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

\NewEnviron

myquote (\dagger) \BODY \NewEnvironmyquote2   \BODY (*) \cftpagenumbersoffpart

On pp-adic LL-functions for symplectic representations of \GL(N)\GL(N) over number fields

Chris Williams
Abstract

Let FF be a number field, and π\pi a regular algebraic cuspidal automorphic representation of GLN(𝔸F)\mathrm{GL}_{N}(\mathbb{A}_{F}) of symplectic type. When π\pi is spherical at all primes 𝔭|p\mathfrak{p}|p, we construct a pp-adic LL-function attached to any regular non-critical spin pp-refinement π~\tilde{\pi} of π\pi to QQ-parahoric level, where QQ is the (n,n)(n,n)-parabolic. More precisely, we construct a distribution Lp(π~)L_{p}(\tilde{\pi}) on the Galois group Galp\mathrm{Gal}_{p} of the maximal abelian extension of FF unramified outside pp\infty, and show that it interpolates all the standard critical LL-values of π\pi at pp (including, for example, cyclotomic and anticyclotomic variation when FF is imaginary quadratic). We show that Lp(π~)L_{p}(\tilde{\pi}) satisfies a natural growth condition; in particular, when π~\tilde{\pi} is ordinary, Lp(π~)L_{p}(\tilde{\pi}) is a (bounded) measure on Galp\mathrm{Gal}_{p}. As a corollary, when π\pi is unitary, has very regular weight, and is QQ-ordinary at all 𝔭|p\mathfrak{p}|p, we deduce non-vanishing L(π×(χNF/),1/2)0L(\pi\times(\chi\circ N_{F/\mathbb{Q}}),1/2)\neq 0 of the twisted central value for all but finitely many Dirichlet characters χ\chi of pp-power conductor.

footnotetext: [email protected]. . 2020 MSC: Primary 11F33, 11F67; Secondary 11R23.

1.  Introduction

The special values of LL-functions are highly important in modern number theory, and are the subject of a vast network of conjectures, including:

  1. (I)

    The Beilinson and Bloch–Kato conjectures, which describe arithmetic data in terms of the complex analytic properties of special values. As a special case, non-vanishing of a special value should force finiteness of an associated Selmer group.

  2. (II)

    Deligne’s conjecture [Del79], which predicts that the special values, ostensibly arbitrary transcendental numbers, are algebraic after scaling by controlled complex ‘periods’.

  3. (III)

    The Coates–Perrin-Riou/Panchishkin conjectures [CPR89, Coa89, Pan94], that predict these special LL-values satisfy deep pp-adic congruences, generalising Kummer’s congruences for the Riemann zeta function.

The congruences predicted by (III) are encoded in the existence of a pp-adic LL-function. Beautiful objects in their own right, pp-adic LL-functions are fundamental in Iwasawa theory, and have led to substantial results towards (I), including special cases of the Birch–Swinnerton-Dyer conjecture.

A very general setting to consider these questions is that of regular algebraic cuspidal automorphic representations (RACARs) π\pi of GLN\operatorname{GL}_{N} over a number field FF. In this paper, we attack (III) by constructing pp-adic LL-functions for those π\pi which are symplectic (RASCARs, for RA-symplectic-CARs).

1.1.  Main result

Let us state our main result more precisely. Let FF be an arbitrary number field, G . . =ResF/𝐐(GLN)G\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\mathrm{Res}_{F/\mathbf{Q}}(\operatorname{GL}_{N}), and π\pi a RASCAR of G(𝐀)G(\mathbf{A}). This forces N=2nN=2n to be even, and ππη\pi\cong\pi^{\vee}\otimes\eta to be essentially self-dual, for a Hecke character η\eta of F×\𝐀F×F^{\times}\backslash\mathbf{A}_{F}^{\times}. Further, π\pi is symplectic if and only if it admits a Shalika model, if and only if it is a functorial transfer from ResF/𝐐GSpin2n+1\mathrm{Res}_{F/\mathbf{Q}}\mathrm{GSpin}_{2n+1} [FJ93, AS06, AS14]. In this setting, the automorphic realisation of Deligne’s conjecture was recently proved – including the period relations at infinity – in work of Jiang–Sun–Tian [JST].

Let pp be a prime, and assume that π\pi is spherical at all primes 𝔭\mathfrak{p} of FF above pp. Then π𝔭=IndBGθ𝔭\pi_{\mathfrak{p}}=\operatorname{Ind}_{B}^{G}\theta_{\mathfrak{p}} is an unramified principal series, the normalised induction of some unramified character θ𝔭=(θ𝔭,1,,θ𝔭,2n)\theta_{\mathfrak{p}}=(\theta_{\mathfrak{p},1},...,\theta_{\mathfrak{p},2n}) (where BB is the upper-triangular Borel). Let QGQ\subset G be the standard parabolic subgroup with Levi H . . =ResF/𝐐(GLn×GLn)H\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\mathrm{Res}_{F/\mathbf{Q}}(\operatorname{GL}_{n}\times\operatorname{GL}_{n}), and let J𝔭GL2n(F𝔭)J_{\mathfrak{p}}\subset\operatorname{GL}_{2n}(F_{\mathfrak{p}}) be the parahoric subgroup of type QQ. If 𝔭|p\mathfrak{p}|p, a regular QQ-refinement to level J𝔭J_{\mathfrak{p}} is a choice of simple Hecke eigenvalue α𝔭\alpha_{\mathfrak{p}} on π𝔭J𝔭\pi_{\mathfrak{p}}^{J_{\mathfrak{p}}}, defined precisely in Definition 3.3. We assume this choice is spin (Definition 3.3), in that it interacts well with the Shalika model; as explained in the introduction of [BGW], we expect this is essential for constructions of pp-adic LL-functions via Shalika models. Write π~\tilde{\pi} for π\pi with such a choice of regular spin QQ-refinement at each 𝔭|p\mathfrak{p}|p. We assume π~\tilde{\pi} to be non-QQ-critical (Definition 7.1). This is satisfied if the integrally normalised eigenvalues α𝔭\alpha_{\mathfrak{p}}^{\circ}, defined in §7.1, have non-QQ-critical slope. In particular if π~\tilde{\pi} is ordinary (i.e. each vp(α𝔭)=0v_{p}(\alpha_{\mathfrak{p}}^{\circ})=0) then it is non-QQ-critical.

We say a Hecke character χ:F×\𝐀F×𝐂×\chi:F^{\times}\backslash\mathbf{A}_{F}^{\times}\to\mathbf{C}^{\times} is critical for π\pi if s=1/2s=1/2 is a Deligne-critical value of L(π×χ,s)L(\pi\times\chi,s), and then we define L(π,χ) . . =L(π×χ,1/2)L(\pi,\chi)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=L(\pi\times\chi,1/2). The critical characters are determined by their infinity type and the weight λ\lambda of π\pi. For general FF, this LL-function can have many different critical regions; see e.g. the pictures in [LLZ15, p.1605] or [BDP13, §4.1]. We focus here on the ‘standard’ critical range111Constructions of pp-adic LL-functions interpolating other ranges would be extremely interesting; see e.g. [BDP13], which gives an example for GL2\operatorname{GL}_{2} over an imaginary quadratic field, assuming π\pi is base-change., corresponding to the ‘balanced weight’ condition in [JST] (and, when FF is imaginary quadratic, region Σ(1)\Sigma^{(1)} in [LLZ15]). We describe this range in Lemma 2.4. For example, when FF is imaginary quadratic, this gives a square of standard critical infinity types (see §7.4). Crucially, these standard critical values admit a representation-theoretic interpretation via branching laws for HGH\subset G, described in Lemma 2.6.

We write Critp(π)\mathrm{Crit}_{p}(\pi) for the set of standard critical characters that have pp-power conductor. Let Galp\operatorname{Gal}_{p} be the Galois group of the maximal abelian extension of FF unramified outside pp\infty. Attached to any χCritp(π)\chi\in\mathrm{Crit}_{p}(\pi) is a canonical character χ[p]\chi_{[p]} on Galp\operatorname{Gal}_{p} (see §2.5).

We construct a pp-adic LL-function attached to π~\tilde{\pi}, in the following sense:

Theorem A.

For any choice of isomorphism ip:𝐂𝐐¯pi_{p}:\mathbf{C}\xrightarrow{\,\smash{\raisebox{-2.79857pt}{$\scriptstyle\sim$}}\,}\overline{\mathbf{Q}}_{p}, there exists a finite extension L/𝐐pL/\mathbf{Q}_{p} and an LL-valued locally analytic distribution Lpip(π~)L_{p}^{i_{p}}(\tilde{\pi}) on Galp\operatorname{Gal}_{p} such that:

  1. (a)

    For any χCritp(π)\chi\in\mathrm{Crit}_{p}(\pi) of conductor 𝔭|p𝔭β𝔭\prod_{\mathfrak{p}|p}\mathfrak{p}^{\beta_{\mathfrak{p}}}, we have

    Lpip(π~,χ)\displaystyle L_{p}^{i_{p}}(\tilde{\pi},\chi) . . =ip1(Galpχ[p]dLpip(π~))\displaystyle\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=i_{p}^{-1}\left(\int_{\operatorname{Gal}_{p}}\chi_{[p]}\cdot dL_{p}^{i_{p}}(\tilde{\pi})\right)
    =Aτ(χf)n𝔭|pe(π𝔭,χ𝔭)L(p)(π,χ)Ωπ,χ,\displaystyle=A\cdot\tau(\chi_{f})^{n}\cdot\prod_{\mathfrak{p}|p}e(\pi_{\mathfrak{p}},\chi_{\mathfrak{p}})\cdot\frac{L^{(p)}(\pi,\chi)}{\Omega_{\pi,\chi_{\infty}}},

    where AA is a constant defined in (5.15), Ωπ,χ𝐂×\Omega_{\pi,\chi_{\infty}}\in\mathbf{C}^{\times} is defined in Definition 3.7, and

    e(π𝔭,χ𝔭)={q𝔭β𝔭(n2n2)α𝔭β𝔭:χ𝔭 ramified,i=n+12n1θ𝔭,i1χ𝔭1(ϖ𝔭)q𝔭1/21θ𝔭,iχ𝔭(ϖ𝔭)q𝔭1/2:χ𝔭 unramified.e(\pi_{\mathfrak{p}},\chi_{\mathfrak{p}})=\left\{\begin{array}[]{cl}q_{\mathfrak{p}}^{\beta_{\mathfrak{p}}\left(\tfrac{n^{2}-n}{2}\right)}\alpha_{\mathfrak{p}}^{-\beta_{\mathfrak{p}}}&:\chi_{\mathfrak{p}}\text{ ramified},\\ \displaystyle\prod_{i=n+1}^{2n}\frac{1-\theta_{\mathfrak{p},i}^{-1}\chi_{\mathfrak{p}}^{-1}(\varpi_{\mathfrak{p}})q_{\mathfrak{p}}^{-1/2}}{1-\theta_{\mathfrak{p},i}\chi_{\mathfrak{p}}(\varpi_{\mathfrak{p}})q_{\mathfrak{p}}^{-1/2}}&:\chi_{\mathfrak{p}}\text{ unramified}.\end{array}\right.
  2. (b)

    Lpip(π~)L_{p}^{i_{p}}(\tilde{\pi}) is admissible of growth (h𝔭)𝔭|p(h_{\mathfrak{p}})_{\mathfrak{p}|p}, where h𝔭 . . =vp(α𝔭)h_{\mathfrak{p}}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=v_{p}(\alpha_{\mathfrak{p}}^{\circ}).

The growth condition is described in Definition 7.4. In particular, if π~\tilde{\pi} is ordinary, then Lpip(π~)L_{p}^{i_{p}}(\tilde{\pi}) is a bounded distribution, that is, a pp-adic measure. Here L(p)L^{(p)} is the LL-function with the Euler factors at pp removed, τ(χf)\tau(\chi_{f}) is a Gauss sum, q𝔭=NF/𝐐(𝔭)q_{\mathfrak{p}}=N_{F/\mathbf{Q}}(\mathfrak{p}), and ϖ𝔭\varpi_{\mathfrak{p}} is a uniformiser. This agrees exactly with the conjectures of Coates–Perrin-Riou/Panchishkin [Pan94, Conj. 6.2], except possibly the term Ωπ,χ\Omega_{\pi,\chi_{\infty}}, which we comment on in Remark 3.8.

If FF is imaginary quadratic, and π~\tilde{\pi} has non-QQ-critical slope, then (a) and (b) determine Lpip(π~)L_{p}^{i_{p}}(\tilde{\pi}) uniquely by [Loe14]. Analogous unicity results for FF totally real are [BDW, Prop. 6.25].

To our knowledge, Theorem A gives the first such construction beyond the special cases of GL2\operatorname{GL}_{2} (over any FF) or FF totally real; see §1.3 for a summary of previous results. The totally real case was handled in [DJR20, BDW]. The construction we give is inspired by those works, particularly the overconvergent approach of [BDW]. There are some additional features in the general number field setting, which we briefly summarise.

  1. (1)

    Notably, whilst Leopoldt’s conjecture predicts that Lp(π~)L_{p}(\tilde{\pi}) is 1-variabled when FF is totally real, in general our pp-adic LL-functions can be many-variabled. For example, when F=𝐐F=\mathbf{Q} we have Galp𝐙p×\operatorname{Gal}_{p}\cong\mathbf{Z}_{p}^{\times}, and one only has cyclotomic variation; when FF is imaginary quadratic, one has a two-variable pp-adic LL-function, with both cyclotomic and anticyclotomic variation.

  2. (2)

    The method uses automorphic cycles/modular symbols. The totally real (resp. GL2\operatorname{GL}_{2}) constructions relied on a certain numerical coincidence; that these cycles have dimension equal to the top degree (resp. bottom degree) in which cuspidal cohomology for GG contributes. This rendered some relevant cohomology groups 1-dimensional, making certain choices unique up to scalar. In the general number field case, we work in the middle of the cuspidal range, where the analogous groups are never 1-dimensional.

We handle (1) by blending the methods of [BDW] with earlier work [Wil17, BW19] of the author and Barrera for GL2\operatorname{GL}_{2}. Problem (2) is more serious, and for this we exploit works of Lin–Tian [LT20] and Jiang–Sun–Tian [JST]. They nailed down good elements in these higher-dimensional cohomology groups and proved not only the non-vanishing hypothesis of attached zeta integrals at infinity, but also the expected period relations at infinity, at least in the cyclotomic direction (see Remark 3.8). We hope that the period relations in all directions can be extracted via similar methods, but do not address that here, instead focusing on the pp-adic interpolation.

We imagine the following special case might be of particular interest. Let \mathcal{F} be a cohomological Siegel modular form on GSp4/𝐐\mathrm{GSp}_{4}/\mathbf{Q}, and base-change it to an imaginary quadratic field FF. Under appropriate assumptions, we may apply our construction to the Langlands transfer of /F\mathcal{F}/F to GL4/F\operatorname{GL}_{4}/F (via [AS06]), giving a two-variable pp-adic spin LL-function for /F\mathcal{F}/F, including anticyclotomic variation. For GL2\operatorname{GL}_{2}, such constructions for base-change modular forms have had important arithmetic consequences, such as proofs of one divisibility of the Iwasawa Main Conjecture [SU14].

1.2.  Application: non-vanishing of central twists

When π~\tilde{\pi} is ordinary at each 𝔭|p\mathfrak{p}|p, then as mentioned above, our construction yields a pp-adic measure. As an immediate application, we get the following generalisation of a result of Dimitrov–Januszewski–Raghuram (who in [DJR20] treated the case FF totally real). Let λ=(λσ)σΣ\lambda=(\lambda_{\sigma})_{\sigma\in\Sigma} denote the weight of π\pi, where λσ=(λσ,1,,λσ,2n)𝐙2n\lambda_{\sigma}=(\lambda_{\sigma,1},...,\lambda_{\sigma,2n})\in\mathbf{Z}^{2n} is dominant and Σ\Sigma is the set of embeddings σ:F𝐂\sigma:F\hookrightarrow\mathbf{C} (see §2.3).

Theorem B.

Let π\pi be a unitary RASCAR of G(𝐀)G(\mathbf{A}), and suppose

λσ,n>λσ,n+1 for all σΣ.\lambda_{\sigma,n}>\lambda_{\sigma,n+1}\qquad\text{ for all }\sigma\in\Sigma. (1.1)

Suppose there exists a rational prime pp such that π𝔭\pi_{\mathfrak{p}} is spherical and QQ-ordinary for all 𝔭|p\mathfrak{p}|p. Then for all but finitely many Dirichlet characters χ\chi of pp-power conductor, we have non-vanishing of the twisted central LL-value

L(π×(χNF/),12)0.L(\pi\times(\chi\circ N_{F/\mathbb{Q}}),\tfrac{1}{2})\neq 0.

Here QQ-ordinarity is in the sense of [Hid98, §1.1]. The unitary assumption ensures s=1/2s=1/2 is a Deligne-critical value of L(π,s)L(\pi,s).

Given Theorem A, the proof of Theorem B is simple. The argument is literally identical to [DJR20, §4.4], so we give only a sketch, and refer the reader there for full details.

Proof.

Since π𝔭\pi_{\mathfrak{p}} is QQ-ordinary for all 𝔭|p\mathfrak{p}|p, by [DJR20, Lem. 4.4] there is a (unique) regular spin ordinary QQ-refinement π~\tilde{\pi} of π\pi. Let Lpip(π~)L_{p}^{i_{p}}(\tilde{\pi}) the pp-adic LL-function from Theorem A. By [JS76] (see [BDW, Lem. 7.4]), the weight condition forces existence of some j0j\neq 0 such that the (non-central) twisted LL-values L(π×(χNF/𝐐),j+1/2)L(\pi\times(\chi\circ N_{F/\mathbf{Q}}),j+1/2) are non-zero for all Dirichlet characters χ\chi.

If χ\chi is a Dirichlet character of pp-power conductor, then χNF/𝐐\chi\circ N_{F/\mathbf{Q}} is a cyclotomic character of Galp\operatorname{Gal}_{p}. The restriction of Lpip(π~)L_{p}^{i_{p}}(\tilde{\pi}) to the cyclotomic line is a bounded rigid analytic function on the disjoint union of a finite number of open unit discs. By the interpolation property and non-vanishing at non-central values, the pp-adic LL-function is non-zero on each disc that contains characters of the form χNF/𝐐\chi\circ N_{F/\mathbf{Q}}. By Weierstrass preparation, it has finitely many zeros in each such disc, hence on the cyclotomic line; so the theorem follows from the interpolation property. ∎

If one drops assumption (1.1), the theorem instead becomes: if there exists a χ\chi as in the theorem such that L(π×(χNF/),1/2)0L(\pi\times(\chi\circ N_{F/\mathbb{Q}}),1/2)\neq 0, then there are infinitely many such χ\chi.

In the GL4\operatorname{GL}_{4} case, after transferring via [AS06], this gives non-vanishing of many twisted central values of spin LL-functions of (very regular weight, cohomological, Klingen-ordinary) genus 2 Siegel modular forms over number fields.

1.3.  Relation to the literature

This work generalises (and is visibly inspired by) many other constructions. We summarise a few. In the case of GL(2)\operatorname{GL}(2), i.e. n=1n=1, every RACAR is a RASCAR, and our main results/methods specialise exactly to those of [BW19]; and the results/methods of that paper in turn followed the earlier works [PS11, Bar18, Wil17] (over 𝐐\mathbf{Q}, totally real, and imaginary quadratic base fields respectively). These methods were later used in [Bel12, BDJ22, BH, BW21a, BW21b] to vary pp-adic LL-functions in families.

In the case of general GL(2n)\operatorname{GL}(2n), when FF is totally real our main result specialises exactly to [BDW, Thm. 6.23]. Earlier constructions in the ordinary situation were given in [AG94, Geh18, DJR20]. This construction was then used in [BDW, BDG+] to construct and study pp-adic families, and – under some technical hypotheses – to vary pp-adic LL-functions over these families.

The variation of the present construction in families would be very interesting. The methods of these earlier papers – which worked in either top or bottom cohomological degree – do not apply. Fundamental difficulties include: controlling the families; constructing classes in the correct cohomological degree; and showing that these classes interpolate the ‘good’ elements from [JST].

Rohrlich [Roh89] proved a GL2\operatorname{GL}_{2} analogue of Theorem B. The case FF totally real is [DJR20]. For GL4/𝐐\operatorname{GL}_{4}/\mathbf{Q}, an analogous result for all weights and unitary CARs was recently proved in [RY].

Finally, in Theorems A and B it should be possible to weaken the assumption that π𝔭\pi_{\mathfrak{p}} is spherical at all 𝔭|p\mathfrak{p}|p to QQ-parahoric-spherical using forthcoming work of Dimitrov–Jorza [DJ].

Acknowledgements

This paper would not exist without my previous collaboration with Daniel Barrera and Mladen Dimitrov, and I thank them wholeheartedly for several years’ worth of discussions. I also thank Andy Graham and Andrei Jorza for many highly relevant discussions during our follow-up collaboration, and David Loeffler for his comments and corrections on an earlier draft. This research was supported by EPSRC Postdoctoral Fellowship EP/T001615/2.

2.  Preliminaries

2.1.  Notation

Let FF be a number field of degree d=r+2sd=r+2s, where FF has rr real embeddings and ss pairs of complex embeddings. Let Σ . . ={σ:F𝐂}=Σ𝐑Σ𝐂\Sigma\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\{\sigma:F\hookrightarrow\mathbf{C}\}=\Sigma_{\mathbf{R}}\sqcup\Sigma_{\mathbf{C}} be the union of the real and complex embeddings. Each σΣ\sigma\in\Sigma has a conjguate cσc\sigma, and σ=cσ\sigma=c\sigma if and only if σ(F)𝐑\sigma(F)\subset\mathbf{R}. Fix a rational prime pp and an isomorphism ip:𝐂𝐐¯pi_{p}:\mathbf{C}\xrightarrow{\,\smash{\raisebox{-2.79857pt}{$\scriptstyle\sim$}}\,}\overline{\mathbf{Q}}_{p}. Let FpF^{p\infty} be the maximal abelian extension of FF unramified outside pp\infty, and let Galp . . =Gal(Fp/F)\operatorname{Gal}_{p}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\mathrm{Gal}(F^{p\infty}/F) be its Galois group. Let ψ\psi be the standard non-trivial additive character of F\𝐀FF\backslash\mathbf{A}_{F} (as e.g. fixed in [DJR20, §4.1]).

If vv is a place of FF, we write FvF_{v} for the completion of FF at vv, 𝒪v\mathcal{O}_{v} for its ring of integers, and 𝐅v\mathbf{F}_{v} for its residue field. We fix a choice of uniformiser ϖv𝒪v\varpi_{v}\in\mathcal{O}_{v}.

Let G=ResF/𝐐GL2nG=\mathrm{Res}_{F/\mathbf{Q}}\operatorname{GL}_{2n}, B=TNB=TN be the upper-triangular Borel sugroup in GG, TT the diagonal torus and NN the unipotent. Let H=ResF/𝐐(GLn×GLn)H=\mathrm{Res}_{F/\mathbf{Q}}(\operatorname{GL}_{n}\times\operatorname{GL}_{n}), with ι:HG\iota:H\hookrightarrow G, (h1,h2)(h1h2)(h_{1},h_{2})\mapsto\left(\begin{smallmatrix}h_{1}&\\ &h_{2}\end{smallmatrix}\right). We may abuse notation and write e.g. B(Fv)B(F_{v}) for the upper-triangular matrices in GL2n(Fv)\operatorname{GL}_{2n}(F_{v}).

Let K=CZG(𝐑)K_{\infty}=C_{\infty}Z_{\infty}\subset G(\mathbf{R}), where ZZ_{\infty} is the center and C=O2n(𝐑)r×U2n(𝐂)sC_{\infty}=\mathrm{O}_{2n}(\mathbf{R})^{r}\times\mathrm{U}_{2n}(\mathbf{C})^{s} is the maximal compact subgroup of G(𝐑)G(\mathbf{R}). If AA is a reductive real Lie group, then AA^{\circ} denotes the connected component of the identity.

Let Q=HNQQ=HN_{Q} be the standard parabolic with Levi HH. For a prime 𝔭|p\mathfrak{p}|p of GG, let J𝔭 . . ={gGL2n(𝒪𝔭):g(mod𝔭)Q(𝐅𝔭)}GL2n(F𝔭)J_{\mathfrak{p}}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\{g\in\operatorname{GL}_{2n}(\mathcal{O}_{\mathfrak{p}}):g\hskip 2.0pt(\mathrm{mod}\hskip 2.0pt\mathfrak{p})\in Q(\mathbf{F}_{\mathfrak{p}})\}\subset\operatorname{GL}_{2n}(F_{\mathfrak{p}}) be the parahoric subgroup of type QQ, and Jp . . =𝔭|pJ𝔭G(𝐐p)J_{p}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\prod_{\mathfrak{p}|p}J_{\mathfrak{p}}\subset G(\mathbf{Q}_{p}).

2.2.  RASCARs

Let π\pi be a regular algebraic cuspidal automorphic representation (RACAR) of G(𝐀)G(\mathbf{A}). If π\pi is essentially-self-dual, i.e. there exists some Hecke character η\eta such that ππη1\pi^{\vee}\cong\pi\otimes\eta^{-1}, then L(π×π,s)=L(π,Sym2η1,s)L(π,2η1,s)L(\pi\times\pi^{\vee},s)=L(\pi,\mathrm{Sym}^{2}\otimes\eta^{-1},s)L(\pi,\wedge^{2}\otimes\eta^{-1},s). This has a simple pole at s=1s=1, which occurs either in the symmetric or exterior square LL-function. Then:

Definition 2.1.

