \(V(\varvec{\xi })^{\pm }\) is guaranteed by Assumption 1.1.3, cf. Section 5 of [7].) If \(\varvec{\xi }=\varvec{f}\) and \(u=2\) the specialisation isomorphism \(\rho _{2}\) identifies \(V(\varvec{f})^{-}\otimes _{2}L\) with the maximal unramified quotient \(V(f)^{-}\) of V(f). If \(\varvec{\xi }=\varvec{g}_{\alpha },\varvec{h}_{\alpha }\) and \(u=1\) we set \(V(\xi )_{\beta }=V(\varvec{\xi })^{+}\otimes _{1}L\) and \(V(\xi )_{\alpha }=V(\varvec{\xi })^{-}\otimes _{1}L\). One has \(V(\xi )=V(\xi )_{\alpha }\oplus {}V(\xi )_{\beta }\), where \(V(\xi )_{\gamma }=V(\xi )^{\text {Frob}_{p}=\gamma _{\xi }}\) for \(\gamma =\alpha ,\beta \) is the submodule of \(V(\xi )\) on which an arithmetic Frobenius \(\text {Frob}_{p}\) acts as multiplication by \(\gamma _{\xi }=\alpha _{\xi },\beta _{\xi }\) (cf. Assumption 1.1.3).

There is a natural \(G_\mathbf{Q }\)-equivariant skew-symmetric perfect pairing

$$\begin{aligned} \pi _{\varvec{\xi }} : V(\varvec{\xi })\otimes _{\mathscr {O}_{\varvec{\xi }}}V(\varvec{\xi }){\mathop {\longrightarrow }\limits ^{}}\mathscr {O}_{\varvec{\xi }}(\chi _{\xi }\cdot {}\chi _{\text {cyc}}^{\varvec{u}-1}), \end{aligned}$$

inducing perfect dualities \(\pi _{\varvec{\xi }} : V(\varvec{\xi })^{\pm }\otimes _{\mathscr {O}_{\varvec{\xi }}}V(\varvec{\xi })^{\mp }{\mathop {\longrightarrow }\limits ^{}}\mathscr {O}_{\varvec{\xi }}(\chi _{\xi }\cdot {}\chi _{\text {cyc}}^{\varvec{u}-1})\). (See Section 5 cf. [7] for the definitions).

Denote by \(\Xi _{\varvec{f}\varvec{g}\varvec{h}}=\chi _{\text {cyc}}^{(4-\varvec{k}-l-\varvec{m})/2} : G_\mathbf{Q }{\mathop {\longrightarrow }\limits ^{}}\mathscr {O}_{\!\varvec{f}\varvec{g}\varvec{h}}^{*}\) the character whose composition with evaluation at (klm) in \(U_{\!\varvec{f}}\times {}U_{\varvec{g}}\times {}U_{\varvec{h}}\cap {}\mathbf{Z} ^{3}\) on \(\mathscr {O}_{\!\varvec{f}\varvec{g}\varvec{h}}\) equals \(\chi _{\text {cyc}}^{(4-k-l-m)/2}\). If \(\cdot \) denotes one of the symbols \(\emptyset ,+\) and −, define

$$\begin{aligned} \varvec{V}^{\cdot }=V(\varvec{f})^{\cdot }{\hat{\otimes }}_{L}V(\varvec{g}_{\alpha }){\hat{\otimes }}V(\varvec{h}_{\alpha })\otimes _{\mathscr {O}_{\!\varvec{f}\varvec{g}\varvec{h}}}\Xi _{\varvec{f}\varvec{g}\varvec{h}}. \end{aligned}$$

Then \(\varvec{V}=V(\varvec{f},\varvec{g}_{\alpha },\varvec{h}_{\alpha })\), resp. \(\varvec{V}^{\pm }=V(\varvec{f},\varvec{g}_{\alpha },\varvec{h}_{\alpha })^{\pm }\) is a free \(\mathscr {O}_{\!\varvec{f}\varvec{g}\varvec{h}}\)-module of rank 8, resp. 4, equipped with a continuous action of \(G_\mathbf{Q ,N}\), resp. \(G_\mathbf{Q _{p}}\). As \(\chi _{g}\cdot {}\chi _{h}=1\) (cf. Assumption 1.1), the product of the perfect dualities \(\pi _{\varvec{\xi }}\), for \(\varvec{\xi }=\varvec{f},\varvec{g}_{\alpha },\varvec{h}_{\alpha }\), yields a perfect skew-symmetric Kummer duality \(\varvec{\pi } : \varvec{V}\otimes _{\mathscr {O}_{\!\varvec{f}\varvec{g}\varvec{h}}}\varvec{V}{\mathop {\longrightarrow }\limits ^{}}\mathscr {O}_{\!\varvec{f}\varvec{g}\varvec{h}}(1)\), inducing a perfect local Kummer duality \(\varvec{\pi } : \varvec{V}^{\pm }\otimes _{\mathscr {O}_{\!\varvec{f}\varvec{g}\varvec{h}}}\varvec{V}^{\mp }{\mathop {\longrightarrow }\limits ^{}} \mathscr {O}_{\!\varvec{f}\varvec{g}\varvec{h}}(1)\). After setting

$$\begin{aligned} V^{\cdot }=V(f,g,h)^{\cdot }=V(f)^{\cdot }\otimes _{L}V(g)\otimes _{L}V(h) \end{aligned}$$

and \(w_{o}=(2,1,1)\), the product \(\rho _{w_{o}}=\rho _{2}{\hat{\otimes }}\rho _{1}{\hat{\otimes }}\rho _{1}\) gives natural isomorphisms

$$\begin{aligned} \rho _{w_{o}} : \varvec{V}^{\cdot }\otimes _{w_{o}}L\simeq {}V^{\cdot } \end{aligned}$$
(10)

(where \(\cdot \otimes _{w_{o}}L\) denotes the base change along evaluation at \(w_{o}\) on \(\mathscr {O}_{\!\varvec{f}\varvec{g}\varvec{h}}\)). Let

$$\begin{aligned} \pi _{fgh} : V\otimes _{L}V{\mathop {\longrightarrow }\limits ^{}}L(1) \end{aligned}$$
(11)

be the specialisation of \(\varvec{\pi }\) via \(\rho _{w_{o}}\), and define \(\pi : V^{\pm }\otimes _{L}V^{\mp }{\mathop {\longrightarrow }\limits ^{}}L(1)\) similarly.

4.1.1.1 Weight one differentials

Define \(D(\varvec{\xi })^{-}=H^{0}(\mathbf{Q} _{p},V(\varvec{\xi })^{-}{\hat{\otimes }}_\mathbf{Q _{p}}{\hat{\mathbf{Q }}}_{p}^{\text {nr}})\), where \({\hat{\mathbf{Q }}}_{p}^{\text {nr}}\) is the p-adic completion of the maximal unramified extension of \(\mathbf{Q} _{p}\) (and as usual \(\varvec{\xi }\) denotes one of \(\varvec{f}\), \(\varvec{g}_{\alpha }\) and \(\varvec{h}_{\alpha }\)). For each u in \(U_{\varvec{\xi }}\cap {}\mathbf{Z} _{\geqslant {}2}\) there is a natural comparison isomorphism between \(D(\varvec{\xi })^{-}\otimes _{u}L\) and the \(\varvec{\xi }_{u}\)-isotypic component of the space of cuspidal modular forms of weight u, level \(\Gamma _{1}(N_{\xi }p)\) and Fourier coefficients in L. Assumption 1.1.3 guarantees that \(D(\varvec{\xi })^{-}\) is free (of rank one) over \(\mathscr {O}_{\varvec{\xi }}\), and admits a basis \(\omega _{\varvec{\xi }}\) whose image in \(D(\varvec{\xi })^{-}\otimes _{u}L\) corresponds to \(\varvec{\xi }_{u}\) under the aforementioned comparison isomorphism, for each u in \(U_{\varvec{\xi }}\cap {}\mathbf{Z} _{\geqslant {}2}\). (We refer to Section 3.1 of [6] and the references therein for more details.)

For \(\varvec{\xi }=\varvec{g}_{\alpha },\varvec{h}_{\alpha }\), the holomorphic weight-one differential

$$\begin{aligned} \omega _{\xi _{\alpha }}\in {}(V(\xi )_{\alpha }\otimes _\mathbf{Q _{p}}\mathbf{Q} _{p}^{\text {nr}})^{G_\mathbf{Q _{p}}} \end{aligned}$$

mentioned in Eq. (5) is defined to be the weight-one specialisation of \(\omega _{\varvec{\xi }}\), viz. the image of \(\omega _{\varvec{\xi }}\) in the quotient \(D(\varvec{\xi })^{-}\otimes _{1}L=D(\xi )_{\alpha }\). The weight-one specialisation of \(\pi _{\varvec{\xi }}\) yields a perfect \(G_\mathbf{Q }\)-equivariant skew-symmetric pairing

$$\begin{aligned} \pi _{\xi } : V(\xi )\otimes _{L}V(\xi ){\mathop {\longrightarrow }\limits ^{}}L(\chi _{\xi }). \end{aligned}$$

Let c be the common conductor of \(\chi _{g}\) and \(\chi _{h}\), and identify \((L(\chi _{\xi })\otimes _\mathbf{Q _{p}}\mathbf{Q} _{p}^{\text {nr}})^{G_\mathbf{Q _{p}}}\) with L via the Gauß sum \(G(\chi _{\xi })=(-c)^{i_{\xi }}\sum _{a\in {}(\mathbf{Z} /c\mathbf{Z} )^{*}}\chi _{\xi }(a)^{-1}\otimes {}e^{2\pi {}ia/c}\), where \(i_{g}=0\) and \(i_{h}=1\) (so that \(G(\chi _{g})\cdot {}G(\chi _{h})=1\) by Assumption 1.1.1). The pairing \(\pi _{\xi }\) then induces a perfect duality \(\left<{}\cdot ,\cdot \right>_{\xi } : D(\xi )_{\alpha }\otimes _{L}D(\xi )_{\beta }{\mathop {\longrightarrow }\limits ^{}}L\), where \(D(\xi )_{\gamma }=(V(\xi )_{\gamma }\otimes _\mathbf{Q _{p}}\mathbf{Q} _{p}^{\text {nr}})^{G_\mathbf{Q _{p}}}\). One defines the antiholomorphic weight-one differential (cf. Eq. (5))

$$\begin{aligned} \eta _{\xi _{\alpha }}\in {}(V(\xi )_{\beta }\otimes _\mathbf{Q _{p}}\mathbf{Q} _{p}^{\text {nr}})^{G_\mathbf{Q _{p}}} \end{aligned}$$

to be the dual of \(\omega _{\xi _{\alpha }}\) under \(\left<{}\cdot ,\cdot \right>_{\xi }\), viz. the element satisfying \(\left<{}\omega _{\xi _{\alpha }},\eta _{\xi _{\alpha }}\right>_{\xi }=1\).

