Hostname: page-component-745bb68f8f-kw2vx Total loading time: 0 Render date: 2025-02-06T04:42:23.495Z Has data issue: false hasContentIssue false

Arithmetic diagonal cycles on unitary Shimura varieties

Published online by Cambridge University Press:  27 October 2020

M. Rapoport
Affiliation:
Mathematisches Institut der Universität Bonn, Endenicher Allee 60, 53115Bonn, Germanyrapoport@math.uni-bonn.de Department of Mathematics, University of Maryland, College Park, MD20742, USA
B. Smithling
Affiliation:
Department of Mathematics, University of Maryland, College Park, MD20742, USAcorgcorgcorg@gmail.com
W. Zhang
Affiliation:
Department of Mathematics, Massachusetts Institute of Technology, 77 Massachusetts Avenue, Cambridge, MA02139, USAweizhang@mit.edu
Rights & Permissions [Opens in a new window]

Abstract

We define variants of PEL type of the Shimura varieties that appear in the context of the arithmetic Gan–Gross–Prasad (AGGP) conjecture. We formulate for them a version of the AGGP conjecture. We also construct (global and semi-global) integral models of these Shimura varieties and formulate for them conjectures on arithmetic intersection numbers. We prove some of these conjectures in low dimension.

Type
Research Article
Copyright
Copyright © The Author(s) 2020

1. Introduction

The theorem of Gross and Zagier [Reference Gross and ZagierGZ86] relates the Néron–Tate heights of Heegner points on modular curves to special values of derivatives of certain $L$-functions. Ever since the appearance of [Reference Gross and ZagierGZ86], the problem of generalizing this fundamental result to higher dimension has attracted considerable attention. The generalization that is most relevant to the present paper is the arithmetic Gan–Gross–Prasad conjecture (AGGP conjecture) [Reference Gan, Gross and PrasadGGP12, § 27]. This conjectural generalization concerns Shimura varieties attached to orthogonal groups of signature $(2, n-2)$, and to unitary groups of signature $(1, n-1)$ (note that modular curves are closely related to Shimura varieties associated to orthogonal groups of signature $(2, 1)$ and to unitary groups of signature $(1, 1)$). In [Reference Gan, Gross and PrasadGGP12, § 27], algebraic cycles of codimension one on such Shimura varieties are defined by exploiting embeddings of Shimura varieties attached to orthogonal groups of signature $(2, n-3)$ (respectively, to unitary groups of signature $(1, n-2)$). By taking the graphs of these embeddings, one obtains cycles in codimension just above half the (odd) dimension of the ambient variety.

For any algebraic variety $X$ smooth and proper of odd dimension over a number field, Beilinson and Bloch have defined a height pairing on the rational Chow group ${{\mathrm {Ch}}}(X)_{\mathbb{Q},0}$ of cohomologically trivial cycles of codimension just above half the dimension. Their definition makes use of some wide open unsolved conjectures on algebraic cycles and the existence of regular proper integral models of $X$. By suitably replacing the graph cycle in the case at hand by a cohomologically trivial avatar, one obtains a linear form on ${{\mathrm {Ch}}}(X)_{\mathbb{Q},0}$, where now $X$ is the product of the two Shimura varieties in question. The AGGP conjecture relates a special value of the derivative of an $L$-function to the non-triviality of the restriction of this linear form to a Hecke eigenspace in ${{\mathrm {Ch}}}(X)_{\mathbb{Q}, 0}$. It is stated in a very succinct way in [Reference Gan, Gross and PrasadGGP12], for orthogonal groups and for unitary groups. In the present paper we restrict ourselves to unitary groups, and one of our aims is to give more details on (a variant of) this conjecture in this case. Our version here is also an improvement of the version of the conjecture in [Reference ZhangZha12a, Reference ZhangZha12b]. One new feature of our version is that we use the standard sign conjecture for unitary Shimura varieties (a theorem of Morel and Suh [Reference Morel and SuhMS19, Theorem 1.2], cf.  Section 6.2) to construct ‘Hecke–Künneth’ projectors that project the total cohomology of our Shimura variety to the odd-degree part.

As indicated above, the AGGP conjecture is based on conjectures of Beilinson and Bloch which seem out of reach at present. As a consequence, the conjecture in [Reference Gan, Gross and PrasadGGP12] has not been proved in a single case of higher dimension.Footnote 1 A variant of the AGGP conjecture, inspired by the relative trace formula of Jacquet and Rallis [Reference Jacquet and RallisJR11], has been proposed by the third author [Reference ZhangZha12a]. This variant relates the Beilinson–Bloch height pairing with distributions that appear in the relative trace formula. The variant leads to local conjectures (on intersection numbers on Rapoport–Zink spaces), namely the arithmetic fundamental lemma (AFL) conjecture and the arithmetic transfer (AT) conjecture (cf. [Reference ZhangZha12a, Reference Rapoport, Smithling and ZhangRSZ18]), and these have been proved in various cases [Reference ZhangZha12a, Reference Rapoport, Terstiege and ZhangRTZ13, Reference MihatschMih16, Reference Rapoport, Smithling and ZhangRSZ17, Reference MihatschMih17, Reference Rapoport, Smithling and ZhangRSZ18]. The second aim of the present paper is to formulate a global conjecture whose proof in various cases is a realistic goal. In the present paper, basing ourselves on our local papers [Reference ZhangZha12a, Reference Rapoport, Smithling and ZhangRSZ18], we prove this conjecture for unitary groups of size $n\leq 3$.

To formulate this conjecture, we define variants of the Shimura varieties appearing in [Reference Gan, Gross and PrasadGGP12] and [Reference ZhangZha12a] which are of PEL type, that is, are related to moduli problems of abelian varieties with polarizations, endomorphisms, and level structures. In fact, we even define integral models of these Shimura varieties, in a global version and a semi-global version. The construction of such models is the third aim of the present paper. Once these models are defined, we replace the Beilinson–Bloch pairing on the cohomologically trivial Chow group by the Gillet–Soulé pairing on the arithmetic Chow group of the (global or semi-global) integral model.

Now that we have formulated the three main goals of this paper, let us be more specific.

Let $F$ be a CM number field, with maximal totally real subfield $F_0$. We fix a CM type $\Phi$ of $F$ and a distinguished element $\varphi _0\in \Phi$. Let $n\geq 2$ and let $r \colon \operatorname {Hom}(F,\mathbb{C}) \to \{0,1,n-1,n\}$, $\varphi \mapsto r_\varphi$ be the function defined by

\[ r_\varphi := \begin{cases} 1, & \varphi = \varphi_0,\\ 0, & \varphi \in \Phi \smallsetminus \{\varphi_0\},\\ n-r_{\overline\varphi}, & \varphi \notin \Phi. \end{cases} \]

Associated to these data, there is the field $E\subset \overline {\mathbb{Q}}$ which is the composite of the reflex field of $r$ and the reflex field of $\Phi$. Then $E$ contains $F$ via $\varphi _0$. We denote by $Z^{\mathbb{Q}}$ the torus

\[ Z^{\mathbb{Q}} := \{z\in \operatorname{Res}_{F/\mathbb{Q}}(\mathbb{G}_m) \,| \,\operatorname{Nm}_{F/F_0}(z)\in\mathbb{G}_m\}. \]

We also fix an $F/F_0$-hermitian vector space $W$ of dimension $n$ with signature

\[ {{\mathrm{sig}}}(W_\varphi)=(r_\varphi, r_{\overline\varphi}), \quad \varphi\in\Phi. \]

Let $G$ be the unitary group of $W$, considered as an algebraic group over $\mathbb{Q}$.Footnote 2 Associated to $(G, r)$ is the Shimura variety of [Reference Gan, Gross and PrasadGGP12]. In the present paper we instead consider the Shimura variety associated to $\widetilde G := Z^{\mathbb{Q}}\times G$. We are able to formulate a PEL moduli problem $M_{K_{\widetilde G}}(\widetilde G)$ of abelian varieties with additional structure (endomorphisms and polarization) which defines a model over $E$ of the Shimura variety

(1.1)\begin{equation} \mathrm{Sh}_{K_{\widetilde G}}(\widetilde G)= M_{K_{\widetilde G}}(\widetilde G)\otimes_E\mathbb{C}. \end{equation}

(In fact, we require that $K_{\widetilde G}=K_{Z^{\mathbb{Q}}}\times K_G$, where $K_{Z^{\mathbb{Q}}}$ is the unique maximal compact subgroup of $Z^{\mathbb{Q}}(\mathbb{A}_f)$ and where $K_G$ is an open compact subgroup of $G(\mathbb{A}_f)$.) The group differs from the group of unitary similitudes $\mathrm {GU}(W)$ by a central isogeny. The Shimura variety corresponding to the latter group is considered by Kottwitz [Reference KottwitzKot92], and he formulates a PEL moduli problem over the reflex field of $r$ which almost defines a model for it, but not quite, because of the possible failure of the Hasse principle for $\mathrm {GU}(W)$. This Shimura variety is also considered by Harris and Taylor [Reference Harris and TaylorHT01]. In the setup of [Reference Harris and TaylorHT01], we have $E=F$ and both their Shimura variety and ours are defined over $F$; however, ours offers a number of technical advantages over theirs.Footnote 3 The definition of our moduli problem is based on a sign invariant ${{\mathrm {inv}}}^{r}_v(A_0, A)\in \{\pm 1\}$ for every non-archimedean place $v$ of $F_0$ which is non-split in $F$. Here $(A_0, \iota _0, \lambda _0)$ is a polarized abelian variety of dimension $d=[F_0:\mathbb{Q}]$ with complex multiplication of CM type $\overline \Phi$ of $F$, and $(A, \iota , \lambda )$ is a polarized abelian variety of dimension $nd$ with complex multiplication of generalized CM type $r$ of $F$. This sign invariant is similar to the one in [Reference Kudla and RapoportKR15, Reference Kudla, Rapoport and ZinkKRZ20], but much simpler. This simplicity is another reflection of the advantage of our Shimura varieties over those considered by Kottwitz [Reference KottwitzKot92].

This sign invariant also allows us to define global integral models of $M_{K_{\widetilde G}}(\widetilde G)$ over $\operatorname {Spec} O_E$ (at least when $F/F_0$ is not everywhere unramified) and semi-global integral models over $\operatorname {Spec} O_{E, (\nu )}$, where $\nu$ is a fixed non-archimedean place of $E$, of residue characteristic $p$. These integral models generalize those in [Reference Bruinier, Howard, Kudla, Rapoport and YangBHKR+20] when $F_0=\mathbb{Q}$ and when $K_G$ is the stabilizer of a self-dual lattice in $W$. Here we allow $K_G$ to be the stabilizer of more general vertex lattices, in the sense of [Reference Kudla and RapoportKR11]. To achieve flatness, we sometimes have to impose conditions on the Lie algebras of the abelian varieties in play that are known in a similar context from our earlier local papers [Reference Rapoport, Smithling and ZhangRSZ17, Reference Rapoport, Smithling and ZhangRSZ18] (the Pappas wedge condition, the spin condition and its refinement, the Eisenstein conditions). However, in contrast to Kottwitz, we do not need any unramifiedness conditions.

Once the model $M_{K_{\widetilde G}}(\widetilde G)$ and its global or semi-global model are defined, we can also create a restriction situation in analogy with [Reference Gan, Gross and PrasadGGP12]. Namely, fixing a totally negative vector $u\in W$ (satisfying additional integrality conditions for the global (respectively, semi-global) integral situation), we define $W^{\flat }$ to be the orthogonal complement of $u$. Then $W^{\flat }$ satisfies the same conditions as $W$, with $n$ replaced by $n-1$. We obtain a finite unramified morphism

(1.2)\begin{equation} M_{K_{\widetilde H}}(\widetilde H)\longrightarrow M_{K_{\widetilde G}}(\widetilde G) \end{equation}

(respectively, their global or semi-global integral versions). Here $\widetilde H=Z^{\mathbb{Q}}\times H$, where $H=\mathrm {U}(W^{\flat })$, considered as an algebraic group over $\mathbb{Q}$. Using the graph of the above morphism, we obtain an element in the rational Chow group,

\[ z_{K_{\widetilde{HG}}}\in {{\mathrm{Ch}}}^{n-1}(M_{K_{\widetilde {HG}}}(\widetilde {HG}))_\mathbb{Q}. \]

Here $M_{K_{\widetilde {HG}}}(\widetilde {HG})$ is the model defined as above for the Shimura variety for the group $Z^{\mathbb{Q}}\times H\times G$. Using the Hecke–Künneth projector, we construct a cohomologically trivial variant $z_{ K_{\widetilde {HG}}, 0}\in {{\mathrm {Ch}}}^{n-1}(M_{K_{\widetilde {HG}}}(\widetilde {HG}))_{0,\mathbb{C}}$ of this element, which, via the Beilinson–Bloch pairing, in turn defines a linear form

\[ \ell_{K_{\widetilde {HG}}}\colon {{\mathrm{Ch}}}^{n-1}(M_{K_{\widetilde {HG}}}(\widetilde {HG}))_{\mathbb{C}, 0}\longrightarrow \mathbb{C}. \]

Our variant of the AGGP conjecture is expressed in terms of this linear form.