We say that π\pi has symplectic type, and call it a RASCAR (RA-symplectic-CAR), if any (hence all) of the following equivalent conditions hold:

  • (1)

    The exterior square LL-function L(π,2η1,s)L(\pi,\wedge^{2}\otimes\eta^{-1},s) has a pole at s=1s=1;

  • (2)

    π\pi is the functorial transfer of an irreducible generic cuspidal automorphic representation Π\Pi of GSpin2n+1(𝐀F)\mathrm{GSpin}_{2n+1}(\mathbf{A}_{F}) with central character η\eta;

  • (3)

    π\pi admits a non-trivial (η,ψ)(\eta,\psi)-Shalika model.

Recall π\pi has a (η,ψ)(\eta,\psi)-Shalika model if there is an intertwining 𝒮ψη:πInd𝒮(𝐀)G(𝐀)(ηψ)\mathcal{S}_{\psi}^{\eta}:\pi\hookrightarrow\operatorname{Ind}_{\mathcal{S}(\mathbf{A})}^{G(\mathbf{A})}(\eta\otimes\psi), where 𝒮={s(h,X)=(hh)(1nX1n):hGLn,XMn}\mathcal{S}=\big{\{}s(h,X)=\left(\begin{smallmatrix}h&\\ &h\end{smallmatrix}\right)\left(\begin{smallmatrix}1_{n}&X\\ &1_{n}\end{smallmatrix}\right):h\in\operatorname{GL}_{n},X\in\mathrm{M}_{n}\big{\}} is the Shalika group and (ηψ)(s(h,X))=η(det(h))ψ(Tr(X))(\eta\otimes\psi)(s(h,X))=\eta(\det(h))\cdot\psi(\mathrm{Tr}(X)). Our conventions are summarised in [BDW, §2.6].

In the definition, (1)\iff(3) was substantially proved in [JS90], and (1)\iff(2) in [AS06, AS14]. For the equivalence in the exact form above, see [JST, Prop. 2.5].

2.3.  Weights

Let X(T)X^{*}(T) be the set of algebraic weights for TT. A general λX(T)\lambda\in X^{*}(T) has form λ=(λσ)σΣ\lambda=(\lambda_{\sigma})_{\sigma\in\Sigma}, where λσ=(λσ,1,,λσ,2n)𝐙2n\lambda_{\sigma}=(\lambda_{\sigma,1},...,\lambda_{\sigma,2n})\in\mathbf{Z}^{2n}. Let X+(T)X_{+}^{*}(T) be the set of dominant weights, where λσ,iλσ,i+1\lambda_{\sigma,i}\geqslant\lambda_{\sigma,i+1} for all σ\sigma and ii. Attached to any λX+(T)\lambda\in X_{+}^{*}(T) is an algebraic representation VλV_{\lambda} of GG of highest weight λ\lambda, with dual VλV_{\lambda}^{\vee}. We can decompose Vλ=σΣVλ,σV_{\lambda}=\otimes_{\sigma\in\Sigma}V_{\lambda,\sigma}, Vλ=σΣVλ,σV_{\lambda}^{\vee}=\otimes_{\sigma\in\Sigma}V_{\lambda,\sigma}^{\vee}, where Vλ,σV_{\lambda,\sigma} is the algebraic GL2n\operatorname{GL}_{2n}-representation of highest weight λσ\lambda_{\sigma}, with GG acting via σ\sigma.

Definition 2.2.

We say λX+(T)\lambda\in X_{+}^{*}(T) is pure if there exists 𝗐𝐙{\sf w}\in\mathbf{Z} such that

λσ,i=λcσ,2n+1i=𝗐for all σ and i.\lambda_{\sigma,i}=\lambda_{c\sigma,2n+1-i}={\sf w}\qquad\text{for all }\sigma\text{ and }i.

We write X0(T)X+(T)X_{0}^{*}(T)\subset X_{+}^{*}(T) for the space of pure dominant weights.

Let π\pi be RASCAR of G(𝐀)G(\mathbf{A}). Let 𝔤 . . =Lie(G(𝐑))\mathfrak{g}_{\infty}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\mathrm{Lie}(G(\mathbf{R})). As in [Clo90], there is a unique λX0(T)\lambda\in X^{*}_{0}(T), the weight of π\pi, such that

H(𝔤,K;πVλ(𝐂))0.\mathrm{H}^{\bullet}(\mathfrak{g}_{\infty},K_{\infty}^{\circ};\pi_{\infty}\otimes V_{\lambda}^{\vee}(\mathbf{C}))\neq 0.

Here λ\lambda is pure by [Clo90, Lem. 4.9]. Since π\pi is a RASCAR, λ\lambda is further restricted: by [JST, Prop. 2.17], for all σΣ\sigma\in\Sigma there exists 𝗐σ𝐙{\sf w}_{\sigma}\in\mathbf{Z} such that λσ,i+λσ,2n+1i=𝗐σ\lambda_{\sigma,i}+\lambda_{\sigma,2n+1-i}={\sf w}_{\sigma} (i.e. each λσ\lambda_{\sigma} is individually pure). We have 𝗐σ=𝗐{\sf w}_{\sigma}={\sf w} for σΣ𝐑\sigma\in\Sigma_{\mathbf{R}}, and 𝗐σ+𝗐cσ=2𝗐{\sf w}_{\sigma}+{\sf w}_{c\sigma}=2{\sf w} for σΣ𝐂\sigma\in\Sigma_{\mathbf{C}}. We let 𝗐¯=(𝗐σ)σΣ𝐙Σ\underline{{\sf w}}=({\sf w}_{\sigma})_{\sigma\in\Sigma}\in\mathbf{Z}^{\Sigma}.

2.4.  Critical LL-values

Let π\pi be a RASCAR with standard LL-function L(π,s)L(\pi,s). For a Hecke character χ:F×\𝐀F×𝐂×\chi:F^{\times}\backslash\mathbf{A}_{F}^{\times}\to\mathbf{C}^{\times}, let π×χ\pi\times\chi denote the RASCAR with G(𝐀)G(\mathbf{A})-action twisted by χ\chi. We write L(π,χ) . . =L(π×χ,1/2)L(\pi,\chi)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=L(\pi\times\chi,1/2).

Definition 2.3.

Let χ\chi be a Hecke character with infinity type 𝐣=(jσ)σΣ𝐙Σ\mathbf{j}=(j_{\sigma})_{\sigma\in\Sigma}\in\mathbf{Z}^{\Sigma}, and let π\pi have weight λ\lambda. We say χ\chi is a standard critical value for L(π,)L(\pi,-) if

λσ,njσλσ,n+1σΣ.-\lambda_{\sigma,n}\leqslant j_{\sigma}\leqslant-\lambda_{\sigma,n+1}\qquad\forall\sigma\in\Sigma. (2.1)

In this case, we say χ\chi is critical for λ\lambda. We write Crit(π)\mathrm{Crit}(\pi) for the set of all such χ\chi.

Lemma 2.4.

If χCrit(π)\chi\in\mathrm{Crit}(\pi), then s=1/2s=1/2 is a Deligne-critical value of L(π×χ,s)L(\pi\times\chi,s).

Proof.

The RACAR π×χ\pi\times\chi has weight λ+𝐣=(λσ+(jσ,,jσ))σ\lambda+\mathbf{j}=(\lambda_{\sigma}+(j_{\sigma},...,j_{\sigma}))_{\sigma}. By [JST, Prop. 2.20], s=1/2s=1/2 is a standard critical value of L(π×χ,s)L(\pi\times\chi,s) if and only if (λσ,n+jσ)0(λσ,n+1+jσ)-(\lambda_{\sigma,n}+j_{\sigma})\leqslant 0\leqslant-(\lambda_{\sigma,n+1}+j_{\sigma}) for all σΣ\sigma\in\Sigma. This is equivalent to (2.1). ∎

Remark 2.5.

For 𝐣\mathbf{j} in this range, λ+𝐣\lambda+\mathbf{j} is a balanced weight in the sense of [JST]. There are other Deligne-critical values, but for these λ+𝐣\lambda+\mathbf{j} is not balanced.

Crucially Crit(π)\mathrm{Crit}(\pi) also admits a representation-theoretic description. For 𝐣1,𝐣2𝐙Σ\mathbf{j}_{1},\mathbf{j}_{2}\in\mathbf{Z}^{\Sigma}, let V𝐣1,𝐣2HV_{\mathbf{j}_{1},\mathbf{j}_{2}}^{H} be the HH-representation of highest weight (𝐣1,𝐣2)(\mathbf{j}_{1},\mathbf{j}_{2}), that is, det1𝐣1det2𝐣2\det_{1}^{\mathbf{j}_{1}}\cdot\det_{2}^{\mathbf{j}_{2}}. The following is proved in the same way as [GR14, Prop. 6.3.1] and [JST, Prop. 2.20], via [Kna01, Thm. 2.1].

Lemma 2.6.

A Hecke character χ\chi of infinity type 𝐣𝐙Σ\mathbf{j}\in\mathbf{Z}^{\Sigma} is in Crit(π)\mathrm{Crit}(\pi) if and only if

HomH(Vλ,V𝐣,𝗐¯𝐣H)0.\mathrm{Hom}_{H}(V_{\lambda}^{\vee},V^{H}_{\mathbf{j},-\underline{{\sf w}}-\mathbf{j}})\neq 0.

In this case, HomH(Vλ,V𝐣,𝗐¯𝐣H)\mathrm{Hom}_{H}(V_{\lambda}^{\vee},V^{H}_{\mathbf{j},-\underline{{\sf w}}-\mathbf{j}}) is a line, generated by some κ𝐣\kappa_{\mathbf{j}}.

We will interpolate the standard critical LL-values ‘at pp’. As such, let

Critp(π) . . ={χCrit(π):cond(χ)=pβ0},\mathrm{Crit}_{p}(\pi)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\{\chi\in\mathrm{Crit}(\pi):\mathrm{cond}(\chi)=p^{\beta_{0}}\}, (2.2)

where pβ0 . . =𝔭|p𝔭β0,𝔭p^{\beta_{0}}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\prod_{\mathfrak{p}|p}\mathfrak{p}^{\beta_{0,\mathfrak{p}}}, for β0=(β0,𝔭)𝐙0{𝔭|p}\beta_{0}=(\beta_{0,\mathfrak{p}})\in\mathbf{Z}_{\geqslant 0}^{\{\mathfrak{p}|p\}} a multiexponent.

2.5.  Hecke characters on ray class and Galois groups

Let χ\chi be a Hecke character with infinity type 𝐣𝐙[Σ]\mathbf{j}\in\mathbf{Z}[\Sigma]. We have an algebraic homomorphism w𝐣:F×𝐐¯×w^{\mathbf{j}}:F^{\times}\longrightarrow\overline{\mathbf{Q}}^{\times} given by w𝐣(x)=σΣσ(x)jσ.w^{\mathbf{j}}(x)=\prod_{\sigma\in\Sigma}\sigma(x)^{j_{\sigma}}. This then induces maps

w𝐣:(F𝐐𝐑)×𝐂×ip𝐐¯p×,wp𝐣:(F𝐐𝐐p)×𝐐¯p×.w^{\mathbf{j}}_{\infty}:(F\otimes_{\mathbf{Q}}\mathbf{R})^{\times}\rightarrow\mathbf{C}^{\times}\xrightarrow{i_{p}}\overline{\mathbf{Q}}_{p}^{\times},\qquad w^{\mathbf{j}}_{p}:(F\otimes_{\mathbf{Q}}\mathbf{Q}_{p})^{\times}\rightarrow\overline{\mathbf{Q}}_{p}^{\times}. (2.3)
Definition 2.7.

We define χ[p]\chi_{[p]} to be the function

χ[p]:𝐀F×𝐐¯p×,x\displaystyle\chi_{[p]}:\mathbf{A}_{F}^{\times}\longrightarrow\overline{\mathbf{Q}}_{p}^{\times},\qquad x χ(x)[w𝐣(x)]1wp𝐣(xp)=χ(sgn(x))χf(x)wp𝐣(xp),\displaystyle\longmapsto\chi(x)\cdot\big{[}w_{\infty}^{\mathbf{j}}(x_{\infty})\big{]}^{-1}\cdot w_{p}^{\mathbf{j}}(x_{p})=\chi_{\infty}(\mathrm{sgn}(x_{\infty}))\cdot\chi_{f}(x)\cdot w_{p}^{\mathbf{j}}(x_{p}),

where sgn(x){±1}Σ𝐑\mathrm{sgn}(x_{\infty})\in\{\pm 1\}^{\Sigma_{\mathbf{R}}}. Via [Wei56] this takes values in some finite extension L/𝐐pL/\mathbf{Q}_{p}. If 𝔣\mathfrak{f} is the conductor of χ\chi, then (as in e.g. [BW19, Prop. 2.4]) χ[p]\chi_{[p]} is a locally analytic character on the pp-adic analytic group

ClF+(𝔣p) . . =F×\𝐀F×/U(𝔣p)F+,\mathrm{Cl}_{F}^{+}(\mathfrak{f}p^{\infty})\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=F^{\times}\backslash\mathbf{A}_{F}^{\times}/U(\mathfrak{f}p^{\infty})F_{\infty}^{+},

where U(𝔣p)U(\mathfrak{f}p^{\infty}) is the group of elements of 𝒪^F×\widehat{\mathcal{O}}_{F}^{\times} that are congruent to 1(mod𝔣p)1\hskip 2.0pt(\mathrm{mod}\hskip 2.0pt\mathfrak{f}p^{\infty}).

Recall Galp\operatorname{Gal}_{p} is the Galois group for the maximal abelian extension of FF unramified outside pp\infty. The structure of this group is described in detail in [BDW, §6.1]; in particular, the Artin reciprocity map induces an isomorphism ClF+(p)Galp\mathrm{Cl}_{F}^{+}(p^{\infty})\xrightarrow{\,\smash{\raisebox{-2.79857pt}{$\scriptstyle\sim$}}\,}\operatorname{Gal}_{p}, and we have an exact sequence

0𝒪¯F,+×(𝒪F𝐙p)×GalpClF+0.0\to\overline{\mathcal{O}}_{F,+}^{\times}\to(\mathcal{O}_{F}\otimes\mathbf{Z}_{p})^{\times}\to\operatorname{Gal}_{p}\to\mathrm{Cl}_{F}^{+}\to 0. (2.4)

Via reciprocity, we may consider χ[p]\chi_{[p]} as a character of Galp\operatorname{Gal}_{p}.

2.6.  Locally symmetric spaces and local systems

If KG(𝐀f)K\subset G(\mathbf{A}_{f}) is open compact, the locally symmetric space of level KK is

SK . . =G(𝐐)\G(𝐀)/KK.S_{K}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=G(\mathbf{Q})\backslash G(\mathbf{A})/KK_{\infty}^{\circ}.

If MM is a left G(𝐐)G(\mathbf{Q})-module, then we have an attached ‘archimedean’ local system \mathcal{M} on SKS_{K} given by the locally constant sections of

G(𝐐)\[G(𝐀)×M]/KKSK,G(\mathbf{Q})\backslash[G(\mathbf{A})\times M]/KK_{\infty}^{\circ}\longrightarrow S_{K}, (2.5)

where γ(g,m)kz=(γgkz,γm)\gamma(g,m)kz=(\gamma gkz,\gamma\cdot m).

Similarly, if MM is a left KK-module, then we have an attached ‘pp-adic’ local system on SKS_{K}, given by the local constant sections of (2.5), but with action γ(g,m)kz=(γgkz,k1m)\gamma(g,m)kz=(\gamma gkz,k^{-1}\cdot m).

If MM is a left G(𝐐p)G(\mathbf{Q}_{p})-module, then G(𝐐)G(\mathbf{Q}) acts via G(𝐐)G(𝐐p)G(\mathbf{Q})\subset G(\mathbf{Q}_{p}), and KK acts via projection to G(𝐐p)G(\mathbf{Q}_{p}). We get two attached local systems \mathcal{M} and \mathscr{M}, and these are isomorphic via the map (g,m)(gp1m)(g,m)\mapsto(g_{p}^{-1}\cdot m) of local systems. For more detail see [BDW, §2.3].

2.7.  Algebraic and analytic coefficient modules

2.7.1.  Algebraic coefficients

If λX(T)\lambda\in X^{\bullet}(T) is a dominant algebraic weight, recall VλV_{\lambda} is the algebraic GG-representation of highest weight λ\lambda. If 𝐐pL𝐐¯p\mathbf{Q}_{p}\subset L\subset\overline{\mathbf{Q}}_{p}, we describe Vλ(L)V_{\lambda}(L) via algebraic induction IndB¯Gλ\mathrm{Ind}_{\overline{B}}^{G}\lambda; namely,

Vλ(L) . . ={f:G(𝐐p)L:f is algebraic,f(b¯g)=λ(b¯)f(g)b¯B¯(𝐐p)},V_{\lambda}(L)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\{f:G(\mathbf{Q}_{p})\to L:f\text{ is algebraic},f(\overline{b}g)=\lambda(\overline{b})f(g)\ \ \forall\overline{b}\in\overline{B}(\mathbf{Q}_{p})\},

where B¯\overline{B} is the opposite Borel. This space carries a natural left action of G(𝐐p)G(\mathbf{Q}_{p}) given by

(gf)(h) . . =f(hg),g,hG(𝐐p),(g\cdot f)(h)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=f(hg),\qquad g,h\in G(\mathbf{Q}_{p}), (2.6)

inducing a dual left action on Vλ(L)V_{\lambda}^{\vee}(L) by gμ(f) . . =μ(g1f)g\cdot\mu(f)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\mu(g^{-1}\cdot f). By §2.6, we have attached local systems 𝒱λ(L)\mathcal{V}_{\lambda}^{\vee}(L) and 𝒱λ(L)\mathscr{V}_{\lambda}^{\vee}(L), and a natural isomorphism Hc(SK,𝒱λ(L))Hc(SK,𝒱λ(L))\mathrm{H}^{\bullet}_{\mathrm{c}}(S_{K},\mathcal{V}_{\lambda}^{\vee}(L))\xrightarrow{\,\smash{\raisebox{-2.79857pt}{$\scriptstyle\sim$}}\,}\mathrm{H}^{\bullet}_{\mathrm{c}}(S_{K},\mathscr{V}_{\lambda}^{\vee}(L)).

For our pp-adic interpolation, the subgroup HGH\subset G plays a crucial role; its importance is that G/HG/H is a spherical variety. It is thus convenient to consider VλV_{\lambda} as a double algebraic induction, namely

VλIndQ¯GIndB¯HHλV_{\lambda}\cong\operatorname{Ind}_{\overline{Q}}^{G}\operatorname{Ind}_{\overline{B}\cap H}^{H}\lambda (2.7)

(see [BDW, Lem. 3.6]). Write VλH=IndB¯HHλV_{\lambda}^{H}=\operatorname{Ind}_{\overline{B}\cap H}^{H}\lambda for the algebraic HH-representation of highest weight λ\lambda, and denote the standard action of hHh\in H on vVλHv\in V_{\lambda}^{H} by hλv\langle h\rangle_{\lambda}\cdot v. Then precisely, (2.7) means we can identify Vλ(L)V_{\lambda}(L) with the space of algebraic f:G(𝐐p)VλH(L)f:G(\mathbf{Q}_{p})\to V_{\lambda}^{H}(L) satisfying

f(n¯hg)=hλf(g)for all n¯N¯Q(𝐐p),hH(𝐐p),gG(𝐐p).f(\overline{n}hg)=\langle h\rangle_{\lambda}\cdot f(g)\qquad\text{for all }\overline{n}\in\overline{N}_{Q}(\mathbf{Q}_{p}),h\in H(\mathbf{Q}_{p}),g\in G(\mathbf{Q}_{p}). (2.8)

2.7.2.  Parahoric analytic coefficients

Recall JpG(𝐐p)J_{p}\subset G(\mathbf{Q}_{p}) is the QQ-parahoric subgroup. The theory of distributions on JpJ_{p} was developed in [BW21c], and – in our setting – described in detail in [BDW, §3]. In (2.7), we replace the second algebraic induction with locally analytic induction, i.e. we let

𝒜λ(L)=LAIndQ¯(𝐙p)JpJpVλH(L),\mathcal{A}_{\lambda}(L)=\mathrm{LAInd}_{\overline{Q}(\mathbf{Z}_{p})\cap J_{p}}^{J_{p}}V_{\lambda}^{H}(L),

whence 𝒜λ(L)\mathcal{A}_{\lambda}(L) is the space of locally analytic f:JpVλH(L)f:J_{p}\to V_{\lambda}^{H}(L) such that

f(n¯hg)=hλf(g)for all n¯N¯Q(𝐐p),hH(𝐐p),gJp.f(\overline{n}hg)=\langle h\rangle_{\lambda}\cdot f(g)\qquad\text{for all }\overline{n}\in\overline{N}_{Q}(\mathbf{Q}_{p}),h\in H(\mathbf{Q}_{p}),g\in J_{p}. (2.9)

Let 𝒟λ(L)=Homcts(𝒜λ(L),L)\mathcal{D}_{\lambda}(L)=\operatorname{Hom}_{\mathrm{cts}}(\mathcal{A}_{\lambda}(L),L) be the continuous dual.

Dualising the natural inclusion Vλ(L)𝒜λ(L)V_{\lambda}(L)\subset\mathcal{A}_{\lambda}(L) yields a specialisation map

rλ:𝒟λ(L)Vλ(L).r_{\lambda}:\mathcal{D}_{\lambda}(L)\twoheadrightarrow V_{\lambda}^{\vee}(L). (2.10)

2.7.3.  The *-action on distributions

The space 𝒜λ(L)\mathcal{A}_{\lambda}(L) carries a left JpJ_{p}-action by

(gf)(h)=f(hg),h,gJp.(g*f)(h)=f(hg),\qquad h,g\in J_{p}.

Recall ϖ𝔭\varpi_{\mathfrak{p}} is a uniformiser for F𝔭F_{\mathfrak{p}}, and let

t=t𝔭 . . =ι(ϖ𝔭In,In)=(ϖ𝔭InIn)GL2n(F𝔭).t=t_{\mathfrak{p}}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\iota(\varpi_{\mathfrak{p}}I_{n},I_{n})=\left(\begin{smallmatrix}\varpi_{\mathfrak{p}}I_{n}&\\ &I_{n}\end{smallmatrix}\right)\in\operatorname{GL}_{2n}(F_{\mathfrak{p}}). (2.11)

If f𝒜λf\in\mathcal{A}_{\lambda}, then the function (t𝔭1f):NQ(𝐙p)VλH(L)(t_{\mathfrak{p}}^{-1}*f):N_{Q}(\mathbf{Z}_{p})\to V_{\lambda}^{H}(L) defined by (t𝔭1f)(n) . . =f(t𝔭nt𝔭1)(t_{\mathfrak{p}}^{-1}*f)(n)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=f(t_{\mathfrak{p}}nt_{\mathfrak{p}}^{-1}) extends uniquely to an element of 𝒜λ\mathcal{A}_{\lambda} (recalling NQ(𝐙p)Q(𝐙p)N_{Q}(\mathbf{Z}_{p})\subset Q(\mathbf{Z}_{p}) is the unipotent radical). We get an induced dual action on 𝒟λ\mathcal{D}_{\lambda} given by (t𝔭μ)(f)=μ(t𝔭1f)(t_{\mathfrak{p}}*\mu)(f)=\mu(t_{\mathfrak{p}}^{-1}*f), as in [BDW, §3.4].

Let ΔpG(𝐐p)\Delta_{p}\subset G(\mathbf{Q}_{p}) (resp. Δp1\Delta_{p}^{-1}) be the semigroup generated by JpJ_{p} and t𝔭t_{\mathfrak{p}} (resp. t𝔭1t_{\mathfrak{p}}^{-1}) for 𝔭|p\mathfrak{p}|p. The actions above extend to actions of Δp1\Delta_{p}^{-1} on 𝒜λ\mathcal{A}_{\lambda} and Δp\Delta_{p} on 𝒟λ\mathcal{D}_{\lambda}.

2.7.4.  Intertwining of actions

The *-action of Δp1\Delta_{p}^{-1} on 𝒜λ(L)\mathcal{A}_{\lambda}(L) preserves the subspace Vλ(L)V_{\lambda}(L). In particular, the Δp\Delta_{p}-action on 𝒟λ(L)\mathcal{D}_{\lambda}(L) descends to the quotient Vλ(L)V_{\lambda}^{\vee}(L). Moreover, this *-action preserves natural integral subspaces in all of these spaces (see [BDW, Rem. 3.13]).

As ΔpG(𝐐p)\Delta_{p}\subset G(\mathbf{Q}_{p}), (2.6) induces a \cdot-action of Δp\Delta_{p} on Vλ(L)V_{\lambda}^{\vee}(L). The \cdot- and *-actions of JpJ_{p} visibly agree; but the actions of t𝔭1t_{\mathfrak{p}}^{-1} differ. From the definitions, if μVλ(L)\mu\in V_{\lambda}^{\vee}(L), then

t𝔭μ=λ(t𝔭)×(t𝔭μ).t_{\mathfrak{p}}*\mu=\lambda(t_{\mathfrak{p}})\times(t_{\mathfrak{p}}\cdot\mu). (2.12)

Since the \cdot- and *-actions of JpJ_{p} agree, if K=KpJpG(𝐀f)K=K^{p}J_{p}\subset G(\mathbf{A}_{f}) is a parahoric-at-pp level group, then both actions yield the same local system 𝒱λ(L)\mathscr{V}_{\lambda}^{\vee}(L) on SKS_{K}. However, since the actions of t𝔭t_{\mathfrak{p}} are different, we get two different Hecke operators on the cohomology Hc(SK,𝒱λ(L))\mathrm{H}^{\bullet}_{\mathrm{c}}(S_{K},\mathscr{V}_{\lambda}^{\vee}(L)): the ‘automorphic’ operator U𝔭U_{\mathfrak{p}}^{\cdot}, which is canonical but may have non-integral eigenvalues; and the ‘pp-adic’ operator U𝔭U_{\mathfrak{p}}^{*}, which is integrally normalised but depends on the choice of ϖ𝔭\varpi_{\mathfrak{p}}.