4.1.1.2 The embeddings \(\gamma _{g}\) and \(\gamma _{h}\)

With the notations of Sect. 1, set \(V_{g}=V_{\varrho _{1}}\) and \(V_{h}=V_{\varrho _{2}}\). Let \(\xi \) denote either g or h. As recalled above, the Artin representation \(V(\xi )=V(\varvec{\xi })\otimes _{1}L\) affords the dual of the p-adic Deligne representation of \(\xi \) with coefficients in L, id est is isomorphic to \(V_{\xi }\otimes _\mathbf{Q (\varrho )}L\). Enlarging L if necessary, we normalise the \(G_\mathbf{Q }\)-equivariant embedding \(\gamma _{\xi } : V_{\xi }{\mathop {\longrightarrow }\limits ^{}}V(\xi )\) (introduced in Eq. (3)) by requiring that the composition \(\pi _{\xi }\circ {}(\gamma _{\xi }\otimes {}\gamma _{\xi })\) takes values in the number field \(\mathbf{Q} (\varrho )\) (via the embedding fixed at the outset).

4.1.2 Selmer complexes

Let \(\mathbf{R} {\tilde{\Gamma }}_{f}(\mathbf{Q} ,V)\) be the Nekovář Selmer complex associated with \((V,V^{+})\) (cf. Section 2.2 of [6]). It is an element of the derived category \(\text {D}_{\text {ft}}^{b}(L)\) of cohomologically bounded complexes of L-modules with cohomology of finite type over L, sitting is an exact triangle

$$\begin{aligned} \mathbf{R} \Gamma _{\text {cont}}(G_\mathbf{Q ,N},V){\mathop {\longrightarrow }\limits ^{p^{-}\circ {}\text {res}_{p}}}\mathbf{R} \Gamma _{\text {cont}}(G_\mathbf{Q _{p}},V^{-}){\mathop {\longrightarrow }\limits ^{}}\mathbf{R} {\tilde{\Gamma }}_{f}(\mathbf{Q} ,V)[1], \end{aligned}$$
(12)

where \(\mathbf{R} \Gamma _{\text {cont}}(G,\cdot )\) is the complex of continuous non-homogeneous cochains of G with values in \(\cdot \), \(\text {res}_{p}\) is the restriction map (induced by the embedding fixed at the outset) and \(p^{-}\) is the map induced by the projection \(V{\mathop {\longrightarrow }\limits ^{}}V^{-}\). Denote by

$$\begin{aligned} \tilde{H}^{\cdot }_{f}(\mathbf{Q} ,V)=H^{\cdot }(\mathbf{R} {\tilde{\Gamma }}_{f}(\mathbf{Q} ,V)) \end{aligned}$$

the cohomology of \(\mathbf{R} {\tilde{\Gamma }}(\mathbf{Q} ,V)\), let \(\text {Sel}(\mathbf{Q} ,V)\) be the Bloch–Kato Selmer group of V over \(\mathbf{Q} \), and let \(i^{+} : V^{+}{\mathop {\longrightarrow }\limits ^{}}V\) be the natural inclusion. Then there is a commutative and exact diagram of L-vector spaces (cf. loc. cit.)

(13)

where the first line arises from the exact triangle (12). In addition there is a unique section \(\imath _{\text {ur}} : \text {Sel}(\mathbf{Q} ,V){\mathop {\longrightarrow }\limits ^{}}\tilde{H}^{1}_{f}(\mathbf{Q} ,V)\) of the above projection such that \(\imath _{\text {ur}}(x)^{+}\) belongs to the Bloch–Kato finite subspace \(H^{1}_{\text {fin}}(\mathbf{Q} _{p},V^{+})\) for each x in \(\text {Sel}(\mathbf{Q} ,V)\). We often use \(\jmath \) and \(\imath _{\text {ur}}\) to identify Nekovář’s extended Selmer group \(\tilde{H}^{1}_{f}(\mathbf{Q} ,V)\) with the naive extended Selmer group \(\text {Sel}^{\dag }(\mathbf{Q} ,V)=H^{0}(\mathbf{Q} _{p},V^{-})\oplus {}\text {Sel}(\mathbf{Q} ,V)\) (cf. Sect. 1).

One similarly associates with \((\varvec{V},\varvec{V}^{+})\) a Selmer complex

$$\begin{aligned} \mathbf{R} {\tilde{\Gamma }}_{f}(\mathbf{Q} ,\varvec{V})\in {}\text {D}_{\text {ft}}^{b}(\mathscr {O}_{\!\varvec{f}\varvec{g}\varvec{h}}) \end{aligned}$$

sitting in an exact triangle analogous to (12). (We refer to loc. cit. for more details.)

4.2 Preliminary lemmas

This section gives a concrete description of the functionals for q in \(H^{0}(\mathbf{Q} _{p},V^{-})\) (cf. Lemma 3.4 below).

4.2.1 Bockstein maps

Let \((\varvec{{\mathcal {C}}},{\mathcal {C}})\) denote one of the pairs

$$\begin{aligned} (\mathbf{R} \Gamma _{p}(\varvec{V}^{-}),\mathbf{R} \Gamma _{p}(V^{-})), (\mathbf{R} \Gamma (\varvec{V}),\mathbf{R} \Gamma (V)) \quad {\mathrm {and}}\quad (\mathbf{R} {\tilde{\Gamma }}_{f} (\mathbf{Q} ,\varvec{V}),\mathbf{R} {\tilde{\Gamma }}_{f}(\mathbf{Q} ,V)), \end{aligned}$$

where \(\mathbf{R} \Gamma _{p}(\cdot )\) and \(\mathbf{R} \Gamma (\cdot )\) are shorthands for \(\mathbf{R} \Gamma _{\text {cont}}(\mathbf{Q} _{p},\cdot )=\mathbf{R} \Gamma _{\text {cont}}(G_\mathbf{Q _{p}},\cdot )\) and \(\mathbf{R} \Gamma _{\text {cont}}(G_\mathbf{Q ,N},\cdot )\) respectively (cf. Sect. 3.1.2). The specialisation maps \(\rho _{w_{o}}\) (cf. Eq. (10)) induce isomorphisms

$$\begin{aligned} \rho _{w_{o}} : \varvec{{\mathcal {C}}}\otimes _{\mathscr {O}_{\!\varvec{f}\varvec{g}\varvec{h}},w_{o}}^\mathbf{L }L\simeq {}{\mathcal {C}}\ \ \ \text {and}\ \ \ \rho _{w_{o}}\otimes \text {id} : \varvec{{\mathcal {C}}}\otimes _{\mathscr {O}_{\!\varvec{f}\varvec{g}\varvec{h}}}^\mathbf{L }\mathscr {I}/\mathscr {I}^{2}[1]\simeq {}{\mathcal {C}}\otimes _{L}\mathscr {I}/\mathscr {I}^{2}[1]. \end{aligned}$$
(14)

Applying \(\varvec{{\mathcal {C}}}\otimes _{\mathscr {O}_{\!\varvec{f}\varvec{g}\varvec{h}}}^\mathbf{L }\cdot \) to the exact triangle

$$\begin{aligned} \mathscr {I}/\mathscr {I}^{2}{\mathop {\longrightarrow }\limits ^{}}\mathscr {O}_{\!\varvec{f}\varvec{g}\varvec{h}}/\mathscr {I}^{2}{\mathop {\longrightarrow }\limits ^{}}L{\mathop {\longrightarrow }\limits ^{}}\mathscr {I}/\mathscr {I}^{2}[1] \end{aligned}$$

(arising from evaluation on \(w_{o}\)) then yields a derived Bockstein map

$$\begin{aligned} \varvec{\beta }_{\varvec{{\mathcal {C}}}/{\mathcal {C}}} : {\mathcal {C}}{\mathop {\longrightarrow }\limits ^{}}{\mathcal {C}}\otimes _{L}\mathscr {I}/\mathscr {I}^{2}[1], \end{aligned}$$

which in turn induces in cohomology a Bockstein map

$$\begin{aligned} \beta _{\varvec{{\mathcal {C}}}/{\mathcal {C}}} : H^{i}({\mathcal {C}}){\mathop {\longrightarrow }\limits ^{}}H^{i+1}({\mathcal {C}})\otimes _{L}\mathscr {I}/\mathscr {I}^{2}. \end{aligned}$$

If no risk of confusion arises, we simply write \(\beta \) for \(\beta _{\varvec{{\mathcal {C}}}/{\mathcal {C}}}\). Let

$$\begin{aligned} \jmath : H^{i}(\mathbf{Q} _{p},V^{-}){\mathop {\longrightarrow }\limits ^{}}\tilde{H}^{i+1}_{f}(\mathbf{Q} ,V) \end{aligned}$$

be the maps arising from the exact triangle (12).

Lemma 3.1

The following diagram commutes.

Proof

For \(M=V,\varvec{V}\) one has an exact triangle (cf. Equation (12))

$$\begin{aligned} \Delta _{M} : \mathbf{R} \Gamma _{\text {cont}}(G_\mathbf{Q ,N},M)[-1]{\mathop {\longrightarrow }\limits ^{p^{-}\circ {}\text {res}_{p}\,}}\mathbf{R} \Gamma _{\text {cont}}(\mathbf{Q} _{p},M^{-})[-1]{\mathop {\longrightarrow }\limits ^{\varvec{\jmath }_{M}\,}} \mathbf{R} {\tilde{\Gamma }}_{f}(\mathbf{Q} ,M). \end{aligned}$$

Moreover \(\Delta _{V}\) is obtained by applying \(\cdot {}\otimes _{\mathscr {O}_{\varvec{f}\varvec{g}\varvec{h}},w_{o}}^\mathbf{L }L\) to \(\Delta _{\varvec{V}}\) (cf. Eq. (14)). It follows from the definition of the derived Bockstein maps \(\varvec{\beta }^{-}\) and \(\varvec{\beta }\) on \(\mathbf{R} \Gamma _{\text {cont}}(\mathbf{Q} _{p},V^{-})\) and \(\mathbf{R} {\tilde{\Gamma }}(\mathbf{Q} ,V)\) respectively that \(\varvec{\jmath }_{V}\otimes \mathscr {I}/\mathscr {I}^{2}[1]\circ {}\varvec{\beta }^{-}\) is equal to \(\varvec{\beta }\circ {}\varvec{\jmath }_{V}\). Since by definition the maps \(\jmath \) are the ones induced in cohomology by \(\varvec{\jmath }_{V}\), the lemma follows. \(\square \)

The following lemma gives a concrete description of \(\beta _{\varvec{{\mathcal {C}}}/{\mathcal {C}}}\).