We similarly define, under certain hypotheses, elements in the rational Chow group (respectively, rational arithmetic Chow group) of the (global or semi-global) integral model $\mathcal {M}_{K_{\widetilde {HG}}}(\widetilde {HG})$. Let us consider the Gillet–Soulé intersection product pairing on the rational arithmetic Chow group of $\mathcal {M}_{K_{\widetilde {HG}}}(\widetilde {HG})$. We have a conjecture on the value of the intersection product of $z_{K_{\widetilde {HG}}}$ and its image under a Hecke correspondence. Let us state the semi-global version, since this is the one for which we can produce concrete evidence. We also have a global version.

Conjecture 1.1 (Semi-global conjecture) Fix a non-split place $v_0$ of $F_0$ over the place $p\leq \infty$ of $\mathbb{Q}$. Let $f=\bigotimes _{\ell } f_\ell \in \mathscr {H}_{{K_{\widetilde {HG}}}}^{p}$ ($\mathscr {H}_{{K_{\widetilde {HG}}}}$ if $p$ is archimedean) be a completely decomposed element of the finite Hecke algebra of $\widetilde {HG}$, and let $f'=\bigotimes _{v}f'_v\in \mathscr {H}(G'(\mathbb{A}_{F_0}))$ be a Gaussian test function in the Hecke algebra of $G'=\operatorname {Res}_{F/F_0}(\mathrm {GL}_{n-1}\times \mathrm {GL}_n)$ such that $\bigotimes _{v < \infty }f'_v$ is a smooth transfer of $f$. Assume that for some place $\lambda$ prime to $p$, the function $f$ has regular support at $\lambda$ in the sense of Definition 8.4 and that $f'$ has regular support at $\lambda$ in the sense of Definition 7.5.

  1. (i) Assume that $v_0$ is non-archimedean of hyperspecial type (cf. § 4.1) and that $f_{v_0}'=\mathbf {1}_{G'(O_{F_0,v_0})}$. Then

    \[ \operatorname{\mathrm{Int}}_{v_0}(f)=-\partial J_{v_0}(f'). \]
  2. (ii) Assume that $v_0$ is archimedean, or non-archimedean of AT type (cf. § 4.4). Then

    \[ \operatorname{\mathrm{Int}}_{v_0}(f)=-\partial J_{v_0}(f')-J(f'_{\mathrm{corr}}[v_0]), \]
    where $f'_{\mathrm {corr}}[v_0]=\bigotimes _{v}f'_{\mathrm {corr}, v}$, with $f'_{\mathrm {corr}, v}=f'_v$ for $v\neq v_0$, is a correction function. Furthermore, $f'$ may be chosen such that $f'_{\mathrm {corr}}[v_0]$ is zero.

We refer to the body of the text for an explanation of the terms used (cf. Conjecture 8.13). Here it suffices to remark that $\operatorname {\mathrm {Int}}_{v_0}(f)$ is the weighted sum of the local contributions of all places $\nu$ of $E$ over $v_0$ to the Gillet–Soulé intersection product of the diagonal subscheme $\mathcal {M}_{K_{\widetilde H}}(\widetilde H)$ of $\mathcal {M}_{K_{\widetilde {HG}}}(\widetilde {HG})$ and its translate by the Hecke correspondence $R(f)$ associated to $f$. We denote by $J$ (respectively, $\partial J_{v_0}$) the distributions (cf. (7.9) and (7.10)) that arise in the twisted trace formula approach to the AGGP conjecture. The regularity assumption on $f$ guarantees that for non-archimedean $v_0$ the intersection of the two cycles has support in characteristic $p$, and the regularity assumption on $f'$ guarantees that the distributions $J$ and $\partial J_{v_0}$ localize.

In the global context, when the Hecke correspondence $R(f)$ satisfies a suitable regularity assumption, the intersection product localizes, that is, is a finite sum of contributions, one from each place $\nu$ of $E$. We group together the local contributions from all places $\nu$ which induce a given place $v_0$ of $F_0$. From this point of view, Conjecture 1.1(i) predicts the contribution of the good places, and Conjecture 1.1(ii) the contributions from the archimedean places and certain bad places.

The conjecture is accessible in certain cases. We prove the following theorem.

Theorem 1.2 Let $v_0$ be a non-archimedean place of $F_0$ that is non-split in $F$. Then Conjecture 1.1 holds true for $n\leq 3$.

The main input is our work in the local case (intersection product on Rapoport–Zink spaces): the proof of the AFL conjecture in the hyperspecial case for $n\leq 3$ by one of us [Reference ZhangZha12a] and the proof of the AT conjecture for $n\leq 3$ in [Reference Rapoport, Smithling and ZhangRSZ17, Reference Rapoport, Smithling and ZhangRSZ18]. The passage from the local statement to a global statement is modeled on the similar passage in [Reference Kudla and RapoportKR14] (which also inspired the similar passage in [Reference ZhangZha12a]). In fact, this similarity is not only formal. Indeed, the definition of Kudla–Rapoport divisors uses in an essential way that the Shimura variety for $\mathrm {GU}(W)$ is replaced by the Shimura variety for $\widetilde G$ (in fact, in [Reference Kudla and RapoportKR14], one uses $Z^{\mathbb{Q}}\times \mathrm {GU}(W)$; as remarked in [Reference Bruinier, Howard, Kudla, Rapoport and YangBHKR+20], the definition can be realized on the Shimura variety for $\widetilde G$). In this way, there is a direct connection between the intersection problem occurring in Conjecture 1.1 and the intersection problem in [Reference Kudla and RapoportKR14].

We also have a theorem for split (hence non-archimedean) places (cf. Proposition 8.12).

Theorem 1.3 Let $v_0$ be a place of $F_0$ that is split in $F$. Let $f$ and $f'$ be as in Conjecture 8.8. Then

\[ \operatorname{\mathrm{Int}}_{v_0}(f)=\partial J_{v_0}(f')=0. \]

The significance of this theorem is that in the global context, again under a regularity assumption on the Hecke correspondence $R(f)$, the contribution of the places $v_0$ which split in $F$ is trivial.

Let us now put the results of this paper in perspective. B. Gross asked for integral models of the Shimura varieties for unitary groups. Here we answer a modified version of his question by constructing integral models of the Shimura varieties for $\widetilde G$. However, we have to pay a price by having to replace the field $F$, over which Gross's Shimura varieties have a model, by the field $E$, over which our Shimura varieties have a model; and $E$ may be strictly larger than $F$. This also causes us to modify our adaptation of the AGGP conjecture. It would be interesting to understand whether the Kisin–Pappas construction of integral models of Shimura varieties of abelian type [Reference Kisin and PappasKP18] yields a solution of Gross's question which can be used to give a variant of the AGGP conjecture which avoids having to replace $F$ by a bigger field.

Our current knowledge of AT conjectures forces on us to be very specific when imposing level structures in our moduli problems. It seems realistic to hope that more cases of AT conjectures than in [Reference Rapoport, Smithling and ZhangRSZ17, Reference Rapoport, Smithling and ZhangRSZ18] can be formulated, and this would allow more flexibility for the level structures. We hope to return to this point.

How realistic is it to hope that the conjectures on the arithmetic intersection pairing can be proved, in cases that go beyond those treated in this paper? The stumbling block seems to be that in higher dimension it is difficult to avoid degenerate intersections—and degenerate intersections seem to be a challenge to currently available techniques, already in the local situationFootnote 4 (intersection on Rapoport–Zink spaces). It might be fruitful to search for intersection problems derived from those considered here which avoid these apparently very difficult problems, in the spirit of B. Howard's papers concerning the Kudla–Rapoport divisor intersection problem (cf. [Reference HowardHow12, Reference HowardHow15]).

As is apparent, automorphic $L$-functions do not appear explicitly in the statements above, contrary to what happens in the AGGP conjecture. $L$-functions are involved implicitly because the distributions $J$ and $\partial J_{v_0}$ are related to them (cf. § 7). However, more analytic work is involved to make this relation more explicit. One of us (W.Z.) hopes to return to this point and explain this issue in more detail.

We also mention the article [Reference ZhangZha18] of the third author, which discusses some of the broader context on the relation of special values of $L$-functions and their derivatives to period integrals and intersections of cycles, into which this paper falls.

The organization of the rest of this paper is as follows. In § 2 we introduce the groups in play and define the concept of matching in this context. In § 3 we introduce the various Shimura varieties mentioned above, the corresponding moduli interpretation over $E$, and the relation between the moduli variety for $W^{\flat }$ and for $W$. In § 4 we define the semi-global integral models of these moduli schemes and the morphisms between them. We do this in several contexts: for the hyperspecial level, split level, Drinfeld level, and AT parahoric level. These levels reflect the possibility of applying the AFL conjecture (respectively, the AT conjecture, or the vanishing theorem (Theorem 1.3) in the split level). The AT parahoric level is complicated by the fact that the morphisms between the semi-global integral models are in some cases not defined in the naive way. In § 5 we give the global integral models, first without level structure and then with Drinfeld level structure. Section 6 is devoted to giving our version of the AGGP conjecture. Section 7 is preparatory for the final section. Here we explain the distributions arising in the context of the relative trace formula and their relation to $L$-functions. In § 8 everything finally comes together. First, we formulate our conjecture on the arithmetic intersection numbers. We do this in the global case without level structure, in the global case with Drinfeld level structure, and in the semi-global case. Second, we give the proofs of the semi-global versions in cases of small dimension (cf. Theorem 1.2). There are two appendices. In Appendix A we define the sign invariant that is used in the formulation of the moduli schemes. In Appendix B we check that in the case of banal signature the relevant local models are trivial in a precise sense.

Notation

Except in § 2, $F$ denotes a CM number field and $F_0$ denotes its (maximal) totally real subfield of index $2$ (in § 2, $F/F_0$ can be any quadratic extension of number fields). We denote by $a\mapsto \overline a$ the non-trivial automorphism of $F/F_0$. We fix a presentation $F = F_0(\sqrt \Delta )$ for some totally negative element $\Delta \in F_0$, and we let $\Phi$ be the CM type for $F$ determined by $\sqrt \Delta$,

(1.3)\begin{equation} \Phi := \{\varphi\colon F \rightarrow \mathbb{C} \,|\, \varphi(\sqrt\Delta) \in \mathbb{R}_{>0} \cdot \sqrt{-1}\}. \end{equation}

Note that by weak approximation, every CM type for $F$ arises in this way for some $F/F_0$-traceless element $\sqrt \Delta \in F^{\times }$.

We use the symbols $v$ and $v_0$ to denote places of $F_0$, and $w$ and $w_0$ to denote places of $F$. We write $F_{0,v}$ for the $v$-adic completion of $F_0$, and we set $F_v := F \otimes _{F_0} F_{0,v}$; thus $F_v$ is isomorphic to $F_{0,v} \times F_{0,v}$ or to a quadratic field extension of $F_{0,v}$ according as $v$ is split or non-split in $F$. We often identify the CM type $\Phi$ (or more precisely, the restrictions of its elements to $F_0$) with the archimedean places of $F_0$. When $v$ is a finite place, we write $\mathfrak {p}_v$ for the maximal ideal in $O_{F_0}$ at $v$, $\varpi _v$ for a uniformizer in $F_{0,v}$, and $\pi _v$ for a uniformizer in $F_v$ (when $v$ splits in $F$ this means an ordered pair of uniformizers on the right-hand side of the isomorphism $F_v \cong F_{0,v} \times F_{0,v}$). We write $O_{F_0,(v)}$ for the localization of $O_{F_0}$ at the maximal ideal $\mathfrak {p}_v$, and $O_{F_0,v} \subset F_{0,v}$ for its $\mathfrak {p}_v$-adic completion. We use analogous notation for other fields in place of $F_0$ and other finite places in place of $v$. In particular, we will often consider the $v$-adic completion $O_{F,v} = O_F \otimes _{O_{F_0}} O_{F_0,v}$ of $O_F$.