2.8.  Automorphic cohomology classes

By [Clo90, p.120], if π\pi has weight λ\lambda then the (𝔤,K)(\mathfrak{g}_{\infty},K_{\infty}^{\circ})-cohomology Hi(𝔤,K;πVλ(𝐂))\mathrm{H}^{i}(\mathfrak{g}_{\infty},K_{\infty}^{\circ};\pi_{\infty}\otimes V_{\lambda}^{\vee}(\mathbf{C})) is non-zero exactly when

rn2+s(2n2n)ir(n2+n1)+s(2n2+n1).rn^{2}+s(2n^{2}-n)\leqslant i\leqslant r(n^{2}+n-1)+s(2n^{2}+n-1). (2.13)

Via cuspidal cohomology (see [Clo90]), for any open compact KG(𝐀f)K\subset G(\mathbf{A}_{f}) there is an injective and Hecke-equivariant map

Hi(𝔤,K;πVλ(𝐂))πfK⸦-→Hci(SK,𝒱λ(𝐂)).\mathrm{H}^{i}(\mathfrak{g}_{\infty},K_{\infty}^{\circ};\pi_{\infty}\otimes V_{\lambda}^{\vee}(\mathbf{C}))\otimes\pi_{f}^{K}\lhook\joinrel\relbar\joinrel\rightarrow\mathrm{H}^{i}_{\mathrm{c}}(S_{K},\mathcal{V}_{\lambda}^{\vee}(\mathbf{C})).

Composing with ip:𝐂𝐐¯pi_{p}:\mathbf{C}\xrightarrow{\,\smash{\raisebox{-2.79857pt}{$\scriptstyle\sim$}}\,}\overline{\mathbf{Q}}_{p}, and the isomorphism from §2.6, we get an injective map

Hi(𝔤,K;πVλ(𝐂))πfK⸦-→Hci(SK,𝒱λ(𝐐¯p))Hci(SK,𝒱λ(𝐐¯p)).\mathrm{H}^{i}(\mathfrak{g}_{\infty},K_{\infty}^{\circ};\pi_{\infty}\otimes V_{\lambda}^{\vee}(\mathbf{C}))\otimes\pi_{f}^{K}\lhook\joinrel\relbar\joinrel\rightarrow\mathrm{H}^{i}_{\mathrm{c}}(S_{K},\mathcal{V}_{\lambda}^{\vee}(\overline{\mathbf{Q}}_{p}))\xrightarrow{\,\smash{\raisebox{-2.79857pt}{$\scriptstyle\sim$}}\,}\mathrm{H}^{i}_{\mathrm{c}}(S_{K},\mathscr{V}_{\lambda}^{\vee}(\overline{\mathbf{Q}}_{p})). (2.14)

3.  Local test data

To get useful cohomology classes from the map (2.14), we must make good choices of local input data, that is, elements φvπv\varphi_{v}\in\pi_{v} for all finite vv, and a class [ω][\omega] in the (𝔤,K)(\mathfrak{g}_{\infty},K_{\infty}^{\circ})-cohomology. We do so here; in §3.1 for vpv\nmid p\infty, in §3.2 for v|pv|p, and §3.3 at infinity.

Since our motivation is global, let π\pi be a RASCAR with an (η,ψ)(\eta,\psi)-Shalika model. Let λ\lambda be the weight of π\pi, with purity weight 𝗐¯𝐙Σ\underline{{\sf w}}\in\mathbf{Z}^{\Sigma}. Let χCritp(π)\chi\in\mathrm{Crit}_{p}(\pi) be a Hecke character of infinity type 𝐣\mathbf{j} (and pp-power conductor). Let Vχ,sHV_{\chi,s}^{H} be the character χ||s(χη||s)1\chi|\cdot|^{s}\otimes(\chi\eta|\cdot|^{s})^{-1} of H(𝐀)H(\mathbf{A}). We write Vχv,sHV_{\chi_{v},s}^{H}, Vχf,sHV_{\chi_{f},s}^{H} and Vχ,sHV_{\chi_{\infty},s}^{H} for the restrictions to H(Fv)H(F_{v}), H(𝐀f)H(\mathbf{A}_{f}) and H(𝐑)H(\mathbf{R}).

For each place vv of FF, fix a local intertwining 𝒮ψvηv\mathcal{S}_{\psi_{v}}^{\eta_{v}} of πv\pi_{v} into its Shalika model. This induces finite and infinite intertwinings 𝒮ψfηf\mathcal{S}_{\psi_{f}}^{\eta_{f}} and 𝒮ψη\mathcal{S}_{\psi_{\infty}}^{\eta_{\infty}}.

3.1.  Friedberg–Jacquet integrals at finite places

First we recap [FJ93, §3] (see also [JST, Prop. 2.10]). Let vv be a finite place, and fix a Haar measure dgvdg_{v} on GLn(Fv)\operatorname{GL}_{n}(F_{v}) such that GLn(𝒪v)\operatorname{GL}_{n}(\mathcal{O}_{v}) has volume 1. The local Friedberg–Jacquet integral is

Zv(φv,χv,s) . . =GLn(Fv)𝒮ψvηv(φv)[(g1)]χv||s12(det(g))dg.Z_{v}(\varphi_{v},\chi_{v},s)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\int_{\operatorname{GL}_{n}(F_{v})}\mathcal{S}_{\psi_{v}}^{\eta_{v}}(\varphi_{v})\left[\begin{pmatrix}g&\\ &1\end{pmatrix}\right]\chi_{v}|\cdot|^{s-\tfrac{1}{2}}\Big{(}\det(g)\Big{)}dg.

This converges absolutely for Re(s)0\mathrm{Re}(s)\gg 0. Its normalisation

Zv(φv,χv,s) . . =1L(πv×χv,s)Z(φv,χv,s)HomH(Fv)(πvVχv,s1/2H,𝐂)Z_{v}^{\circ}(\varphi_{v},\chi_{v},s)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\frac{1}{L(\pi_{v}\times\chi_{v},s)}Z(\varphi_{v},\chi_{v},s)\in\mathrm{Hom}_{H(F_{v})}\Big{(}\pi_{v}\otimes V_{\chi_{v},s-1/2}^{H},\ \mathbf{C}\Big{)}

admits analytic continuation to 𝐂\mathbf{C}.

The following are proved in Propositions 3.1 and 3.2 of [FJ93] respectively (cf. [DJR20, Prop. 3.3]). Let δv\delta_{v} be the valuation of the different of Fv/𝐐F_{v}/\mathbf{Q}_{\ell}, where v|v|\ell, and let qvq_{v} be the size of 𝐅v\mathbf{F}_{v}.

Proposition 3.1 (Friedberg–Jacquet).
  • (i)

    There exists a test vector φvFJπv\varphi_{v}^{\mathrm{FJ}}\in\pi_{v} such that

    Zv(φvFJ,χv,s)=(qvs1/2χv(ϖv)1)nδvZ_{v}^{\circ}(\varphi_{v}^{\mathrm{FJ}},\chi_{v},s)=(q_{v}^{s-1/2}\chi_{v}(\varpi_{v})^{-1})^{n\delta_{v}} (3.1)

    for all ss and all unramified χv\chi_{v}.

  • (ii)

    If πv\pi_{v} is spherical, the spherical vector is such a test vector.

Note we do not get 1 on the right-hand side of (3.1), as our ψv\psi_{v} need not have conductor 𝒪v\mathcal{O}_{v}.

If φfπf\varphi_{f}\in\pi_{f} can be written as φf=vφv\varphi_{f}=\otimes_{v}\varphi_{v}, then define Zf(φf,χf,s) . . =vZv(φv,χv,s)Z_{f}^{\circ}(\varphi_{f},\chi_{f},s)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\prod_{v\nmid\infty}Z_{v}^{\circ}(\varphi_{v},\chi_{v},s) (and similarly Zf=ZvZ_{f}=\prod Z_{v}). Note that

Zf(,χf,s)HomH(𝐀F,f)(πfVχf,s1/2H,𝐂).Z_{f}^{\circ}(-,\chi_{f},s)\in\mathrm{Hom}_{H(\mathbf{A}_{F,f})}(\pi_{f}\otimes V_{\chi_{f},s-1/2}^{H},\mathbf{C}).

Let φfFJ,(p)=vpφvFJπf(p)\varphi_{f}^{\mathrm{FJ},(p)}=\otimes_{v\nmid p\infty}\varphi_{v}^{\mathrm{FJ}}\in\pi_{f}^{(p)}. If χCritp(π)\chi\in\mathrm{Crit}_{p}(\pi) and vpv\nmid p\infty, then χv\chi_{v} is unramified, hence:

Proposition 3.2.

For 𝔭|p\mathfrak{p}|p, let φ𝔭π𝔭\varphi_{\mathfrak{p}}\in\pi_{\mathfrak{p}} be arbitrary, and let φf=φfFJ,(p)(𝔭|pφ𝔭)πf\varphi_{f}=\varphi_{f}^{\mathrm{FJ},(p)}\otimes(\otimes_{\mathfrak{p}|p}\varphi_{\mathfrak{p}})\in\pi_{f}. Then for any χCritp(π)\chi\in\mathrm{Crit}_{p}(\pi), and for Re(s)0\mathrm{Re}(s)\gg 0, we have

Zf(φf,χf,12)=(vpχ(ϖvδv)n)(𝔭|pZ𝔭(φ𝔭,χ𝔭,12))L(p)(πfχf,12),\textstyle Z_{f}(\varphi_{f},\chi_{f},\tfrac{1}{2})=\left(\prod_{v\nmid p\infty}\chi(\varpi_{v}^{\delta_{v}})^{-n}\right)\cdot\left(\prod_{\mathfrak{p}|p}Z_{\mathfrak{p}}\big{(}\varphi_{\mathfrak{p}},\chi_{\mathfrak{p}},\tfrac{1}{2}\big{)}\right)\cdot L^{(p)}\big{(}\pi_{f}\otimes\chi_{f},\tfrac{1}{2}\big{)}, (3.2)

where L(p)L^{(p)} is the LL-function without the Euler factors at 𝔭|p\mathfrak{p}|p.

3.2.  QQ-refinements and local choices at 𝔭|p\mathfrak{p}|p

For pp-adic interpolation, it is essential to treat the primes 𝔭|p\mathfrak{p}|p separately. In this case, good local vectors have been pinned down in [DJR20, §3], [BDW, §2.7], [BDG+, §9], which we summarise here. For ease of notation, for §3.2 only, we will largely drop subscripts 𝔭\mathfrak{p}, i.e. write π=π𝔭\pi=\pi_{\mathfrak{p}}, ϖ=ϖ𝔭\varpi=\varpi_{\mathfrak{p}}, J=J𝔭J=J_{\mathfrak{p}}, etc.).

Assume that π\pi is spherical. We write π=IndBGθ\pi=\operatorname{Ind}_{B}^{G}\theta, with θ=(θ1,,θ2n)\theta=(\theta_{1},...,\theta_{2n}) as in [BDG+, §6.1]. Recall t=t𝔭t=t_{\mathfrak{p}} from (2.11). On πJ\pi^{J}, we have the Hecke operator U𝔭:=[JtJ]U_{\mathfrak{p}}:=[JtJ].

Definition 3.3.

A QQ-refinement π~=(π,α)\tilde{\pi}=(\pi,\alpha) of π\pi is a choice of U𝔭U_{\mathfrak{p}}-eigenvalue α\alpha on πJ\pi^{J}. We say π~\tilde{\pi} is regular if α\alpha is a simple eigenvalue.

Recall π\pi is the functorial transfer of a representation Π\Pi of GSpin2n+1(F𝔭)\mathrm{GSpin}_{2n+1}(F_{\mathfrak{p}}). Let 𝒬GSpin2n+1\mathcal{Q}\subset\mathrm{GSpin}_{2n+1} be the parabolic associated to QQ by [BGW, §2.1]. We say π~\tilde{\pi} is spin if α\alpha is an eigenvalue of a Hecke operator 𝒰𝔭\mathcal{U}_{\mathfrak{p}} on the 𝒬\mathcal{Q}-parahoric invariant vectors in Π\Pi; see [BGW, §3] for more details (noting this is called QQ-spin there).

A regular spin QQ-refinement was called ‘QQ-regular’ in [DJR20].

Let π~=(π,α)\tilde{\pi}=(\pi,\alpha) be a regular spin QQ-refinement. Let q=q𝔭 . . =#𝐅𝔭q=q_{\mathfrak{p}}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\#\mathbf{F}_{\mathfrak{p}}. Up to reordering the θi\theta_{i}, by [Che04, Lem. 4.8.4] (cf. [BDG+, Prop. 2.5]) we may assume

α=qn2/2θn+1(ϖ)θ2n(ϖ).\alpha=q^{-n^{2}/2}\theta_{n+1}(\varpi)\cdots\theta_{2n}(\varpi). (3.3)

The relevant Friedberg–Jacquet integral at 𝔭\mathfrak{p} is a twisted one, as is familiar in the theory of pp-adic LL-functions; see §5. The following ‘local Birch lemma’ is [BDG+, Prop. 9.3]. Let u . . =(1wn01)GL2n(F𝔭)u\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\left(\begin{smallmatrix}1&w_{n}\\ 0&1\end{smallmatrix}\right)\in\operatorname{GL}_{2n}(F_{\mathfrak{p}}), where wnw_{n} is the longest Weyl element for GLn\operatorname{GL}_{n}. Recall δ=δ𝔭\delta=\delta_{\mathfrak{p}} is the valuation of the different of F𝔭F_{\mathfrak{p}}. For a smooth character χ\chi of F𝔭×F_{\mathfrak{p}}^{\times}, let τ(χ)\tau(\chi) be the local Gauss sum (normalised as in [BDG+, Prop. 9.3]).

Proposition 3.4.

There exists an eigenvector φαπJ[U𝔭α]\varphi^{\alpha}\in\pi^{J}[U_{\mathfrak{p}}-\alpha] such that for any smooth character χ\chi of F𝔭×F_{\mathfrak{p}}^{\times} of conductor 𝔭β0𝒪𝔭\mathfrak{p}^{\beta_{0}}\subset\mathcal{O}_{\mathfrak{p}}, we have

Z𝔭((u1tβ)φα,χ,12)=qn(q1)nqδ(n2n)/2αδQ(π,χ),Z_{\mathfrak{p}}\Big{(}(u^{-1}t^{\beta})\cdot\varphi^{\alpha},\chi,\tfrac{1}{2}\Big{)}=\frac{q^{n}}{(q-1)^{n}}\cdot\frac{q^{\delta(n^{2}-n)/2}}{\alpha^{\delta}}\cdot Q(\pi,\chi),

where β=max(1,β0)\beta=\mathrm{max}(1,\beta_{0}) and

Q(π,χ)={qβ(n2n2)τ(χ)n:χ ramified,χ(ϖ)nδαi=n+12n1θi1χ1(ϖ)q1/21θiχ(ϖ)q1/2:χ unramified.Q(\pi,\chi)=\left\{\begin{array}[]{cl}q^{\beta\left(\tfrac{-n^{2}-n}{2}\right)}\cdot\tau(\chi)^{n}&:\chi\text{ ramified},\\ \displaystyle{\chi(\varpi)^{-n\delta}\cdot\alpha\cdot\prod_{i=n+1}^{2n}\frac{1-\theta_{i}^{-1}\chi^{-1}(\varpi)q^{-1/2}}{1-\theta_{i}\chi(\varpi)q^{-1/2}}}&:\chi\text{ unramified}.\end{array}\right.
Proof.

This is slightly rearranged from [BDG+]; the only significant difference in the ramified case is that we exploit (3.3). In the unramified case, we have pulled a factor of θiχ(ω)q1/2\theta_{i}\chi(\omega)q^{1/2} out of the product in Q(π,χ)Q(\pi,\chi) op. cit., and again used (3.3). ∎

If π~=(π,α)\tilde{\pi}=(\pi,\alpha) is not spin, then by contrast we expect this zeta integral vanishes identically on the α\alpha-eigenspace (see [BGW, §8]).

3.3.  Classes at infinity

We now choose a class [ω][\omega] at infinity, via the ‘zeta integral at infinity’. The choice connects closely to the automorphic cycles in §4 (cf. [JST, §4.1]).

3.3.1.  The zeta integral at infinity

Recall ι:HG\iota:H\hookrightarrow G and KK_{\infty} from §2.1, let L . . =H(𝐑)KL_{\infty}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=H(\mathbf{R})\cap K_{\infty}, and let 𝒳H . . =H(𝐑)/L\mathcal{X}_{H}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=H(\mathbf{R})^{\circ}/L_{\infty}^{\circ}. Let 𝔛H . . =Lie(𝒳H)\mathfrak{X}_{H}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\mathrm{Lie}(\mathcal{X}_{H}), and let t . . =dim𝐑(𝔛H)t\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\mathrm{dim}_{\mathbf{R}}(\mathfrak{X}_{H}). Letting 𝔛H𝐂 . . =𝔛H𝐑𝐂\mathfrak{X}_{H}^{\mathbf{C}}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\mathfrak{X}_{H}\otimes_{\mathbf{R}}\mathbf{C}, we have t𝔛H𝐂=𝐂\wedge^{t}\mathfrak{X}_{H}^{\mathbf{C}}=\mathbf{C}. Given a t𝔛H𝐂\wedge^{t}\mathfrak{X}_{H}^{\mathbf{C}}-valued Haar measure μ\mu on GLn(F𝐑)\operatorname{GL}_{n}(F\otimes\mathbf{R}), the attached Friedberg–Jacquet integral at infinity

Zμ(,χ,s)HomH(𝐑)((t𝔛H𝐂)πVχ,s1/2H,𝐂)Z_{\mu}^{\circ}(-,\chi_{\infty},s)\in\mathrm{Hom}_{H(\mathbf{R})^{\circ}}\Big{(}(\wedge^{t}\mathfrak{X}_{H}^{\mathbf{C}})^{*}\otimes\pi_{\infty}\otimes V_{\chi_{\infty},s-1/2}^{H},\ \mathbf{C}\Big{)}

is defined when Re(s)0\mathrm{Re}(s)\gg 0 by

Zμ(\displaystyle Z_{\mu}^{\circ}( ωφ1,χ,s) . . =\displaystyle\omega\otimes\varphi_{\infty}\otimes 1,\chi_{\infty},s)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=
1L(πχ,s)GLn(F𝐑)𝒮ψη(φ)[(g1)]χ||s12(det(g))dμ,ωg.\displaystyle\frac{1}{L(\pi_{\infty}\otimes\chi_{\infty},s)}\int_{\operatorname{GL}_{n}(F\otimes\mathbf{R})}\mathcal{S}_{\psi_{\infty}}^{\eta_{\infty}}(\varphi_{\infty})\left[\begin{pmatrix}g&\\ &1\end{pmatrix}\right]\chi_{\infty}|\cdot|^{s-\tfrac{1}{2}}\Big{(}\det(g)\Big{)}d\langle\mu,\omega\rangle g.

3.3.2.  Modular symbols at infinity

The following is [JST, §4.3]. As H(𝐑)H(\mathbf{R})^{\circ}-modules, we have

VχH . . =Vχ,0H=χη1χ1V𝐣,𝗐¯𝐣H(𝐂).V_{\chi_{\infty}}^{H}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=V_{\chi_{\infty},0}^{H}=\chi_{\infty}\otimes\eta_{\infty}^{-1}\chi_{\infty}^{-1}\cong V_{\mathbf{j},-\underline{{\sf w}}-\mathbf{j}}^{H}(\mathbf{C}).

In particular, our choice κ𝐣:VλV𝐣,𝗐¯𝐣H\kappa_{\mathbf{j}}:V_{\lambda}^{\vee}\to V^{H}_{\mathbf{j},-\underline{{\sf w}}-\mathbf{j}} (from Lemma 2.6) induces an H(𝐑)H(\mathbf{R})^{\circ}-module map Vλ(𝐂)VχHV_{\lambda}^{\vee}(\mathbf{C})\to V_{\chi_{\infty}}^{H}, and hence an H(𝐑)H(\mathbf{R})^{\circ}-module map Vλ(𝐂)(VχH)𝐂V_{\lambda}^{\vee}(\mathbf{C})\otimes(V_{\chi_{\infty}}^{H})^{\vee}\to\mathbf{C}. We thus obtain a map

Zμ(,χ,1/2)κ𝐣:[πVχH][Vλ(𝐂)(VχH)]t𝔛H𝐂.Z_{\mu}^{\circ}(-,\chi_{\infty},1/2)\otimes\kappa_{\mathbf{j}}:[\pi_{\infty}\otimes V_{\chi_{\infty}}^{H}]\otimes[V_{\lambda}^{\vee}(\mathbf{C})\otimes(V_{\chi_{\infty}}^{H})^{\vee}]\longrightarrow\wedge^{t}\mathfrak{X}_{H}^{\mathbf{C}}.
Definition 3.5.

Let 𝔥=Lie(H(𝐑))\mathfrak{h}=\mathrm{Lie}(H(\mathbf{R})). The modular symbol at infinity is the composition

𝒫χ:Ht(𝔤,K;πVλ(𝐂))\displaystyle\mathcal{P}_{\chi_{\infty}}:\mathrm{H}^{t}(\mathfrak{g},K_{\infty}^{\circ};\pi_{\infty}\otimes V_{\lambda}^{\vee}(\mathbf{C})) ιHt(𝔥,L;πVλ(𝐂))\displaystyle\xrightarrow{\ \iota^{*}\ }\mathrm{H}^{t}(\mathfrak{h},L_{\infty}^{\circ};\pi_{\infty}\otimes V_{\lambda}^{\vee}(\mathbf{C}))
=Ht(𝔥,L;[πVχH][Vλ(𝐂)(VχH)])\displaystyle=\mathrm{H}^{t}(\mathfrak{h},L_{\infty}^{\circ};[\pi_{\infty}\otimes V_{\chi_{\infty}}^{H}]\otimes[V_{\lambda}^{\vee}(\mathbf{C})\otimes(V_{\chi_{\infty}}^{H})^{\vee}])
Zμ(,χ,1/2)κ𝐣Ht(𝔥,L;t𝔛H𝐂)=𝐂.\displaystyle\xrightarrow{\ \ Z_{\mu}^{\circ}(-,\chi_{\infty},1/2)\otimes\kappa_{\mathbf{j}}\ \ }\mathrm{H}^{t}(\mathfrak{h},L_{\infty}^{\circ};\wedge^{t}\mathfrak{X}_{H}^{\mathbf{C}})=\mathbf{C}.
Theorem 3.6 (Lin–Tian, Jiang–Sun–Tian).

There exists a class [ω]Ht(𝔤,K;πVλ(𝐂))[\omega]\in\mathrm{H}^{t}(\mathfrak{g},K_{\infty}^{\circ};\pi_{\infty}\otimes V_{\lambda}^{\vee}(\mathbf{C})) such that 𝒫χ([ω])0\mathcal{P}_{\chi_{\infty}}([\omega])\neq 0 for all χCrit(π)\chi\in\mathrm{Crit}(\pi).

Proof.

After replacing π\pi (globally) with π×χ~\pi\times\widetilde{\chi} for some (fixed) χ~Crit(π)\widetilde{\chi}\in\mathrm{Crit}(\pi), and modifying λ\lambda accordingly, we may assume the weight of π\pi is balanced in the sense of [JST] (since then j=0j=0 is a critical integer). Let [ω][\omega] be the class chosen by Jiang–Sun–Tian after Proposition 4.9 of [JST]. The required non-vanishing is shown op. cit. for χ\chi of the form χ=χ0||j\chi=\chi_{0}|\cdot|^{j}, with χ0\chi_{0} finite-order and j𝐙j\in\mathbf{Z}; that is, for 𝐣\mathbf{j} parallel. If FF has a real embedding, this accounts for all possible critical χ\chi, and we are done.

If FF has no real embedding, then χ\chi can have non-parallel infinity type. To see the more general non-vanishing we state here, we look at the proof in [JST]. The choice of [ω][\omega] stems from the choice of map ϕ0\phi_{0} in Lemma 5.12 op. cit. Once this choice is fixed, non-vanishing follows from non-vanishing of the composition of three maps in Lemma 5.13 op. cit. The first two maps are independent of χ\chi_{\infty}, so are non-vanishing as op. cit. Non-vanishing of the third follows from non-vanishing of Zμ(,χ,1/2)Z^{\circ}_{\mu}(-,\chi_{\infty},1/2) on the minimal KK-type in π\pi_{\infty}. But this was shown by Lin–Tian in [LT20, Thm. 1.2]. We are in situation (2) of that theorem since χ\chi is a standard critical character. ∎

Definition 3.7.

Define the period at χ\chi_{\infty} (attached to the choice [ω][\omega]) to be Ωπ,χ . . =𝒫χ([ω])1\Omega_{\pi,\chi_{\infty}}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\mathcal{P}_{\chi_{\infty}}([\omega])^{-1}.

Remark 3.8.

Once [ω][\omega] is fixed, clearly Ωπ,χ\Omega_{\pi,\chi_{\infty}} depends only on χ\chi_{\infty}, which in turn depends only on 𝐣\mathbf{j} and a sign ϵχ{±1}Σ(𝐑)\epsilon_{\chi}\in\{\pm 1\}^{\Sigma(\mathbf{R})} (notation as in [BW19, §2.2.1]). We further expect that for certain normalised choices of κ𝐣\kappa_{\mathbf{j}} that Ωπ,χ\Omega_{\pi,\chi_{\infty}} should have very light, explicit dependence on 𝐣\mathbf{j}.