Lemma 3.2

Let \((\varvec{{\mathcal {C}}},{\mathcal {C}})\) be as above, let z be a 1-cocycle in \({\mathcal {C}}\), let Z be a 1-cochain in \(\varvec{{\mathcal {C}}}\), and let \(Z_{\varvec{k}}, Z_{l}\) and \(Z_{\varvec{m}}\) be 2-cochains in \(\varvec{{\mathcal {C}}}\) such that

$$\begin{aligned} \rho _{w_{o}}(Z)=z\ \ \text {and}\ \ dZ=Z_{\varvec{k}}\cdot {}(\varvec{k}-2)+Z_{l}\cdot {}(l-1)+Z_{\varvec{m}}\cdot {}(\varvec{m}-1). \end{aligned}$$

Then \(z_{\cdot }=\rho _{w_{o}}(Z_{\cdot })\) is a 2-cocycle for \(\cdot =\varvec{k},l,\varvec{m}\), and one has the equality

$$\begin{aligned} -\beta _{\varvec{{\mathcal {C}}}/{\mathcal {C}}}(cl(z))=cl(z_{\varvec{k}})\cdot {}(\varvec{k}-2)+cl(z_{l})\cdot {}(l-1)+cl(z_{\varvec{m}})\cdot {}(\varvec{m}-1) \end{aligned}$$

in \(H^{2}({\mathcal {C}})\otimes _{L}\mathscr {I}/\mathscr {I}^{2}\), where \(cl(\cdot )\) is the class in \(H^{i}({\mathcal {C}})\) represented by the i-cocycle \(\cdot \).

Proof

The proof is very similar to that of [16, Lemma 5.5]. We omit it. \(\square \)

4.2.2 Local and global duality

Nekovář’s generalised Poitou–Tate duality associates with the perfect duality \(\pi _{fgh}\) introduced in Eq. (11) a global cup-product pairing (cf. Section 2.4 of [6])

$$\begin{aligned} \left<{}\cdot ,\cdot \right>_{\text {Nek}} : \tilde{H}^{2}_{f}(\mathbf{Q} ,V)\otimes _{L}\tilde{H}^{1}_{f}(\mathbf{Q} ,V){\mathop {\longrightarrow }\limits ^{}}L. \end{aligned}$$
(15)

The pairing \(\pi _{fgh}\) induces a Kummer duality \(V^{-}\otimes _{L}V^{+}{\mathop {\longrightarrow }\limits ^{}}L(1)\) and we denote by

$$\begin{aligned} \left<{}\cdot ,\cdot \right>_{\text {Tate}} : H^{1}(\mathbf{Q} _{p},V^{-})\otimes _{L}H^{1}\left( \mathbf{Q} _{p},V^{+}\right) {\mathop {\longrightarrow }\limits ^{}}L \end{aligned}$$
(16)

the induced local Tate duality pairing. Recall finally the map

$$\begin{aligned} \cdot ^{+} : \tilde{H}_{f}^{1}(\mathbf{Q} ,V){\mathop {\longrightarrow }\limits ^{}}H^{1}(\mathbf{Q} _{p},V^{+}) \end{aligned}$$

introduced in diagram (13).

Lemma 3.3

For each \(\zeta \) in \(H^{1}(\mathbf{Q} _{p},V^{-})\) and \(\xi \) in \(\tilde{H}^{1}_{f}(\mathbf{Q} ,V)\) one has

$$\begin{aligned} \left<{}\jmath (\zeta ),\xi \right>_{\text {Nek}}=\langle \zeta ,\xi ^{+}\rangle _{\text {Tate}}. \end{aligned}$$

Proof

This is proved as in [16, Lemma 5.7]. \(\square \)

4.2.3 The Garrett–Nekovář p-adic height pairing

Set

$$\begin{aligned} {\tilde{\beta }}_{\varvec{f}\varvec{g}_{\alpha }\varvec{h}_{\alpha }} =\beta _\mathbf{R {\tilde{\Gamma }}_{f}(\mathbf{Q} ,\varvec{V})/\mathbf{R} {\tilde{\Gamma }}_{f}(\mathbf{Q} ,V)} : \tilde{H}^{1}_{f}(\mathbf{Q} ,V){\mathop {\longrightarrow }\limits ^{}}\tilde{H}^{2}_{f}(\mathbf{Q} ,V)\otimes _{L}\mathscr {I}/\mathscr {I}^{2}. \end{aligned}$$

After identifying \(\tilde{H}^{1}_{f}(\mathbf{Q} ,V)\) with \(\text {Sel}^{\dag }(\mathbf{Q} ,V)\) (cf. Sect. 3.1.2), the canonical height introduced in Sect.  is defined by (cf. [6, Section 2])

for each x and y in \(\tilde{H}^{1}_{f}(\mathbf{Q} ,V)\), where we write again \(\left<{}\cdot ,\cdot \right>_{\text {Nek}}\) for the \(\mathscr {I}/\mathscr {I}^{2}\)-base change of Nekovář’s cup-product (15). Lemmas 3.1 and 3.3 give the following

Lemma 3.4

For each q in \(H^{0}(\mathbf{Q} _{p},V^{-})\) one has

as \(\mathscr {I}/\mathscr {I}^{2}\)-valued maps on \(\tilde{H}^{1}_{f}(\mathbf{Q} ,V)\), where \(\beta _{\varvec{f}\varvec{g}_{\alpha }\varvec{h}_{\alpha }}^{-}=\beta _\mathbf{R \Gamma _{p}(\varvec{V}^{-})/\mathbf{R} \Gamma _{p}(V^{-})}\) (and we write again \(\left<{}\cdot ,\cdot \right>_{\text {Tate}}\) for the \(\mathscr {I}/\mathscr {I}^{2}\)-base change of the local Tate pairing (16)).

4.3 Computation of

Assume in this subsection \(\alpha _{f}=\alpha _{g}\cdot {}\alpha _{h}\), so that \(H^{0}(\mathbf{Q} _{p},V^{-})\) is generated over L by the periods

$$\begin{aligned} q_{\alpha \alpha }=\sqrt{m_{p}}\cdot {}q(f)\otimes \omega _{g_{\alpha }}\otimes \omega _{h_{\alpha }}\ \ \ \text {and}\ \ \ q_{\beta \beta }=\sqrt{m_{p}}\cdot {}q(f)\otimes \eta _{g_{\alpha }}\otimes \eta _{h_{\alpha }}. \end{aligned}$$

Recall that \(\chi _{\text {cyc}} : G_\mathbf{Q }{\mathop {\longrightarrow }\limits ^{}}\mathbf{Z} _{p}^{*}\) denotes the p-adic cyclotomic character. Fix a lift \(\varvec{q}_{\beta \beta }\) in \(\varvec{V}^{-}\) of \(q_{\beta \beta }\) under \(\rho _{w_{o}}\). Since (cf. Sect. 3.1.1)

and \(V(\xi )_{\beta }=V(\varvec{\xi }_{\alpha })^{+}\otimes _{1}L\) for \(\xi =g,h\), we can choose \(\varvec{q}_{\beta \beta }\) in the \(G_\mathbf{Q _{p}}\)-submodule

(cf. Sect. 3.1.1). By Eq. (9) one has

$$\begin{aligned} d\varvec{q}_{\beta \beta }=\Phi \cdot {}\varvec{q}_{\beta \beta }, \end{aligned}$$
(17)

where d denotes the differentials of the complex \(\mathbf{R} \Gamma _{\text {cont}}(\mathbf{Q} _{p},\varvec{V}^{-})\) and

$$\begin{aligned} \Phi =\frac{{\check{a}}_{p}(\varvec{f})}{{\check{a}}_{p}(\varvec{g}_{\alpha })\cdot {}{\check{a}}_{p}(\varvec{h}_{\alpha })}\cdot {}\chi _{\text {cyc}}^{(l+\varvec{m}-\varvec{k})/2}-1 : G_\mathbf{Q _{p}}{\mathop {\longrightarrow }\limits ^{}}\mathscr {O}_{\!\varvec{f}\varvec{g}\varvec{h}}. \end{aligned}$$

The assumption \(\alpha _{f}=\alpha _{g}\cdot {}\alpha _{h}\) implies that \(\Phi \) takes value in \(\mathscr {I}\), and that its composition \(\Phi ^{\prime }\) with the projection \(\mathscr {I}{\mathop {\longrightarrow }\limits ^{}}\mathscr {I}/\mathscr {I}^{2}\) is of the form

$$\begin{aligned} \Phi ^{\prime }=\varphi _{\varvec{k}}\cdot {}(\varvec{k}-2)+\varphi _{l}\cdot {}(l-1)+\varphi _{\varvec{m}}\cdot {}(\varvec{m}-1) \end{aligned}$$

with \(\varphi _{\varvec{u}}\) in \(H^{1}(\mathbf{Q} _{p},\mathbf{Q} _{p})\) for \(\varvec{u}=\varvec{k},l,\varvec{m}\). Identify \(H^{1}(\mathbf{Q} _{p},\mathbf{Q} _{p})\) with the \(\mathbf{Q} _{p}\)-vector space \(\text {Hom}_{}(\mathbf{Q} _{p}^{*},\mathbf{Q} _{p})\) of continuous morphisms of groups from \(\mathbf{Q} _{p}^{*}\) to \(\mathbf{Q} _{p}\) via the local reciprocity map \(\text {rec}_{p} : \mathbf{Q} _{p}^{*}{\mathop {\longrightarrow }\limits ^{}}G_\mathbf{Q _{p}}^{\text {ab}}\), normalised by requiring \(\text {rec}_{p}(p^{-1})\) to be an arithmetic Frobenius. By local class field theory, for each p-adic unit u one has

$$\begin{aligned} \varphi _{\varvec{k}}(u)=\frac{\partial }{\partial \varvec{k}}\left. \left( \left<{}u\right>^{(l+\varvec{m}-\varvec{k})/2}-1\right) \right| _{w_{o}}=-\frac{1}{2}\cdot {}\log _{p}(u), \end{aligned}$$

where \(\left<{}\cdot \right> : \mathbf{Z} _{p}^{*}{\mathop {\longrightarrow }\limits ^{}}1+p\mathbf{Z} _{p}\) denotes the projection to principal units, and

$$\begin{aligned} \varphi _{\varvec{k}}(p)=\frac{\partial }{\partial \varvec{k}}\left. \left( \frac{a_{p}(\varvec{g}_{\alpha })\cdot {}a_{p}(\varvec{h}_{\alpha })}{a_{p}(\varvec{f})}-1\right) \right| _{w_{o}}= \frac{1}{2}\cdot {}{\mathfrak {L}}_{\varvec{f}}^{\text {an}} \end{aligned}$$