We write $\mathbb{A}$, $\mathbb{A}_{F_0}$, and $\mathbb{A}_F$ for the adele rings of $\mathbb{Q}$, $F_0$, and $F$, respectively. We systematically use a subscript $f$ for the ring of finite adeles, and a superscript $p$ for the adeles away from the prime number $p$.

We take all hermitian forms to be linear in the first variable and conjugate-linear in the second, and we assume that they are non-degenerate unless we say otherwise. For $k$ any field, $A$ an étale $k$-algebra of degree $2$, and $W$ a finite free $A$-module equipped with an $A/k$-hermitian form, we write $\det W \in k^{\times }/ \operatorname {Nm}_{A/k}A^{\times }$ for the class of $\det J$, where $J$ is any hermitian matrix (relative to the choice of an $A$-basis for $W$) representing the form. Note that this value group is trivial when $A \simeq k \times k$. We also write $-W$ for the same $A$-module as $W$, but with the hermitian form multiplied by $-1$. Of course $W$ and $-W$ have the same unitary groups. When $W$ is an $F/F_0$-hermitian space of dimension $n$ and $v$ is a place of $F_0$, we write $W_v$ for the induced $F_v/F_{0,v}$-hermitian space $W \otimes _{F_0} F_{0,v}$, and we define

(1.4)\begin{equation} {{\mathrm{inv}}}_v(W_v) := (-1)^{n(n-1)/2} \det W_v \in F_{0,v}^{\times} / \operatorname{Nm} F_v^{\times}. \end{equation}

We say that $W_v$ is split at a finite place $v$ if ${{\mathrm {inv}}}_v(W_v) = 1$; under our normalization, the antidiagonal unit matrix always defines a split hermitian form. When $v$ is an archimedean place, the form on $W_v$ is isometric to $\operatorname {diag}(1^{(r)},(-1)^{(s)})$ for some $r+s=n$, and we write ${{\mathrm {sig}}}_v(W_v) := (r,s)$ (the signature). In the local setting, isometry classes of $n$-dimensional $F_v/F_{0,v}$-hermitian spaces are classified by ${{\mathrm {inv}}}_v$ when $v$ is a finite place, and by ${{\mathrm {sig}}}_v$ when $v$ is an archimedean place. By the Hasse principle, two global hermitian spaces are isometric if and only if they are isometric at every place $v$, that is, they have the same invariants at each finite place and the same signatures at each archimedean place. Given a global space $W$ as above, the product formula for the norm residue symbol for the extension $F/F_0$ gives

(1.5)\begin{equation} \prod_v {{\mathrm{inv}}}_v(W_v) = 1, \end{equation}

where $v$ ranges through the places of $F_0$, and where we identify $F_{0,v}^{\times } / \operatorname {Nm} F_v^{\times } \subset \{\pm 1\}$. Conversely, Landherr's theorem asserts that a collection $(W_v)_v$ of $F_v/F_{0,v}$-hermitian spaces arises as the set of local completions of a (unique, by the Hasse principle) global $F/F_0$-hermitian space exactly when ${{\mathrm {inv}}}_v(W_v) = 1$ for all but finitely many $v$ and the product formula (1.5) holds. Given an embedding $\varphi \colon F \to \mathbb{C}$, we write $W_\varphi := W \otimes _{F,\varphi }\mathbb{C}$ for the induced hermitian space over $\mathbb{C}$.

For any abelian scheme $A$ over a base scheme $S$, we denote by $A^{\vee }$ the dual abelian scheme. When $S$ is locally noetherian and the prime number $p$ is invertible on it, we write $T_p(A)$ for the $p$-adic Tate module of $A$ (regarded as a smooth $\mathbb{Z}_p$-sheaf on $S$) and $V_p(A) := T_p(A) \otimes \mathbb{Q}$ for the rational $p$-adic Tate module (regarded as a smooth $\mathbb{Q}_p$-sheaf on $S$). When $S$ is a $\mathbb{Z}_{(p)}$-scheme, we similarly write $\widehat V^{p}(A)$ for the rational prime-to-$p$ Tate module of $A$. When $S$ is a scheme in characteristic zero, we write $\widehat T(A)$ and $\widehat V(A)$ for the respective full Tate and full rational Tate modules of $A$.

We use a superscript $\circ$ to denote the operation $-\otimes _\mathbb{Z} \mathbb{Q}$ on groups of homomorphisms of abelian schemes, so that, for example, $\operatorname {Hom}^{\circ }(A,A') := \operatorname {Hom}(A,A')\otimes _\mathbb{Z} \mathbb{Q}$.

Given modules $M$ and $N$ over a ring $R$, we write $M \subset ^{r} N$ to indicate that $M$ is an $R$-submodule of $N$ of finite colength $r$. Typically $R$ will be $O_{F,v}$ for $v$ a finite place of $F_0$. When $\Lambda$ is an $O_F$-lattice in an $F/F_0$-hermitian space, we denote the dual lattice with respect to the hermitian form by $\Lambda ^{*}$. We use the same notation when $\Lambda$ is an $O_{F,v}$-lattice in an $F_v/F_{0,v}$-hermitian space, and we call $\Lambda$ a vertex lattice of type $r$ if $\Lambda \subset ^{r} \Lambda ^{*} \subset \pi _v^{-1} \Lambda$. Note that this terminology differs slightly from [Reference Kudla and RapoportKR11, Reference Rapoport, Terstiege and WilsonRTW14], for example.Footnote 5 A vertex lattice is a vertex lattice of type $r$ for some $r$. Let us single out the following special cases. A self-dual lattice is, of course, a vertex lattice of type $0$. An almost self-dual lattice is a vertex lattice of type 1. At the other extreme, a vertex lattice $\Lambda$ is $\pi _v$-modular if $\Lambda ^{*} = \pi _v^{-1} \Lambda$, and almost $\pi _v$-modular if $\Lambda \subset \Lambda ^{*} \subset ^{1} \pi _v^{-1}\Lambda$.

Given a discretely valued field $L$, we denote by $\breve L$ the completion of a maximal unramified extension of $L$.

We write $1_n$ for the $n \times n$ identity matrix. We use a subscript $S$ to denote base change to a scheme (or other object) $S$, and when $S = \operatorname {Spec} A$, we often use a subscript $A$ instead.

2. Group-theoretic setup

In this section we introduce the groups and linear-algebraic objects which will be in play throughout the paper. Let $F/F_0$ be a quadratic extension of number fields.

2.1 Similitude groups and variants

We begin by introducing algebraic groups over $F_0$:

\begin{gather*} G' :=\operatorname{Res}_{F/F_0}(\mathrm{GL}_{n-1} \times \mathrm{GL}_n), \\ H_1' := \operatorname{Res}_{F/F_0} \mathrm{GL}_{n-1}, \quad H_2' := \mathrm{GL}_{n-1}\times\mathrm{GL}_n, \quad H_{1,2}' := H_1' \times H_2'. \end{gather*}

Next let $W$ be a non-degenerate $F/F_0$-hermitian space of dimension $n \geq 2$. We fix a non-isotropic vector $u \in W$, which we call the special vector. We denote by $W^{\flat }$ the orthogonal complement of $u$ in $W$. We define another four algebraic groups over $F_0$:

\begin{align*} G &:= \mathrm{U}(W), \quad H := \mathrm{U}(W^{\flat}),\\ G_W&:= H \times G, \quad H_W := H \times H. \end{align*}

We systematically use the symbol $c$ to denote the similitude factor of a point on a unitary similitude group. We further define the following algebraic groups over $\mathbb{Q}$:

\begin{align*} Z^{\mathbb{Q}} &:= \{ z \in \operatorname{Res}_{F/\mathbb{Q}} \mathbb{G}_m \,|\, \operatorname{Nm}_{F/F_0}(z) \in \mathbb{G}_m\},\\ H^{\mathbb{Q}} &:= \{ h \in \operatorname{Res}_{F_0/\mathbb{Q}} \mathrm{GU}(W^{\flat}) \,|\, c(h)\in \mathbb{G}_m \},\\ G^{\mathbb{Q}} &:= \{ g \in \operatorname{Res}_{F_0/\mathbb{Q}} \mathrm{GU}(W) \,|\, c(g)\in \mathbb{G}_m \},\\ \widetilde H &:= Z^{\mathbb{Q}} \times_{\mathbb{G}_m} H^{\mathbb{Q}} = \{(z,h) \in Z^{\mathbb{Q}} \times H^{\mathbb{Q}} \,|\, \operatorname{Nm}_{F/F_0}(z) = c(h)\},\\ \widetilde G &:= Z^{\mathbb{Q}} \times_{\mathbb{G}_m} G^{\mathbb{Q}} = \{(z,g) \in Z^{\mathbb{Q}} \times G^{\mathbb{Q}} \,|\, \operatorname{Nm}_{F/F_0}(z) = c(g)\},\\ \widetilde{HG} &:= \widetilde H \times_{Z^{\mathbb{Q}}} \widetilde G = Z^{\mathbb{Q}} \times_{\mathbb{G}_m} H^{\mathbb{Q}} \times_{\mathbb{G}_m} G^{\mathbb{Q}} \\ &\phantom{:}= \{(z,h,g) \in Z^{\mathbb{Q}} \times H^{\mathbb{Q}} \times G^{\mathbb{Q}} \,|\, \operatorname{Nm}_{F/F_0}(z) = c(h) = c(g)\}. \end{align*}

Note that $Z^{\mathbb{Q}}$ is naturally a central subgroup of $H^{\mathbb{Q}}$ and $G^{\mathbb{Q}}$, and these inclusions give rise to product decompositions

(2.1)

We also record that the decomposition $W = W^{\flat } \oplus Fu$ gives rise to natural closed embeddings of algebraic groups:

(2.2)

Thus, in terms of the product decompositions in (2.1), the embeddings $\widetilde H {\hookrightarrow} \widetilde G$ and $\widetilde H {\hookrightarrow} \widetilde {HG}$ in (

2.2

) are obtained by applying the functor $Z^{\mathbb{Q}} \times -$ to the embeddings $\operatorname {Res}_{F_0/\mathbb{Q}} H {\hookrightarrow} \operatorname {Res}_{F_0/\mathbb{Q}} G$ and $\operatorname {Res}_{F_0/\mathbb{Q}} H {\hookrightarrow} \operatorname {Res}_{F_0/\mathbb{Q}} (H \times G)$, respectively.

2.2 Orbit matching

The following lemma is obvious.

Lemma 2.1 The natural projections in (2.1) induce isomorphisms

\[ \widetilde G / \widetilde H {\tilde{\to}} \operatorname{Res}_{F_0/\mathbb{Q}} G/ \operatorname{Res}_{F_0/\mathbb{Q}} H \]

and

\[ \widetilde H \backslash \widetilde{HG} / \widetilde H {\tilde{\to}} \operatorname{Res}_{F_0/\mathbb{Q}} H \backslash \operatorname{Res}_{F_0/\mathbb{Q}} G_W / \operatorname{Res}_{F_0/\mathbb{Q}} H. \]

There is a natural injection of orbit spaces of regular semisimple elements,

\[ H(F_{0, v})\backslash G_W(F_{0, v})_\mathrm{rs}/H(F_{0, v}) {\hookrightarrow} H'_1(F_{0,v})\backslash G'(F_{0, v})_\mathrm{rs}/H'_2(F_{0, v}), \]

for any place $v$ of $F_0$ (see [Reference Rapoport, Smithling and ZhangRSZ17, § 2] for the case where $v$ is a non-archimedean place not split in $F$; the case of archimedean places is completely analogous). If $v$ is split in $F$, we define matching as in [Reference ZhangZha14a, § 2]. In brief, we identify $H(F_{0,v})$ with $\mathrm {GL}_{n-1}(F_{0,v})$ and $G(F_{0,v})$ with $\mathrm {GL}_{n}(F_{0,v})$. This gives a natural way of matching regular semisimple elements, a homogeneous version of the matching relation of [Reference ZhangZha12a].

Using Lemma 2.1, we obtain an injection for every prime number $p$,

(2.3)\begin{equation} \widetilde H(\mathbb{Q}_p) \backslash \widetilde{HG}(\mathbb{Q}_p)_\mathrm{rs} / \widetilde H(\mathbb{Q}_p) {\hookrightarrow} \prod_{v\mid p} H'_1(F_{0, v})\backslash G'(F_{0, v})_\mathrm{rs}/ H'_2(F_{0, v}). \end{equation}

3. The Shimura varieties

For the rest of the paper we take $F$ to be a CM field over $\mathbb{Q}$ and $F_0$ to be its totally real subfield of index $2$. We recall from the Introduction that we fix a totally imaginary element $\sqrt \Delta \in F^{\times }$, and we denote by $\Phi$ the induced CM type for $F$ given in (1.3).