To be more precise: fix χ0\chi_{0} finite order, and let χj . . =χ0||j\chi_{j}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\chi_{0}|\cdot|^{j}. In [JST] the κ𝐣\kappa_{\mathbf{j}} are normalised (in Lemma 4.8) to send (1wnwn1)v0\left(\begin{smallmatrix}1&-w_{n}\\ w_{n}&1\end{smallmatrix}\right)\cdot v_{0}^{\vee} to 1, where v0v_{0}^{\vee} is a highest weight vector in VλV_{\lambda}^{\vee}. One of their main technical results is then that there exists ϵ{±1}\epsilon\in\{\pm 1\} such that ϵjijn[F:𝐐]𝒫χj,\epsilon^{j}\cdot i^{-jn[F:\mathbf{Q}]}\cdot\mathcal{P}_{\chi_{j,\infty}} is independent of jj. This leads to their proof of the period relations for L(π×χ0,s)L(\pi\times\chi_{0},s) as ss ranges over critical integers jj.

We will later, via a different method, specify choices of κ𝐣\kappa_{\mathbf{j}} via a different method that allows pp-adic interpolation. Naturally one should expect that our choices are related to those of [JST], and that such a relation should imply the value of our pp-adic LL-function at any special value on the cyclotomic line would satisfy exactly the interpolation predicted by Panchishkin (including the factor at infinity). If FF has a real embedding, this accounts for all special values. When FF has no real embedding, it is natural to hope that an adaptation of the methods in [JST] would provide an analogous result for the more general 𝒫χ\mathcal{P}_{\chi_{\infty}}, which would mean our pp-adic LL-functions have the correct interpolation factor at \infty at all special values.

3.4.  Summary

We collect together our choices.

Test Data 3.9.

Let π\pi be a RASCAR of weight λ\lambda spherical at every 𝔭|p\mathfrak{p}|p. Then:

  • For each 𝔭|p\mathfrak{p}|p, fix a regular spin QQ-refinement π~𝔭=(π𝔭,α𝔭)\tilde{\pi}_{\mathfrak{p}}=(\pi_{\mathfrak{p}},\alpha_{\mathfrak{p}}) of π𝔭\pi_{\mathfrak{p}}. Write π~=(π,α)\tilde{\pi}=(\pi,\alpha) for the collection of these choices, for α=(α𝔭)𝔭|p\alpha=(\alpha_{\mathfrak{p}})_{\mathfrak{p}|p}. Let φ𝔭α𝔭π𝔭\varphi_{\mathfrak{p}}^{\alpha_{\mathfrak{p}}}\in\pi_{\mathfrak{p}} be as in Proposition 3.4.

  • Away from pp, let φfFJ,(p)=vpφvFJπf(p)\varphi_{f}^{\mathrm{FJ},(p)}=\otimes_{v\nmid p\infty}\varphi_{v}^{\mathrm{FJ}}\in\pi_{f}^{(p)}, for φvFJ\varphi_{v}^{\mathrm{FJ}} as in Proposition 3.1.

  • At \infty, let [ω]Ht(𝔤,K;πVλ(𝐂))[\omega]\in\mathrm{H}^{t}(\mathfrak{g},K_{\infty}^{\circ};\pi_{\infty}\otimes V_{\lambda}^{\vee}(\mathbf{C})) be the class from Theorem 3.6.

Write φfFJ,α=φfFJ,(p)(𝔭|pφ𝔭α𝔭)\varphi_{f}^{\mathrm{FJ},\alpha}=\varphi_{f}^{\mathrm{FJ},(p)}\otimes(\otimes_{\mathfrak{p}|p}\varphi_{\mathfrak{p}}^{\alpha_{\mathfrak{p}}}).

Let u1tpβ=𝔭|pu1t𝔭β𝔭G(𝐐p)u^{-1}t_{p}^{\beta}=\prod_{\mathfrak{p}|p}u^{-1}t_{\mathfrak{p}}^{\beta_{\mathfrak{p}}}\in G(\mathbf{Q}_{p}). Then for any χCritp(π)\chi\in\mathrm{Crit}_{p}(\pi), we have computed 𝒫χ([ω])Zf(u1tpβφfFJ,α,χ,1/2)\mathcal{P}_{\chi_{\infty}}([\omega])\cdot Z_{f}^{\circ}(u^{-1}t_{p}^{\beta}\cdot\varphi_{f}^{\mathrm{FJ},\alpha},\chi,1/2) in terms of L(π×χ,1/2)L(\pi\times\chi,1/2), explicit terms at pp, and a non-zero period.

4.  Critical LL-values via cohomology

We now describe a cohomological interpretation of the Friedberg–Jacquet integrals, and hence the standard critical LL-values, following [GR14] and [JST]. We begin to tailor our approach towards pp-adic interpolation, however, with features from [DJR20, BDW, BDG+].

Throughout, let π\pi be a RASCAR with an (η,ψ)(\eta,\psi)-Shalika model. Let λ\lambda be the weight of π\pi, with purity weight 𝗐¯\underline{{\sf w}}. Let χCritp(π)\chi\in\mathrm{Crit}_{p}(\pi) be a Hecke character of infinity type 𝐣\mathbf{j} and conductor pβ0=𝔭|p𝔭β0,𝔭p^{\beta_{0}}=\prod_{\mathfrak{p}|p}\mathfrak{p}^{\beta_{0,\mathfrak{p}}}, where β0=(β0,𝔭)𝐙0𝔭|p\beta_{0}=(\beta_{0,\mathfrak{p}})\in\mathbf{Z}_{\geqslant 0}^{\mathfrak{p}|p} is a multiexponent. For technical reasons (cf. Proposition 3.4), we let β𝔭=max(1,β0,𝔭)\beta_{\mathfrak{p}}=\max(1,\beta_{0,\mathfrak{p}}) and let β=(β𝔭)𝐙1𝔭|p\beta=(\beta_{\mathfrak{p}})\in\mathbf{Z}_{\geqslant 1}^{\mathfrak{p}|p}.

4.1.  Automorphic cycles

One of the main results of [FJ93, §2], reinterpreted in [JST, Prop. 4.1], is that the product

ZμZf(,χ,s):(t𝔛H𝐂)πVχ,s1/2H𝐂,Z_{\mu}^{\circ}\cdot Z_{f}^{\circ}(-,\chi,s):(\wedge^{t}\mathfrak{X}_{H}^{\mathbf{C}})^{*}\otimes\pi\otimes V_{\chi,s-1/2}^{H}\longrightarrow\mathbf{C},

introduced here in §3, can be computed via period integrals for HGH\subset G. In particular, they showed that there exists a Haar measure μ\mu such that

[Zμ(,χ,s)Zf(,χf,s)](ωφ1)L(π×χ)=Xβφω,[Z_{\mu}^{\circ}(-,\chi_{\infty},s)\cdot Z_{f}^{\circ}(-,\chi_{f},s)\big{]}(\omega\otimes\varphi\otimes 1)\cdot L(\pi\times\chi)=\int_{X_{\beta}}\varphi\cdot\omega,

where XβX_{\beta} is an automorphic cycle for HH that we define below. We will explain this, and interpret the right-hand side via cohomology.

For the rest of §4, let KG(𝐀f)K\subset G(\mathbf{A}_{f}) be an open compact subgroup fixing φfFJ,α\varphi_{f}^{\mathrm{FJ},\alpha} from Test Data 3.9. Recall L=H(𝐑)KL_{\infty}=H(\mathbf{R})\cap K_{\infty}. (Here we take intersections with HH with respect to ι\iota).

Definition 4.1.

Let LH(𝐀f)KL\subset H(\mathbf{A}_{f})\cap K be open compact. The automorphic cycle of level LL is

XL . . =H(𝐐)\H(𝐀)/LL.X_{L}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=H(\mathbf{Q})\backslash H(\mathbf{A})/LL_{\infty}^{\circ}.

By [Ash80, Lemma 2.7], ι:HG\iota:H\hookrightarrow G induces a proper map XLSKX_{L}\hookrightarrow S_{K}, which we also denote ι\iota.

In §5, we will choose LβH(𝐀f)L_{\beta}\subset H(\mathbf{A}_{f}) at β\beta, which (as in [BDW, §4.1]) is sufficiently small that:

  • χ\chi and η\eta are trivial on LβL_{\beta},

  • Xβ . . =XLβX_{\beta}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=X_{L_{\beta}} is a real manifold,

  • and H(𝐐)hLβLh1=ZG(𝐐)LβLH(\mathbf{Q})\cap hL_{\beta}L^{\circ}_{\infty}h^{-1}=Z_{G}(\mathbf{Q})\cap L_{\beta}L_{\infty}^{\circ} for all hH(𝐀)h\in H(\mathbf{A}).

Recall t=dim𝐑(𝒳H)=dim𝐑(Xβ)t=\mathrm{dim}_{\mathbf{R}}(\mathcal{X}_{H})=\mathrm{dim}_{\mathbf{R}}(X_{\beta}), and FF has rr (resp. ss) real (resp. complex) places.

Lemma 4.2.

We have t=dim𝐑(Xβ)=r(n2+n1)+s(2n21).t=\mathrm{dim}_{\mathbf{R}}(X_{\beta})=r(n^{2}+n-1)+s(2n^{2}-1).

Proof.

At each real place, we get a contribution of [GLn(𝐑)×GLn(𝐑)]/[On(𝐑)×On(𝐑)]𝐑×[\operatorname{GL}_{n}(\mathbf{R})\times\operatorname{GL}_{n}(\mathbf{R})]/[\mathrm{O}_{n}(\mathbf{R})\times\mathrm{O}_{n}(\mathbf{R})]\mathbf{R}^{\times}, which has dimension 2(n2)2(n2n)/21=n2+n12(n^{2})-2(n^{2}-n)/2-1=n^{2}+n-1.

At each complex place, we get [GLn(𝐂)×GLn(𝐂)]/[Un(𝐂)×Un(𝐂)]𝐂×[\operatorname{GL}_{n}(\mathbf{C})\times\operatorname{GL}_{n}(\mathbf{C})]/[\mathrm{U}_{n}(\mathbf{C})\times\mathrm{U}_{n}(\mathbf{C})]\mathbf{C}^{\times}, which has dimension 2(2n2)[2(n2)+21]=2n212(2n^{2})-[2(n^{2})+2-1]=2n^{2}-1, noting dim𝐑(𝐂×[Un(𝐂)×Un(𝐂)])=1\mathrm{dim}_{\mathbf{R}}(\mathbf{C}^{\times}\cap[\mathrm{U}_{n}(\mathbf{C})\times\mathrm{U}_{n}(\mathbf{C})])=1. ∎

4.2.  Evaluation maps, I

Still summarising [JST], we now give a global version. Recall κ𝐣:VλV𝐣,𝗐¯𝐣H\kappa_{\mathbf{j}}:V_{\lambda}^{\vee}\to V_{\mathbf{j},-\underline{{\sf w}}-\mathbf{j}}^{H} from Lemma 2.6. Let Ev𝐣JST\mathrm{Ev}_{\mathbf{j}}^{\mathrm{JST}} be the composition

Ev𝐣JST:Hct(SK,𝒱λ(𝐂))\displaystyle\mathrm{Ev}_{\mathbf{j}}^{\mathrm{JST}}:\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathcal{V}_{\lambda}^{\vee}(\mathbf{C})) H0(Xβ,𝒱𝐣,𝗐¯+𝐣H(𝐂))\displaystyle\otimes\mathrm{H}^{0}(X_{\beta},\mathcal{V}_{-\mathbf{j},\underline{{\sf w}}+\mathbf{j}}^{H}(\mathbf{C})) (4.1)
ιidHct(Xβ,𝒱λ(𝐂))H0(Xβ,𝒱𝐣,𝗐¯+𝐣H(𝐂))\displaystyle\xrightarrow{\ \iota^{*}\otimes\mathrm{id}\ }\mathrm{H}^{t}_{\mathrm{c}}(X_{\beta},\mathcal{V}_{\lambda}^{\vee}(\mathbf{C}))\otimes\mathrm{H}^{0}(X_{\beta},\mathcal{V}_{-\mathbf{j},\underline{{\sf w}}+\mathbf{j}}^{H}(\mathbf{C}))
κ𝐣idHct(Xβ,𝒱𝐣,𝗐¯𝐣H(𝐂))H0(Xβ,𝒱𝐣,𝗐¯+𝐣H(𝐂))\displaystyle\xrightarrow{\ \kappa_{\mathbf{j}}\otimes\mathrm{id}\ }\mathrm{H}^{t}_{\mathrm{c}}(X_{\beta},\mathcal{V}_{\mathbf{j},-\underline{{\sf w}}-\mathbf{j}}^{H}(\mathbf{C}))\otimes\mathrm{H}^{0}(X_{\beta},\mathcal{V}_{-\mathbf{j},\underline{{\sf w}}+\mathbf{j}}^{H}(\mathbf{C}))
Hct(Xβ,𝐂)𝐂.\displaystyle\xrightarrow{\ \cup\ }\mathrm{H}^{t}_{\mathrm{c}}(X_{\beta},\mathbf{C})\xrightarrow{\,\smash{\raisebox{-2.79857pt}{$\scriptstyle\sim$}}\,}\mathbf{C}.

The final arrow is an integration map Xβ\int_{X_{\beta}} (that we will make precise in (5.4)).

As in [JST, (4.12)], the Lie algebra cohomology yields a natural injective map (and class)

Θχ:VχfH . . =Vχf,0H⸦-→H0(Xβ,\displaystyle\Theta_{\chi}:V_{\chi_{f}}^{H}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=V_{\chi_{f},0}^{H}\lhook\joinrel\relbar\joinrel\rightarrow\mathrm{H}^{0}(X_{\beta}, 𝒱𝐣,𝗐¯+𝐣H(𝐂)),\displaystyle\mathcal{V}_{-\mathbf{j},\underline{{\sf w}}+\mathbf{j}}^{H}(\mathbf{C})),
[VχfH] . . =Θχ(1).\displaystyle[V_{\chi_{f}}^{H}]\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\Theta_{\chi}(1).

The choice of [ω][\omega] (from Theorem 3.6) yields, via (2.14), an injective map

Θ[ω]K:πfK⸦-→Hct(SK,𝒱λ(𝐂)).\Theta_{[\omega]}^{K}:\pi_{f}^{K}\lhook\joinrel\relbar\joinrel\rightarrow\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathcal{V}_{\lambda}^{\vee}(\mathbf{C})). (4.2)

The following is the commutative diagram after [JST, Lem. 4.11]. The term vol(Lβ)\mathrm{vol}(L_{\beta}) arises because of our normalisation of Xβ\int_{X_{\beta}}, compared to their period integral in (4.7) op. cit.

Proposition 4.3.

We have a commutative diagram

πfKVχf\textstyle{\pi_{f}^{K}\otimes V_{\chi_{f}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Θ[ω]KΘχ\scriptstyle{\Theta_{[\omega]}^{K}\otimes\Theta_{\chi}}Zf(,χf,1/2)\scriptstyle{Z_{f}^{\circ}\big{(}-,\chi_{f},1/2\big{)}}𝐂\textstyle{\mathbf{C}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}vol(Lβ)1L(π×χ,1/2)Ωπ,χ\scriptstyle{\mathrm{vol}(L_{\beta})^{-1}\cdot\frac{L\big{(}\pi\times\chi,1/2\big{)}}{\Omega_{\pi,\chi_{\infty}}}}Hct(SK,𝒱λ(𝐂))H0(Xβ,𝒱𝐣,𝗐¯+𝐣H(𝐂))\textstyle{\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathcal{V}_{\lambda}^{\vee}(\mathbf{C}))\otimes\mathrm{H}^{0}(X_{\beta},\mathcal{V}_{-\mathbf{j},\underline{{\sf w}}+\mathbf{j}}^{H}(\mathbf{C}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ev𝐣JST\scriptstyle{\mathrm{Ev}_{\mathbf{j}}^{\mathrm{JST}}}𝐂.\textstyle{\mathbf{C}.}

Finally, we rephrase in language we will consider in the next sections, and summarise all the above results. Let

EvχJST:Hct(SK,𝒱λ(𝐂))𝐂,ϕEv𝐣JST(ϕ[VχfH]).\mathrm{Ev}_{\chi}^{\mathrm{JST}}:\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathcal{V}_{\lambda}^{\vee}(\mathbf{C}))\longrightarrow\mathbf{C},\qquad\phi\longmapsto\mathrm{Ev}_{\mathbf{j}}^{\mathrm{JST}}\big{(}\phi\otimes[V_{\chi_{f}}^{H}]\big{)}. (4.3)

From Proposition 3.2, we see immediately that:

Corollary 4.4.

If φf=φfFJ,(p)(𝔭|pφ𝔭)πfK\varphi_{f}=\varphi_{f}^{\mathrm{FJ},(p)}\otimes(\otimes_{\mathfrak{p}|p}\varphi_{\mathfrak{p}})\in\pi_{f}^{K}, then for all χCrit(π)\chi\in\mathrm{Crit}(\pi), we have

EvχJSTΘ[ω]K(φf)=vol(Lβ)1vpχ(ϖvδv)n𝔭|pZ𝔭(φ𝔭,χ𝔭,12)L(p)(π×χ,12)Ωπ,χ.\mathrm{Ev}_{\chi}^{\mathrm{JST}}\circ\Theta_{[\omega]}^{K}(\varphi_{f})=\mathrm{vol}(L_{\beta})^{-1}\cdot\prod_{v\nmid p\infty}\chi(\varpi_{v}^{\delta_{v}})^{-n}\cdot\prod_{\mathfrak{p}|p}Z_{\mathfrak{p}}\big{(}\varphi_{\mathfrak{p}},\chi_{\mathfrak{p}},\tfrac{1}{2}\big{)}\cdot\frac{L^{(p)}\big{(}\pi\times\chi,\tfrac{1}{2}\big{)}}{\Omega_{\pi,\chi_{\infty}}}. (4.4)

5.  Evaluation maps

We now give abstract generalisations of EvχJST\mathrm{Ev}_{\chi}^{\mathrm{JST}}. To connect the above to the study in [BDW], we first give a more pedestrian reinterpretation of EvχJST\mathrm{Ev}_{\chi}^{\mathrm{JST}}, before describing analogues with pp-adic local systems. As in the last section, π\pi will be a RASCAR with an (η,ψ)(\eta,\psi)-Shalika model, of weight λ\lambda, with purity weight 𝗐¯\underline{{\sf w}}. Let χ\chi be a Hecke character of conductor pβ0p^{\beta_{0}}, with β0=(β0,𝔭)𝔭|p𝐙0𝔭|p\beta_{0}=(\beta_{0,\mathfrak{p}})_{\mathfrak{p}|p}\in\mathbf{Z}^{\mathfrak{p}|p}_{\geqslant 0}, and infinity type 𝐣\mathbf{j} critical for π\pi. Again set β𝔭=max(1,β0,𝔭)\beta_{\mathfrak{p}}=\max(1,\beta_{0,\mathfrak{p}}) and let β=(β𝔭)𝐙1𝔭|p\beta=(\beta_{\mathfrak{p}})\in\mathbf{Z}_{\geqslant 1}^{\mathfrak{p}|p}.

We want to work with pp-adic, rather than complex, coefficients; so recall we fixed an isomorphism ip:𝐂𝐐¯pi_{p}:\mathbf{C}\to\overline{\mathbf{Q}}_{p}. Via this isomorphism, we identify Vλ(𝐂)V_{\lambda}^{\vee}(\mathbf{C}) and Vλ(𝐐¯p)V_{\lambda}^{\vee}(\overline{\mathbf{Q}}_{p}), and let

Θ[ω]K,ip=ipΘ[ω]K:πfK⸦-→Hct(SK,𝒱λ(𝐐¯p))\Theta_{[\omega]}^{K,i_{p}}=i_{p}\circ\Theta_{[\omega]}^{K}:\pi_{f}^{K}\lhook\joinrel\relbar\joinrel\rightarrow\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathcal{V}_{\lambda}^{\vee}(\overline{\mathbf{Q}}_{p})) (5.1)

for Θ[ω]K\Theta_{[\omega]}^{K} the map from (4.2).

5.1.  Automorphic cycles revisited

5.1.1.  Level structures

We now specify the levels LβL_{\beta}. Let K=vKvG(𝐀f)K=\prod_{v}K_{v}\subset G(\mathbf{A}_{f}), such that:

  • (i)

    For vpv\nmid p\infty, we take KvK_{v} fixing φvFJ\varphi_{v}^{\mathrm{FJ}} (from §3.1);

  • (ii)

    For 𝔭|p\mathfrak{p}|p, we take K𝔭=J𝔭K_{\mathfrak{p}}=J_{\mathfrak{p}} parahoric (as in §2.1).

Let uG(𝐀f)u\in G(\mathbf{A}_{f}) be the matrix with uv=1u_{v}=1 if vpv\nmid p, and u𝔭=(1wn01)u_{\mathfrak{p}}=\left(\begin{smallmatrix}1&w_{n}\\ 0&1\end{smallmatrix}\right) for all 𝔭|p\mathfrak{p}|p. For a multiexponent β=(β𝔭)𝔭|p\beta=(\beta_{\mathfrak{p}})_{\mathfrak{p}|p}, let Lβ=L(p)𝔭|pL𝔭β𝔭H(𝐀f)L_{\beta}=L^{(p)}\prod_{\mathfrak{p}|p}L_{\mathfrak{p}}^{\beta_{\mathfrak{p}}}\subset H(\mathbf{A}_{f}), where (taking all intersections with HH with respect to ι\iota):

  • (i)

    away from pp, L(p)G(𝐀f(p))L^{(p)}\subset G(\mathbf{A}_{f}^{(p)}) is the principal congruence subgroup of some (fixed, suppressed) prime-to-pp ideal 𝔪𝒪F\mathfrak{m}\subset\mathcal{O}_{F}, chosen so that L(p)H(𝐀f(p))K(p)L^{(p)}\subset H(\mathbf{A}_{f}^{(p)})\cap K^{(p)};

  • (ii)

    and L𝔭β𝔭=H(𝐙p)K𝔭(u𝔭1t𝔭β𝔭K𝔭t𝔭β𝔭u𝔭).L_{\mathfrak{p}}^{\beta_{\mathfrak{p}}}=H(\mathbf{Z}_{p})\cap K_{\mathfrak{p}}\cap\Big{(}u_{\mathfrak{p}}^{-1}t_{\mathfrak{p}}^{\beta_{\mathfrak{p}}}\cdot K_{\mathfrak{p}}\cdot t_{\mathfrak{p}}^{-\beta_{\mathfrak{p}}}u_{\mathfrak{p}}\Big{)}.

We take 𝔪\mathfrak{m} large enough that the three conditions after Definition 4.1 are satisfied, and as op. cit., we let Xβ . . =XLβX_{\beta}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=X_{L_{\beta}}, the automorphic cycle of level pβp^{\beta}.

Remark 5.1.

The matrix u𝔭u_{\mathfrak{p}} is an open-orbit representative of the spherical variety G/HG/H (that is, B(F𝔭)u𝔭H(F𝔭)B(F_{\mathfrak{p}})\cdot u_{\mathfrak{p}}\cdot H(F_{\mathfrak{p}}) is open in G(F𝔭)G(F_{\mathfrak{p}})). The matrix t𝔭t_{\mathfrak{p}} induces the action of the U𝔭U_{\mathfrak{p}} Hecke operator.

5.1.2.  Connected components and fundamental classes

Lemma 5.2.

The map (det1/det2,1/det2)(\det_{1}/\det_{2},1/\det_{2}) descends to a surjective map

Lβ-↠(1+pβ𝔪𝒪^F)×(1+𝔪𝒪^F)𝐀F××𝐀F×.L_{\beta}\relbar\joinrel\twoheadrightarrow\big{(}1+p^{\beta}\mathfrak{m}\widehat{\mathcal{O}}_{F}\big{)}\times\big{(}1+\mathfrak{m}\widehat{\mathcal{O}}_{F}\big{)}\subset\mathbf{A}_{F}^{\times}\times\mathbf{A}_{F}^{\times}.
Proof.

Identical to [DJR20, Lem. 2.1] and [BDG+, Lem. 4.5]. ∎

By strong approximation for HH, the connected components of XβX_{\beta} are indexed by

π0(Xβ) . . =ClF+(pβ𝔪)×ClF+(𝔪),\pi_{0}(X_{\beta})\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\mathrm{Cl}_{F}^{+}(p^{\beta}\mathfrak{m})\times\mathrm{Cl}_{F}^{+}(\mathfrak{m}), (5.2)

where ClF+(I)\mathrm{Cl}_{F}^{+}(I) is the narrow ray class group of conductor II. For δH(𝐀f)\delta\in H(\mathbf{A}_{f}), we write [δ][\delta] for its associated class in π0(Xβ)\pi_{0}(X_{\beta}) and denote the corresponding connected component

Xβ[δ] . . =H(𝐐)\H(𝐐)δLβH(𝐑)/LβL.X_{\beta}[\delta]\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=H(\mathbf{Q})\backslash H(\mathbf{Q})\delta L_{\beta}H(\mathbf{R})^{\circ}/L_{\beta}L_{\infty}^{\circ}.