(cf. Eq. (7)). As a consequence \(-2\cdot {}\varphi _{\varvec{k}}\) is equal to

$$\begin{aligned} \log _{\varvec{f}}=\log _{p}-{\mathfrak {L}}_{\varvec{f}}^{\text {an}}\cdot {}\text {ord}_{p}\in {}H^{1}(\mathbf{Q} _{p},\mathbf{Q} _{p}) \end{aligned}$$

(where the p-adic valuation \(\text {ord}_{p} : \mathbf{Q} _{p}^{*}{\mathop {\longrightarrow }\limits ^{}}\mathbf{Q} _{p}\) is normalised by \(\text {ord}_{p}(p)=1\)). Similarly one shows that \(2\cdot {}\varphi _{l}\) and \(2\cdot {}\varphi _{m}\) are equal to the logarithms \(\log _{\varvec{g}_{\alpha }}=\log _{p}-{\mathfrak {L}}_{\varvec{g}_{\alpha }}^{\text {an}}\cdot {}\text {ord}_{p}\) and \(\log _{\varvec{h}_{\alpha }}=\log _{p}-{\mathfrak {L}}_{\varvec{g}_{\alpha }}^{\text {an}}\cdot {}\text {ord}_{p}\). It then follows from Eq. (17) and Lemma 3.2 that

$$\begin{aligned} 2\cdot {}\beta _{\varvec{f}\varvec{g}_{\alpha }\varvec{h}_{\alpha }}^{-}(q_{\beta \beta })=\left( \log _{\varvec{f}}\cdot {}(\varvec{k}-2)-\log _{\varvec{g}_{\alpha }}\cdot {}(l-1)-\log _{\varvec{h}_{\alpha }}\cdot {}(\varvec{m}-1)\right) \otimes {}q_{\beta \beta } \end{aligned}$$
(18)

in \(H^{1}(\mathbf{Q} _{p},V^{-})\otimes _{L}\mathscr {I}/\mathscr {I}^{2}\), where (with the notations introduced in Sect. 3.2.1) one writes \(\beta ^{-}_{\varvec{f}\varvec{g}_{\alpha }\varvec{h}_{\alpha }}\) for the Bockstein map \(\beta _{\varvec{{\mathcal {C}}}/{\mathcal {C}}}\) associated with \(\varvec{{\mathcal {C}}}=\mathbf{R} \Gamma _{p}(\varvec{V}^{-})\). Note that

$$\begin{aligned} V(f)^{-}_{\beta \beta }=V(f)^{-}\otimes _\mathbf{Q _{p}}V(g)_{\beta }\otimes _{L}V(h)_{\beta } \end{aligned}$$

is an \(L[G_\mathbf{Q _{p}}]\)-direct summand of \(V^{-}\) on which \(G_\mathbf{Q _{p}}\) acts trivially, so that \(\log _{\varvec{\xi }}\otimes {}q_{\beta \beta }\) (for \(\varvec{\xi }=\varvec{f},\varvec{g}_{\alpha },\varvec{h}_{\alpha }\)) belongs to the direct summand

$$\begin{aligned} H^{1}(\mathbf{Q} _{p},V(f)^{-}_{\beta \beta })=H^{1}(\mathbf{Q} _{p},\mathbf{Q} _{p})\otimes _\mathbf{Q _{p}}V(f)^{-}_{\beta \beta } \end{aligned}$$

of the local cohomology group \(H^{1}(\mathbf{Q} _{p},V^{-})\). Similarly

$$\begin{aligned} V(f)^{+}_{\alpha \alpha }=V(f)^{+}\otimes _\mathbf{Q _{p}}V(g)_{\alpha }\otimes _{L}V(h)_{\alpha } \end{aligned}$$

is an \(L[G_\mathbf{Q _{p}}]\)-direct summand of \(V^{+}\) isomorphic to \(\mathbf{Q} _{p}(1)\), hence

$$\begin{aligned} H^{1}(\mathbf{Q} _{p},V(f)^{+}_{\alpha \alpha })=H^{1}(\mathbf{Q} _{p},\mathbf{Q} _{p}(1))\otimes _\mathbf{Q _{p}}V(f)^{+}_{\alpha \alpha }(-1) \end{aligned}$$
(19)

is a direct summand of \(H^{1}(\mathbf{Q} _{p},V^{+})\). The local Tate pairing \(\left<{}\cdot ,\cdot \right>_{\text {Tate}}\) introduced in Sect. 3.2.2 induces a perfect duality (denoted by the same symbol) between \(H^{1}(\mathbf{Q} _{p},V(f)^{-}_{\beta \beta })\) and \(H^{1}(\mathbf{Q} _{p},V(f)^{+}_{\alpha \alpha })\), and identifying \(H^{1}(\mathbf{Q} _{p},\mathbf{Z} _{p}(1))\) with the p-adic completion \({\hat{\mathbf{Q }}}_{p}^{*}\) of \(\mathbf{Q} _{p}^{*}\) via the local Kummer map, local class field theory gives

$$\begin{aligned} \langle \varphi \otimes {}v^{-},u\otimes {}v^{+}\rangle _{\text {Tate}}=\varphi (u)\cdot {}\pi _{fgh}(-1)(v^{+}\otimes {}v^{-}) \end{aligned}$$
(20)

for each \(\varphi \) in \(H^{1}(\mathbf{Q} _{p},\mathbf{Q} _{p})\), u in \(H^{1}(\mathbf{Q} _{p},\mathbf{Q} _{p}(1))\), \(v^{-}\) in \(V(f)^{-}_{\beta \beta }\) and \(v^{+}\) in \(V(f)^{+}_{\alpha \alpha }\). Here

$$\begin{aligned} \pi _{fgh}(-1) : V(f)^{+}_{\alpha \alpha }(-1)\otimes _{L}V(f)^{-}_{\beta \beta }{\mathop {\longrightarrow }\limits ^{}}L \end{aligned}$$

is the composition of \(\pi _{fgh}\otimes {}\mathbf{Q} _{p}(-1)\) with the evaluation pairing \(L(1)\otimes _{L}L(-1){\mathop {\longrightarrow }\limits ^{}}L\).

Recall that we identify \(H^{0}(\mathbf{Q} _{p},V^{-})\) with a submodule of \(\tilde{H}^{1}_{f}(\mathbf{Q} ,V)\) via the embedding \(\jmath \) introduced in Diagram (13). Lemma 3.4 and Eqs. (18) and (20) give

(21)

for each z in \(\tilde{H}^{1}_{f}(\mathbf{Q} ,V)\), where \(\varvec{\xi }=\varvec{f},\varvec{g}_{\alpha },\varvec{h}_{\alpha }\), \(u_{o}=2,1,1\) is the centre of \(U_{\varvec{\xi }}\), and

$$\begin{aligned} z_{\alpha \alpha }^{+}\in {}H^{1}(\mathbf{Q} _{p},\mathbf{Q} _{p}(1))={\hat{\mathbf{Q }}}_{p}^{*}\otimes _\mathbf{Z _{p}}\mathbf{Q} _{p} \end{aligned}$$

is defined as follows. Let \(\text {pr}_{\alpha \alpha }\) denote the projection onto the direct summand \(H^{1}(\mathbf{Q} _{p},V(f)^{+}_{\alpha \alpha })\) of the local cohomology group \(H^{1}(\mathbf{Q} _{p},V^{+})\), and let \(q_{\beta \beta }^{*}\) be the generator of \(V(f)^{+}_{\alpha \alpha }(-1)\) dual to \(q_{\beta \beta }\) under \(\pi _{fgh}(-1)\), namely satisfying

$$\begin{aligned} \pi _{fgh}(-1)(q_{\beta \beta }^{*}\otimes {}q_{\beta \beta })=1. \end{aligned}$$

Then \(z_{\alpha \alpha }^{+}\) is defined (via the natural isomorphism (19)) by the identity

$$\begin{aligned} \text {pr}_{\alpha \alpha }(z^{+})=z_{\alpha \alpha }^{+}\otimes {}q_{\beta \beta }^{*}. \end{aligned}$$
(22)