3.1 The Shimura data

In this subsection we define Shimura data for some of the groups introduced in § 2. We assume that the hermitian space $W$ has the following signatures at the archimedean places of $F_0$: for a distinguished element $\varphi _0 \in \Phi$ the signature of $W_{\varphi _0}$ is $(1,n-1)$, and for all other $\varphi \in \Phi$ the signature of $W_\varphi$ is $(0,n)$. We also assume that the special vector $u$ is totally negative, that is, that $(u, u)_\varphi <0$ for all $\varphi$.

We first define Shimura data $(G^{\mathbb{Q}}, \{h_{G^{\mathbb{Q}}}\})$ and $(H^{\mathbb{Q}}, \{h_{ H^{\mathbb{Q}}}\})$ (cf. [Reference Pappas and RapoportPR09, § 1.1]). Using the canonical inclusions $G^{\mathbb{Q}}_\mathbb{R}\subset \prod _{\varphi \in \Phi } \mathrm {GU}(W_\varphi )$ and $H^{\mathbb{Q}}_\mathbb{R}\subset \prod _{\varphi \in \Phi } \mathrm {GU}(W^{\flat }_\varphi )$, it suffices to define the components $h_{G^{\mathbb{Q}}, \varphi }$ of $h_{G^{\mathbb{Q}}}$ and $h_{H^{\mathbb{Q}}, \varphi }$ of $h_{H^{\mathbb{Q}}}$. We define matrices

\[ J_\varphi := \begin{cases} \operatorname{diag}(1, (-1)^{(n-1)}), & \varphi=\varphi_0,\\ \operatorname{diag}( -1, -1, \dotsc, -1), & \varphi\in\Phi\smallsetminus\{\varphi_0\}, \end{cases} \]

and we define the matrix $J_\varphi ^{\flat }$ by removing a $-1$ from $J_\varphi$. We may then choose bases $W_\varphi \simeq \mathbb{C}^{n}$ and $W^{\flat }_\varphi \simeq \mathbb{C}^{n-1}$ such that $u \otimes 1 \in W_\varphi$ identifies with a multiple of $(0^{(n-1)},1)$, such that the inclusion $W_\varphi ^{\flat } \subset W_\varphi$ is compatible with the inclusion $\mathbb{C}^{n-1} \subset \mathbb{C}^{n-1} \oplus \mathbb{C} = \mathbb{C}^{n}$, and such that the hermitian forms on $W_\varphi$ and $W_\varphi ^{\flat }$ have respective normal forms

\[ (x, y)_\varphi = {^t}{x} J_\varphi \overline y \quad\text{and}\quad (x^\flat, y^\flat)_\varphi = {^t}{x}{^\flat} J^\flat_\varphi \overline y^\flat. \]

We then define the component maps

\[ h_{G^{\mathbb{Q}},\varphi}\colon \mathbb{C}^{\times}\longrightarrow \mathrm{GU}(W_\varphi)(\mathbb{R}) \quad\text{and}\quad h_{H^{\mathbb{Q}},\varphi}\colon \mathbb{C}^{\times}\longrightarrow \mathrm{GU}(W_\varphi^{\flat})(\mathbb{R}) \]

to be induced by the respective $\mathbb{R}$-algebra homomorphisms

By definition of $\Phi$, the form $x,y \mapsto \operatorname {tr}_{\mathbb{C}/\mathbb{R}} \varphi (\sqrt \Delta ){{}^{-1}}(h_{G^{\mathbb{Q}},\varphi }(\sqrt {-1}) x, y)_\varphi$ is symmetric and positive definite on $W_\varphi$ for each $\varphi \in \Phi$, and similarly for $W^{\flat }_\varphi$.

We next define Shimura data $(\widetilde H, \{h_{\widetilde H}\})$, $(\widetilde G, \{h_{\widetilde G}\})$, and $(\widetilde {HG}, \{h_{\widetilde {HG}}\})$. For this, note that $\Phi$ induces an identification

\[ Z^{\mathbb{Q}}(\mathbb{R}) \cong \{(z_\varphi)\in (\mathbb{C}^{\times})^{\Phi} \,|\, |z_\varphi| = |z_{\varphi'}| \ \text{for all}\ \varphi,\varphi' \in \Phi \}. \]

In terms of this identification, we define $h_{Z^{\mathbb{Q}}}\colon \mathbb{C}^{\times } \to Z^{\mathbb{Q}}(\mathbb{R})$ to be the diagonal embedding, pre-composed with complex conjugation. We then obtain the desired Shimura data by defining the Shimura homomorphisms

\[ h_{\widetilde H}\colon \mathbb{C}^{\times} {\buildrel {(h_{Z^{\mathbb Q}},h_{H^{\mathbb Q}})} \over \longrightarrow} \widetilde H(\mathbb{R}), \quad h_{\widetilde G}\colon \mathbb{C}^{\times} {\buildrel {(h_{Z^{\mathbb Q}},h_{G^{\mathbb Q}})} \over \longrightarrow} \widetilde G(\mathbb{R}), \quad h_{\widetilde{HG}}\colon \mathbb{C}^{\times} {\buildrel {(h_{Z^{\mathbb Q}},h_{H^{\mathbb Q}},h_{G^{\mathbb Q}})} \over \longrightarrow} \widetilde{HG}(\mathbb{R}). \]

It is easy to see that $(\widetilde H, \{h_{\widetilde H}\})$, $(\widetilde G, \{h_{\widetilde G}\})$, and $(\widetilde {HG}, \{h_{\widetilde {HG}}\})$ have common reflex field $E \subset \mathbb{C}$ characterized by

(3.1)\begin{equation} \operatorname{Aut}(\mathbb{C}/E) = \{\sigma\in\operatorname{Aut}(\mathbb{C})\mid \sigma \circ\Phi=\Phi \text{ and } \sigma\circ\varphi_0=\varphi_0 \}. \end{equation}

We therefore obtain canonical models over $E$ of the Shimura varieties

\[ \mathrm{Sh}_{K_{\widetilde H}}(\widetilde H, \{h_{\widetilde H}\}), \quad \mathrm{Sh}_{K_{\widetilde G}}(\widetilde G, \{h_{\widetilde G}\}), \quad \mathrm{Sh}_{K_{\widetilde{HG}}}(\widetilde{HG}, \{h_{\widetilde{HG}}\}) , \]

where ${K_{\widetilde H}}$ (respectively ${K_{\widetilde G}}$, ${K_{\widetilde {HG}}}$) varies through the open compact subgroups of $\widetilde H(\mathbb{A}_f)$ (respectively $\widetilde G(\mathbb{A}_f)$, $\widetilde {HG}(\mathbb{A}_f)$).

Remark 3.1 When $F$ is of the form $K F_0$ for an imaginary quadratic field $K/\mathbb{Q}$ and $\Phi$ is the unique CM type induced from $K$ containing $\varphi _0$ (the case taken by Harris and Taylor in [Reference Harris and TaylorHT01]), $\varphi _0$ identifies $F {\tilde{\to}} E$. In general, $E$ is the join of $F$ (embedded in $\overline{\mathbb{Q}}$ via $\varphi _0$) and the reflex field $E_\Phi$ of $\Phi$ (the latter is the fixed field in $\mathbb{C}$ of the group $\{\sigma \in \operatorname {Aut}(\mathbb{C}) \mid \sigma \circ \Phi = \Phi \}$). In particular, $\varphi _0$ always embeds $F$ into $E$, but $E$ may be larger.

The morphisms (2.2) are obviously compatible with the Shimura data $\{h_{\widetilde H}\}$ and $\{h_{\widetilde G}\}$, (respectively, $\{h_{\widetilde H}\}$ and $\{h_{\widetilde {HG}}\}$). We therefore obtain injective morphisms of Shimura varieties, that is, injective morphisms of pro-varieties, in the sense of [Reference DeligneDel71, Proposition 1.15]:

(3.2)\begin{equation} \mathrm{Sh}(\widetilde H, \{h_{\widetilde H}\}) \hookrightarrow \mathrm{Sh}(\widetilde G, \{h_{\widetilde G}\}) \quad\text{and}\quad \mathrm{Sh}(\widetilde H, \{h_{\widetilde H}\}) \hookrightarrow \mathrm{Sh}(\widetilde{HG}, \{h_{\widetilde{HG}}\}). \end{equation}

Remark 3.2 The above Shimura varieties are related to other Shimura varieties as follows.

(i) The pair $(Z^{\mathbb{Q}}, \{h_{Z^{\mathbb{Q}}}\})$ is a Shimura datum, and there are morphisms of Shimura data

\[ (\widetilde H, \{h_{\widetilde H}\})\longrightarrow (Z^{\mathbb{Q}}, \{h_{Z^{\mathbb{Q}}}\}), \quad (\widetilde G, \{h_{\widetilde G}\})\longrightarrow (Z^{\mathbb{Q}}, \{h_{Z^{\mathbb{Q}}}\}), \quad (\widetilde{HG}, \{h_{\widetilde{HG}}\})\longrightarrow (Z^{\mathbb{Q}}, \{h_{Z^{\mathbb{Q}}}\}) \]

induced by the natural projections to $Z^{\mathbb{Q}}$. These induce morphisms of Shimura varieties

(3.3)\begin{equation} \begin{gathered} \mathrm{Sh}(\widetilde H, \{h_{\widetilde H}\})\longrightarrow \mathrm{Sh}(Z^{\mathbb{Q}}, \{h_{Z^{\mathbb{Q}}}\}),\\ \mathrm{Sh}(\widetilde G, \{h_{\widetilde G}\})\longrightarrow \mathrm{Sh}(Z^{\mathbb{Q}}, \{h_{Z^{\mathbb{Q}}}\}),\\ \mathrm{Sh}(\widetilde{HG}, \{h_{\widetilde{HG}}\})\longrightarrow \mathrm{Sh}(Z^{\mathbb{Q}}, \{h_{Z^{\mathbb{Q}}}\}), \end{gathered} \end{equation}

which identify

\[ \mathrm{Sh}(\widetilde{HG}, \{h_{\widetilde{HG}}\}) \cong \mathrm{Sh}(\widetilde H, \{h_{\widetilde H}\}) \times_{\mathrm{Sh}(Z^{\mathbb{Q}}, \{h_{Z^{\mathbb{Q}}}\})} \mathrm{Sh}(\widetilde G, \{h_{\widetilde G}\}). \]

The reflex field of $(Z^{\mathbb{Q}}, \{h_{Z^{\mathbb{Q}}}\})$ is the subfield $E_\Phi$ of $E$.

(ii) There are morphisms of Shimura data

\[ (\widetilde H, \{h_{\widetilde H}\})\longrightarrow (H^{\mathbb{Q}}, \{h_{H^{\mathbb{Q}}}\}) \quad\text{and}\quad (\widetilde G, \{h_{\widetilde G}\})\longrightarrow (G^{\mathbb{Q}}, \{h_{G^{\mathbb{Q}}}\}), \]

both induced by the natural projections, which induce morphisms of Shimura varieties

\[ \mathrm{Sh}(\widetilde H, \{h_{\widetilde H}\})\longrightarrow \mathrm{Sh}(H^{\mathbb{Q}}, \{h_{H^{\mathbb{Q}}}\}) \quad\text{and}\quad \mathrm{Sh}(\widetilde G, \{h_{\widetilde G}\})\longrightarrow \mathrm{Sh}(G^{\mathbb{Q}}, \{h_{G^{\mathbb{Q}}}\}). \]

(iii) One may also introduce Shimura data

\[ (\operatorname{Res}_{F_0/\mathbb{Q}} H, \{h_{H}\}), \quad (\operatorname{Res}_{F_0/\mathbb{Q}} G, \{h_{ G}\}), \quad (\operatorname{Res}_{F_0/\mathbb{Q}}(H\times G), \{h_{H\times G}\}), \]

where $h_H$ (respectively $h_G$, $h_{H \times G}$) is defined by composing $h_{\widetilde H}$ (respectively $h_{\widetilde G}$, $h_{\widetilde {HG}}$) with the projection to the second factor in (2.1). Note that the restrictions of $h_H$, $h_G$, and $h_{H\times G}$ to the subgroup $\mathbb{R}^{\times }$ of $\mathbb{C}^{\times }$ are trivial. In particular, the corresponding Shimura varieties are not of PEL type. These are the Shimura varieties that appear in Gan, Gross, and Prasad [Reference Gan, Gross and PrasadGGP12, § 27]. The product decompositions in (2.1) induce product decompositions of Shimura data,