We can describe these components as arithmetic quotients of the symmetric space 𝒳H=H(𝐑)/L\mathcal{X}_{H}=H(\mathbf{R})^{\circ}/L_{\infty}^{\circ} from §3.3. For β𝐙1\beta\in\mathbf{Z}_{\geqslant 1} and δH(𝐀f)\delta\in H(\mathbf{A}_{f}), define a congruence subgroup

Γβ,δ . . =H(𝐐)δLβH(𝐑)δ1.\Gamma_{\beta,\delta}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=H(\mathbf{Q})\cap\delta L_{\beta}H(\mathbf{R})^{\circ}\delta^{-1}. (5.3)

The group Γβ,δ\Gamma_{\beta,\delta} acts on 𝒳H\mathcal{X}_{H} by left translation via Γβ,δH(𝐑)\Gamma_{\beta,\delta}\hookrightarrow H(\mathbf{R})^{\circ}, and there is an isomorphism

cδ:Γβ,δ\𝒳HXβ[δ]Xβ,[h]δ[δh],c_{\delta}:\Gamma_{\beta,\delta}\backslash\mathcal{X}_{H}\xrightarrow{\,\smash{\raisebox{-2.79857pt}{$\scriptstyle\sim$}}\,}X_{\beta}[\delta]\subset X_{\beta},\ \ \ [h_{\infty}]_{\delta}\mapsto[\delta h_{\infty}],

where if [h]𝒳H[h_{\infty}]\in\mathcal{X}_{H}, we write [h]δ[h_{\infty}]_{\delta} for its image in Γβ,δ\𝒳H\Gamma_{\beta,\delta}\backslash\mathcal{X}_{H}.

Since Γβ,δ\𝒳H\Gamma_{\beta,\delta}\backslash\mathcal{X}_{H} is a connected orientable manifold of dimension tt, we have an isomorphism

(θδ):Hct(Γβ,δ\𝒳H,𝐙)𝐙,(-\cap\theta_{\delta}):\mathrm{H}^{t}_{\mathrm{c}}(\Gamma_{\beta,\delta}\backslash\mathcal{X}_{H},\mathbf{Z})\xrightarrow{\,\smash{\raisebox{-2.79857pt}{$\scriptstyle\sim$}}\,}\mathbf{Z},

where θδHtBM(Γβ,δ\𝒳H,𝐙)𝐙\theta_{\delta}\in\mathrm{H}_{t}^{\mathrm{BM}}(\Gamma_{\beta,\delta}\backslash\mathcal{X}_{H},\mathbf{Z})\cong\mathbf{Z} is a fundamental class, chosen via the process described in [DJR20, §2.2.5] and [BDW, §4.2.3]. Our choices, and the maps cδc_{\delta}, induce an integration map promised in (4.1):

Xβ . . =δ(θδ)cδ:Hct(Xβ,𝐙)𝐙\int_{X_{\beta}}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\sum_{\delta}(-\cap\theta_{\delta})\circ c_{\delta}^{*}:\mathrm{H}^{t}_{\mathrm{c}}(X_{\beta},\mathbf{Z})\longmapsto\mathbf{Z} (5.4)

5.2.  The map EvχJST\mathrm{Ev}_{\chi}^{\mathrm{JST}} revisited

Now let 𝐐pL𝐐¯p\mathbf{Q}_{p}\subset L\subset\overline{\mathbf{Q}}_{p} containing 𝐐(π,η)\mathbf{Q}(\pi,\eta). For pp-adic interpolation we must twist the map ι:XβSK\iota:X_{\beta}\to S_{K}. Define

ιβ:XβSK,[h][hu1tpβ],hH(𝐀),tp=𝔭|pt𝔭β𝔭.\iota_{\beta}:X_{\beta}\longrightarrow S_{K},\qquad[h]\mapsto[hu^{-1}t_{p}^{\beta}],\qquad h\in H(\mathbf{A}),\ \ t_{p}=\prod_{\mathfrak{p}|p}t_{\mathfrak{p}}^{\beta_{\mathfrak{p}}}.

One may check ιβ𝒱λ=ι𝒱λ\iota_{\beta}^{*}\mathcal{V}_{\lambda}^{\vee}=\iota^{*}\mathcal{V}_{\lambda}^{\vee}, so for δH(𝐀)\delta\in H(\mathbf{A}), we have maps

Hct(SK,𝒱λ(L))ιβHct(Xβ,ι𝒱λ(L))cδHct(Γβ,δ\𝒳H,cδι𝒱λ(L)),\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathcal{V}_{\lambda}^{\vee}(L))\xrightarrow{\ \iota_{\beta}^{*}\ }\mathrm{H}^{t}_{\mathrm{c}}(X_{\beta},\iota^{*}\mathcal{V}_{\lambda}^{\vee}(L))\xrightarrow{\ c_{\delta}^{*}\ }\mathrm{H}^{t}_{\mathrm{c}}(\Gamma_{\beta,\delta}\backslash\mathcal{X}_{H},c_{\delta}^{*}\iota^{*}\mathcal{V}_{\lambda}^{\vee}(L)),

where cδι𝒱λc_{\delta}^{*}\iota^{*}\mathcal{V}_{\lambda}^{\vee} can be checked to be the local system given by locally constant sections of Γβ,δ\[𝒳H×Vλ(L)]Γβ,δ\𝒳H\Gamma_{\beta,\delta}\backslash[\mathcal{X}_{H}\times V_{\lambda}^{\vee}(L)]\to\Gamma_{\beta,\delta}\backslash\mathcal{X}_{H} with action γ([h],v)=([γh],γv)\gamma([h_{\infty}],v)=([\gamma h_{\infty}],\gamma\cdot v) (cf. §2.6).

If MM is a left Γβ,δ\Gamma_{\beta,\delta}-module, let MΓβ,δ . . =M/{mγm:mM,γΓβ,δ}M_{\Gamma_{\beta,\delta}}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=M/\{m-\gamma\cdot m:m\in M,\gamma\in\Gamma_{\beta,\delta}\} be the coinvariants. The quotient map Vλ(L)(Vλ(L))Γβ,δV_{\lambda}^{\vee}(L)\to(V_{\lambda}^{\vee}(L))_{\Gamma_{\beta,\delta}} trivialises the local system, inducing a map

coinvβ,δ:Hct(Γβ,δ\𝒳H,cδι𝒱λ(L))Hct(Γβ,δ\𝒳H,𝐙)𝐙(Vλ(L))Γβ,δ.\mathrm{coinv}_{\beta,\delta}:\mathrm{H}^{t}_{\mathrm{c}}(\Gamma_{\beta,\delta}\backslash\mathcal{X}_{H},c_{\delta}^{*}\iota^{*}\mathcal{V}_{\lambda}^{\vee}(L))\longrightarrow\mathrm{H}^{t}_{\mathrm{c}}(\Gamma_{\beta,\delta}\backslash\mathcal{X}_{H},\mathbf{Z})\otimes_{\mathbf{Z}}(V_{\lambda}^{\vee}(L))_{\Gamma_{\beta,\delta}}.
Definition 5.3.

Define

Evβ,δ:Hct(SK,𝒱λ(L))\displaystyle\mathrm{Ev}_{\beta,\delta}:\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathcal{V}_{\lambda}^{\vee}(L)) cδ(ιβ)Hct(Γβ,δ\𝒳H,cδι𝒱λ(L))\displaystyle\xrightarrow{\ c_{\delta}^{*}\circ(\iota_{\beta})^{*}\ }\mathrm{H}^{t}_{\mathrm{c}}(\Gamma_{\beta,\delta}\backslash\mathcal{X}_{H},c_{\delta}^{*}\iota^{*}\mathcal{V}_{\lambda}^{\vee}(L)) (5.5)
coinvβ,δHct(Γβ,δ\𝒳H,𝐙)𝐙(Vλ(L))Γβ,δ\displaystyle\xrightarrow{\mathrm{coinv}_{\beta,\delta}}\mathrm{H}^{t}_{\mathrm{c}}(\Gamma_{\beta,\delta}\backslash\mathcal{X}_{H},\mathbf{Z})\otimes_{\mathbf{Z}}(V_{\lambda}^{\vee}(L))_{\Gamma_{\beta,\delta}} (5.6)
θδ(Vλ(L))Γβ,δ.\displaystyle\xrightarrow[\,\smash{\raisebox{2.79857pt}{$\scriptstyle\sim$}}\,]{-\cap\theta_{\delta}}(V_{\lambda}^{\vee}(L))_{\Gamma_{\beta,\delta}}.

Recall V𝐣,𝗐¯𝐣HV^{H}_{\mathbf{j},-\underline{{\sf w}}-\mathbf{j}} is the algebraic HH-representation of highest weight (𝐣,𝗐¯𝐣)(\mathbf{j},-\underline{{\sf w}}-\mathbf{j}). Recalling wp𝐣w_{p}^{\mathbf{j}} from (2.3), V𝐣,𝗐¯𝐣H(L)V^{H}_{\mathbf{j},-\underline{{\sf w}}-\mathbf{j}}(L) is an H(𝐐p)H(\mathbf{Q}_{p})-representation, with h=(h1,h2)H(𝐐p)h=(h_{1},h_{2})\in H(\mathbf{Q}_{p}) acting by

wp𝐣(det(h1))wp𝗐¯𝐣(det(h2))= . . (wp𝐣wp𝗐¯𝐣)(h).w_{p}^{\mathbf{j}}(\det(h_{1}))w_{p}^{-\underline{{\sf w}}-\mathbf{j}}(\det(h_{2}))=\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}(w_{p}^{\mathbf{j}}\otimes w_{p}^{-\underline{{\sf w}}-\mathbf{j}})(h). (5.7)

Recall we chose (via Lemma 2.6) a non-zero element κ𝐣HomH(𝐐p)(Vλ(L),V𝐣,𝗐¯𝐣H(L))\kappa_{\mathbf{j}}\in\operatorname{Hom}_{H(\mathbf{Q}_{p})}(V_{\lambda}^{\vee}(L),V^{H}_{\mathbf{j},-\underline{{\sf w}}-\mathbf{j}}(L)).

Lemma 5.4.

The map κ𝐣\kappa_{\mathbf{j}} induces a map κ𝐣:(Vλ)Γβ,δV𝐣,𝗐¯𝐣H(L)\kappa_{\mathbf{j}}:(V_{\lambda}^{\vee})_{\Gamma_{\beta,\delta}}\to V^{H}_{\mathbf{j},-\underline{{\sf w}}-\mathbf{j}}(L).

Proof.

It suffices to prove Γβ,δ\Gamma_{\beta,\delta} acts trivially on the target. If γ=(γ1,γ2)Γβ,δ\gamma=(\gamma_{1},\gamma_{2})\in\Gamma_{\beta,\delta}, then det(γ1),det(γ2)𝒪F,+×\det(\gamma_{1}),\det(\gamma_{2})\in\mathcal{O}_{F,+}^{\times}. Moreover det(γ1/γ2)1(modpβ)\det(\gamma_{1}/\gamma_{2})\equiv 1\hskip 2.0pt(\mathrm{mod}\hskip 2.0ptp^{\beta}) and det(γ2)1(mod𝔪)\det(\gamma_{2})\equiv 1\hskip 2.0pt(\mathrm{mod}\hskip 2.0pt\mathfrak{m}) (by Lemma 5.2).

By assumption there exists a Hecke character χ\chi of infinity type 𝐣\mathbf{j} and conductor dividing pβp^{\beta}; this forces wp𝐣(det(γ1/γ2))=w𝐣(det(γ1/γ2))=1w_{p}^{\mathbf{j}}(\det(\gamma_{1}/\gamma_{2}))=w^{\mathbf{j}}(\det(\gamma_{1}/\gamma_{2}))=1, as in [BW19, §5.2.3]. Similarly, existence of η\eta with infinity type 𝗐¯\underline{{\sf w}} and conductor dividing 𝔪\mathfrak{m} forces wp𝗐¯(det(γ2))=1w_{p}^{-\underline{{\sf w}}}(\det(\gamma_{2}))=1. Thus (wp𝐣wp𝗐¯𝐣)(γ)=1(w_{p}^{\mathbf{j}}\otimes w_{p}^{-\underline{{\sf w}}-\mathbf{j}})(\gamma)=1, i.e. γ\gamma acts trivially on V𝐣,𝗐¯𝐣H(L)V^{H}_{\mathbf{j},-\underline{{\sf w}}-\mathbf{j}}(L), as required. ∎

Definition 5.5.

Let

Evχ:Hct(SK,𝒱λ(L))\displaystyle\mathrm{Ev}_{\chi}:\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathcal{V}_{\lambda}^{\vee}(L)) L(χ),\displaystyle\longrightarrow L(\chi),
ϕ\displaystyle\phi δπ0(Xβ)(χχ1η1)(δ)[κ𝐣Evβ,δ(ϕ)].\displaystyle\longmapsto\sum_{\delta\in\pi_{0}(X_{\beta})}(\chi\otimes\chi^{-1}\eta^{-1})(\delta)\cdot\Big{[}\kappa_{\mathbf{j}}\circ\mathrm{Ev}_{\beta,\delta}(\phi)\Big{]}.

Here we have chosen a basis u𝐣u_{\mathbf{j}} of V𝐣,𝗐¯𝐣H(L)V_{\mathbf{j},-\underline{{\sf w}}-\mathbf{j}}^{H}(L) to identify it with L(χ)L(\chi). Via ipi_{p}, we can do this compatibly with the choices in §4.2.

By considering how all these maps behave on the local systems, we see that actually this ‘new’ construction is just a twisted version of the map EvχJST\mathrm{Ev}_{\chi}^{\mathrm{JST}} from (4.3) (hence it is independent of the choices of δ\delta). Precisely, recall

u1tpβ=𝔭|pu1t𝔭β𝔭G(𝐐p), and let Kβ . . =u1tpβKtpβu.u^{-1}t_{p}^{\beta}=\textstyle\prod_{\mathfrak{p}|p}u^{-1}t_{\mathfrak{p}}^{\beta_{\mathfrak{p}}}\subset G(\mathbf{Q}_{p}),\qquad\text{ and let }\qquad K_{\beta}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=u^{-1}t_{p}^{\beta}Kt_{p}^{-\beta}u.

Note LβKβL_{\beta}\subset K_{\beta}, so we can apply EvχJST\mathrm{Ev}_{\chi}^{\mathrm{JST}} at level KβK_{\beta}. If φfπK\varphi_{f}\in\pi^{K}, then u1tpβφfπfKβu^{-1}t_{p}^{\beta}\cdot\varphi_{f}\in\pi_{f}^{K_{\beta}}, and:

Lemma 5.6.

The following diagram commutes:

πfK\textstyle{\pi_{f}^{K}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Θ[ω]K,ip\scriptstyle{\Theta_{[\omega]}^{K,i_{p}}}φfu1tpβφf\scriptstyle{\varphi_{f}\mapsto u^{-1}t_{p}^{\beta}\cdot\varphi_{f}}Hct(SK,𝒱λ(𝐐¯p))\textstyle{\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathcal{V}_{\lambda}^{\vee}(\overline{\mathbf{Q}}_{p}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Evχ\scriptstyle{\mathrm{Ev}_{\chi}}𝐐¯p\textstyle{\overline{\mathbf{Q}}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ip1\scriptstyle{i_{p}^{-1}}πfKβ\textstyle{\pi_{f}^{K_{\beta}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Θ[ω]Kβ\scriptstyle{\Theta_{[\omega]}^{K_{\beta}}}Hct(SKβ,𝒱λ(𝐂))\textstyle{\mathrm{H}^{t}_{\mathrm{c}}(S_{K_{\beta}},\mathcal{V}_{\lambda}^{\vee}(\mathbf{C}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}EvχJST\scriptstyle{\mathrm{Ev}_{\chi}^{\mathrm{JST}}}𝐂.\textstyle{\mathbf{C}.}

Note that the twisting matrix u1tpβu^{-1}t_{p}^{\beta} is exactly what appeared in Proposition 3.4.

5.3.  Abstract evaluation maps

So far we have only considered evaluation maps only with coefficients in 𝒱λ\mathcal{V}_{\lambda}^{\vee}, the local system attached to VλV_{\lambda}^{\vee} with its G(𝐐)G(\mathbf{Q})-action. Ultimately we will consider coefficients in pp-adic distributions, for which only the KK-local systems make sense. We now present a version of the above maps for KK-local systems, with abstract coefficient modules, generalising those constructed in [BDW, §4] (when FF is totally real) and [BW19, §10.1] (for GL2\operatorname{GL}_{2}).

The constructions/proofs of [BDW] all go through exactly as op. cit., so we omit details.

Recall ΔpG(𝐐p)\Delta_{p}\subset G(\mathbf{Q}_{p}) from §2.7.3. Let MM be a left Δp\Delta_{p}-module, with action denoted \bullet. (When M=𝒟λM=\mathcal{D}_{\lambda}, this will be the *-action. When M=VλM=V_{\lambda}^{\vee}, recall from §2.7.4 that we have two such actions, the \cdot- and *-actions, and we shall need to consider both).

The level KK acts on MM via its projection to Jp=𝔭|pJ𝔭ΔpJ_{p}=\prod_{\mathfrak{p}|p}J_{\mathfrak{p}}\subset\Delta_{p}, giving a local system \mathscr{M} on SKS_{K} via §2.6. If γΓβ,δ\gamma\in\Gamma_{\beta,\delta} (from (5.3)), then δ1γδLβK\delta^{-1}\gamma\delta\in L_{\beta}\subset K, so Γβ,δ\Gamma_{\beta,\delta} acts on MM via γΓβ,δm . . =(δ1γδ)fm\gamma\bullet_{\Gamma_{\beta,\delta}}m\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=(\delta^{-1}\gamma\delta)_{f}\bullet m. Hence we may take coinvariants for this action.

The evaluation maps for \mathscr{M} are inspired by those in Definition 5.3 for 𝒱λ\mathcal{V}_{\lambda}^{\vee}. A crucial difference is that ιβ\iota_{\beta}^{*}\mathscr{M} is not ι\iota^{*}\mathscr{M} any more, but there is a map

τβ:ιβι,(h,m)(h,u1tpβm).\tau_{\beta}^{\bullet}:\iota_{\beta}^{*}\mathscr{M}\to\iota^{*}\mathscr{M},\qquad(h,m)\mapsto(h,u^{-1}t_{p}^{\beta}\bullet m).
Definition 5.7.

The evaluation map for (M,)(M,\bullet) of level pβp^{\beta} at δ\delta is the composition

𝒱β,δM,[]:Hct(SK,)τβ(ιβ)Hct(Xβ,\displaystyle\mathscr{E}{\scriptstyle\mathscr{V}}_{\beta,\delta}^{M,[\bullet]}:\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathscr{M})\xrightarrow{\ \tau_{\beta}^{\bullet}\circ(\iota_{\beta})^{*}\ }\mathrm{H}^{t}_{\mathrm{c}}(X_{\beta}, ι)cδHtc(Γβ,δ\𝒳H,cδι)\displaystyle\iota^{*}\mathscr{M})\xrightarrow{c_{\delta}^{*}}\mathrm{H}^{t}_{\mathrm{c}}(\Gamma_{\beta,\delta}\backslash\mathcal{X}_{H},c_{\delta}^{*}\iota^{*}\mathscr{M}) (5.8)
coinvβ,δHct(Γβ,δ\𝒳H,𝐙)MΓβ,δθδMΓβ,δ.\displaystyle\xrightarrow{\mathrm{coinv}_{\beta,\delta}}\mathrm{H}^{t}_{\mathrm{c}}(\Gamma_{\beta,\delta}\backslash\mathcal{X}_{H},\mathbf{Z})\otimes M_{\Gamma_{\beta,\delta}}\xrightarrow[\,\smash{\raisebox{2.79857pt}{$\scriptstyle\sim$}}\,]{-\cap\theta_{\delta}}M_{\Gamma_{\beta,\delta}}.

We track dependence on M,δ,βM,\delta,\beta. The following are proved exactly as in [BDW, §4.3].

Lemma 5.8.

(Variation in MM; [BDW, Lem. 4.6]) Let κ:MN\kappa:M\rightarrow N be a Δp\Delta_{p}-module map. There is a commutative diagram

Hct(SK,)\textstyle{\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathscr{M})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒱β,δM,[]\scriptstyle{\mathscr{E}{\scriptstyle\mathscr{V}}_{\beta,\delta}^{M,[\bullet]}}κ\scriptstyle{\kappa_{*}}MΓβ,δ\textstyle{M_{\Gamma_{\beta,\delta}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}κ\scriptstyle{\kappa}Hct(SK,𝒩)\textstyle{\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathscr{N})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒱β,δN,[]\scriptstyle{\mathscr{E}{\scriptstyle\mathscr{V}}_{\beta,\delta}^{N,[\bullet]}}NΓβ,δ.\textstyle{N_{\Gamma_{\beta,\delta}}.}
Proposition 5.9.

(Variation in δ\delta; [BDW, Prop. 4.9]) Let NN be a left H(𝐀)H(\mathbf{A})-module, with action \bullet, such that H(𝐐)H(\mathbf{Q}) and H(𝐑)H(\mathbf{R})^{\circ} act trivially. Let κ:MN\kappa:M\to N be a map of LβL_{\beta}-modules. Then

𝒱β,[δ]M,[],κ . . =δ[κ𝒱β,δM,[]]:Hct(SK,)N\mathscr{E}{\scriptstyle\mathscr{V}}_{\beta,[\delta]}^{M,[\bullet],\kappa}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\delta\bullet\left[\kappa\circ\mathscr{E}{\scriptstyle\mathscr{V}}_{\beta,\delta}^{M,[\bullet]}\right]:\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathscr{M})\longrightarrow N

is well-defined and independent of the representative δ\delta of [δ][\delta].

Fix 𝔭|p\mathfrak{p}|p, and define β=(β𝔮)𝔮|p\beta^{\prime}=(\beta_{\mathfrak{q}}^{\prime})_{\mathfrak{q}|p}, where β𝔭=β𝔭+1\beta_{\mathfrak{p}}^{\prime}=\beta_{\mathfrak{p}}+1 and β𝔮=β𝔮\beta_{\mathfrak{q}}^{\prime}=\beta_{\mathfrak{q}} for 𝔮𝔭\mathfrak{q}\neq\mathfrak{p}. We have natural projections

prβ,𝔭:XβXβ,prβ,𝔭:π0(Xβ)π0(Xβ).\mathrm{pr}_{\beta,\mathfrak{p}}:X_{\beta^{\prime}}\longrightarrow X_{\beta},\ \ \ \mathrm{pr}_{\beta,\mathfrak{p}}:\pi_{0}(X_{\beta^{\prime}})\rightarrow\pi_{0}(X_{\beta}).

The action of t𝔭Δpt_{\mathfrak{p}}\in\Delta_{p} on MM yields an action of U𝔭 . . =[Kt𝔭K]U_{\mathfrak{p}}^{\bullet}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=[Kt_{\mathfrak{p}}K] on Hct(SK,)\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathscr{M}), via [BDW, §2.3.2].

Proposition 5.10.

(Variation in β\beta; [BDW, Prop. 4.10]) Let NN and κ\kappa be as in Proposition 5.9. If β>0\beta>0, then as maps Hct(SK,)N\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathscr{M})\to N we have

[η]prβ,𝔭1([δ])𝒱β,[η]M,[],κ=𝒱β,[δ]M,[],κU𝔭.\sum_{[\eta]\in\mathrm{pr}_{\beta,\mathfrak{p}}^{-1}([\delta])}\mathscr{E}{\scriptstyle\mathscr{V}}_{\beta^{\prime},[\eta]}^{M,[\bullet],\kappa}=\mathscr{E}{\scriptstyle\mathscr{V}}_{\beta,[\delta]}^{M,[\bullet],\kappa}\circ U_{\mathfrak{p}}^{\bullet}.

5.4.  Classical evaluation maps

We now use the above to construct linear functionals

𝒱χ[]:Hct(SK,𝒱λ(L))L(χ),\mathscr{E}{\scriptstyle\mathscr{V}}_{\chi}^{[\bullet]}:\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathscr{V}_{\lambda}^{\vee}(L))\longrightarrow L(\chi),

analogous to Evχ\mathrm{Ev}_{\chi} with 𝒱λ\mathcal{V}_{\lambda}^{\vee}. We have

β,δVλ,[]:Hct(SK,𝒱λ(L))(Vλ(L))Γβ,δ.\mathscr{E}_{\beta,\delta}^{V_{\lambda}^{\vee},[\bullet]}:\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathscr{V}_{\lambda}^{\vee}(L))\longrightarrow(V_{\lambda}^{\vee}(L))_{\Gamma_{\beta,\delta}}.

Lemma 5.4 says κ𝐣\kappa_{\mathbf{j}} factors through the Γβ,δ\Gamma_{\beta,\delta}-coinvariants. That lemma used the \cdot-action on VλV_{\lambda}^{\vee}, but by the same proof it is also true of the *-action.

To apply Proposition 5.9 to κ𝐣\kappa_{\mathbf{j}}, we now extend V𝐣,𝗐¯𝐣H(L)V^{H}_{\mathbf{j},-\underline{{\sf w}}-\mathbf{j}}(L) to an H(𝐀)H(\mathbf{A})-module in such a way that H(𝐐)H(\mathbf{Q}) and H(𝐑)H(\mathbf{R})^{\circ} act trivially (as required by the statement of that proposition).

Definition 5.11.

Let Vχ,[p]H(L)V_{\chi,[p]}^{H}(L) be the 1-dimensional H(𝐀)H(\mathbf{A})-module χ[p]χ[p]1η[p]1\chi_{[p]}\otimes\chi_{[p]}^{-1}\eta_{[p]}^{-1}, recalling χ[p]\chi_{[p]} and η[p]\eta_{[p]} are the ray class characters attached to χ\chi and η\eta in §2.5. Precisely, it is the space L(χ)L(\chi), with h=(h1,h2)H(𝐀)h=(h_{1},h_{2})\in H(\mathbf{A}) acting as

hv=[χ[p](det(h1h2))η[p](det(1h2))]v.h\cdot v=\Big{[}\chi_{[p]}\left(\mathrm{det}\left(\tfrac{h_{1}}{h_{2}}\right)\right)\cdot\eta_{[p]}\left(\mathrm{det}\left(\tfrac{1}{h_{2}}\right)\right)\Big{]}v.

By construction, H(𝐐)H(\mathbf{Q}) and H(𝐑)H(\mathbf{R})^{\circ} act trivially on Vχ,[p]HV_{\chi,[p]}^{H}.

Lemma 5.12.