We now determine \(z_{\alpha \alpha }^{+}\) for \(z=\jmath (q_{\alpha \alpha })\). By definition \(\jmath (q_{\alpha \alpha })\) is represented by

$$\begin{aligned} c_{\alpha \alpha }=(0,d\tilde{q}_{\alpha \alpha },\tilde{q}_{\alpha \alpha })\in {}\tilde{\text {C}}^{1}_{f}(\mathbf{Q} ,V), \end{aligned}$$

where \(\tilde{q}_{\alpha \alpha }\) in V is a lift of \(q_{\alpha \alpha }\) under the the projection \(V{\mathop {\longrightarrow }\limits ^{}}V^{-}\), and where

$$\begin{aligned} d\tilde{q}_{\alpha \alpha } : G_\mathbf{Q _{p}}{\mathop {\longrightarrow }\limits ^{}}V^{+} \end{aligned}$$

is its image under the differential in \(\mathbf{R} \Gamma _{\text {cont}}(\mathbf{Q} _{p},V)\). By construction \(d\tilde{q}_{\alpha \alpha }\) represents the class \(q_{\alpha \alpha }^{+}=\jmath (q_{\alpha \alpha })^{+}\) in \(H^{1}(\mathbf{Q} _{p},V^{+})\). Since \(V(\xi )\) is the direct sum of \(V(\xi )_{\alpha }\) and \(V(\xi )_{\beta }\) for \(\xi =g,h\), we can (and will) choose \(\tilde{q}_{\alpha \alpha }\) of the form

$$\begin{aligned} \tilde{q}_{\alpha \alpha }=\sqrt{m_{p}}\cdot {}\tilde{q}(f)\otimes \omega _{g_{\alpha }}\otimes \omega _{h_{\alpha }} \end{aligned}$$

for a lift \(\tilde{q}(f)\) of q(f) under the projection \(V(f){\mathop {\longrightarrow }\limits ^{}}V(f)^{-}\), so that \(d\tilde{q}_{\alpha \alpha }\) represents the image of \(q_{\alpha \alpha }\) under the connecting morphism

$$\begin{aligned} \delta _{\alpha \alpha } : V(f)^{-}_{\alpha \alpha }{\mathop {\longrightarrow }\limits ^{}}H^{1}(\mathbf{Q} _{p},V(f)^{+}_{\alpha \alpha }) \end{aligned}$$

arising from the short exact sequence of \(G_\mathbf{Q _{p}}\)-modules

$$\begin{aligned} 0{\mathop {\longrightarrow }\limits ^{}}V(f)^{+}_{\alpha \alpha }{\mathop {\longrightarrow }\limits ^{}}V(f)_{\alpha \alpha }{\mathop {\longrightarrow }\limits ^{}}V(f)^{-}_{\alpha \alpha }{\mathop {\longrightarrow }\limits ^{}}0, \end{aligned}$$

where \(V(f)^{\cdot }_{\alpha \alpha }\) is the \(L[G_\mathbf{Q _{p}}]\)-direct summand \(V(f)^{\cdot }\otimes _\mathbf{Q _{p}}V(g)_{\alpha }\otimes _{L}V(h)_{\alpha }\) of \(V^{\cdot }\). Let \(q_{A}\) in \(p\mathbf{Z} _{p}\) be the Tate period of \(A_\mathbf{Q _{p}}\). Tate’s theory gives a rigid analytic isomorphisms between the base change \(E_\mathbf{Q _{p}^{2}}\) of the Tate curve \(E=\mathbf{G} _{m,\mathbf{Q} _{p}}^{rig}/q_{A}^\mathbf{Z }\) to the quadratic unramified extension \(\mathbf{Q} _{p^{2}}\) of \(\mathbf{Q} _{p}\) and \(A_\mathbf{Q _{p^{2}}}\). Set \(V_{p}(E)=H_\mathrm{{\acute{e}t}}^{1}(E_{{\bar{\mathbf{Q }}}_{p}},\mathbf{Q} _{p}(1))\) and let \(\wp _{\text {Tate}} : V_{p}(E)\simeq {}V_{p}(A)\) be the isomorphisms of \(G_\mathbf{Q _{p^{2}}}\)-modules induced by the Tate uniformisation. There is a short exact sequence of \(\mathbf{Q} _{p}[G_\mathbf{Q _{p}}]\)-modules

$$\begin{aligned} 0{\mathop {\longrightarrow }\limits ^{}}\mathbf{Q} _{p}(1){\mathop {\longrightarrow }\limits ^{a\,}}V_{p}(E){\mathop {\longrightarrow }\limits ^{b\,}}\mathbf{Q} _{p}{\mathop {\longrightarrow }\limits ^{}}0, \end{aligned}$$
(23)

where \(a(\zeta _{p^{\infty }})=(\zeta _{p^{n}}\cdot {}q_{A}^\mathbf{Z })_{n\geqslant {}1}\) for each compatible system \(\zeta _{p^{\infty }}=(\zeta _{p^{n}})_{n{}\geqslant {}1}\) of \(p^{n}\)-th roots of unity, and b is the \(\mathbf{Q} _{p}\)-linear extension of the inverse limit of (canonical) maps

$$\begin{aligned} b_{n} : E({\bar{\mathbf{Q }}}_{p})_{p^{n}}=({\bar{\mathbf{Q }}}_{p}^{*}/q_{A}^\mathbf{Z })_{p^{n}}{\mathop {\longrightarrow }\limits ^{}}\mathbf{Z} /p^{n}\mathbf{Z} \end{aligned}$$

defined by \(b_{n}(x\cdot {}q_{A}^\mathbf{Z })=\frac{p^{n}\cdot {}\text {ord}_{p}(x)}{\text {ord}_{p}(q_{A})}+p^{n}\cdot {}\mathbf{Z} \). By definition \(q(A)=\wp _{\text {Tate}}^{-}(1)\), where \(\wp _{\text {Tate}}^{-}\circ {}b\) is the composition of \(\wp _{\text {Tate}}\) and the projection \(V_{p}(A){\mathop {\longrightarrow }\limits ^{}}V_{p}(A)^{-}\) onto the maximal \(G_\mathbf{Q _{p}}\)-unramified quotient, and

$$\begin{aligned} \tilde{q}(f)=\wp _{\infty }^{-1}\circ {}\wp _{\text {Tate}}(\!\root p^{\infty } \of {q_{A}}\,) \end{aligned}$$

is the image of a compatible system \(\root p^{\infty } \of {q_{A}}\) of \(p^{n}\)-th roots of the Tate period \(q_{A}\) under the composition of \(\wp _{\text {Tate}}\) and the inverse of the isomorphism \(\wp _{\infty } : V(f)\simeq {}V_{p}(A)\) induced by the fixed modular parametrisation \(\wp _{\infty } : X_{1}(N_{f}){\mathop {\longrightarrow }\limits ^{}}A\). As a consequence 1 in \(\mathbf{Q} _{p}\) maps to \(q_{A}{\hat{\otimes }}1\) under the connecting map \(\mathbf{Q} _{p}{\mathop {\longrightarrow }\limits ^{}}H^{1}(\mathbf{Q} _{p},\mathbf{Q} _{p}(1))=\mathbf{Q} _{p}^{*}{\hat{\otimes }}\mathbf{Q} _{p}\) associated with the short exact sequence (23), hence

$$\begin{aligned} \jmath (q_{\alpha \alpha })^{+}=cl(d\tilde{q}_{\alpha \alpha })=\delta _{\alpha \alpha }(q_{\alpha \alpha })=\sqrt{m_{p}}\cdot {} (\wp _{\infty *}^{-1}\circ {}\wp _{\text {Tate}})_{*}^{+}(q_{A}{\hat{\otimes }}1)\otimes {} \omega _{g_{\alpha }}\otimes \omega _{h_{\alpha }} \end{aligned}$$
(24)

in

$$\begin{aligned} H^{1}(\mathbf{Q} _{p},V(f)^{+}_{\alpha \alpha })=H^{0}\big (\text {Gal}(\mathbf{Q} _{p^{2}}/\mathbf{Q} ),H^{1}(\mathbf{Q} _{p^{2}},V(f)^{+})\otimes _\mathbf{Q _{p}}V(g)_{\alpha }\otimes _{L}V(h)_{\alpha }\big ), \end{aligned}$$

where

$$\begin{aligned} (\wp _{\infty }^{-1}\circ {}\wp _{\text {Tate}})_{*}^{+} : \mathbf{Q} _{p^{2}}^{*}{\hat{\otimes }}\mathbf{Q} _{p}\simeq {}H^{1}(\mathbf{Q} _{p^{2}},V(f)^{+}) \end{aligned}$$

is the map induced in cohomology by the composition of \(\wp _{\infty }^{-1}\) and

$$\begin{aligned} \wp _{\text {Tate}}^{+}=\wp _{\text {Tate}}\circ {}a. \end{aligned}$$

If \({\mathcal {A}}\) denotes either A or E, denote by

$$\begin{aligned} \pi _{{\mathcal {A}}} : V_{p}({\mathcal {A}})(-1)\otimes _\mathbf{Q _{p}}V_{p}({\mathcal {A}}){\mathop {\longrightarrow }\limits ^{}}\mathbf{Q} _{p} \end{aligned}$$

the composition of the evaluation pairing \(\mathbf{Q} _{p}(1)\otimes _\mathbf{Q _{p}}\mathbf{Q} _{p}(-1){\mathop {\longrightarrow }\limits ^{}}\mathbf{Q} _{p}\) with the base change of the Weil pairing on \(V_{p}({\mathcal {A}})\) by \(\mathbf{Q} _{p}(-1)\). Set

$$\begin{aligned} q(A)^{*}=\wp _{\text {Tate}}^{+}(\zeta _{p^{\infty }})\otimes {}\zeta _{p^{\infty }}^{*}\in {}V_{p}(A)^{+}(-1), \end{aligned}$$

where \(\zeta _{p^{\infty }}\) is a generator of \(\mathbf{Q} _{p}(1)\) and \(\zeta _{p^{\infty }}^{*}\) in \(\mathbf{Q} _{p}(-1)\) is its dual basis, and set

$$\begin{aligned} q(f)^{*}=\deg (\wp _{\infty })\cdot {}\wp _{\infty }^{-1}(q(A)^{*})\in {}V(f)^{+}(-1). \end{aligned}$$

As \(\pi _{E}((a(y)\otimes {}z)\otimes {}x)=b(x)\cdot {}z(y)\) for each x in \(V_{p}(E)\), y in \(\mathbf{Q} _{p}(1)\) and z in \(\mathbf{Q} _{p}(-1)\), the functoriality of the Poincaré duality under finite morphisms yields

$$\begin{aligned} \pi _{f}(q(f)^{*}\otimes {}q(f)) =\pi _{A}(q(A)^{*}\otimes {}q(A))=\pi _{E}\big ((a(\zeta _{p^{\infty }})\otimes {}\zeta _{p^{\infty }}^{*})\otimes {}\!\!\root p^{\infty } \of {q_{A}}\,\big )=1, \end{aligned}$$

then (by the definition of the weight-one differentials \(\eta _{\xi _{\alpha }}\), cf. Sect. 3.1.1)

$$\begin{aligned} q_{\beta \beta }^{*}=\frac{1}{\sqrt{m_{p}}}\cdot {}q(f)^{*}\otimes {}\omega _{g_{\alpha }}\otimes \omega _{h_{\alpha }}. \end{aligned}$$

Together with Eq. (24) this gives

$$\begin{aligned} \jmath (q_{\alpha \alpha })^{+}=\frac{m_{p}}{\deg (\wp _{\infty })}\cdot {}(q_{A}{\hat{\otimes }}1)\otimes {}q_{\beta \beta }^{*}, \end{aligned}$$
(25)

id est

$$\begin{aligned} \jmath (q_{\alpha \alpha })^{+}_{\alpha \alpha }=\frac{m_{p}}{\deg (\wp _{\infty })}\cdot {}q_{A}{\hat{\otimes }}1. \end{aligned}$$
(26)

According to Theorem 3.18 of [9] \({\mathfrak {L}}_{\varvec{f}}^{\text {an}}=\frac{\log _{p}(q_{A})}{\text {ord}_{p}(q_{A})}\), so that

(27)

by Eqs. (21) and (26).