(3.4)\begin{align} (\widetilde H, \{h_{\widetilde H}\}) &= (Z^{\mathbb{Q}},\{h_{Z^{\mathbb{Q}}}\}) \times (\operatorname{Res}_{F/F_0} H,\{h_H\}),\notag\\ (\widetilde G, \{h_{\widetilde G}\}) &= (Z^{\mathbb{Q}},\{h_{Z^{\mathbb{Q}}}\}) \times (\operatorname{Res}_{F/F_0} G,\{h_G\}),\\ (\widetilde{HG}, \{h_{\widetilde{HG}}\}) &= (Z^{\mathbb{Q}},\{h_{Z^{\mathbb{Q}}}\}) \times (\operatorname{Res}_{F_0/\mathbb{Q}}(H\times G), \{h_{H\times G}\}).\notag \end{align}

In [Reference Gan, Gross and PrasadGGP12, § 27], the Shimura variety $\mathrm {Sh}(\operatorname {Res}_{F_0/\mathbb{Q}}G, \{h_{G}\})$ is considered over its reflex field, which is $F$, embedded into $\mathbb{C}$ via $\varphi _0$. By contrast, in the present paper, we consider the Shimura variety $\mathrm {Sh}(\widetilde G, \{h_{\widetilde G}\})$ over $E$, which is the join of the reflex fields of the two factors in the product decomposition (3.4). The natural projections then induce morphisms of Shimura varieties,

(3.5)\begin{equation} \begin{gathered} \mathrm{Sh}(\widetilde H, \{h_{\widetilde H}\})\longrightarrow \mathrm{Sh}(\operatorname{Res}_{F_0/\mathbb{Q}}H, \{h_H\}),\\ \mathrm{Sh}(\widetilde G, \{h_{\widetilde G}\})\longrightarrow \mathrm{Sh}(\operatorname{Res}_{F_0/\mathbb{Q}}G, \{h_{G}\}),\\ \mathrm{Sh}(\widetilde{HG}, \{h_{\widetilde{HG}}\})\longrightarrow \mathrm{Sh}(\operatorname{Res}_{F_0/\mathbb{Q}}(H\times G), \{h_{H\times G}\}). \end{gathered} \end{equation}

Remark 3.3 Let us finally make our Shimura varieties more concrete. We consider the case of $\mathrm {Sh}(\widetilde G, \{h_{\widetilde G}\})$; the other Shimura varieties are analogous. In terms of the product decomposition $\widetilde G_\mathbb{R} \cong Z^{\mathbb{Q}}_\mathbb{R} \times \prod _{\varphi \in \Phi } \mathrm {U}(W_\varphi )$ induced by (2.1), the conjugacy class $\{h_{\widetilde G}\}$ is the product of $\{h_{Z^{\mathbb{Q}}}\}$ with the $\mathrm {U}(W_\varphi )(\mathbb{R})$-conjugacy class $\{h_{G,\varphi }\}$ for each $\varphi \in \Phi$, where $h_{G,\varphi }$ denotes the $\varphi$-component of the cocharacter $h_G$ defined in Remark 3.2(iii). The conjugacy class $\{h_{Z^{\mathbb{Q}}}\}$ consists of a single element; so does $\{h_{G,\varphi }\}$ for $\varphi \neq \varphi _0$, since in this case $h_{G,\varphi }$ is the trivial cocharacter. For $\varphi = \varphi _0$, in terms of the basis for $W_{\varphi _0}$ chosen above, $h_{G,\varphi _0}$ is the cocharacter

\[ h_{G,\varphi_0} \colon z \mapsto \operatorname{diag}(z/\overline z, 1,\ldots, 1). \]

The conjugacy class $\{h_{G,\varphi _0}\}$ then identifies with the open subset ${{\mathcal {D}}}_{\varphi _0} \subset \mathbb{P}(W_{\varphi _0})(\mathbb{C})$ of positive-definite lines for the hermitian form (send $h \in \{h_{G,\varphi _0}\}$ to the $-1$-eigenspace of $h({\sqrt {-1}})$; we remark that ${{\mathcal {D}}}_{\varphi _0}$ is also isomorphic to the open unit ball in $\mathbb{C}^{n-1}$). Thus, for $K_{\widetilde G} \subset \widetilde G(\mathbb{A}_f)$ an open compact subgroup, we obtain the presentation

\[ \mathrm{Sh}_{K_{\widetilde G}}(\widetilde G, \{h_{\widetilde G}\}) (\mathbb{C}) = \widetilde G(\mathbb{Q}) \backslash [{{\mathcal{D}}}_{\varphi_0}\times \widetilde G(\mathbb{A}_f)/K_{\widetilde G}] , \]

where the action of $\widetilde G(\mathbb{Q})$ is diagonal by the translation action on $\widetilde G(\mathbb{A}_f)$ and by the action on ${{\mathcal {D}}}_{\varphi _0}$ given via

\[ \widetilde G(\mathbb{Q})\longrightarrow \widetilde G_{\mathrm{ad}}(\mathbb{R})\longrightarrow \textrm{PU}(W_{\varphi_0})(\mathbb{R}). \]

3.2 The moduli problem over $E$

In this subsection we define moduli problems on the category of schemes over $\operatorname {Spec} E$ for the three Shimura varieties above. Since these are almost identical in each case, let us do this for $\mathrm {Sh}(\widetilde G, \{h_{\widetilde G}\})$, and only indicate briefly the modifications needed for the other two (mostly for $\mathrm {Sh}(\widetilde {HG}, \{h_{\widetilde {HG}}\})$). We will only consider open compact subgroups ${K_{\widetilde G}}\subset \widetilde G(\mathbb{A}_f)$ which, with respect to the product decomposition (2.1), are of the form

(3.6)\begin{equation} K_{\widetilde G}= K_{Z^{\mathbb{Q}}} \times K_G, \end{equation}

where $K_G\subset G(\mathbb{A}_{F_0,f})$ is an open compact subgroup and $K_{Z^{\mathbb{Q}}}\subset {Z^{\mathbb{Q}}}(\mathbb{A}_f)$ is the unique maximal compact subgroup

(3.7)\begin{equation} K_{Z^{\mathbb{Q}}} := Z^{\mathbb{Q}}(\widehat{\mathbb{Z}}) = \{z\in (O_F\otimes \widehat{\mathbb{Z}})^{\times} \,|\, \operatorname{Nm}_{F/F_0}(z)\in \widehat{\mathbb{Z}}^{\times}\}. \end{equation}

(Note that $Z^{\mathbb{Q}}$ is defined over $\operatorname {Spec} \mathbb{Z}$ in an obvious way.)

Before doing this, let us first introduce an auxiliary moduli problem $M_0$ over $E$. In fact, for use in the construction of integral models later, we will define a moduli problem $\mathcal {M}_0$ over $O_E$ whose generic fiber will be $M_0$. For a locally noetherian $O_E$-scheme $S$, we define $\mathcal {M}_0(S)$ to be the groupoid of triples $(A_0, \iota _0, \lambda _0)$, where

  • $A_0$ is an abelian variety over $S$ with an $O_F$-action $\iota _0\colon O_F \to \operatorname {End}(A_0)$, which satisfies the Kottwitz condition of signature $((0, 1)_{\varphi \in \Phi })$, that is,

    (3.8)\begin{equation} \operatorname{char}(\iota(a)\mid\operatorname{Lie} A_0) = \prod_{\varphi\in\Phi}(T-\overline\varphi(a)) \quad\text{for all}\ a\in O_F; \end{equation}
    and
  • $\lambda _0$ is a principal polarization of $A_0$ whose Rosati involution induces on $O_F$, via $\iota _0$, the non-trivial Galois automorphism of $F/F_0$.

A morphism between two objects $(A_0,\iota _0,\lambda _0)$ and $(A'_0,\iota '_0,\lambda '_0)$ is an $O_F$-linear isomorphism $\mu _0\colon A_0 \to A'_0$ under which $\lambda _0'$ pulls back to $\lambda _0$. Then $\mathcal {M}_0$ is a Deligne–Mumford (DM) stack, finite and étale over $\operatorname {Spec} O_E$ (cf. [Reference HowardHow12, Proposition 3.1.2]).Footnote 6 (In fact, in [Reference HowardHow12, Proposition 3.1.2], $\mathcal {M}_0$ is defined over the ring of integers $O_{E_\Phi }$ in the reflex field of $\Phi$, which is contained in $O_E$; cf. Remark 3.2(i).) We let $M_0$ denote the generic fiber of $\mathcal {M}_0$.

Unfortunately, it may happen that $\mathcal {M}_0$ is empty. In order to circumvent this problem, we also introduce the following variant of $\mathcal {M}_0$ (cf. [Reference HowardHow12, Definition 3.1.1]). Fix a non-zero ideal $\mathfrak {a}$ of $O_{F_0}$. Then we introduce the DM stack $\mathcal {M}_0^{\mathfrak{a}}$ of triples $(A_0, \iota _0, \lambda _0)$ as before, except that we replace the condition that $\lambda _0$ is principal by the condition that $\ker \lambda _0=A_0[\mathfrak {a}]$. Then, again, $\mathcal {M}_0^{\mathfrak{a}}$ is finite and étale over $\operatorname {Spec} O_E$ (cf. [Reference HowardHow12, Proposition 3.1.2]).

If $\mathcal {M}_0^{\mathfrak{a}}$ is non-empty, then the complex fiber $\mathcal {M}^{\mathfrak{a}}_0 \otimes _{O_E} \mathbb{C}$ is isomorphic to a finite number of copies of $\mathrm {Sh}_{K_{Z^{\mathbb{Q}}}}(Z^{\mathbb{Q}}, \{h_{Z^{\mathbb{Q}}}\})$. More precisely, let $\mathcal {L}_\Phi ^{\mathfrak{a}}$ be the set of isomorphism classes of pairs $(\Lambda _0,\langle \text {,}\rangle _0)$ consisting of a locally free $O_F$-module $\Lambda _0$ of rank $1$ equipped with a non-degenerate alternating form $\langle \text {,}\rangle _0\colon \Lambda _0 \times \Lambda _0 \to \mathbb{Z}$ such that $\langle ax, y \rangle _0 = \langle x, \overline a y\rangle _0$ for all $x,y \in \Lambda _0$ and $a \in O_F$, such that $x \mapsto \langle \sqrt \Delta x, x \rangle _0$ is a negative-definite quadratic form on $\Lambda _0$, and such that the dual lattice $\Lambda _0^{\vee }$ of $\Lambda _0$ inside $\Lambda _0 \otimes _\mathbb{Z} \mathbb{Q}$ equals $\mathfrak {a}{{}^{-1}} \Lambda _0$. Then $\mathcal {L}_\Phi ^{\mathfrak{a}}$ is a finite set.Footnote 7 If $(\Lambda _0,\langle \text {,}\rangle _0)$ defines a class in $\mathcal {L}_\Phi ^{\mathfrak{a}}$, then using $\Phi$ to identify $F \otimes _\mathbb{Q} \mathbb{R} {\tilde{\to}} \mathbb{C}^{\Phi }$, and hence to define a complex structure on $\Lambda _0 \otimes _\mathbb{Z} \mathbb{R}$, the quotient $(\Lambda _0 \otimes _\mathbb{Z} \mathbb{R}) / \Lambda _0$ is a complex torus which defines a $\mathbb{C}$-point on $\mathcal {M}_0^{\mathfrak{a}}$. On the other hand, given a $\mathbb{C}$-point $(A_0,\iota _0,\lambda _0)$ on $\mathcal {M}_0^{\mathfrak{a}}$, the first homology group $H_1(A_0(\mathbb{C}),\mathbb{Z})$, endowed with its Riemann form induced by $\lambda _0$, defines a class in $\mathcal {L}_\Phi ^{\mathfrak{a}}$. In this way $\mathcal {L}_\Phi ^{\mathfrak{a}}$ is in bijection with the isomorphism classes of objects in $\mathcal {M}_0^{\mathfrak{a}}(\mathbb{C})$.

These inverse processes give rise to a decomposition of $\mathcal {M}_0^{\mathfrak{a}}$ into a union of Shimura varieties, in terms of the following equivalence relation on $\mathcal {L}_\Phi ^{\mathfrak{a}}$: define $\Lambda _0 \sim \Lambda _0'$ if $\Lambda _0 \otimes _\mathbb{Z} \widehat{\mathbb{Z}}$ and $\Lambda _0' \otimes _\mathbb{Z} \widehat{\mathbb{Z}}$ are $\widehat O_F$-linearly similar up to a factor in $\widehat{\mathbb{Z}}^{\times }$, and if $\Lambda _0 \otimes _\mathbb{Z} \mathbb{Q}$ and $\Lambda _0' \otimes _\mathbb{Z} \mathbb{Q}$ are $F$-linearly similar up to a factor in $\mathbb{Q}^{\times }$. We have the following lemma.