The identity map is an isomorphism of 1-dimensional LβL_{\beta}-modules

V𝐣,𝗐¯𝐣H(L)Vχ,[p]H(L)L(χ).V^{H}_{\mathbf{j},-\underline{{\sf w}}-\mathbf{j}}(L)\xrightarrow{\,\smash{\raisebox{-2.79857pt}{$\scriptstyle\sim$}}\,}V_{\chi,[p]}^{H}(L)\cong L(\chi).

Here LβL_{\beta} acts on V𝐣,𝗐¯𝐣HV^{H}_{\mathbf{j},-\underline{{\sf w}}-\mathbf{j}} via projection to H(𝐐p)H(\mathbf{Q}_{p}), and on Vχ,[p]HV_{\chi,[p]}^{H} by restricting the H(𝐀)H(\mathbf{A})-action.

Proof.

Via (5.7), =(1,2)Lβ\ell=(\ell_{1},\ell_{2})\in L_{\beta} acts on V𝐣,𝗐¯𝐣H(L)V^{H}_{\mathbf{j},-\underline{{\sf w}}-\mathbf{j}}(L) as

v=(wp𝐣wp𝗐¯𝐣)(p)v.\ell\cdot v=(w_{p}^{\mathbf{j}}\otimes w_{p}^{-\underline{{\sf w}}-\mathbf{j}})(\ell_{p})\ v. (5.9)

By definition of χ[p]\chi_{[p]} and η[p]\eta_{[p]}, the action on Vχ,[p]HV_{\chi,[p]}^{H} is by

v=χ(121)η(21)×(wp𝐣wp𝗐¯𝐣)(p)v.\ell\cdot v=\chi(\ell_{1}\ell_{2}^{-1})\eta(\ell_{2}^{-1})\times(w_{p}^{\mathbf{j}}\otimes w_{p}^{-\underline{{\sf w}}-\mathbf{j}})(\ell_{p})\ v. (5.10)

Now det(121)1(modpβ𝒪^F)\det(\ell_{1}\ell_{2}^{-1})\equiv 1\hskip 2.0pt(\mathrm{mod}\hskip 2.0ptp^{\beta}\widehat{\mathcal{O}}_{F}) by Lemma 5.2, so as χ\chi has conductor dividing pβp^{\beta}, we see χ(121) . . =χ(det(121))=1\chi(\ell_{1}\ell_{2}^{-1})\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\chi(\det(\ell_{1}\ell_{2}^{-1}))=1. Similarly det(2)1(mod𝔪)\det(\ell_{2})\equiv 1\hskip 2.0pt(\mathrm{mod}\hskip 2.0pt\mathfrak{m}), and the conductor of η\eta divides 𝔪\mathfrak{m}, we have η(21)=1\eta(\ell_{2}^{-1})=1. In particular, (5.9) and (5.10) agree, proving the lemma. ∎

Corollary 5.13.

For any δH(𝐀f)\delta\in H(\mathbf{A}_{f}), the map

𝒱χ,[δ][]:Hct(SK,𝒱λ(L))𝒱β,δVλ,[](Vλ(L))Γβ,δκ𝐣V𝐣,𝗐¯𝐣H(L)Vχ,[p]HδVχ,[p]HL(χ)\mathscr{E}{\scriptstyle\mathscr{V}}_{\chi,[\delta]}^{[\bullet]}:\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathscr{V}_{\lambda}^{\vee}(L))\xrightarrow{\ \mathscr{E}{\scriptstyle\mathscr{V}}_{\beta,\delta}^{V_{\lambda}^{\vee},[\bullet]}\ }\big{(}V_{\lambda}^{\vee}(L)\big{)}_{\Gamma_{\beta,\delta}}\xrightarrow{\kappa_{\mathbf{j}}}V^{H}_{\mathbf{j},-\underline{{\sf w}}-\mathbf{j}}(L)\xrightarrow{\,\smash{\raisebox{-2.79857pt}{$\scriptstyle\sim$}}\,}V_{\chi,[p]}^{H}\xrightarrow{\ \cdot\delta\ }V_{\chi,[p]}^{H}\cong L(\chi)

is well-defined and depends only on the class [δ]π0(Xβ)[\delta]\in\pi_{0}(X_{\beta}).

Proof.

Immediate by combining Proposition 5.9 and Lemma 5.12. ∎

Definition 5.14.

The classical evaluation map attached to χ\chi and the action \bullet is the map

𝒱χ[] . . =δπ0(Xβ)𝒱χ,[δ][]:Hct(SK,𝒱λ(L))L.\mathscr{E}{\scriptstyle\mathscr{V}}_{\chi}^{[\bullet]}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\sum_{\delta\in\pi_{0}(X_{\beta})}\mathscr{E}{\scriptstyle\mathscr{V}}_{\chi,[\delta]}^{[\bullet]}:\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathscr{V}_{\lambda}^{\vee}(L))\longrightarrow L.

It depends on the choices of κ𝐣\kappa_{\mathbf{j}} and a basis u𝐣u_{\mathbf{j}} of the line Vχ,[p]HL(χ)V_{\chi,[p]}^{H}\cong L(\chi), each unique up to scalar.

As mentioned previously, there are two natural choices of action \bullet on Vλ(L)V_{\lambda}^{\vee}(L): the standard \cdot-action (which connects more cleanly to the theory from §4), and the *-action inherited as a quotient of 𝒟λ\mathcal{D}_{\lambda}. We get two evaluation maps 𝒱χ[]\mathscr{E}{\scriptstyle\mathscr{V}}_{\chi}^{[\cdot]} and 𝒱χ[]\mathscr{E}{\scriptstyle\mathscr{V}}_{\chi}^{[*]}. Since the only dependence on \bullet is in the map τβ\tau_{\beta}^{\bullet}, from (2.12) we have (recalling χ\chi has conductor related to pβp^{\beta})

𝒱χ[]=λ(tpβ)×𝒱χ[].\mathscr{E}{\scriptstyle\mathscr{V}}_{\chi}^{[*]}=\lambda(t_{p}^{\beta})\times\mathscr{E}{\scriptstyle\mathscr{V}}_{\chi}^{[\cdot]}. (5.11)

5.5.  Comparison between Evχ\mathrm{Ev}_{\chi} and 𝒱χ[]\mathscr{E}{\scriptstyle\mathscr{V}}_{\chi}^{[\cdot]}

We now connect the evaluation with =\bullet=\cdot to Definition 5.5 and hence to §4. Let ϕHct(SK,𝒱λ(L))\phi\in\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathcal{V}_{\lambda}^{\vee}(L)), and let υ\upsilon denote the natural isomorphism

υ:Hct(SK,𝒱λ(L))Hct(SK,𝒱λ(L))\upsilon:\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathcal{V}_{\lambda}^{\vee}(L))\xrightarrow{\,\smash{\raisebox{-2.79857pt}{$\scriptstyle\sim$}}\,}\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathscr{V}_{\lambda}^{\vee}(L))

from §2.6. Then, taking \cdot-actions everywhere:

Proposition 5.15.

We have Evχ(ϕ)=𝒱χ[](υ(ϕ)).\mathrm{Ev}_{\chi}(\phi)=\mathscr{E}{\scriptstyle\mathscr{V}}_{\chi}^{[\cdot]}(\upsilon(\phi)).

Proof.

We have a commutative diagram (cf. [DJR20, Prop. 4.6] and [BDJ22, Prop. 4.1])

ϕ\textstyle{\phi\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\in}υ(ϕ)\textstyle{\upsilon(\phi)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\in}Hct(SK,𝒱λ(L))\textstyle{\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathcal{V}_{\lambda}^{\vee}(L))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ιβ\scriptstyle{\iota_{\beta}^{*}}(g,v)(g,gp1v)\scriptstyle{(g,v)\longmapsto(g,g_{p}^{-1}\cdot v)}Hct(SK,𝒱λ(L))\textstyle{\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathscr{V}_{\lambda}^{\vee}(L))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τβιβ\scriptstyle{\tau_{\beta}^{\cdot}\circ\iota_{\beta}^{*}}Hct(Xβ,ι𝒱λ(L))\textstyle{\mathrm{H}^{t}_{\mathrm{c}}(X_{\beta},\iota^{*}\mathcal{V}_{\lambda}^{\vee}(L))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(h,v)(h,hp1v)\scriptstyle{(h,v)\longmapsto(h,h_{p}^{-1}\cdot v)}cδ\scriptstyle{c_{\delta}^{*}}Hct(Xβ,ι𝒱λ(L))\textstyle{\mathrm{H}^{t}_{\mathrm{c}}(X_{\beta},\iota^{*}\mathscr{V}_{\lambda}^{\vee}(L))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}cδ\scriptstyle{c_{\delta}^{*}}Hct(Γβ,δ\𝒳H,cδι𝒱λ(L))\textstyle{\mathrm{H}^{t}_{\mathrm{c}}(\Gamma_{\beta,\delta}\backslash\mathcal{X}_{H},c_{\delta}^{*}\iota^{*}\mathcal{V}_{\lambda}^{\vee}(L))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}([h],v)([h],δp1v)\scriptstyle{([h_{\infty}],v)\mapsto([h_{\infty}],\delta_{p}^{-1}\cdot v)}(θδ)coinvβ,δ\scriptstyle{(-\cap\theta_{\delta})\circ\mathrm{coinv}_{\beta,\delta}}Hct(Γβ,δ\𝒳H,cδι𝒱λ(L))\textstyle{\mathrm{H}^{t}_{\mathrm{c}}(\Gamma_{\beta,\delta}\backslash\mathcal{X}_{H},c_{\delta}^{*}\iota^{*}\mathscr{V}_{\lambda}^{\vee}(L))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(θδ)coinvβ,δ\scriptstyle{(-\cap\theta_{\delta})\circ\mathrm{coinv}_{\beta,\delta}}(Vλ(L))Γβ,δ\textstyle{(V_{\lambda}^{\vee}(L))_{\Gamma_{\beta,\delta}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}vδp1v\scriptstyle{v\longmapsto\delta_{p}^{-1}\cdot v}(Vλ(L))Γβ,δ,\textstyle{(V_{\lambda}^{\vee}(L))_{\Gamma_{\beta,\delta}},}

where the horizontal maps are induced by the stated maps of local systems. In particular,

δp1Evβ,δ(ϕ)=𝒱β,δ[](υ(ϕ))\delta_{p}^{-1}\cdot\mathrm{Ev}_{\beta,\delta}(\phi)=\mathscr{E}{\scriptstyle\mathscr{V}}_{\beta,\delta}^{[\cdot]}(\upsilon(\phi))

as elements of (Vλ(L))Γβ,δ(V_{\lambda}^{\vee}(L))_{\Gamma_{\beta,\delta}}. Composing with the H(𝐐p)H(\mathbf{Q}_{p})-module map κ𝐣\kappa_{\mathbf{j}} gives

(wp𝐣wp𝗐¯+𝐣)(δp)[κ𝐣Evβ,δ(ϕ)]=[κ𝐣𝒱β,δ[](υ(ϕ))].(w_{p}^{-\mathbf{j}}\otimes w_{p}^{\underline{{\sf w}}+\mathbf{j}})(\delta_{p})\Big{[}\kappa_{\mathbf{j}}\circ\mathrm{Ev}_{\beta,\delta}(\phi)\Big{]}=\Big{[}\kappa_{\mathbf{j}}\circ\mathscr{E}{\scriptstyle\mathscr{V}}_{\beta,\delta}^{[\cdot]}(\upsilon(\phi))\Big{]}. (5.12)

Now consider both sides as elements of Vχ,[p]HV_{\chi,[p]}^{H} by Lemma 5.12, and act by δ\delta on both sides (via the H(𝐀)H(\mathbf{A}) action on Vχ,[p]HV_{\chi,[p]}^{H}). On the left-hand side, the factor (wp𝐣wp𝗐¯𝐣)(δp)(w_{p}^{\mathbf{j}}\otimes w_{p}^{-\underline{{\sf w}}-\mathbf{j}})(\delta_{p}) in χ[p]χ[p]1η[p]1\chi_{[p]}\otimes\chi_{[p]}^{-1}\eta_{[p]}^{-1} cancels with the left-most term in (5.12), so we have

δ((wp𝐣wp𝗐¯+𝐣)(δp)[κ𝐣Evβ,δ(ϕ)])=(χχ1η1)(δ)[κ𝐣Evβ,δ(ϕ)].\delta\cdot\left((w_{p}^{-\mathbf{j}}\otimes w_{p}^{\underline{{\sf w}}+\mathbf{j}})(\delta_{p})\Big{[}\kappa_{\mathbf{j}}\circ\mathrm{Ev}_{\beta,\delta}(\phi)\Big{]}\right)=(\chi\otimes\chi^{-1}\eta^{-1})(\delta)\cdot\Big{[}\kappa_{\mathbf{j}}\circ\mathrm{Ev}_{\beta,\delta}(\phi)\Big{]}.

On the right-hand side we get, by definition, 𝒱χ,[δ][](υ(ϕ)).\mathscr{E}{\scriptstyle\mathscr{V}}_{\chi,[\delta]}^{[\cdot]}(\upsilon(\phi)). Summing both sides over δπ0(Xβ)\delta\in\pi_{0}(X_{\beta}) completes the proof. ∎

The following theorem, now using the *-action, summarises the last three sections. As in the last line of the proof of [DJR20, Thm. 4.7]), the global Gauss sum is

τ(χf) . . =χv unramifiedχv(ϖv)δvχv ramifiedτ(χv);\tau(\chi_{f})\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\prod_{\chi_{v}\text{ unramified}}\chi_{v}(\varpi_{v})^{-\delta_{v}}\cdot\prod_{\chi_{v}\text{ ramified}}\tau(\chi_{v}); (5.13)

this is different from some treatments, as our ψ\psi has conductor 𝔡1\mathfrak{d}^{-1}, rather than 𝒪^F\widehat{\mathcal{O}}_{F}.

Theorem 5.16.

Let π\pi and φfFJ,α\varphi_{f}^{\mathrm{FJ},\alpha} be as in Test Data 3.9. For all χCrit(π)\chi\in\mathrm{Crit}(\pi), we have

ip1𝒱χ[]Θ[ω]K,ip(φfFJ,α)=Aλ(tpβ)τ(χf)n𝔭|pQ(π𝔭,χ𝔭)L(p)(π×χ,12)Ωπ,χ,i_{p}^{-1}\circ\mathscr{E}{\scriptstyle\mathscr{V}}_{\chi}^{[*]}\circ\Theta_{[\omega]}^{K,i_{p}}(\varphi_{f}^{\mathrm{FJ},\alpha})=A\cdot\lambda(t_{p}^{\beta})\cdot\tau(\chi_{f})^{n}\cdot\prod_{\mathfrak{p}|p}Q^{\prime}(\pi_{\mathfrak{p}},\chi_{\mathfrak{p}})\cdot\frac{L^{(p)}\big{(}\pi\times\chi,\tfrac{1}{2}\big{)}}{\Omega_{\pi,\chi_{\infty}}}, (5.14)

where

Q(π𝔭,χ𝔭)={q𝔭β𝔭(n2n2):χ𝔭 ramified,α𝔭i=n+12n1θ𝔭,i1χ𝔭1(ϖ𝔭)q𝔭1/21θ𝔭,iχ𝔭(ϖ𝔭)q𝔭1/2:χ𝔭 unramified,Q^{\prime}(\pi_{\mathfrak{p}},\chi_{\mathfrak{p}})=\left\{\begin{array}[]{cl}q_{\mathfrak{p}}^{\beta_{\mathfrak{p}}\left(\tfrac{n^{2}-n}{2}\right)}&:\chi_{\mathfrak{p}}\text{ ramified},\\ \displaystyle\alpha_{\mathfrak{p}}\cdot\prod_{i=n+1}^{2n}\frac{1-\theta_{\mathfrak{p},i}^{-1}\chi_{\mathfrak{p}}^{-1}(\varpi_{\mathfrak{p}})q_{\mathfrak{p}}^{-1/2}}{1-\theta_{\mathfrak{p},i}\chi_{\mathfrak{p}}(\varpi_{\mathfrak{p}})q_{\mathfrak{p}}^{-1/2}}&:\chi_{\mathfrak{p}}\text{ unramified},\end{array}\right.

and

A . . =1vol(L1)𝔭|p(q𝔭n(q𝔭1)nq𝔭δ𝔭(n2n)/2α𝔭δ𝔭)A\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\frac{1}{\mathrm{vol}(L_{1})}\cdot\prod_{\mathfrak{p}|p}\left(\frac{q_{\mathfrak{p}}^{n}}{(q_{\mathfrak{p}}-1)^{n}}\cdot q_{\mathfrak{p}}^{\delta_{\mathfrak{p}}(n^{2}-n)/2}\cdot\alpha_{\mathfrak{p}}^{-\delta_{\mathfrak{p}}}\right) (5.15)

is a constant independent of χ\chi. Here L1 . . =L(1,,1)L_{1}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=L_{(1,...,1)} (i.e. we take β𝔭=1\beta_{\mathfrak{p}}=1 for all 𝔭\mathfrak{p}).

Proof.

By (5.11) (to pass from * to \cdot evaluations, introducing λ(tpβ)\lambda(t_{p}^{\beta})), Corollary 4.4 and Lemma 5.6, noting that u1tpβu^{-1}t_{p}^{\beta} is trivial outside primes above pp, we have

ip1𝒱χ[]Θ[ω]K,ip(φfFJ,α)=λ(tpβ)\displaystyle i_{p}^{-1}\circ\mathscr{E}{\scriptstyle\mathscr{V}}_{\chi}^{[*]}\circ\Theta_{[\omega]}^{K,i_{p}}(\varphi_{f}^{\mathrm{FJ},\alpha})=\lambda(t_{p}^{\beta}) vol(Lβ)1vpχv(ϖv)nδv\displaystyle\cdot\mathrm{vol}(L_{\beta})^{-1}\cdot\prod_{v\nmid p\infty}\chi_{v}(\varpi_{v})^{-n\delta_{v}}
×𝔭|pZ𝔭([u1t𝔭β𝔭φ𝔭α𝔭],χ𝔭,12)L(p)(π×χ,12)Ωπ,χ.\displaystyle\times\prod_{\mathfrak{p}|p}Z_{\mathfrak{p}}\big{(}[u^{-1}t_{\mathfrak{p}}^{\beta_{\mathfrak{p}}}\cdot\varphi_{\mathfrak{p}}^{\alpha_{\mathfrak{p}}}],\chi_{\mathfrak{p}},\tfrac{1}{2}\big{)}\cdot\frac{L^{(p)}\big{(}\pi\times\chi,\tfrac{1}{2}\big{)}}{\Omega_{\pi,\chi_{\infty}}}.

By [BDG+, Lem. 4.4] (cf. Lemma 5.2) we have

vol(Lβ)=vol(L1)δB(tpβ)=vol(L1)𝔭|pq𝔭β𝔭n2.\textstyle\mathrm{vol}(L_{\beta})=\mathrm{vol}(L_{1})\cdot\delta_{B}(t_{p}^{\beta})=\mathrm{vol}(L_{1})\cdot\prod_{\mathfrak{p}|p}q_{\mathfrak{p}}^{-\beta_{\mathfrak{p}}n^{2}}.

The twisted local zeta integrals at pp were evaluated222This is where we want to use β𝐙1𝔭|p\beta\in\mathbf{Z}_{\geqslant 1}^{\mathfrak{p}|p}, rather than the ‘true’ conductor pβ0p^{\beta_{0}}; for pp-adic interpolation it is necessary to twist non-trivially at each 𝔭\mathfrak{p}, even when the conductor β0,𝔭\beta_{0,\mathfrak{p}} itself is trivial. in Proposition 3.4. Combining all of this, we get the factor Q(π𝔭,α𝔭)Q^{\prime}(\pi_{\mathfrak{p}},\alpha_{\mathfrak{p}}) and χ𝔭(ϖ𝔭δ𝔭)\chi_{\mathfrak{p}}(\varpi_{\mathfrak{p}}^{-\delta_{\mathfrak{p}}}) (resp. τ(χ𝔭)\tau(\chi_{\mathfrak{p}})) if χ𝔭\chi_{\mathfrak{p}} is unramified (resp. ramified). Finally we conclude by the identity (5.13). ∎

6.  pp-adic interpolation of evaluation maps

We will interpolate the LL-values appearing on the right-hand side of Theorem 5.16 by interpolating the left-hand side, i.e. 𝒱χ[]\mathscr{E}{\scriptstyle\mathscr{V}}_{\chi}^{[*]}, as χ\chi varies. The main result of §6 is Proposition 6.12.

6.1.  Alignment of branching laws

The evaluation maps of the previous section depended on choices of bases

κ𝐣HomH(𝐐p)(Vλ(L),V𝐣,𝗐¯𝐣H(L)),u𝐣V𝐣,𝗐¯𝐣H(L)L.\kappa_{\mathbf{j}}\in\operatorname{Hom}_{H(\mathbf{Q}_{p})}(V_{\lambda}^{\vee}(L),V^{H}_{\mathbf{j},-\underline{{\sf w}}-\mathbf{j}}(L)),\qquad u_{\mathbf{j}}\in V^{H}_{\mathbf{j},-\underline{{\sf w}}-\mathbf{j}}(L)\cong L. (6.1)

The choices can be combined into a single choice κ𝐣:Vλ(L)L\kappa_{\mathbf{j}}^{\circ}:V_{\lambda}^{\vee}(L)\to L defined by

κ𝐣(μ)=κ𝐣(μ)u𝐣μVλ(L).\kappa_{\mathbf{j}}(\mu)=\kappa_{\mathbf{j}}^{\circ}(\mu)\cdot u_{\mathbf{j}}\qquad\forall\mu\in V_{\lambda}^{\vee}(L). (6.2)

In Theorem 5.16, these choices manifested themselves on the left-hand side in the definition of 𝒱χ[]\mathscr{E}{\scriptstyle\mathscr{V}}_{\chi}^{[*]}, and on the right-hand side (via ipi_{p}) in the zeta integral at infinity. To interpolate the 𝒱χ[]\mathscr{E}{\scriptstyle\mathscr{V}}_{\chi}^{[*]} requires a careful alignment of the choices as 𝐣\mathbf{j} varies, which we carry out here. The main idea is that we can collapse all the different choices (as 𝐣\mathbf{j} varies) of branching law κ𝐣\kappa_{\mathbf{j}} for HGH\subset G onto a single choice of branching law for Gn . . =ResF/Q(GLn)HG_{n}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\mathrm{Res}_{F/Q}(\operatorname{GL}_{n})\subset H, diagonally embedded.

Write λ=(λ,λ′′)\lambda=(\lambda^{\prime},\lambda^{\prime\prime}), where λ,λ′′\lambda^{\prime},\lambda^{\prime\prime} are two weights for GnG_{n}. Then as HH-representations VλHVλGnVλ′′GnV_{\lambda}^{H}\cong V_{\lambda^{\prime}}^{G_{n}}\otimes V_{\lambda^{\prime\prime}}^{G_{n}}. We can restrict this under the diagonal embedding of GnG_{n}, obtaining:

Lemma 6.1.

The restriction VλH(L)|GnV_{\lambda}^{H}(L)|_{G_{n}} to a diagonal copy of GnG_{n} contains the GnG_{n}-representation wp𝗐¯detw_{p}^{\underline{{\sf w}}}\circ\det with multiplicity 1.

Proof.

At each σΣ\sigma\in\Sigma, since λσ,i+λσ,2n+1i=𝗐σ\lambda_{\sigma,i}+\lambda_{\sigma,2n+1-i}={\sf w}_{\sigma}, we have λ′′=(λ)+(𝗐σ,,𝗐σ)\lambda^{\prime\prime}=(\lambda^{\prime})^{\vee}+({\sf w}_{\sigma},...,{\sf w}_{\sigma}), so

VλσGLnVλσ′′GLnVλ,σGLn(Vλ,σGLn)det𝗐σ,V_{\lambda^{\prime}_{\sigma}}^{\operatorname{GL}_{n}}\otimes V_{\lambda^{\prime\prime}_{\sigma}}^{\operatorname{GL}_{n}}\cong V_{\lambda^{\prime},\sigma}^{\operatorname{GL}_{n}}\otimes(V_{\lambda^{\prime},\sigma}^{\operatorname{GL}_{n}})^{\vee}\otimes\mathrm{det}^{{\sf w}_{\sigma}},

which contains det𝗐σ\det^{{\sf w}_{\sigma}} with multiplicity 1. Since VλGnVλ′′GnV_{\lambda^{\prime}}^{G_{n}}\otimes V_{\lambda^{\prime\prime}}^{G_{n}} is just the tensor product over all σΣ\sigma\in\Sigma, and wp𝗐¯(det(g))=σdetwσ(gσ)w_{p}^{\underline{{\sf w}}}(\det(g))=\otimes_{\sigma}\det^{w_{\sigma}}(g_{\sigma}), the result follows. ∎

Fix a generator vλHwp𝗐¯detVλH(L)|Gnv^{H}_{\lambda}\in w_{p}^{\underline{{\sf w}}}\circ\det\subset V_{\lambda}^{H}(L)|_{G_{n}}. Recall the description Vλ=IndQ¯GVλHV_{\lambda}=\operatorname{Ind}_{\overline{Q}}^{G}V_{\lambda}^{H} from (2.7). Let

NQ×(𝐙p) . . ={(1X01)NQ(𝐙p):XGn(𝐙p)}.N_{Q}^{\times}(\mathbf{Z}_{p})\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\left\{\left(\begin{smallmatrix}1&X\\ 0&1\end{smallmatrix}\right)\in N_{Q}(\mathbf{Z}_{p}):X\in G_{n}(\mathbf{Z}_{p})\right\}. (6.3)
Proposition 6.2.
  1. (i)

    For each 𝐣\mathbf{j} critical for λ\lambda, there exists a unique

    [v𝐣:G(𝐙p)VλH(L)]Vλ(L)\big{[}v_{\mathbf{j}}:G(\mathbf{Z}_{p})\to V_{\lambda}^{H}(L)\big{]}\in V_{\lambda}(L)

    with

    v𝐣[(1X01)]=wp𝐣(det(wnX))((X1)λvλH)v_{\mathbf{j}}\left[\left(\begin{smallmatrix}1&X\\ 0&1\end{smallmatrix}\right)\right]=w_{p}^{\mathbf{j}}(\det(w_{n}X))\cdot\Big{(}\left\langle\left(\begin{smallmatrix}X&\\ &1\end{smallmatrix}\right)\right\rangle_{\lambda}\cdot v_{\lambda}^{H}\Big{)}

    for all (1X01)NQ×(𝐙p)\left(\begin{smallmatrix}1&X\\ 0&1\end{smallmatrix}\right)\in N_{Q}^{\times}(\mathbf{Z}_{p}).