4.4 Computation of

Assume in this subsection \(\alpha _{f}=\beta _{g}\cdot {}\alpha _{h}\), so that \(H^{0}(\mathbf{Q} _{p},V^{-})\) is generated by the p-adic periods

$$\begin{aligned} q_{\alpha \beta }=\sqrt{m_{p}}\cdot {}q(f)\otimes \omega _{g_{\alpha }}\otimes \eta _{h_{\alpha }}\ \ \ \text {and}\ \ \ q_{\beta \alpha }=\sqrt{m_{p}}\cdot {}q(f)\otimes \eta _{g_{\alpha }}\otimes \omega _{h_{\alpha }}. \end{aligned}$$

For \(\gamma \delta =\alpha \beta ,\beta \alpha \) and \(\cdot =\emptyset ,\pm \), define \(V(f)^{\cdot }_{\gamma \delta }=V(f)^{\cdot }\otimes _\mathbf{Q _{p}}V(g)_{\gamma } \otimes {}V(h)_{\delta }\). Then

$$\begin{aligned} H^{0}(\mathbf{Q} _{p},V^{-})=V(f)^{-}_{\alpha \beta }\oplus {}V(f)^{-}_{\beta \alpha }, \end{aligned}$$

\(G_\mathbf{Q _{p}}\) acts on \(V(f)^{+}_{\alpha \beta }\) and \(V(f)^{+}_{\beta \alpha }\) via the p-adic cyclotomic character, and the local Tate pairing \(\left<{}\cdot ,\cdot \right>_{\text {Tate}}\) introduced in Sect. 3.2.2 induces a perfect duality (denoted by the same symbol) between \(H^{1}(\mathbf{Q} _{p},V(f)^{-}_{\alpha \beta })\) and \(H^{1}(\mathbf{Q} _{p},V(f)^{+}_{\beta \alpha })\). The argument of the proof of Eq. (25) shows that

$$\begin{aligned} \jmath (q_{\beta \alpha })^{+}=\frac{m_{p}}{\deg (\wp _{\infty })}\cdot {}(q_{A}{\hat{\otimes }}1)\otimes {}q_{\alpha \beta }^{*} \end{aligned}$$
(28)

in the direct summand \(H^{1}(\mathbf{Q} _{p},V(f)^{+}_{\beta \alpha })=\mathbf{Q} _{p}^{*}{\hat{\otimes }}V(f)^{+}_{\beta \alpha }(-1)\) of \(H^{1}(\mathbf{Q} _{p},V^{+})\), where

$$\begin{aligned} q_{\alpha \beta }^{*}=\frac{1}{\sqrt{m_{p}}}\cdot {}q(f)^{*}\otimes \eta _{g_{\alpha }}\otimes \omega _{h_{\alpha }}\ \ \text {satisfies}\ \ \pi _{fgh}(-1)(q_{\alpha \beta }^{*}\otimes {}q_{\alpha \beta })=1. \end{aligned}$$
(29)

Let \(\text {pr}_{\alpha \beta } : H^{1}(\mathbf{Q} _{p},V^{-}){\mathop {\longrightarrow }\limits ^{}}H^{1}(\mathbf{Q} _{p},\mathbf{Q} _{p})\otimes _\mathbf{Q _{p}}V(f)^{-}_{\alpha \beta }\) denote the projection, and write

$$\begin{aligned} \text {pr}_{\alpha \beta }\otimes {}\mathscr {I}/\mathscr {I}^{2}\circ {}\beta ^{-}_{\varvec{f}\varvec{g}_{\alpha }\varvec{h}_{\alpha }}(q_{\alpha \beta })= \sum _{\varvec{u}}\gamma _{\varvec{u}}\otimes {}q_{\alpha \beta }\cdot {}(\varvec{u}-u_{o}) \end{aligned}$$
(30)

with \(\gamma _{\varvec{u}}\) in \(H^{1}(\mathbf{Q} _{p},\mathbf{Q} _{p})=\text {Hom}_{}(\mathbf{Q} _{p}^{*},\mathbf{Q} _{p})\) for \(\varvec{u}=\varvec{k},l,\varvec{m}\), where (with the notations introduced in Sect. 3.2.1) \(\beta _{\varvec{f}\varvec{g}_{\alpha }\varvec{h}_{\alpha }}^{-}\) is a shorthand for

$$\begin{aligned} \beta _\mathbf{R \Gamma _{\text {cont}}(\mathbf{Q} _{p},\varvec{V}^{-})/\mathbf{R} \Gamma _{\text {cont}}(\mathbf{Q} _{p},V^{-})} : H^{0}(\mathbf{Q} _{p},V^{-}){\mathop {\longrightarrow }\limits ^{}}H^{1}(\mathbf{Q} _{p},V^{-})\otimes _{L}\mathscr {I}/\mathscr {I}^{2}, \end{aligned}$$

and \(u_{o}=2\) if \(\varvec{u}=\varvec{k}\) and \(u_{o}=1\) if \(\varvec{u}=l,\varvec{m}\). Then (cf. Eq. (21))

(31)

where the last equality follows from Eq. (29) and the analogue of Eq. (20) obtained by replacing \(\alpha \alpha \) and \(\beta \beta \) with \(\beta \alpha \) and \(\alpha \beta \) respectively. It then remains to compute \(\gamma _{\varvec{u}}\) for \(\varvec{u}\) equal to \(\varvec{k}\), \(l\) and \(\varvec{m}\).

For \(\varvec{\xi }=\varvec{f},\varvec{g}_{\alpha },\varvec{h}_{\alpha }\), fix \(\mathscr {O}_{\varvec{\xi }}\)-bases \(b_{\varvec{\xi }}^{\pm }\) of \(V(\varvec{\xi })^{\pm }\). After identifying \(V(\varvec{\xi })\) with \(\mathscr {O}_{\varvec{\xi }}\oplus {}\mathscr {O}_{\varvec{\xi }}\) via the \(\mathscr {O}_{\varvec{\xi }}\)-basis \((b_{\varvec{\xi }}^{+},b_{\varvec{\xi }}^{-})\), the action of \(G_\mathbf{Q _{p}}\) on \(V(\varvec{\xi })\) is given by (cf. Eq. (9))

for a continuous map \(c_{\varvec{\xi }} : G_\mathbf{Q _{p}}{\mathop {\longrightarrow }\limits ^{}}\mathscr {O}_{\varvec{\xi }}\). Without loss of generality, assume that

$$\begin{aligned} \varvec{q}_{\alpha \beta }=b_{\varvec{f}}^{-}{\hat{\otimes }}b_{\varvec{g}_{\alpha }}^{-}{\hat{\otimes }}b_{\varvec{h}_{\alpha }}^{+}\otimes {}1 \end{aligned}$$

in \(\varvec{V}^{-}=V(\varvec{f})^{-}{\hat{\otimes }}_{L}V(\varvec{g}_{\alpha }){\hat{\otimes }}_{L}V(\varvec{h}_{\alpha })\otimes _{\mathscr {O}_{\!\varvec{f}\varvec{g}\varvec{h}}}\Xi _{\varvec{f}\varvec{g}\varvec{h}}\) maps to

$$\begin{aligned} q_{\alpha \beta }\in {}V(f)^{-}_{\alpha \beta }=V(f)^{-}\otimes _\mathbf{Q _{p}}V(g)_{\alpha }\otimes _{L}V(h)_{\beta } \end{aligned}$$

under \(\rho _{w} : \varvec{V}^{-}{\mathop {\longrightarrow }\limits ^{}}V^{-}\). (Recall that \(V(\xi )=V(\varvec{\xi }_{\alpha })\otimes _{1}L\) is the direct sum of the modules \(V(\xi )_{\alpha }=V(\varvec{\xi }_{\alpha })^{-}\otimes _{1}L\) and \(V(\xi )_{\beta }=V(\varvec{\xi }_{\alpha })^{+}\otimes _{1}L\) for \(\xi =g,h\), cf. Sect. 3.1.1.) Then

$$\begin{aligned} d\varvec{q}_{\alpha \beta }=\Gamma \cdot {}\varvec{q}_{\alpha \beta }+\Delta \cdot {}\varvec{q}_{\beta \beta }, \end{aligned}$$
(32)

where \(\varvec{q}_{\beta \beta }=b_{\varvec{f}}^{-}{\hat{\otimes }}b_{\varvec{g}_{\alpha }}^{+}{\hat{\otimes }}b_{\varvec{h}_{\alpha }}^{+}\otimes {}1\), where

$$\begin{aligned} \Gamma =\frac{{\check{a}}_{p}(\varvec{f})\cdot {}{\check{a}}_{p}(\varvec{g}_{\alpha })}{{\check{a}}_{p}(\varvec{h}_{\alpha })}\cdot {}\chi _{h}\cdot {}\chi _{\text {cyc}}^{(\varvec{m}-\varvec{k}-l+2)/2}-1 \end{aligned}$$

and where

$$\begin{aligned} \Delta ={\check{a}}_{p}(\varvec{f})\cdot {}{\check{a}}_{p}(\varvec{h}_{\alpha })^{-1}\cdot {}\chi _{h}\cdot {}\chi _{\text {cyc}}^{(\varvec{m}-\varvec{k}-l+2)/2}\cdot {}c_{\varvec{g}_{\alpha }}. \end{aligned}$$