Lemma 3.4 The stack $\mathcal {M}_0^{\mathfrak{a}}$ admits a decomposition into open and closed substacks,

(3.9)\begin{equation} \mathcal{M}_0^{\mathfrak{a}} = \coprod_{\xi \in \mathcal{L}_\Phi^{\mathfrak{a}}/\sim} \mathcal{M}_0^{\mathfrak{a},\xi}, \end{equation}

induced by sending an object $(A_0,\iota _0,\lambda _0)$ in $\mathcal {M}_0^{\mathfrak{a}}(\mathbb{C})$ to the $\sim$-class of $H_1(A_0(\mathbb{C}),\mathbb{Z})$ endowed with its Riemann form.

Proof. Since $\mathcal {M}_0^{\mathfrak{a}}$ is finite and étale over $\operatorname {Spec} O_E$, it suffices to demonstrate the asserted decomposition in the generic fiber $M_0^{\mathfrak{a}}$,

(3.10)\begin{equation} M_0^{\mathfrak{a}} = \coprod_{\xi \in \mathcal{L}_\Phi^{\mathfrak{a}}/\sim} M_0^{\mathfrak{a},\xi}. \end{equation}

To establish (3.10), again by étaleness, it suffices to show that the $\sim$-class of $H_1(A_0(\mathbb{C}),\mathbb{Z})$ is constant on the $\operatorname {Aut}(\mathbb{C}/E)$-orbit of each object $(A_0,\iota _0,\lambda _0) \in M_0^{\mathfrak{a}}(\mathbb{C})$. So let $\sigma \in \operatorname {Aut}(\mathbb{C}/E)$. Then $\sigma$ identifies the Tate modules $\widehat T (A_0) {\tilde{\to}} \widehat T (A_0^{\sigma })$ compatibly with the Weil pairings on the two sides up to the similitude factor given by the image of $\sigma$ under the cyclotomic character $\operatorname {Aut}(\mathbb{C}) \to \widehat{\mathbb{Z}}^{\times }$. Hence, by the compatibility between singular homology and the Tate module, $H_1(A_0(\mathbb{C}),\widehat{\mathbb{Z}})$ and $H_1(A_0^{\sigma }(\mathbb{C}),\widehat{\mathbb{Z}})$ are $\widehat O_F$-linearly similar up to a factor in $\widehat{\mathbb{Z}}^{\times }$. On the other hand, the CM abelian variety $A_0$ is isomorphic to one defined over the algebraic closure of $E$ in $\mathbb{C}$, and hence by [Reference Chai, Conrad and OortCCO14, Theorem A.2.3.5] there exists a quasi-isogeny of complex abelian varieties $A_0 \to A_0^{\sigma }$ pulling $\lambda _0^{\sigma }$ back to a $\mathbb{Q}_{>0}$-multiple of $\lambda _0$. Hence $H_1(A_0(\mathbb{C}),\mathbb{Q})$ and $H_1(A_0^{\sigma }(\mathbb{C}),\mathbb{Q})$ are $F$-linearly similar up to a factor in $\mathbb{Q}^{\times }$, as desired.

If $\mathcal {M}_0^{\mathfrak{a}} \neq \emptyset$, then for fixed $\xi \in \mathcal {L}_\Phi ^{\mathfrak{a}}/{\sim }$, the complex fiber $\mathcal {M}_0^{\mathfrak {a},\xi } \otimes _{O_E} \mathbb{C}$ of the summand in (3.9) is canonically isomorphic to the Shimura variety $\mathrm {Sh}_{K_{Z^{\mathbb{Q}}}}(Z^{\mathbb{Q}}, \{h_{Z^{\mathbb{Q}}}\})$; this can be checked similarly to [Reference Kudla and RapoportKR14, Propositions 4.3, 4.4], or see [Reference DeligneDel71, (4.12), (4.18)–(4.20)].

Remark 3.5 (i) Given finitely many prime numbers $p_1, \ldots , p_r$, there always exists an ideal $\mathfrak {a}$ relatively prime to $p_1,\ldots , p_r$ such that $\mathcal {M}_0^{\mathfrak{a}}$ is non-empty.

(ii) If $F/F_0$ is ramified at some finite place, then $\mathcal {M}_0^{\mathfrak{a}}$ is non-empty for any $\mathfrak {a}$ (cf. [Reference HowardHow12, proof of Proposition 3.1.6]). A special case of this is when $F= K F_0$, where $K$ is an imaginary quadratic field and the discriminants of $K/\mathbb{Q}$ and $F_0/\mathbb{Q}$ are relatively prime. We further remark that in the context of the global integral models we define in § 5 below, we will eventually impose conditions on the hermitian spaces $W$ and $W^{\flat }$ that force $F/F_0$ to be ramified at some finite place (cf. Remark 5.3).

(iii) When $F_0 = \mathbb{Q}$, the set $\mathcal {L}_\Phi ^{\mathfrak{a}}/{\sim }$ is a singleton, so that the decomposition (3.9) is trivial.

(iv) The decomposition (3.9) in fact holds over $\operatorname {Spec} O_{E_\Phi }$, via the same proof. We have worked over $\operatorname {Spec} O_E$ only for notational simplicity throughout the rest of the paper.

(v) Let $(A_0,\iota _0,\lambda _0)\in \mathcal {M}_0^{\mathfrak{a}} (k)$, where $k$ is a perfect field of positive characteristic $p$. Here is the recipe to determine on which summand $\mathcal {M}_0^{\mathfrak {a}, \xi }$ this point lies. Let $(\widetilde A_0,\widetilde \iota _0,\widetilde \lambda _0)$ be the canonical lift of $(A_0,\iota _0,\lambda _0)$ over the ring of $p$-typical Witt vectors $W(k)$. Then $(\widetilde A_0,\widetilde \iota _0,\widetilde \lambda _0)$ is defined over a subring $R \subset W(k)$ which is finitely generated over $\mathbb{Z}$, and after choosing an embedding $R \to \mathbb{C}$, we obtain a point of $\mathcal {M}_0^{\mathfrak {a}}(\mathbb{C})$ which determines $\xi \in \mathcal {L}_\Phi ^{\mathfrak{a}}/{\sim }$ as in Lemma 3.4. This is the label of the summand containing $(A_0,\iota _0,\lambda _0)$.

In the following, we fix an ideal $\mathfrak {a}$ such that $\mathcal {M}_0^{\mathfrak{a}}$ is non-empty. We continue to denote the generic fiber of this stack by $M_0^{\mathfrak{a}}$. We also fix $\xi \in \mathcal {L}_\Phi ^{\mathfrak{a}}/{\sim }$. We now define a groupoid $M_{K_{\widetilde G}}(\widetilde G)$ fibered over the category of locally noetherian schemes over $E$. Here, for ease of notation, we have suppressed the ideal $\mathfrak {a}$ and the element $\xi$. For such a scheme $S$, the objects of $M_{K_{G}}(\widetilde G)(S)$ are collections $(A_0,\iota _0,\lambda _0,A,\iota ,\lambda ,\overline \eta )$, where

  • $(A_0, \iota _0, \lambda _0)$ is an object of $M^{\mathfrak {a}, \xi }_0(S)$;

  • $A$ is an abelian scheme over $S$ with an $F$-action $\iota \colon F\to \operatorname {End}^{\circ }(A)$ satisfying the Kottwitz condition of signature $((1, n-1)_{\varphi _0}, (0, n)_{\varphi \in \Phi \smallsetminus \{\varphi _0\}})$, that is,

    (3.11)\begin{equation} \operatorname{char}(\iota(a)\mid\operatorname{Lie} A) = (T-\varphi_0(a))(T-\varphi_0(\bar a))^{n-1}\prod_{\varphi\in\Phi\smallsetminus \{\varphi_0\}}(T-\varphi(\bar a))^{n} \quad\text{for all}\ a\in O_F; \end{equation}
  • $\lambda$ is a polarization $A \to A^{\vee }$ whose Rosati involution induces on $F$, via $\iota$, the non-trivial Galois automorphism of $F/F_0$; and

  • $\overline \eta$ is a ${K_{\widetilde G}}$-level structure, by which we mean a ${K_{G}}$-orbit (equivalently, a ${K_{\widetilde G}}$-orbit, where $K_{\widetilde G}$ acts through its projection $K_{\widetilde G}\to K_G$) of $\mathbb{A}_{F, f}$-linear isometries

    (3.12)\begin{equation} \eta\colon \widehat V(A_0, A)\simeq -W\otimes_F\mathbb{A}_{F, f}. \end{equation}

Here we regard the right-hand side of (3.12) as a constant smooth $\mathbb{A}_{F,f}$-sheaf on $S$, and

\[ \widehat V(A_0,A) := \operatorname{Hom}_{F}(\widehat V(A_0), \widehat V(A)) , \]

endowed with its natural $\mathbb{A}_{F, f}$-valued hermitian form $h_A$, defined by

(3.13)\begin{equation} h_A(x, y) := \lambda_0^{-1}\circ y^{\vee}\circ\lambda\circ x\in\operatorname{End}_{\mathbb{A}_{F, f}}(\widehat V(A_0))=\mathbb{A}_{F, f} \end{equation}

(cf. [Reference Kudla and RapoportKR14, § 2.3]). Over each connected component of $S$, upon choosing a geometric point $\overline s \to S$, we may view $\widehat V (A_0, A)$ as the space $\operatorname {Hom}_{\mathbb{A}_{F,f}}(\widehat V(A_{0,\overline s}), \widehat V(A_{\overline s}))$ endowed with its $\pi _1(S,\overline s)$-action, and in this way we require the orbit $\overline \eta$ to be $\pi _1(S,\overline s)$-invariant (cf. [Reference Kudla and RapoportKR14, Remark 4.2]). We note that the notion of level structure is independent of the choice of $\overline s$ on each connected component of $S$.

A morphism between two objects $(A_0,\iota _0,\lambda _0,A,\iota ,\lambda ,\overline \eta )$ and $(A'_0,\iota '_0,\lambda '_0,A',\iota ',\lambda ',\overline \eta ')$ is given by an isomorphism $\mu _0\colon (A_0, \iota _0, \lambda _0) {\tilde{\to}} (A'_0, \iota '_0, \lambda '_0)$ in $M_0^{\mathfrak {a},\xi }(S)$ and an $F$-linear quasi-isogeny $\mu \colon A \to A'$ pulling $\lambda '$ back to $\lambda$ and $\overline \eta '$ back to $\overline \eta$.

Remark 3.6 (i) Let $r \colon \operatorname {Hom}(F,\mathbb{C}) \to \{0,1,n-1,n\}$, $\varphi \mapsto r_\varphi$ be the function defined by

(3.14)\begin{equation} r_\varphi := \begin{cases} 1, & \varphi = \varphi_0,\\ 0, & \varphi \in \Phi \smallsetminus \{\varphi_0\},\\ n-r_{\overline\varphi}, & \varphi \notin \Phi. \end{cases} \end{equation}

Then the Kottwitz condition (3.11) is

\[ \operatorname{char}(\iota(a)\mid\operatorname{Lie} A) = \prod_{\varphi\in\operatorname{Hom}(F,\mathbb{C})} (T-\varphi(a))^{r_\varphi} \quad\text{for all}\ a \in O_F \]

(cf. [Reference Rapoport and ZinkRZ17, (8.4)]).

(ii) The intervention of the sign on the right-hand side of (3.12) arises from our conventions on the signatures of $W$ at the archimedean places given at the beginning of § 3.1 and on the signatures in the Kottwitz conditions (3.8) and (3.11); cf. the proof of Proposition 3.7 below. If one took the opposite signatures for $W$ and in the two Kottwitz conditions, then no sign would be needed.

The following proposition is a special case of Deligne's description of Shimura varieties of PEL type.

Proposition 3.7 $M_{K_{\widetilde G}}(\widetilde G)$ is a DM stack smooth of relative dimension $n-1$ over $\operatorname {Spec} E$. The coarse moduli scheme of $M_{K_{\widetilde G}}(\widetilde G)$ is a quasi-projective scheme over $\operatorname {Spec} E$, naturally isomorphic to the canonical model of $\mathrm {Sh}_{K_{\widetilde G}}(\widetilde G, \{h_{\widetilde G}\})$. For ${K_{\widetilde G}}$ sufficiently small, the forgetful morphism $M_{K_{\widetilde G}}(\widetilde G)\to M^{\mathfrak {a}, \xi }_0$ is relatively representable.