  2. (ii)

    v𝐣v_{\mathbf{j}} generates V𝐣,𝗐¯+𝐣H(L)Vλ(L)|H(𝐐p).V^{H}_{-\mathbf{j},\underline{{\sf w}}+\mathbf{j}}(L)\subset V_{\lambda}(L)|_{H(\mathbf{Q}_{p})}.

Proof.

This follows [BDW, Lem. 5.11]. Take some choices of κ𝐣\kappa_{\mathbf{j}} and u𝐣u_{\mathbf{j}}. These define dual bases

κ𝐣HomH(𝐐p)(V𝐣,𝗐¯+𝐣H(L),Vλ(L)),u𝐣V𝐣,𝗐¯+𝐣H(L).\kappa_{\mathbf{j}}^{\vee}\in\operatorname{Hom}_{H(\mathbf{Q}_{p})}(V_{-\mathbf{j},\underline{{\sf w}}+\mathbf{j}}^{H}(L),V_{\lambda}(L)),\qquad u_{\mathbf{j}}^{\vee}\in V^{H}_{-\mathbf{j},\underline{{\sf w}}+\mathbf{j}}(L). (6.4)

Thus κ𝐣(u𝐣)Vλ(L)\kappa_{\mathbf{j}}^{\vee}(u_{\mathbf{j}}^{\vee})\in V_{\lambda}(L); and via (2.7), we may define

vλ,𝐣H . . =κ𝐣(u𝐣)[(1n1n01n)]VλH(L).v_{\lambda,\mathbf{j}}^{H}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\kappa_{\mathbf{j}}^{\vee}(u_{\mathbf{j}}^{\vee})\left[\left(\begin{smallmatrix}1_{n}&1_{n}\\ 0&1_{n}\end{smallmatrix}\right)\right]\in V_{\lambda}^{H}(L).

Arguing exactly as in [BDW, Lem. 5.8], we see vλ,𝐣Hv_{\lambda,\mathbf{j}}^{H} is a non-zero element of wp𝗐¯detVλH(L)|Gnw_{p}^{\underline{{\sf w}}}\circ\det\subset V_{\lambda}^{H}(L)|_{G_{n}}, and hence – up to rescaling κ𝐣\kappa_{\mathbf{j}} – we may assume vλ,𝐣H=vλHv_{\lambda,\mathbf{j}}^{H}=v_{\lambda}^{H}, independent of 𝐣\mathbf{j}.

For these rescaled choices, define v𝐣 . . =κ𝐣(u𝐣)v_{\mathbf{j}}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\kappa_{\mathbf{j}}^{\vee}(u_{\mathbf{j}}^{\vee}). Then (i) follows exactly as in [BDW, Lem. 5.7], with uniqueness following from Zariski-density of N¯QHNQ×(𝐙p)G(𝐙p)\overline{N}_{Q}HN_{Q}^{\times}(\mathbf{Z}_{p})\subset G(\mathbf{Z}_{p}) (as in [BDW, Lem. 5.11(i)]). Part (ii) is then identical to [BDW, Lem. 5.11(ii)]. ∎

Definition 6.3.

We fix the choice of κ𝐣:Vλ(L)L\kappa_{\mathbf{j}}^{\circ}:V_{\lambda}^{\vee}(L)\to L by setting κ𝐣(μ)=μ(v𝐣)\kappa_{\mathbf{j}}^{\circ}(\mu)=\mu(v_{\mathbf{j}}). Note this corresponds to the choices of κ𝐣\kappa_{\mathbf{j}} and u𝐣u_{\mathbf{j}} in the proof, after rescaling so that the attached vλ,𝐣H=vλHv_{\lambda,\mathbf{j}}^{H}=v_{\lambda}^{H}.

6.2.  Branching laws for distributions

We now construct a ‘master branching law’ κλ\kappa_{\lambda}, interpolating all the κ𝐣\kappa_{\mathbf{j}} above. Let 𝒜(Galp,L)\mathcal{A}(\operatorname{Gal}_{p},L) be the space of locally analytic functions on Galp\operatorname{Gal}_{p}, with an H(𝐀)H(\mathbf{A})-action by

(h1,h2)f(x) . . =η[p](h2)f(h11h2x).(h_{1},h_{2})*f(x)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\eta_{[p]}(h_{2})f(h_{1}^{-1}h_{2}x).

This induces a dual left-action on 𝒟(Galp,L) . . =Homcts(𝒜(Galp,L),L)\mathcal{D}(\operatorname{Gal}_{p},L)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\mathrm{Hom}_{\mathrm{cts}}(\mathcal{A}(\operatorname{Gal}_{p},L),L). Recall if χCritp(π)\chi\in\mathrm{Crit}_{p}(\pi), then χ[p]\chi_{[p]} induces a character on Galp\operatorname{Gal}_{p} via §2.5. The following is immediate.

Lemma 6.4.
  1. (i)

    H(𝐑)H(\mathbf{R})^{\circ} and H(𝐐)H(\mathbf{Q}) act trivially on 𝒟(Galp,L)\mathcal{D}(\operatorname{Gal}_{p},L).

  2. (ii)

    The map μμ(χ[p])\mu\mapsto\mu(\chi_{[p]}) defines an H(𝐀)H(\mathbf{A})-module map 𝒟(Galp,L)Vχ,[p]H(L)\mathcal{D}(\operatorname{Gal}_{p},L)\to V_{\chi,[p]}^{H}(L).

Recall from (2.4) there is a natural map (𝒪F𝐙p)×Galp(\mathcal{O}_{F}\otimes\mathbf{Z}_{p})^{\times}\to\operatorname{Gal}_{p}, which we denote ȷ\jmath. Given f𝒜(Galp,L)f\in\mathcal{A}(\operatorname{Gal}_{p},L) and x(𝒪F𝐙p)×x\in(\mathcal{O}_{F}\otimes\mathbf{Z}_{p})^{\times}, abusing notation we write f(x) . . =f(ȷ(x))f(x)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=f(\jmath(x)). For such ff, define a map

vλ(f):NQ×(𝐙p)\displaystyle v_{\lambda}(f):N_{Q}^{\times}(\mathbf{Z}_{p}) VλH(L),\displaystyle\longrightarrow V_{\lambda}^{H}(L), (6.5)
(1X01)\displaystyle\left(\begin{smallmatrix}1&X\\ 0&1\end{smallmatrix}\right) f(det(wnX))((X1)λvλH),\displaystyle\longmapsto f(\det(w_{n}X))\left(\langle\left(\begin{smallmatrix}X&\\ &1\end{smallmatrix}\right)\rangle_{\lambda}\cdot v_{\lambda}^{H}\right),

noting det(X)(𝒪F𝐙p)×\det(X)\in(\mathcal{O}_{F}\otimes\mathbf{Z}_{p})^{\times} by the definition (6.3) of NQ×(𝐙p)N_{Q}^{\times}(\mathbf{Z}_{p}). Extending by 0, vλ(f)v_{\lambda}(f) is a locally analytic function NQ(𝐙p)VλH(L)N_{Q}(\mathbf{Z}_{p})\to V_{\lambda}^{H}(L), whence it defines an element vλ(f)𝒜λv_{\lambda}(f)\in\mathcal{A}_{\lambda} via the parahoric transformation law (2.9).

If there was a function wp𝐣𝒜(Galp,L)w_{p}^{\mathbf{j}}\in\mathcal{A}(\operatorname{Gal}_{p},L), then comparing Proposition 6.2 with (6.5), we would formally have vλ(wp𝐣)(n)=v𝐣(n)v_{\lambda}(w_{p}^{\mathbf{j}})(n)=v_{\mathbf{j}}(n) for nNQ×(𝐙p)n\in N_{Q}^{\times}(\mathbf{Z}_{p}). Unfortunately the function wp𝐣w_{p}^{\mathbf{j}} on (F𝐐p)×(F\otimes\mathbf{Q}_{p})^{\times} does not induce a function on Galp\operatorname{Gal}_{p} in general (since it is not 𝐐×\mathbf{Q}^{\times}-invariant). However, if χCritp(π)\chi\in\mathrm{Crit}_{p}(\pi) with infinity type 𝐣\mathbf{j} and conductor dividing pβp^{\beta}, then χ[p]\chi_{[p]} is a function on Galp\operatorname{Gal}_{p}, and χ[p](x)=wp𝐣(x)\chi_{[p]}(x)=w_{p}^{\mathbf{j}}(x) for x𝐀F×x\in\mathbf{A}_{F}^{\times} with x1(modpβ)x\equiv 1\hskip 2.0pt(\mathrm{mod}\hskip 2.0ptp^{\beta}). As such, recalling that β=(β𝔭)𝐙1𝔭|p\beta=(\beta_{\mathfrak{p}})\in\mathbf{Z}_{\geqslant 1}^{\mathfrak{p}|p}, let:

  • 𝒰β=ȷ(1+pβ𝒪F𝐙p)Galp\mathscr{U}_{\beta}=\jmath(1+p^{\beta}\mathcal{O}_{F}\otimes\mathbf{Z}_{p})\subset\operatorname{Gal}_{p},

  • 𝒜(𝒰β,L)𝒜(Galp,L)\mathcal{A}(\mathscr{U}_{\beta},L)\subset\mathcal{A}(\operatorname{Gal}_{p},L) be the space of functions supported on 𝒰β\mathscr{U}_{\beta},

  • NQβ(𝐙p) . . ={nNQ(𝐙p):n(1wn01n)(modpβ)}NQ×(𝐙p),N^{\beta}_{Q}(\mathbf{Z}_{p})\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\{n\in N_{Q}(\mathbf{Z}_{p}):n\equiv\left(\begin{smallmatrix}1&w_{n}\\ 0&1_{n}\end{smallmatrix}\right)\hskip 2.0pt(\mathrm{mod}\hskip 2.0ptp^{\beta})\}\subset N_{Q}^{\times}(\mathbf{Z}_{p}),

  • Jpβ . . =JpN¯Q(𝐙p)H(𝐙p)NQβ(𝐙p)JpJ_{p}^{\beta}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=J_{p}\cap\overline{N}_{Q}(\mathbf{Z}_{p})H(\mathbf{Z}_{p})N_{Q}^{\beta}(\mathbf{Z}_{p})\subset J_{p},

  • 𝒜λβ𝒜λ\mathcal{A}_{\lambda}^{\beta}\subset\mathcal{A}_{\lambda} (resp. 𝒟λβ𝒟λ\mathcal{D}_{\lambda}^{\beta}\subset\mathcal{D}_{\lambda}) be the space of functions (resp. distributions) supported on JpβJ_{p}^{\beta}.

If n=(1X01)NQβ(𝐙p)n=\left(\begin{smallmatrix}1&X\\ 0&1\end{smallmatrix}\right)\in N_{Q}^{\beta}(\mathbf{Z}_{p}), then Xwn(modpβ)X\equiv w_{n}\hskip 2.0pt(\mathrm{mod}\hskip 2.0ptp^{\beta}), so det(wnX)1(modpβ)\det(w_{n}X)\equiv 1\hskip 2.0pt(\mathrm{mod}\hskip 2.0ptp^{\beta}). If f𝒜(𝒰β,L)f\in\mathcal{A}(\mathscr{U}_{\beta},L), we can thus define a function vλβ(f):𝐍Qβ(𝐙p)VλH(L)v_{\lambda}^{\beta}(f):\mathbf{N}_{Q}^{\beta}(\mathbf{Z}_{p})\to V_{\lambda}^{H}(L) exactly as in (6.5). After extending by 0 to NQ(𝐙p)N_{Q}(\mathbf{Z}_{p}), as above we obtain a function

vλβ:𝒜(𝒰β,L)𝒜λβ.v_{\lambda}^{\beta}:\mathcal{A}(\mathscr{U}_{\beta},L)\longrightarrow\mathcal{A}_{\lambda}^{\beta}.

Then, recalling that χ\chi has conductor dividing pβp^{\beta} and infinity type 𝐣\mathbf{j}:

Lemma 6.5.

If gJpβg\in J_{p}^{\beta}, then vλβ(χ[p])(g)=v𝐣(g)v_{\lambda}^{\beta}(\chi_{[p]})(g)=v_{\mathbf{j}}(g).

Proof.

As both vλβ(χ[p])v_{\lambda}^{\beta}(\chi_{[p]}) and v𝐣v_{\mathbf{j}} are functions JpVλH(L)J_{p}\to V_{\lambda}^{H}(L) satisfying the same parahoric transformation law, it suffices to show this for g=(1X01)NQβ(𝐙p)g=\left(\begin{smallmatrix}1&X\\ 0&1\end{smallmatrix}\right)\in N_{Q}^{\beta}(\mathbf{Z}_{p}). Then det(wnX)1(modpβ)\det(w_{n}X)\equiv 1\hskip 2.0pt(\mathrm{mod}\hskip 2.0ptp^{\beta}), so χ[p](det(wnX))=wp𝐣(det(wnX))\chi_{[p]}(\det(w_{n}X))=w_{p}^{\mathbf{j}}(\det(w_{n}X)). Thus (6.5) recovers the formula for v𝐣v_{\mathbf{j}} in Proposition 6.2. ∎

Dualising vλβv_{\lambda}^{\beta} gives a map κλβ:𝒟λβ𝒟(𝒰β,L)𝒟(Galp,L)\kappa_{\lambda}^{\beta}:\mathcal{D}_{\lambda}^{\beta}\to\mathcal{D}(\mathscr{U}_{\beta},L)\subset\mathcal{D}(\operatorname{Gal}_{p},L).

Proposition 6.6.

If χCritp(π)\chi\in\mathrm{Crit}_{p}(\pi) has conductor dividing pβp^{\beta} and infinity type 𝐣\mathbf{j}, we have a commutative diagram

𝒟λβ(L)\textstyle{\mathcal{D}_{\lambda}^{\beta}(L)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}rλ\scriptstyle{r_{\lambda}}κλβ\scriptstyle{\kappa_{\lambda}^{\beta}}𝒟(Galp,L)\textstyle{\mathcal{D}(\operatorname{Gal}_{p},L)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μμ(χ[p])\scriptstyle{\mu\mapsto\mu(\chi_{[p]})}Vλ(L)\textstyle{V_{\lambda}^{\vee}(L)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}κ𝐣\scriptstyle{\kappa_{\mathbf{j}}^{\circ}}L.\textstyle{L.}
Proof.

Let μ𝒟λβ\mu\in\mathcal{D}_{\lambda}^{\beta}. Then

κλβ(μ)(χ[p])=Galpχ[p]𝑑κλβ(μ)=Galpvλβ(χ[p])𝑑μ=Jpβvλβ(χ[p])𝑑μ=Jpβv𝐣𝑑μ=κ𝐣rλ(μ),\kappa_{\lambda}^{\beta}(\mu)(\chi_{[p]})=\int_{\operatorname{Gal}_{p}}\chi_{[p]}\cdot d\kappa_{\lambda}^{\beta}(\mu)=\int_{\operatorname{Gal}_{p}}v_{\lambda}^{\beta}(\chi_{[p]})\cdot d\mu=\int_{J_{p}^{\beta}}v_{\lambda}^{\beta}(\chi_{[p]})\cdot d\mu=\int_{J_{p}^{\beta}}v_{\mathbf{j}}\cdot d\mu=\kappa_{\mathbf{j}}^{\circ}\circ r_{\lambda}(\mu),

where we have used that μ\mu has support on JpβJ_{p}^{\beta}, and μ(v𝐣)=μ(rλ(v𝐣))\mu(v_{\mathbf{j}})=\mu(r_{\lambda}(v_{\mathbf{j}})) since v𝐣Vλv_{\mathbf{j}}\in V_{\lambda}. ∎

6.3.  Overconvergent evaluation maps

Recall KK (hence LβL_{\beta}) acts on 𝒟λ\mathcal{D}_{\lambda} via §2.7.3 by projection to KpJpK_{p}\subset J_{p}, giving a local system 𝒟λ\mathscr{D}_{\lambda} on SKS_{K} by §2.6.

Lemma 6.7.

The action of LβKL_{\beta}\subset K preserves 𝒟λβ\mathcal{D}_{\lambda}^{\beta}.

Proof.

If f𝒜λβf\in\mathcal{A}_{\lambda}^{\beta}, and =(1,2)Lβ\ell=(\ell_{1},\ell_{2})\in L_{\beta}, then for any (1X01)NQ(𝐙p)\left(\begin{smallmatrix}1&X\\ 0&1\end{smallmatrix}\right)\in N_{Q}(\mathbf{Z}_{p}), we have (f)(1X01)=Hf(111X201)(\ell*f)\left(\begin{smallmatrix}1&X\\ 0&1\end{smallmatrix}\right)=\langle\ell\rangle_{H}\cdot f\left(\begin{smallmatrix}1&\ell_{1}^{-1}X\ell_{2}\\ 0&1\end{smallmatrix}\right). By the proof of [BDG+, Lem. 4.5], we have 2wn1wn(modpβ)\ell_{2}\equiv w_{n}\ell_{1}w_{n}\hskip 2.0pt(\mathrm{mod}\hskip 2.0ptp^{\beta}); so 11X2wn(modpβ)\ell_{1}^{-1}X\ell_{2}\equiv w_{n}\hskip 2.0pt(\mathrm{mod}\hskip 2.0ptp^{\beta}) if and only if X2wn21wn(modpβ)X\equiv\ell_{2}w_{n}\ell_{2}^{-1}\equiv w_{n}\hskip 2.0pt(\mathrm{mod}\hskip 2.0ptp^{\beta}). Thus f\ell*f has support on NQβ(𝐙p)N_{Q}^{\beta}(\mathbf{Z}_{p}) if ff does; so LβL_{\beta} preserves 𝒜λβ\mathcal{A}_{\lambda}^{\beta}, hence 𝒟λβ\mathcal{D}_{\lambda}^{\beta}. ∎

To make Proposition 6.6 useful, we need the following support condition.

Lemma 6.8.

The map

𝒱β,δ𝒟λ,[]:Hct(SK,𝒟λ)(𝒟λ)Γβ,δ\mathscr{E}{\scriptstyle\mathscr{V}}_{\beta,\delta}^{\mathcal{D}_{\lambda},[*]}:\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathscr{D}_{\lambda})\to(\mathcal{D}_{\lambda})_{\Gamma_{\beta,\delta}}

has image in (𝒟λβ)Γβ,δ(𝒟λ)Γβ,δ(\mathcal{D}_{\lambda}^{\beta})_{\Gamma_{\beta,\delta}}\subset(\mathcal{D}_{\lambda})_{\Gamma_{\beta,\delta}}.

Proof.

Identical to [BDG+, Lem. 12.4]. ∎

The following, and Lemma 6.4, allows us to use Proposition 5.9.

Lemma 6.9.

The map κλβ:𝒟λβ𝒟(Galp,L)\kappa_{\lambda}^{\beta}:\mathcal{D}_{\lambda}^{\beta}\to\mathcal{D}(\operatorname{Gal}_{p},L) is a map of LβL_{\beta}-modules.

Proof.

It suffices to prove that vλβ:𝒜(𝒰β,L)𝒜λβv_{\lambda}^{\beta}:\mathcal{A}(\mathscr{U}_{\beta},L)\to\mathcal{A}_{\lambda}^{\beta} is a map of LβL_{\beta}-modules. This is almost identical to [BDW, Prop. 6.11(ii)]. We observe that if f𝒜(Galp,L)f\in\mathcal{A}(\operatorname{Gal}_{p},L) and =(1,2)Lβ\ell=(\ell_{1},\ell_{2})\in L_{\beta}, then (f)(x)=wp(det(2,p))𝗐¯f(det(1,p12,p))(\ell*f)(x)=w_{p}(\det(\ell_{2,p}))^{\underline{{\sf w}}}f(\det(\ell_{1,p}^{-1}\ell_{2,p})), since η(det(2))=1\eta(\det(\ell_{2}))=1 (as in Lemma 5.4) and the prime-to-pp part of 112\ell_{1}^{-1}\ell_{2} lies in 𝒰(p)\mathscr{U}(p^{\infty}), i.e. acts trivially on ClF+(p)\mathrm{Cl}_{F}^{+}(p^{\infty}). The rest of the proof proceeds exactly as op. cit. ∎

Proposition 6.10.

For any δH(𝐀f)\delta\in H(\mathbf{A}_{f}), the map

𝒱β,[δ]:Hct(SK,𝒟λ(L))𝒱β,δ𝒟λ,[](𝒟λβ(L))Γβ,δκλβ𝒟(Galp,L)δ𝒟(Galp,L)\mathscr{E}{\scriptstyle\mathscr{V}}_{\beta,[\delta]}^{\dagger}:\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathscr{D}_{\lambda}^{\vee}(L))\xrightarrow{\ \mathscr{E}{\scriptstyle\mathscr{V}}_{\beta,\delta}^{\mathcal{D}_{\lambda},[*]}\ }\big{(}\mathcal{D}_{\lambda}^{\beta}(L)\big{)}_{\Gamma_{\beta,\delta}}\xrightarrow{\kappa_{\lambda}^{\beta}}\mathcal{D}(\operatorname{Gal}_{p},L)\xrightarrow{\ \cdot\delta\ }\mathcal{D}(\operatorname{Gal}_{p},L)

is well-defined and depends only on the class [δ]π0(Xβ)[\delta]\in\pi_{0}(X_{\beta}).

Proof.

Immediate by combining Proposition 5.9 with Lemmas 6.4(i) and 6.9. ∎

Definition 6.11.

The overconvergent evaluation map of level β\beta is the map

𝒱β . . =δπ0(Xβ)𝒱β,[δ]:Hct(SK,𝒟λ(L))𝒟(Galp,L).\mathscr{E}{\scriptstyle\mathscr{V}}_{\beta}^{\dagger}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\sum_{\delta\in\pi_{0}(X_{\beta})}\mathscr{E}{\scriptstyle\mathscr{V}}_{\beta,[\delta]}^{\dagger}:\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathscr{D}_{\lambda}^{\vee}(L))\longrightarrow\mathcal{D}(\operatorname{Gal}_{p},L).

6.4.  Interpolation of classical evaluation maps

The map rλr_{\lambda} from (2.10) is by definition a Δp\Delta_{p}-module map when VλV_{\lambda}^{\vee} is given the *-action, hence induces a UpU_{p}^{*}-equivariant map

rλ:Hct(SK,𝒟λ)Hct(SK,𝒱λ).r_{\lambda}:\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathscr{D}_{\lambda})\longrightarrow\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathscr{V}_{\lambda}^{\vee}). (6.6)
Proposition 6.12.

For any χ\chi of conductor dividing pβp^{\beta}, we have a commutative diagram

Hct(SK,𝒟λ(L))\textstyle{\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathscr{D}_{\lambda}(L))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒱β\scriptstyle{\mathscr{E}{\scriptscriptstyle\mathscr{V}}_{\beta}^{\dagger}}rλ\scriptstyle{r_{\lambda}}𝒟(Galp,L)\textstyle{\mathcal{D}(\operatorname{Gal}_{p},L)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μμ(χ[p])\scriptstyle{\mu\mapsto\mu(\chi_{[p]})}Hct(SK,𝒱λ(L))\textstyle{\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathscr{V}_{\lambda}^{\vee}(L))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒱χ[]\scriptstyle{\mathscr{E}{\scriptscriptstyle\mathscr{V}}_{\chi}^{[*]}}L.\textstyle{L.}
Proof.