The exceptional zero condition \(\alpha _{f}=\beta _{g}\cdot {}\alpha _{h}\) and the self duality condition \(\chi _{g}\cdot {}\chi _{h}=1\) imply that \(\Gamma \) takes values in \(\mathscr {I}\). Moreover, since the \(G_\mathbf{Q _{p}}\)-module \(V(g)=V(\varvec{g}_{\alpha })\otimes _{1}L\) splits as the direct sum of \(V(g)_{\beta }=V(\varvec{g}_{\alpha })^{+}\otimes _{1}L\) and \(V(g)_{\alpha }=V(\varvec{g}_{\alpha })^{-}\otimes _{1}L\), the map \(c_{\varvec{g}_{\alpha }}\) takes values in \((l-1)\cdot {}\mathscr {O}_{\varvec{g}}\), hence \(\Delta \) takes values in \(\mathscr {I}\). Because by construction \(\varvec{q}_{\beta \beta }\) maps to an element of \(V(f)^{-}_{\beta \beta }\) under the specialisation map \(\rho _{w_{o}} : \varvec{V}^{-}{\mathop {\longrightarrow }\limits ^{}}V^{-}\), Lemma 3.2 and Eqs. (30) and (32) yield the identities

$$\begin{aligned} \gamma _{\varvec{u}}=-\frac{\partial }{\partial \varvec{u}}\Gamma (\cdot )(w_{o}), \end{aligned}$$

hence (as in the previous subsection) a direct computation gives

$$\begin{aligned} \gamma _{\varvec{k}}=\frac{1}{2}\cdot {}\log _{\varvec{f}},\ \ \gamma _{l}=\frac{1}{2}\cdot {}\log _{\varvec{g}_{\alpha }}\ \ \text {and} \ \ \gamma _{\varvec{m}}=-\frac{1}{2}\cdot {}\log _{\varvec{h}_{\alpha }}. \end{aligned}$$
(33)

Recalling that \(\log _{\varvec{f}}(q_{A})=0\) by [9, Theorem 3.18], Eq. (31) finally proves

(34)

4.5 Proof of equation (8)

Assume in this subsection that \((A,\varrho )\) is exceptional at p, and fix a Selmer class x in \(\text {Sel}(\mathbf{Q} ,V(f,g,h))\). Let

$$\begin{aligned} \tilde{x}=\imath _{\text {ur}}(x)\in {}\tilde{H}^{1}_{f}(\mathbf{Q} ,V(f,g,h)) \end{aligned}$$

be the corresponding extended Selmer class (cf. Sect. 3.1.2). By construction \(\tilde{x}^{+}\) belongs to the finite subspace of \(H^{1}(\mathbf{Q} _{p},V^{+})\), and its image under the natural map \(i^{+} : H^{1}_{\text {fin}}(\mathbf{Q} _{p},V^{+}){\mathop {\longrightarrow }\limits ^{}}H^{1}_{\text {fin}}(\mathbf{Q} _{p},V)\) equals the restriction of x at p:

$$\begin{aligned} \text {res}_{p}(x)=i^{+}(\tilde{x}^{+}). \end{aligned}$$
(35)

The Galois group \(G_\mathbf{Q _{p}}\) acts on \(V(f)^{+}_{\natural }\) via the p-adic cyclotomic character, hence

$$\begin{aligned} H^{1}_{\text {fin}}(\mathbf{Q} _{p},V(f)^{+}_{\natural })=\mathbf{Z} _{p}^{*}\otimes _\mathbf{Z _{p}}V(f)^{+}_{\natural }(-1) \end{aligned}$$

by Kummer theory. If \(q_{\flat }^{*}\) in \(V(f)^{+}_{\natural }\) denotes (as in the previous subsections) the dual basis of \(q_{\flat }\) in \(V(f)_{\natural }^{-}\) under the pairing \(\pi _{fgh}\), and if one writes

$$\begin{aligned} \text {pr}_{\natural }(\tilde{x}^{+})=\tilde{x}^{+}_{\natural }\otimes {}q_{\flat }^{*}\in {}H^{1}_{\text {fin}}(\mathbf{Q} _{p},V(f)^{+}_{\natural }) \end{aligned}$$

for some \(\tilde{x}^{+}_{\natural }\) in \(\mathbf{Z} _{p}^{*}\otimes _\mathbf{Z _{p}}L\), then Eq. (35) yields the equality

$$\begin{aligned} \log _{\natural }(\text {res}_{p}(x))=\langle \log _{p}^{+}(\tilde{x}^{+}),q_{\flat }\rangle _{fgh}= \langle \log _{p}(\tilde{x}^{+}_{\natural })\otimes {}q_{\flat }^{*},q_{\flat }\rangle _{fgh}=\log _{p}(\tilde{x}^{+}_{\natural }), \end{aligned}$$
(36)

where \(\log _{p}^{+} : H^{1}_{\text {fin}}(\mathbf{Q} _{p},V^{+})\simeq {}D_{\text {dR}}(V^{+})\) is the Bloch–Kato logarithm and (with a slight abuse of notation) we denote again by \(\log _{p} : \mathbf{Z} _{p}^{*}\otimes _\mathbf{Z _{p}}L{\mathop {\longrightarrow }\limits ^{}}L\) the L-linear extension of the p-adic logarithm. In the previous equation we used the functoriality of the Bloch–Kato logarithm and the fact that (by construction) the linear form \(\left<{}\cdot ,q_{\flat }\right>_{fgh}\) on \(D_{\text {dR}}(V^{+})\) factors through the projection onto \(D_{\text {dR}}(V(f)^{+}_{\natural })=V(f)^{+}_{\natural }(-1)\).

Assume (\(\alpha _{f}=\alpha _{g}\cdot {}\alpha _{h}\) and) \(q_{\flat }=q_{\beta \beta }\). According to Eqs. (21) and (36)

(37)

thus proving Eq. (8) in this case.

Assume \(q_{\flat }=q_{\alpha \beta }\). Since (with the notations of Section 3.4) \(\Delta \) takes values in \((l-1)\cdot {}\mathscr {O}_{\!\varvec{f}\varvec{g}\varvec{h}}\), it follows from Lemma 3.2 and Eqs. (32) and (33) that

$$\begin{aligned} 2\cdot {}\beta _{\varvec{f}\varvec{g}_{\alpha }\varvec{h}_{\alpha }}^{-}(q_{\alpha \beta })=\sum _{\varvec{\xi }}\varepsilon _{\varvec{\xi }}\cdot {}\log _{\varvec{\xi }}\otimes {}q_{\alpha \beta }\cdot {}(\varvec{u}-u_{o})+\vartheta \cdot {}(l-1) \end{aligned}$$
(38)

for some cohomology class \(\vartheta \) in \(H^{1}(\mathbf{Q} _{p},V(f)^{-}_{\beta \beta })\), where \(\varepsilon _{\varvec{h}_{\alpha }}=-1\) and \(\varepsilon _{\varvec{\xi }}=+1\) for \(\varvec{\xi }=\varvec{f},\varvec{g}_{\alpha }\). One has then

(39)

thus proving Eq. (8) when \(q_{\flat }=q_{\alpha \beta }\). Switching the roles of the Hida families \(\varvec{g}_{\alpha }\) and \(\varvec{h}_{\alpha }\), this also proves Eq. (8) when \(q_{\flat }=q_{\beta \alpha }\).

Assume finally \(q_{\flat }=q_{\alpha \alpha }\). With the notations of Sect. 3.4, let \((b_{\varvec{\xi }}^{+},b_{\varvec{\xi }}^{-})\) be \(\mathscr {O}_{\varvec{\xi }}\)-bases of \(V(\varvec{\xi })\) such that \(\varvec{q}_{\alpha \alpha }=b_{\varvec{f}}^{-}{\hat{\otimes }}b_{\varvec{g}_{\alpha }}^{-}{\hat{\otimes }}b_{\varvec{h}_{\alpha }}^{-}\otimes {}1\) is a lift of \(q_{\alpha \alpha }\) under the specialisation map \(\rho _{w_{o}} : \varvec{V}^{-}{\mathop {\longrightarrow }\limits ^{}}V^{-}\). Since \(c_{\varvec{\xi }}\) takes values in \((\varvec{u}-u_{o})\cdot {}\mathscr {O}_{\varvec{\xi }}\) for \(\varvec{\xi }=\varvec{g}_{\alpha },\varvec{h}_{\alpha }\), one has

$$\begin{aligned} d\varvec{q}_{\alpha \alpha }\equiv {} \Big (\chi _{\text {cyc}}^{(4-\varvec{k}-l-\varvec{m})/2}\cdot {}\prod _{\varvec{\xi }}{\check{a}}_{p}(\varvec{\xi })-1\Big ) \cdot {}\varvec{q}_{\alpha \alpha }\ \big (\text {mod}\ (l-1,\varvec{m}-1)\cdot {}\text {C}_{\text {cont}}^{1}(\mathbf{Q} _{p},\varvec{V}^{-})\big ), \end{aligned}$$

hence Lemma 3.2 and a direct computation give

$$\begin{aligned} 2\cdot {}\beta ^{-}_{\varvec{f}\varvec{g}_{\alpha }\varvec{h}_{\alpha }}(q_{\alpha \alpha })= \log _{\varvec{f}}\otimes {}q_{\alpha \alpha }\cdot {}(\varvec{k}-2)+\vartheta \cdot {}(l-1)+\vartheta ^{\prime }\cdot {}(\varvec{m}-1) \end{aligned}$$
(40)

for some local cohomology classes \(\vartheta \) and \(\vartheta ^{\prime }\) in \(H^{1}(\mathbf{Q} _{p},V^{-})\). As in (39) one deduces Eq. (8) for \(q_{\flat }=q_{\alpha \alpha }\) from Lemma 3.4 and Eqs. (36) and (40).