Proof. We will content ourselves with exhibiting a map, which turns out to be an isomorphism,

\[ M_{K_{\widetilde G}}(\widetilde G)\otimes_{E}\mathbb{C} \longrightarrow \mathrm{Sh}_{K_{\widetilde G}}(\widetilde G, \{h_{\widetilde G}\}). \]

By Remark 3.2(iii), the target is the product of Shimura varieties

\[ \mathrm{Sh}_{K_{Z^{\mathbb{Q}}}}(Z^{\mathbb{Q}},\{h_{Z^{\mathbb{Q}}}\}) \times \mathrm{Sh}_{K_G}(\operatorname{Res}_{F/F_0} G,\{h_G\}). \]

For the map into the first factor, we simply compose the forgetful map $M_{K_{\widetilde G}}(\widetilde G)\otimes _{E}\mathbb{C} \to M_0^{\mathfrak {a}, \xi }\otimes _{E}\mathbb{C}$ with the isomorphism $M_0^{\mathfrak {a}, \xi }\otimes _{E}\mathbb{C} \simeq \mathrm {Sh}_{K_{Z^{\mathbb{Q}}}}(Z^{\mathbb{Q}},\{h_{Z^{\mathbb{Q}}}\})$.

To explain the map into the second factor, let $(A_0,\iota _0,\lambda _0,A,\iota ,\lambda ,\overline \eta )$ be a $\mathbb{C}$-valued point of $M_{K_{\widetilde G}}(\widetilde G)$. Let $\mathcal {H} := H_1(A, \mathbb{Q})$ and $\mathcal {H}_0 := H_1(A_0, \mathbb{Q})$. The polarization $\lambda$ endows $\mathcal {H}$ with a $\mathbb{Q}$-valued alternating form $\langle \text {,}\rangle$ satisfying $\langle \iota (a)x,y\rangle = \langle x, \iota (\overline a)y\rangle$ for all $a \in F$, and such that the induced form $x,y \mapsto \langle \sqrt {-1} \cdot x, y \rangle$ on $\mathcal {H} \otimes _\mathbb{Q} \mathbb{R}$ is symmetric and positive definite, where multiplication by $\sqrt {-1}$ is defined in terms of the right-hand side of the canonical isomorphism $\mathcal {H}_\mathbb{R} \cong \operatorname {Lie} A$. Similarly, $\lambda _0$ endows $\mathcal {H}_0$ with a Riemann form $\langle \text {,}\rangle _0$.

Let $V(A_0, A):=\operatorname {Hom}_F(\mathcal {H}_0, \mathcal {H})$. Then $V(A_0, A)$ is an $F$-vector space of the same dimension as $W$, and we make it into an $F/F_0$-hermitian space by defining the pairing $(\alpha ,\beta )$ to be the composite

where the checks denote $\mathbb{Q}$-linear duals and the last arrow is the inverse of $y \mapsto \langle -, y \rangle _0$; this composite is an $F$-linear endomorphism of the one-dimensional $F$-vector space $\mathcal {H}_0$, and hence identifies with an element in $F$. Clearly $V(A_0, A) \otimes _F \mathbb{A}_{F,f} \cong \widehat V(A_0, A)$ as hermitian spaces. Hence, by the existence of a level structure, $V(A_0, A) \otimes _F \mathbb{A}_{F,f} \simeq -W \otimes _F \mathbb{A}_{F,f}$. Furthermore, it is easy to see that the Kottwitz conditions (3.8) and (3.11) imply that $V(A_0, A)_\varphi$ has signature $(n-1,1)$ if $\varphi =\varphi _0$ and $(n,0)$ if $\varphi \in \Phi \smallsetminus \{\varphi _0\}$. Hence, by the Hasse principle for hermitian spaces, $V(A_0, A)$ and $-W$ are isomorphic. Choose an isometry $j\colon V(A_0, A){\tilde{\to}} -W$. Using the complex structures on $\mathcal {H}_{0,\mathbb{R}}$ and $\mathcal {H}_\mathbb{R}$, let $z \in \mathbb{C}^{\times }$ act on $V(A_0, A)$ by sending the $F$-linear map $\alpha$ to $z\alpha z{{}^{-1}}$. This defines a homomorphism $\mathbb{C}^{\times } \longrightarrow \mathrm {U}(V(A_0, A))(\mathbb{R})$, and composing this with $j_*\colon \mathrm {U}(V(A_0, A))(\mathbb{R}) {\tilde{\to}} \mathrm {U}(-W)(\mathbb{R}) = \mathrm {U}(W)(\mathbb{R})$ gives an element in $\{h_G\}$. The level structure $\overline \eta$ corresponds to an element of $(\operatorname {Res}_{F_0/\mathbb{Q}} G)(\mathbb{A}_f)/K_{G}$, and eliminating the choice of $j$ corresponds to dividing out by the action of $(\operatorname {Res}_{F_0/\mathbb{Q}} G)(\mathbb{Q})$.

An analogous description holds for the model $M_{K_{\widetilde H}}(\widetilde H)$ of the Shimura variety $\mathrm {Sh}_{K_{\widetilde H}}(\widetilde H, \{h_{\widetilde H}\})$ (replace $n$ by $n-1$, and $W$ by $W^{\flat }$).

There is also an analog for the Shimura variety $\mathrm {Sh}_{K_{\widetilde {HG}}}(\widetilde {HG}, \{h_{\widetilde {HG}}\})$. In this case we take the level subgroup to be of the form

(3.15)\begin{equation} K_{\widetilde{HG}} = K_{Z^{\mathbb{Q}}} \times K_H \times K_G, \end{equation}

where as always $K_{Z^{\mathbb{Q}}}$ is the subgroup (3.7), and $K_H \subset H(\mathbb{A}_{F_0,f})$ and $K_G \subset G(\mathbb{A}_{F_0,f})$ are open compact subgroups. The value of the corresponding moduli functor $M_{K_{\widetilde {HG}}}(\widetilde {HG})$ on a locally noetherian scheme $S$ over $E$ is the set of isomorphism classes of tuples

\[ (A_0,\iota_0,\lambda_0,A^{\flat},\iota^{\flat},\lambda^{\flat}, A, \iota,\lambda, \overline{(\eta^{\flat}, \eta)}) , \]

where the last entry is a pair of $\mathbb{A}_{F, f}$-linear isometries

\[ \eta^{\flat} \colon \widehat V(A_0, A^{\flat})\simeq -W^{\flat}\otimes_F\mathbb{A}_{F, f} \quad\text{and}\quad \eta\colon \widehat V(A_0, A)\simeq -W\otimes_F\mathbb{A}_{F, f} , \]

modulo the action of ${K_{H}\times K_G}$. In other words, the moduli functor $M_{K_{\widetilde {HG}}}(\widetilde {HG})$ is simply the fibered product $M_{K_{\widetilde H}}(\widetilde H) \times _{M^{\mathfrak {a},\xi }_0} M_{K_{\widetilde G}}(\widetilde G)$.

In terms of these moduli problems, the injective morphisms (3.2) can be described as follows. After possibly scaling the special vector $u \in W$, we may and do assume that the norm $(u,u)$ is in $O_{F_0}$. Further, assume that ${K_{H}}\subset H(\mathbb{A}_{F_0,f})\cap {K_{ G}}$. Then the first morphism of Shimura varieties in (3.2) arises by base change from $E$ to $\mathbb{C}$ from the functor morphism

(3.16)

Here $\lambda _0(u)= -(u,u) \lambda _0$, which is indeed a polarization by [Reference ZinkZin82, Lemma 1.4] since $-(u,u)$ is totally positive. Furthermore, the isomorphism $\eta \colon \widehat V(A_0,A^{\flat }\times A_0) \simeq -W \otimes _F \mathbb{A}_{F, f} \bmod K_{G}$ is given, with respect to the decomposition

\[ \operatorname{Hom}_F( \widehat V(A_0), \widehat V(A^{\flat} \times A_0)) \cong \operatorname{Hom}_F(\widehat V(A_0), \widehat V(A^{\flat}) \oplus \widehat V(A_0)) \cong \widehat V(A_0,A^{\flat}) \oplus { \mathbb{A}}_{F, f}, \]

by the trivialization

\[ \eta^{\flat}\colon \widehat V(A_0,A^{\flat}) \simeq -W^{\flat} \otimes_F \mathbb{A}_{F, f} \bmod K_{ H} \]

in the first summand, and by identifying the basis element $1$ in the second summand with $u \otimes 1 \in -W \otimes _F \mathbb{A}_{F,f}$. The morphism (

3.16

) is finite and unramified.

The second morphism in (3.2) then arises from the graph morphism of (3.16),

(3.17)\begin{equation} M_{K_{\widetilde H}} (\widetilde H) \longrightarrow M_{K_{\widetilde{HG}}}(\widetilde{HG}) = M_{K_{\widetilde H}}(\widetilde H) \times_{M^{\mathfrak{a},\xi}_0} M_{K_{\widetilde G}}(\widetilde G). \end{equation}

The morphism (3.17) is a closed embedding.

Remark 3.8 For $F_0 \neq \mathbb{Q}$, the Shimura varieties above are compact. For $F_0 = \mathbb{Q}$, it may happen that the Shimura variety $M_{K_{\widetilde {HG}}}(\widetilde {HG})$ is non-compact. In fact, this will be automatic when $n\geq 3$. In this case, we will need to use its canonical toroidal compactification (cf. [Reference HowardHow15, § 2]).

4. Semi-global integral models

Fix a prime number $p$ and an embedding $\widetilde \nu \colon \overline{\mathbb{Q}} \longrightarrow \overline{\mathbb{Q}}_p$. This determines a $p$-adic place $\nu$ of $E$ and, via $\varphi _0$, places $v_0$ of $F_0$ and $w_0$ of $F$. In this section we will define ‘semi-global’ integral models over $O_{E,(\nu )}$ of the moduli spaces introduced in § 3, in the case of various level structures at $p$. We denote by $\mathcal {V}_p$ the set of places $v$ of $F_0$ over $p$. Throughout this section, we assume that the ideal $\mathfrak {a}$ occurring in the definition of $\mathcal {M}_0^{\mathfrak{a}}$ is prime to $p$ (cf. Remark 3.5(i)). We also fix $\xi \in \mathcal {L}_\Phi ^{\mathfrak{a}}/{\sim }$.

4.1 Hyperspecial level at $v_0$

In this case we assume that the place $v_0$ is unramified over $p$, and that $v_0$ either splits in $F$ or is inert in $F$ and the hermitian space $W_{v_0}$ is split. We also assume $p \neq 2$ if there is any $v \in \mathcal {V}_p$ which is non-split in $F$.Footnote 8 We will define smooth models over $O_{E,(\nu )}$.

For each $v\in \mathcal {V}_{p}$, choose a vertex lattice $\Lambda _{v}$ in the $F_v/F_{0,v}$-hermitian space $W_{v}$. By our assumptions on $v_0$ in the previous paragraph, we may and do take $\Lambda _{v_0}$ to be self-dual. Recalling the subgroup $K_{\widetilde G} = K_{Z^{\mathbb{Q}}} \times K_G$ from (3.6), we take $K_G$ to be of the form

\[ K_G = K_G^{p} \times K_{G,p}, \]

where $K_G^{p} \subset G(\mathbb{A}_{F_0,f}^{p})$ is arbitrary, and where the level subgroup at $p$ is the product

(4.1)\begin{equation} K_{G,p} :=\prod_{v\in \mathcal{V}_p}K_{G, v} \subset G(F_0\otimes \mathbb{Q}_p)=\prod_{v\in \mathcal{V}_p}G(F_{0, v}), \end{equation}

with $K_{G, v}$ the stabilizer of $\Lambda _v$ in $G(F_{0,v})$.

We formulate a moduli problem over $\operatorname {Spec} O_{E,(\nu )}$ as follows. To each locally noetherian $O_{E,(\nu )}$-scheme $S$, we associate the groupoid of tuples $(A_0,\iota _0,\lambda _0,A,\iota ,\lambda ,\overline \eta ^{p})$, where $(A_0,\iota _0,\lambda _0)$ is an object of $\mathcal {M}^{\mathfrak {a}, \xi }_0(S)$. Furthermore:

  • $(A,\iota )$ is an abelian scheme over $S$ with an $O_{F}\otimes \mathbb{Z}_{(p)}$-action $\iota$ satisfying the Kottwitz condition (3.11) of signature $((1, n-1)_{\varphi _0}, (0, n)_{\varphi \in \Phi \smallsetminus \{\varphi _0\}})$.