It suffices to check every square in

Hct(SK,𝒟λ(L))\textstyle{\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathscr{D}_{\lambda}(L))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒱β,δ𝒟λ,[]\scriptstyle{\mathscr{E}{\scriptscriptstyle\mathscr{V}}_{\beta,\delta}^{\mathcal{D}_{\lambda},[*]}}rλ\scriptstyle{r_{\lambda}}(𝒟λβ(L))Γβ,δ\textstyle{(\mathcal{D}_{\lambda}^{\beta}(L))_{\Gamma_{\beta,\delta}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}κλ\scriptstyle{\kappa_{\lambda}}rλ\scriptstyle{r_{\lambda}}𝒟(Galp,L)\textstyle{\mathcal{D}(\operatorname{Gal}_{p},L)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μμ(χ[p])\scriptstyle{\mu\mapsto\mu(\chi_{[p]})}δ\scriptstyle{\delta*}𝒟(Galp,L)\textstyle{\mathcal{D}(\operatorname{Gal}_{p},L)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μμ(χ[p])\scriptstyle{\mu\mapsto\mu(\chi_{[p]})}Hct(SK,𝒱λ(L))\textstyle{\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathscr{V}_{\lambda}^{\vee}(L))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒱β,δVλ,[]\scriptstyle{\mathscr{E}{\scriptscriptstyle\mathscr{V}}_{\beta,\delta}^{V_{\lambda}^{\vee},[*]}}(Vλ(L))Γβ,δ\textstyle{(V_{\lambda}^{\vee}(L))_{\Gamma_{\beta,\delta}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}κ𝐣\scriptstyle{\kappa_{\mathbf{j}}^{\circ}}L\textstyle{L\ignorespaces\ignorespaces\ignorespaces\ignorespaces}δ\scriptstyle{\delta*}L\textstyle{L}

commutes, since the top row is 𝒱β\mathscr{E}{\scriptstyle\mathscr{V}}^{\dagger}_{\beta} and the bottom row χ[]\mathscr{E}_{\chi}^{[*]}. The left-most square commutes by Lemma 5.8 (noting the top arrow is well-defined by Lemma 6.8). The middle square commutes by Proposition 6.6. The right-most square commutes by Lemma 6.4(ii), noting that δ\delta acts on the bottom row via the identification LVχ,[p]H(L)L\cong V_{\chi,[p]}^{H}(L). ∎

7.  The pp-adic LL-function

7.1.  Construction

Let now π\pi be a RASCAR of weight λ\lambda, spherical at all 𝔭|p\mathfrak{p}|p, and let π~=(π,{α𝔭}𝔭}\tilde{\pi}=(\pi,\{\alpha_{\mathfrak{p}}\}_{\mathfrak{p}}\} be a regular spin QQ-refinement as in §3.2. Let φfFJ,απf\varphi_{f}^{\mathrm{FJ},\alpha}\in\pi_{f} and [ω][\omega] be chosen as in Test Data 3.9, and let

ϕπ~ . . =Θ[ω]K,ip(φfFJ,α)Hct(SK,𝒱λ(𝐐¯p)).\phi_{\tilde{\pi}}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\Theta_{[\omega]}^{K,i_{p}}(\varphi_{f}^{\mathrm{FJ},\alpha})\in\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathscr{V}_{\lambda}^{\vee}(\overline{\mathbf{Q}}_{p})). (7.1)

There exists some finite extension L/𝐐pL/\mathbf{Q}_{p} such that ϕπ~\phi_{\tilde{\pi}} is defined over LL, rather than 𝐐¯p\overline{\mathbf{Q}}_{p}. The map Θ[ω]K,ip\Theta_{[\omega]}^{K,i_{p}} is Hecke-equivariant when we give the right-hand side the \cdot-action, hence

U𝔭ϕπ~=α𝔭ϕπ~,U𝔭ϕπ~=λ(t𝔭)α𝔭ϕπ~,U_{\mathfrak{p}}^{\cdot}\phi_{\tilde{\pi}}=\alpha_{\mathfrak{p}}\phi_{\tilde{\pi}},\qquad U_{\mathfrak{p}}^{*}\phi_{\tilde{\pi}}=\lambda(t_{\mathfrak{p}})\alpha_{\mathfrak{p}}\phi_{\tilde{\pi}},

the second via (2.12). Let α𝔭 . . =λ(t𝔭)α𝔭\alpha_{\mathfrak{p}}^{\circ}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\lambda(t_{\mathfrak{p}})\alpha_{\mathfrak{p}}, the U𝔭U_{\mathfrak{p}}^{*}-eigenvalue.

Definition 7.1.

Recall the map rλr_{\lambda} from (6.6). We say π~\tilde{\pi} is non-QQ-critical if rλr_{\lambda} becomes an isomorphism after restricting to the π~\tilde{\pi}-generalised eigenspaces for the Hecke algebra (precisely, the algebra \mathcal{H} from [BDW, Def. 2.10]). We say π~\tilde{\pi} has non-QQ-critical slope if

e𝔭vp(α𝔭)<minσΣ(𝔭)(1+λσ,nλσ,n+1) for all 𝔭|p,e_{\mathfrak{p}}\cdot v_{p}(\alpha_{\mathfrak{p}}^{\circ})<\mathrm{min}_{\sigma\in\Sigma(\mathfrak{p})}(1+\lambda_{\sigma,n}-\lambda_{\sigma,n+1})\qquad\text{ for all $\mathfrak{p}|p$},

where Σ(𝔭)\Sigma(\mathfrak{p}) is the set of σ:F𝐂𝐐¯p\sigma:F\hookrightarrow\mathbf{C}\xrightarrow{\,\smash{\raisebox{-2.79857pt}{$\scriptstyle\sim$}}\,}\overline{\mathbf{Q}}_{p} that extend to F𝔭𝐐¯pF_{\mathfrak{p}}\hookrightarrow\overline{\mathbf{Q}}_{p}, and e𝔭e_{\mathfrak{p}} is the ramification index of 𝔭|p\mathfrak{p}|p. (If π~\tilde{\pi} is ordinary – i.e. vp(α𝔭)=0v_{p}(\alpha_{\mathfrak{p}}^{\circ})=0 for all 𝔭\mathfrak{p} – then it has non-QQ-critical slope.)

The following is [BW21c, Thm. 4.4] (see also [BDW, Thm. 3.17]).

Theorem 7.2.

If π~\tilde{\pi} has non-QQ-critical slope, then it is non-QQ-critical.

If π~\tilde{\pi} is non-QQ-critical, then there exists a unique Hecke eigenclass Φπ~Hct(SK,𝒟λ)\Phi_{\tilde{\pi}}\in\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathscr{D}_{\lambda}) with rλ(Φπ~)=ϕπ~r_{\lambda}(\Phi_{\tilde{\pi}})=\phi_{\tilde{\pi}}. If β𝐙1𝔭|p\beta\in\mathbf{Z}_{\geqslant 1}^{\mathfrak{p}|p}, then let (αp)β=𝔭|p(α𝔭)β𝔭(\alpha_{p}^{\circ})^{\beta}=\prod_{\mathfrak{p}|p}(\alpha_{\mathfrak{p}}^{\circ})^{\beta_{\mathfrak{p}}}.

Definition 7.3.

Let β𝐙1𝔭|p\beta\in\mathbf{Z}_{\geqslant 1}^{\mathfrak{p}|p}. The pp-adic LL-function attached to π~\tilde{\pi} and ipi_{p} is

Lpip(π~) . . =(αp)β𝒱β(Φπ~)𝒟(Galp,L).L_{p}^{i_{p}}(\tilde{\pi})\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=(\alpha_{p}^{\circ})^{-\beta}\mathscr{E}{\scriptstyle\mathscr{V}}_{\beta}^{\dagger}(\Phi_{\tilde{\pi}})\in\mathcal{D}(\operatorname{Gal}_{p},L).

By Proposition 5.10 (with =\bullet=*), this is independent of β\beta. Dependence on ipi_{p} is (7.1) (via (5.1)).

7.2.  Growth conditions

The following is an adaptation of [Loe14, Def. 2.14] for our setting. For 𝐦=(m𝔭)𝐙1𝔭|p\mathbf{m}=(m_{\mathfrak{p}})\in\mathbf{Z}_{\geqslant 1}^{\mathfrak{p}|p}, let p𝐦=𝔭m𝔭p^{\mathbf{m}}=\prod\mathfrak{p}^{m_{\mathfrak{p}}}. We have an exact sequence

0𝒰𝐦GalpClF+(p𝐦)0,0\to\mathscr{U}_{\mathbf{m}}\to\operatorname{Gal}_{p}\to\mathrm{Cl}_{F}^{+}(p^{\mathbf{m}})\to 0,

where 𝒰𝐦\mathscr{U}_{\mathbf{m}} is the image of 1+p𝐦𝒪F𝐙p1+p^{\mathbf{m}}\mathcal{O}_{F}\otimes\mathbf{Z}_{p} under the natural map (𝒪F𝐙p)×Galp(\mathcal{O}_{F}\otimes\mathbf{Z}_{p})^{\times}\to\operatorname{Gal}_{p} from (2.4). Let

𝒜𝐦(Galp,L) . . ={f𝒜(Galp,L):f is analytic on every translate of 𝒰𝐦}.\mathcal{A}_{\mathbf{m}}(\operatorname{Gal}_{p},L)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\{f\in\mathcal{A}(\operatorname{Gal}_{p},L):f\text{ is analytic on every translate of $\mathscr{U}_{\mathbf{m}}$}\}.

Then 𝒜(Galp,L)=lim𝐦𝒜𝐦(Galp,L)\mathcal{A}(\operatorname{Gal}_{p},L)=\varinjlim_{\mathbf{m}}\mathcal{A}_{\mathbf{m}}(\operatorname{Gal}_{p},L) is the union of all the 𝒜𝐦\mathcal{A}_{\mathbf{m}}’s. Moreover each 𝒜𝐦\mathcal{A}_{\mathbf{m}} is a Banach LL-space with respect to a discretely valued norm ||||𝐦||\cdot||_{\mathbf{m}}. Dualising gives a family of norms

μ𝐦\displaystyle||\mu||_{\mathbf{m}} . . =supf𝒜𝐦(Galp,L)|μ(f)|f𝐦\displaystyle\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\mathrm{sup}_{f\in\mathcal{A}_{\mathbf{m}}(\operatorname{Gal}_{p},L)}\tfrac{|\mu(f)|}{||f||_{\mathbf{m}}}
=supf𝐦1|μ(f)|\displaystyle=\mathrm{sup}_{||f||_{\mathbf{m}}\leqslant 1}|\mu(f)| (7.2)

on 𝒟(Galp,L)\mathcal{D}(\operatorname{Gal}_{p},L), which thus obtains the structure of a Fréchet module.

Definition 7.4.

Let 𝐡=(h𝔭)𝐐0𝔭|p\mathbf{h}=(h_{\mathfrak{p}})\in\mathbf{Q}_{\geqslant 0}^{\mathfrak{p}|p}. We say μ𝒟(Galp,L)\mu\in\mathcal{D}(\operatorname{Gal}_{p},L) is admissible of growth 𝐡\mathbf{h} if there exists C0C\geqslant 0 such that for each 𝐦𝐙1𝔭|p,\mathbf{m}\in\mathbf{Z}_{\geqslant 1}^{\mathfrak{p}|p}, we have μ𝐦p𝐡𝐦C,||\mu||_{\mathbf{m}}\leqslant p^{\mathbf{h}\mathbf{m}}C, where 𝐡𝐦=(h𝔭m𝔭)𝔭\mathbf{h}\mathbf{m}=(h_{\mathfrak{p}}m_{\mathfrak{p}})_{\mathfrak{p}}.

Note that if μ\mu is admissible of growth (0)𝔭|p(0)_{\mathfrak{p}|p}, then it defines a bounded measure. Indeed, if XGalpX\subset\operatorname{Gal}_{p} is open compact, define its volume to be μ(1X)\mu(1_{X}), where 1X1_{X} is the indicator function. For 𝐦\mathbf{m} such that 1X𝒜𝐦(Galp,L)1_{X}\in\mathcal{A}_{\mathbf{m}}(\operatorname{Gal}_{p},L), we see |μ(1X)|=|μ(1X)|1X𝐦μ𝐦C|\mu(1_{X})|=\tfrac{|\mu(1_{X})|}{||1_{X}||_{\mathbf{m}}}\leqslant||\mu||_{\mathbf{m}}\leqslant C, so μ\mu is bounded.

Proposition 7.5.

Let ΦHct(SK,𝒟λ)\Phi\in\mathrm{H}^{t}_{\mathrm{c}}(S_{K},\mathscr{D}_{\lambda}) be a Hecke eigenclass with U𝔭Φ=α𝔭ΦU_{\mathfrak{p}}^{*}\Phi=\alpha_{\mathfrak{p}}^{\circ}\Phi for all 𝔭\mathfrak{p}. The distribution 𝒱β(Φ)𝒟(Galp,L)\mathscr{E}{\scriptstyle\mathscr{V}}_{\beta}^{\dagger}(\Phi)\in\mathcal{D}(\operatorname{Gal}_{p},L) is admissible of growth 𝐡p . . =(vp(α𝔭))𝔭|p\mathbf{h}_{p}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=(v_{p}(\alpha_{\mathfrak{p}}^{\circ}))_{\mathfrak{p}|p}.

Proof.

This is a higher-variable version of [BDW, Prop. 6.20] and [BDJ22, Prop. 3.11], and it is proved essentially identically. We indicate changes of notation. One defines β𝐦=(m𝔭e𝔭)𝔭|p\beta_{\mathbf{m}}=(m_{\mathfrak{p}}e_{\mathfrak{p}})_{\mathfrak{p}|p}, and replaces all instances of mm op. cit. (a positive integer) with 𝐦\mathbf{m}. After (6.21) op. cit., gg should instead be analytic on NQβ(𝐙p)N_{Q}^{\beta}(\mathbf{Z}_{p}), and ξ\xi there replaced with u1u^{-1} here. The space Galp[prβm([δ])]\operatorname{Gal}_{p}[\mathrm{pr}_{\beta_{m}}([\delta])] there is the translate of 𝒰𝐦\mathscr{U}_{\mathbf{m}} by δ\delta here. The rest of the proof is identical. ∎

7.3.  Proof of Theorem A

Finally we assemble our results and prove Theorem A from the introduction. We have constructed the distribution Lpip(π~)L_{p}^{i_{p}}(\tilde{\pi}) in Definition 7.3. It has the required growth property by Proposition 7.5. For the interpolation, if χ\chi has conductor dividing pβp^{\beta}, we have

Galpχ[p]𝑑Lpip(π~)=(αp)βGalpχ[p]𝑑𝒱β(Φπ~)=(αp)β𝒱χ[](rλ(Φπ~)),\int_{\operatorname{Gal}_{p}}\chi_{[p]}\cdot dL_{p}^{i_{p}}(\tilde{\pi})=(\alpha_{p}^{\circ})^{-\beta}\int_{\operatorname{Gal}_{p}}\chi_{[p]}\cdot d\mathscr{E}{\scriptstyle\mathscr{V}}_{\beta}^{\dagger}(\Phi_{\tilde{\pi}})=(\alpha_{p}^{\circ})^{-\beta}\cdot\mathscr{E}{\scriptstyle\mathscr{V}}_{\chi}^{[*]}(r_{\lambda}(\Phi_{\tilde{\pi}})),

by Proposition 6.12. Since rλ(Φπ~)=ϕπ~=Θ[ω]K,ip(φfFJ,α)r_{\lambda}(\Phi_{\tilde{\pi}})=\phi_{\tilde{\pi}}=\Theta_{[\omega]}^{K,i_{p}}(\varphi_{f}^{\mathrm{FJ},\alpha}), this gives (αp)β(\alpha_{p}^{\circ})^{-\beta} times the right-hand side of Theorem 5.16, which has a factor λ(tpβ)\lambda(t_{p}^{\beta}). By definition of α𝔭\alpha_{\mathfrak{p}}^{\circ}, we have

λ(tpβ)(αp)β=αpβ=𝔭|pα𝔭β𝔭.\lambda(t_{p}^{\beta})(\alpha_{p}^{\circ})^{-\beta}=\alpha_{p}^{-\beta}=\prod_{\mathfrak{p}|p}\alpha_{\mathfrak{p}}^{-\beta_{\mathfrak{p}}}.

If χ𝔭\chi_{\mathfrak{p}} is unramified, then β𝔭=1\beta_{\mathfrak{p}}=1, and this factor of α𝔭\alpha_{\mathfrak{p}} cancels the one appearing in Q(π𝔭,χ𝔭)Q^{\prime}(\pi_{\mathfrak{p}},\chi_{\mathfrak{p}}) in Theorem 5.16, leaving exactly e(π𝔭,χ𝔭)e(\pi_{\mathfrak{p}},\chi_{\mathfrak{p}}) from Theorem A. If χ𝔭\chi_{\mathfrak{p}} is ramified, then this factor of α𝔭β\alpha_{\mathfrak{p}}^{-\beta} is exactly the difference between Q(π𝔭,χ𝔭)Q^{\prime}(\pi_{\mathfrak{p}},\chi_{\mathfrak{p}}) and e(π𝔭,χ𝔭)e(\pi_{\mathfrak{p}},\chi_{\mathfrak{p}}). This proves the interpolation formula, and completes the proof of Theorem A. ∎

7.4.  Unicity in the imaginary quadratic case

Now let FF be imaginary quadratic. The purity condition on λ\lambda (Definition 2.2) implies that λσ,nλσ,n+1=λcσ,nλcσ,n+1=k\lambda_{\sigma,n}-\lambda_{\sigma,n+1}=\lambda_{c\sigma,n}-\lambda_{c\sigma,n+1}=k, say, so the range of standard critical infinity types for λ\lambda forms the (k+1)×(k+1)(k+1)\times(k+1) square {(jσ,jcσ)𝐙2:λσ,n+1jσλσ,n,λcσ,n+1jcσλcσ,n\{(j_{\sigma},j_{c\sigma})\in\mathbf{Z}^{2}:-\lambda_{\sigma,n+1}\geqslant j_{\sigma}\geqslant-\lambda_{\sigma,n},-\lambda_{c\sigma,n+1}\geqslant j_{c\sigma}\geqslant-\lambda_{c\sigma,n}. The non-QQ-critical slope bound says, then, that ‘the growth is strictly less than the number of critical values’. Let π~\tilde{\pi} be as in Theorem A. The following is proved in [Loe14] (cf. [Wil17, §7.5]).

Proposition 7.6.

If π~\tilde{\pi} is non-QQ-critical slope, then Lpip(π~)L_{p}^{i_{p}}(\tilde{\pi}) is uniquely determined by its growth and interpolation properties.

When FF is totally real, the analogous unicity results are discussed in [BDW, Prop. 6.25]. If FF is not imaginary quadratic or totally real, unicity remains mysterious; see e.g. [BW19, §13].

References

  • [AG94] Avner Ash and David Ginzburg. pp-adic LL-functions for GL(2n){\rm GL}(2n). Invent. Math., 116(1-3):27–73, 1994.
  • [AS06] Mahdi Asgari and Freydoon Shahidi. Generic transfer for general spin groups. Duke Math. J., 132(1):137–190, 2006.
  • [AS14] Mahdi Asgari and Freydoon Shahidi. Image of functoriality for general spin groups. Manuscripta Math., 144(3-4):609–638, 2014.
  • [Ash80] Avner Ash. Non-square-integrable cohomology of arithmetic groups. Duke Math. J., 47(2):435–449, 1980.
  • [Bar18] Daniel Barrera Salazar. Overconvergent cohomology of Hilbert modular varieties and pp-adic LL-functions. Ann. I. Fourier, 68(5):2177–2213, 2018.
  • [BDG+] Daniel Barrera Salazar, Mladen Dimitrov, Andrew Graham, Andrei Jorza, and Chris Williams. On the GL(2n) eigenvariety: branching laws, Shalika families and pp-adic L{L}-functions. Preprint: https://arxiv.org/abs/2211.08126.
  • [BDJ22] Daniel Barrera, Mladen Dimitrov, and Andrei Jorza. pp-adic LL-functions of Hilbert cusp forms and the trivial zero conjecture. J. Eur. Math. Soc. (JEMS), 24(10):3439–3503, 2022.
  • [BDP13] Massimo Bertolini, Henri Darmon, and Kartik Prasanna. Generalized Heegner cycles and pp-adic Rankin LL-series. Duke Math. J., 162(6):1033–1148, 2013. With an appendix by Brian Conrad.
  • [BDW] Daniel Barrera Salazar, Mladen Dimitrov, and Chris Williams. On pp-adic LL-functions for GL(2n)\mathrm{{GL}}(2n) in finite slope Shalika families. Preprint: https://arxiv.org/abs/2103.10907.
  • [Bel12] Joël Bellaïche. Critical pp-adic LL-functions. Invent. Math., 189:1 – 60, 2012.
  • [BGW] Daniel Barrera Salazar, Andrew Graham, and Chris Williams. On the classical symplectic locus in the GL2n\mathrm{GL}_{2n}-eigenvariety. Preprint.
  • [BH] John Bergdall and David Hansen. On pp-adic LL-functions for Hilbert modular forms. Mem. Amer. Math. Soc. To appear. https://arxiv.org/abs/1710.05324.
  • [BW19] Daniel Barrera Salazar and Chris Williams. pp-adic LL-functions for GL2{\rm GL}_{2}. Canad. J. Math., 71(5):1019–1059, 2019.
  • [BW21a] Daniel Barrera Salazar and Chris Williams. Families of Bianchi modular symbols: critical base-change pp-adic LL-functions and pp-adic Artin formalism. Selecta Math. (N.S.), 27(82), 2021. Appendix by Carl Wang-Erickson.
  • [BW21b] Daniel Barrera Salazar and Chris Williams. Overconvergent cohomology, pp-adic LL-functions and families for GL(2)\rm GL(2) over CM fields. J. Théor. Nombres Bordeaux, 33(3, part 1):659–701, 2021.
  • [BW21c] Daniel Barrera Salazar and Chris Williams. Parabolic eigenvarieties via overconvergent cohomology. Math. Z., 299(1-2):961–995, 2021.
  • [Che04] Gaëtan Chenevier. Familles pp-adiques de formes automorphes pour GLn{\rm GL}_{n}. J. Reine Angew. Math., 570:143–217, 2004.
  • [Clo90] Laurent Clozel. Motifs et formes automorphes: applications du principe de fonctorialité. In Automorphic forms, Shimura varieties, and LL-functions, Vol. I (Ann Arbor, MI, 1988), volume 10 of Perspect. Math., pages 77–159. Academic Press, Boston, MA, 1990.
  • [Coa89] John Coates. On pp-adic LL-functions attached to motives over 𝐐{\bf Q}. II. Bol. Soc. Brasil. Mat. (N.S.), 20(1):101–112, 1989.
  • [CPR89] John Coates and Bernadette Perrin-Riou. On pp-adic LL-functions attached to motives over 𝐐{\bf Q}. In Algebraic number theory, volume 17 of Adv. Stud. Pure Math., pages 23–54. Academic Press, Boston, MA, 1989.
  • [Del79] Pierre Deligne. Valeurs de fonctions LL et périodes d’intégrales. Proc. Symp. Pure Math., 33:313–346, 1979.
  • [DJ] Mladen Dimitrov and Andrei Jorza. Parahoric level pp-adic LL-functions for automorphic representations of GL(2n) with Shalika models. In preparation.
  • [DJR20] Mladen Dimitrov, Fabian Januszewski, and A. Raghuram. L{L}-functions of GL(2n)\mathrm{GL}(2n): pp-adic properties and nonvanishing of twists. Compositio Math., 156(12):2437–2468, 2020.
  • [FJ93] Solomon Friedberg and Hervé Jacquet. Linear periods. J. Reine Angew. Math., 443:91–139, 1993.
  • [Geh18] Lennart Gehrmann. On Shalika models and pp-adic LL-functions. Israel J. Math., 226(1):237–294, 2018.
  • [GR14] Harald Grobner and A. Raghuram. On the arithmetic of Shalika models and the critical values of LL-functions for GL2n{\rm GL}_{2n}. Amer. J. Math., 136(3):675–728, 2014. With an appendix by Wee Teck Gan.
  • [Hid98] Haruzo Hida. Automorphic induction and Leopoldt type conjectures for GL(n){\rm GL}(n). volume 2, pages 667–710. 1998. Mikio Sato: a great Japanese mathematician of the twentieth century.
  • [JS76] Hervé Jacquet and Joseph A. Shalika. A non-vanishing theorem for zeta functions of GLn{\rm GL}_{n}. Invent. Math., 38(1):1–16, 1976.
  • [JS90] Hervé Jacquet and Joseph Shalika. Exterior square LL-functions. In Automorphic forms, Shimura varieties, and LL-functions, Vol. II (Ann Arbor, MI, 1988), volume 11 of Perspect. Math., pages 143–226. Academic Press, Boston, MA, 1990.
  • [JST] Dihua Jiang, Binyong Sun, and Fangyang Tian. Period relations for standard LL-functions of symplectic type. Preprint: https://arxiv.org/abs/1909.03476.
  • [Kna01] A. W. Knapp. Branching theorems for compact symmetric spaces. Represent. Theory, 5:404–436, 2001.
  • [LLZ15] Antonio Lei, David Loeffler, and Sarah Livia Zerbes. Euler systems for modular forms over imaginary quadratic fields. Compos. Math., 151(9):1585–1625, 2015.
  • [Loe14] David Loeffler. PP-adic integration on ray class groups and non-ordinary pp-adic LL-functions. In T. Bouganis and O. Venjakob, editors, Iwasawa 2012: State of the art and recent advances, volume 7 of Contributions in Mathematical and Computational Sciences, pages 357 – 378. Springer, 2014.
  • [LT20] Bingchen Lin and Fangyang Tian. Archimedean non-vanishing, cohomological test vectors, and standard LL-functions of GL2n{\rm GL}_{2n}: complex case. Adv. Math., 369:107189, 51, 2020.
  • [Pan94] Alexei A. Panchishkin. Motives over totally real fields and pp-adic LL-functions. Ann. Inst. Fourier (Grenoble), 44(4):989–1023, 1994.
  • [PS11] Robert Pollack and Glenn Stevens. Overconvergent modular symbols and pp-adic LL-functions. Ann. Sci. Éc. Norm. Supér. (4), 44(1):1–42, 2011.
  • [Roh89] David E. Rohrlich. Nonvanishing of LL-functions for GL(2){\rm GL}(2). Invent. Math., 97(2):381–403, 1989.
  • [RY] Maksym Radziwiłł and Liyang Yang. Non-vanishing of twists of GL4(𝐀𝐐)\mathrm{{GL}}_{4}(\mathbf{{A}}_{\mathbf{{Q}}}) L{L}-functions. Preprint: https://arxiv.org/abs/2304.09171.
  • [SU14] Christopher Skinner and Eric Urban. The Iwasawa Main Conjectures for GL(2)(2). Invent. Math., 195 (1):1–277, 2014.
  • [Wei56] André Weil. On a certain type of characters of the idele-class group of an algebraic number-field. In Proceedings of the international symposium on algebraic number theory, 1956.
  • [Wil17] Chris Williams. PP-adic LL-functions of Bianchi modular forms. Proc. Lond. Math. Soc. (3), 114(4):614–656, 2017.