5 Proof of theorem 2.1

Let \({\Pi }_{f}\), \({\Pi }_{g}\) and \({\Pi }_{h}\) be the improving planes in \(U_{\!\varvec{f}}\times {}U_{\varvec{g}}\times {}U_{\varvec{h}}\) defined respectively by the equations \(\varvec{k}=l+\varvec{m}\), \(\varvec{k}=l-\varvec{m}+2\) and \(\varvec{k}=\varvec{m}-l+2\). For \(\xi =f,g,h\) define

$$\begin{aligned} {\mathcal {E}}_{\xi }=1-{\bar{\chi }}_{\xi }(p)\cdot {}\frac{a_{p}(\varvec{\xi })}{a_{p}(\varvec{\xi }^{\prime })\cdot {}a_{p}(\varvec{\xi }^{\prime \!\prime })} \end{aligned}$$

in \(\mathscr {O}_{\!\varvec{f}\varvec{g}\varvec{h}}\), where \(\{\varvec{\xi },\varvec{\xi }^{\prime },\varvec{\xi }^{\prime \!\prime }\}=\{\varvec{f},\varvec{g}_{\alpha },\varvec{h}_{\alpha }\}\). Lemma 9.8 of [7] implies that

$$\begin{aligned} \mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )|_{{\Pi }_{\xi }}={\mathcal {E}}_{\xi }|_{{\Pi }_{\xi }}\cdot {}\mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )_{\xi }^{\star } \end{aligned}$$
(41)

for an improved p-adic L-function \(\mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )_{\xi }^{\star }\) in \({\mathcal {O}}({\Pi }_{\xi })\). Indeed loc. cit. (together with its analogue obtained by switching the roles of g and h) proves that the meromorphic function \(\mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )_{\xi }^{\star }\) on \({\Pi }_{\xi }\) defined by the previous equation is (bounded, hence) regular at \(w_{o}\). Shrinking the discs \(U_{\varvec{\xi }}\) if necessary, we then conclude that the improved p-adic L-function \(\mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )_{\xi }^{\star }\) is analytic on \({\Pi }_{\xi }\), as claimed.

Assume first \(\alpha _{f}=\alpha _{g}\cdot {}\alpha _{h}\), so that

$$\begin{aligned} 2\cdot {}{\mathcal {E}}_{f}\!\!\!\!\pmod {\mathscr {I}^{2}}={\mathfrak {L}}_{\varvec{f}}^{\text {an}}\cdot {}(\varvec{k}-2)-{\mathfrak {L}}_{\varvec{g}_{\alpha }}^{\text {an}}\cdot {}(l-1) -{\mathfrak {L}}_{\varvec{h}_{\alpha }}^{\text {an}}\cdot {}(\varvec{m}-1). \end{aligned}$$
(42)

According to Theorem A and Proposition 9.3 of [7], the partial derivative of \(\mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )\) with respect to \(\varvec{k}\) vanishes at \(w_{o}\), hence

$$\begin{aligned} 2\cdot {}\mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )\!\!\!\pmod {\mathscr {I}^{2}} \end{aligned}$$

is equal to

$$\begin{aligned} \big (({\mathfrak {L}}_{\varvec{f}}^{\text {an}} -{\mathfrak {L}}^{\text {an}}_{g_{\alpha }})\cdot {}(l-1)+ ({\mathfrak {L}}^{\text {an}}_{\varvec{f}}-{\mathfrak {L}}^{\text {an}}_{\varvec{h}_{\alpha }})\cdot {}(\varvec{m}-1)\big )\cdot {}\mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )_{f}^{\star }(w_{o}) \end{aligned}$$

by Eqs. (41) and (42). Moreover, with the notations introduced before the statement of Theorem 2.1, one has \(\mathsf {L}={\Pi }_{f}\cap {}{\Pi }_{g}\) and \(\mathscr {E}_{f}={\mathcal {E}}_{f}|_{\mathsf {L}}\), thus

$$\begin{aligned} \mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )_{f}^{\star }(w_{o})=\mathscr {E}_{g}(w_{o})\cdot {}\mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )^{\star }(w_{o}). \end{aligned}$$

Noting that \(\mathscr {E}_{g}(w_{o})=1-\beta _{h}/\alpha _{h}\) (when \(\alpha _{f}=\alpha _{g}\cdot {}\alpha _{h}\)), the previous discussion and Eq. (27) conclude the proof of Theorem 2.1 when \(\alpha _{f}=\alpha _{g}\cdot {}\alpha _{h}\).

Assume now \(\alpha _{f}=\beta _{g}\cdot {}\alpha _{h}\). In this case, for \(\xi =g,h\), one has

$$\begin{aligned} 2\cdot {}{\mathcal {E}}_{\xi }\!\!\!\pmod {\mathscr {I}^{2}}={\mathfrak {L}}^{\text {an}}_{\varvec{\xi }_{\alpha }}\cdot {}(\varvec{u}-1)-{\mathfrak {L}}^{\text {an}}_{\varvec{f}}\cdot {}(\varvec{k}-2) -{\mathfrak {L}}^{\text {an}}_{\varvec{\xi }^{\prime }_{\alpha }}\cdot {}(\varvec{u}^{\prime }-1), \end{aligned}$$
(43)

where \(\{(\varvec{\xi }_{\alpha },\varvec{u}),(\varvec{\xi }_{\alpha }^{\prime },\varvec{u}^{\prime })\}=\{(\varvec{g}_{\alpha },l),(\varvec{h}_{\alpha },\varvec{m})\}\), and

$$\begin{aligned} -\mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )_{h}^{\star }(w_{o})=\mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )_{g}^{\star }(w_{o}) =\mathscr {E}_{f}(w_{o})\cdot {}\mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )^{\star }(w_{o}). \end{aligned}$$
(44)

The second equality in the previous equation follows as above from the definitions, according to which \(\mathsf {L}={\Pi }_{f}\cap {}{\Pi }_{g}\) and \(\mathscr {E}_{g}={\mathcal {E}}_{g}|_{\mathsf {L}}\). The first equality follows by noting that the restrictions of \({\mathcal {E}}_{g}\) and \({\mathcal {E}}_{h}\) to the line \({\Pi }_{g}\cap {}{\Pi }_{h}\) satisfy

$$\begin{aligned} {\mathcal {E}}_{g}|_{{\Pi }_{g}\cap {}{\Pi }_{h}}=-\left. \frac{{\bar{\chi }}_{g}(p)\cdot {}a_{p}(\varvec{g}_{\alpha })}{a_{p}(\varvec{f})\cdot {}a_{p}(\varvec{h}_{\alpha })}\right| _{{\Pi }_{g}\cap {}{\Pi }_{h}}\cdot {} {\mathcal {E}}_{h}|_{{\Pi }_{g}\cap {}{\Pi }_{h}} \end{aligned}$$

(as \(a_{p}(\varvec{f})|_{{\Pi }_{g}\cap {}{\Pi }_{h}}=\alpha _{f}=\alpha _{f}^{-1}\) and \(\chi _{g}\cdot {}\chi _{h}=1\) by Assumption 1.1.1) with

$$\begin{aligned} -\frac{{\bar{\chi }}_{g}(p)\cdot {}a_{p}(\varvec{g}_{\alpha })}{a_{p}(\varvec{f})\cdot {}a_{p}(\varvec{h}_{\alpha })}(w_{o})=-1. \end{aligned}$$

(In other words \({\mathcal {E}}_{g}|_{{\Pi }_{g}\cap {}{\Pi }_{h}}\) and \(-{\mathcal {E}}_{h}|_{{\Pi }_{g}\cap {}{\Pi }_{h}}\) have the same leading term at \(w_{o}\), which together with the equality \({\mathcal {E}}_{g}\cdot {}\mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )_{g}^{\star }|_{{\Pi }_{g}\cap {}{\Pi }_{h}}= {\mathcal {E}}_{h}\cdot {}\mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )_{h}^{\star }|_{{\Pi }_{g}\cap {}{\Pi }_{h}}\) implies the first identity in Eq. (44).) Write

$$\begin{aligned} 2\cdot {}\mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )\!\!\!\pmod {\mathscr {I}^{2}}=a\cdot {}(\varvec{k}-2)+b\cdot {}(l-1)+c\cdot {}(\varvec{m}-1) \end{aligned}$$

with a, b and c in L. Equations (41) and (43) with \(\xi =g\) and Eq. (44) give

$$\begin{aligned} a+b=\mathscr {E}_{f}(w_{o})\cdot {}\big ({\mathfrak {L}}_{\varvec{g}_{\alpha }}^{\text {an}}-{\mathfrak {L}}_{\varvec{f}}^{\text {an}}\big )\cdot {} \mathscr {L}_{p}^{\star }(w_{o})\ \ \ \text {and}\ \ \ c-a=\mathscr {E}_{f}(w_{o})\cdot {}\big ({\mathfrak {L}}_{\varvec{f}}^{\text {an}}-{\mathfrak {L}}_{\varvec{h}_{\alpha }}^{\text {an}}\big )\cdot {}\mathscr {L}_{p}^{\star }(w_{o}), \end{aligned}$$

where \(\mathscr {L}_{p}^{\star }\) is a shorthand for \(\mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )^{\star }\). Similarly

$$\begin{aligned} b-a=\mathscr {E}_{f}(w_{o})\cdot {}\big ({\mathfrak {L}}_{\varvec{g}_{\alpha }}^{\text {an}}-{\mathfrak {L}}_{\varvec{f}}^{\text {an}}\big )\cdot {}\mathscr {L}_{p}^{\star }(w_{o}) \ \ \text {and}\ \ a+c=\mathscr {E}_{f}(w_{o})\cdot {}\big ({\mathfrak {L}}_{\varvec{f}}^{\text {an}}-{\mathfrak {L}}_{\varvec{h}_{\alpha }}^{\text {an}}\big )\cdot {}\mathscr {L}_{p}^{\star }(w_{o}) \end{aligned}$$

by Eqs. (41) and (43) with \(\xi =h\) and Eq. (44). As a consequence

$$\begin{aligned} -2\cdot {}\mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )\!\!\!\pmod {\mathscr {I}^{2}} \end{aligned}$$

equals

$$\begin{aligned} \mathscr {E}_{f}(w_{o})\cdot {}\big ( ({\mathfrak {L}}_{\varvec{f}}^{\text {an}}-{\mathfrak {L}}_{\varvec{g}_{\alpha }}^{\text {an}})\cdot {}(l-1) -({\mathfrak {L}}_{\varvec{f}}^{\text {an}}-{\mathfrak {L}}_{\varvec{h}_{\alpha }}^{\text {an}})\cdot {}(\varvec{m}-1)\big )\cdot {}\mathscr {L}_{p}^{\alpha \alpha }(A,\varrho )^{\star }(w_{o}). \end{aligned}$$

Noting that \(\mathscr {E}_{f}(w_{o})=1-\frac{\beta _{h}}{\alpha _{h}}\) (when \(\alpha _{f}=\beta _{g}\cdot {}\alpha _{h}\)), the previous discussion and Eq. (34) prove Theorem  2.1 when \(\alpha _{f}=\beta _{g}\cdot {}\alpha _{h}\).