  • $\lambda$ is a polarization on $A$ whose Rosati involution induces on $O_{F}\otimes \mathbb{Z}_{(p)}$ the non-trivial Galois automorphism of $F/F_0$, subject to the following condition. Note that the action of the ring $O_{F_0}\otimes \mathbb{Z}_p \cong \prod _{v \in \mathcal {V}_p} O_{F_0,v}$ on the $p$-divisible group $A[p^{\infty }]$ induces a decomposition of $p$-divisible groups,

    (4.2)\begin{equation} A[p^{\infty}] = \prod_{v \in \mathcal{V}_p} A[v^{\infty}]. \end{equation}
    Since $\operatorname {Ros}_\lambda$ is trivial on $O_{F_0}$, $\lambda$ induces a polarization $\lambda _v \colon A[v^{\infty }] \to A^{\vee }[v^{\infty }] \cong A[v^{\infty }]^{\vee }$ for each $v$. The condition we impose is that $\ker \lambda _v$ is contained in $A[\iota (\pi _v)]$ of rank $\#(\Lambda _v^{*}/ \Lambda _v)$ for each $v \in \mathcal {V}_p$.
  • $\overline \eta ^{p}$ is a $K_G^{p}$-orbit of $\mathbb{A}_{F,f}^{p}$-linear isometries

    (4.3)\begin{equation} \eta^{p}\colon \widehat V^{p}(A_0,A) \simeq -W \otimes_F \mathbb{A}_{F,f}^{p}, \end{equation}
    where
    \[ \widehat V^{p}(A_0,A) := \operatorname{Hom}_{F}(\widehat V^{p}(A_0), \widehat V^{p}(A)), \]
    and the hermitian form on $\widehat V^{p}(A_0,A)$ is the evident prime-to-$p$ analog of (3.13).

We also impose for each $v\neq v_0$ over $p$ the sign condition and the Eisenstein condition. Let us explain these conditions.

The sign condition at $v$ is only non-empty when $v$ does not split in $F$, in which case it requires that at every point $s$ of $S$,

(4.4)\begin{equation} {{\mathrm{inv}}}^{r}_v(A_{0,s}, \iota_{0,s}, \lambda_{0,s}, A_s, \iota_s, \lambda_s) = {{\mathrm{inv}}}_v(-W_v). \end{equation}

Here the left-hand side is the sign factor defined in (A.5) and (A.8) in Appendix A (in the definition of (A.8), one may use the embedding $\widetilde \nu$ fixed at the beginning of this section). The right-hand side is the Hasse invariant of the hermitian space $-W_v$ defined above in (1.4). Note that by Proposition A.1, the left-hand side of (4.4) is a locally constant function in $s$.

The Eisenstein condition is only non-empty when the base scheme $S$ has non-empty special fiber. In this case, we may even base change via $\widetilde \nu \colon O_{E, (\nu )}\to \overline{\mathbb{Z}}_p$ (the ring of integers in $\overline{\mathbb{Q}}_p$) and pass to completions and assume that $S$ is a scheme over $\operatorname {Spf} \overline{\mathbb{Z}}_p$. Similarly to (4.2), there is a decomposition of the $p$-divisible group $A[p^{\infty }]$,

(4.5)\begin{equation} A[p^{\infty}] = \prod_{w\mid p} A[w^{\infty}] , \end{equation}

where the indices range over the places $w$ of $F$ lying over $p$. Since we assume that $p$ is locally nilpotent on $S$, there is a natural isomorphism

\begin{equation*} \operatorname{Lie} A\cong \operatorname{Lie} A[p^{\infty}]=\bigoplus_{w\mid p} \operatorname{Lie} A[w^{\infty}]. \end{equation*}

For each place $w$, by the Kottwitz condition (3.11), the $p$-divisible group $A[w^{\infty }]$ is of height $n \cdot [F_{w}:\mathbb{Q}_p]$ and dimension

(4.6)\begin{equation} \dim A[w^{\infty}] = \sum_{\varphi \in \operatorname{Hom}(F_w, \overline{\mathbb{Q}}_p)} r_\varphi. \end{equation}

Here $r_\varphi$ is as in (3.14), and we have used the embedding $\widetilde \nu \colon \overline{\mathbb{Q}} \to \overline{\mathbb{Q}}_p$ to identify

(4.7)\begin{equation} \operatorname{Hom}_\mathbb{Q}(F, \overline{\mathbb{Q}}) \simeq \operatorname{Hom}_\mathbb{Q}(F, \overline{\mathbb{Q}}_p) , \end{equation}

which in turn identifies

(4.8)\begin{equation} \{\varphi \in \operatorname{Hom}_{\mathbb{Q}}(F, \overline{\mathbb{Q}}) \,|\, w_\varphi=w \} \simeq \operatorname{Hom}_{\mathbb{Q}_p}(F_w, \overline{\mathbb{Q}}_p), \end{equation}

where $w_\varphi$ denotes the $p$-adic place in $F$ induced by $\widetilde \nu \circ \varphi$.

Now suppose that $w$ lies over a place $v$ different from $v_0$. Then the action of $F$ on $A[w^{\infty }]$ is of a banal signature type, in the sense that each integer $r_\varphi$ occurring in (4.6) is equal to $0$ or $n$ (cf. Appendix B). Let $\pi = \pi _w$ be a uniformizer in $F_w$, and let $F_w^{t} \subset F_w$ be the maximal unramified subextension of $\mathbb{Q}_p$. For each $\psi \in \operatorname {Hom}_{\mathbb{Q}_p}(F^{t}_w, \overline{\mathbb{Q}}_p)$, let

\[ A_\psi := \{\varphi\in \operatorname{Hom}_{\mathbb{Q}_p}(F_w, \overline{\mathbb{Q}}_p) \,|\, \varphi|_{F_w^{t}}=\psi \text{ and } r_\varphi=n \}. \]

Set

\[ Q_{A_\psi}(T):=\prod_{\varphi\in A_\psi}(T-\varphi(\pi))\in \overline{\mathbb{Z}}_p[T]. \]

Since we assume that $S$ is a scheme over $\operatorname {Spf} \overline{\mathbb{Z}}_p$, there is a natural isomorphism

\[ O_{F_w^{t}} \otimes_{\mathbb{Z}_p} \mathcal{O}_S {\tilde{\to}} \prod_{\psi \in \operatorname{Hom}_{\mathbb{Q}_p}(F_w^{t}, \overline{\mathbb{Q}}_p)} \mathcal{O}_S, \]

whose $\psi$-component is $\psi \otimes \mathrm {id}$. Hence the $O_{F_w^{t}}$-action on $\operatorname {Lie} A[w^{\infty }]$ induces a decomposition

(4.9)\begin{equation} \operatorname{Lie} A[w^{\infty}] \cong \bigoplus_{\psi \in \operatorname{Hom}_{\mathbb{Q}_p}(F_w^{t}, \overline{\mathbb{Q}}_p)} \operatorname{Lie}_\psi A[w^{\infty}]. \end{equation}

The Eisenstein condition at $v$ demands the identity of endomorphisms of $\operatorname {Lie}_\psi A[w^{\infty }]$,

(4.10)\begin{equation} Q_{A_\psi}(\iota(\pi) \mid \operatorname{Lie}_{\psi} A[w^{\infty}])=0 \quad \text{for each} \quad w \mid v \quad \text{and each} \quad \psi\in \operatorname{Hom}_{\mathbb{Q}_p}(F^{t}_w, \overline{\mathbb{Q}}_p). \end{equation}

This condition is the analog in our context of the condition with the same name in [Reference Rapoport and ZinkRZ17, (8.2)]. The Kottwitz condition implies that the Eisenstein condition at $v$ is automatically satisfied when the one or two places $w$ over $v$ are unramified over $p$ (cf. Lemma B.3).

A morphism between two objects $(A_0, \iota _0, \lambda _0, A, \iota , \lambda , \overline \eta ^{p})$ and $(A_0', \iota _0', \lambda _0', A', \iota ', \lambda ', \overline \eta '^{p})$ is given by an isomorphism $(A_0,\iota _0,\lambda _0) {\tilde{\to}} (A_0',\iota _0',\lambda _0')$ in $\mathcal {M}^{\mathfrak {a},\xi }_0(S)$ and a quasi-isogeny $A \to A'$ which induces an isomorphism

\[ A[p^{\infty}] {\tilde{\to}} A'[p^{\infty}] , \]

compatible with $\iota$ and $\iota '$, with $\lambda$ and $\lambda '$, and with $\overline \eta ^{p}$ and $\overline \eta ^{p\prime }$.

Theorem 4.1 The moduli problem just formulated is representable by a DM stack $\mathcal {M}_{K_{\widetilde G}}(\widetilde G)$ smooth over $\operatorname {Spec} O_{E,(\nu )}$. For $K_G^{p}$ small enough, $\mathcal {M}_{K_{\widetilde G}}(\widetilde G)$ is relatively representable over $\mathcal {M}^{\mathfrak {a},\xi }_0$. Furthermore, the generic fiber $\mathcal {M}_{K_{\widetilde G}} \times _{\operatorname {Spec} O_{E,(\nu )}} \operatorname {Spec} E$ is canonically isomorphic to $M_{K_{\widetilde G}}(\widetilde G)$.

Proof. Representability and relative representability are standard (cf. [Reference KottwitzKot92, p. 391]). Smoothness follows as usual from the theory of local models (cf. [Reference Pappas, Rapoport and SmithlingPRS13]). More precisely, the local model for $\mathcal {M}_{K_{\widetilde G}}(\widetilde G)$ decomposes into a product of local models, one for each of the abelian schemes $A_0$ and $A$ in the moduli problem. The local model corresponding to $A_0$ is étale because $\mathcal {M}_0^{\mathfrak {a},\xi }$ is. Now, under the identifications (4.7) and (4.8), we have

(4.11)\begin{equation} \operatorname{Hom}_\mathbb{Q}(F,\overline{\mathbb{Q}}) \simeq \bigsqcup_{v \in \mathcal{V}_p} \operatorname{Hom}_{\mathbb{Q}_p}(F_v, \mathbb{Q}_p). \end{equation}

In this way the completion $E_\nu$ identifies with the join of the local reflex fields $E_{\Phi _v}$ and $E_{r|_v}$ in $\overline{\mathbb{Q}}_p$ as $v$ varies through $\mathcal {V}_p$, where $\Phi _v := \Phi \cap \operatorname {Hom}_{\mathbb{Q}_p}(F_v,\overline{\mathbb{Q}}_p)$ in terms of the identification (4.11), and where $r|_v\colon \operatorname {Hom}_{\mathbb{Q}_p}(F_v,\overline{\mathbb{Q}}_p) \longrightarrow \mathbb{Z}$ denotes the restriction of the function $r$ to the $v$-summand on the right-hand side of (4.11). The local model $M$ corresponding to $A$ then decomposes as

(4.12)\begin{equation} M = \prod_{v \in \mathcal{V}_p} M_v \times_{\operatorname{Spec} O_{E_{r|_v}}} \operatorname{Spec} O_{E_\nu}. \end{equation}

Here for $v \neq v_0$, by the Kottwitz condition (3.11), $M_v = M(F_v/F_{0,v},r|_v,\Lambda _v)$ is a banal local model, that is, of the form defined in Appendix B. Hence $M_v = \operatorname {Spec} O_{E_{r|_v}}$ by Lemmas B.1 and B.4. (The same is true for the local model corresponding to $A_0$ at every $v \in \mathcal {V}_p$, by these lemmas and Remark B.2) The local model $M_{v_0}$ is smooth by [Reference GörtzGör01].

It remains to prove the last assertion. Let $S$ be a scheme over $E$, and let $(A_0, \iota _0, \lambda _0, A, \iota , \lambda , \overline \eta ^{p})$ be a point of $\mathcal {M}_{K_{\widetilde G}}(S)$. We want to associate to this a point of $M_{K_{\widetilde G}}(S)$, that is, we want to add the $p$-component of $\overline \eta$. The product of hermitian spaces $W \otimes _\mathbb{Q} \mathbb{Q}_p=\prod _{v \in \mathcal {V}_p} W_v$ contains the lattice $\prod _{v \in \mathcal {V}_p}\Lambda _v$, where $\Lambda _v$ is a vertex lattice in $W_v$. By assumption on the polarization $\lambda$, the product of hermitian spaces

\[ V_p(A_0, A) := \operatorname{Hom}_F(V_p(A_0), V_p(A)) = \prod_{v \in \mathcal{V}_p} V_v(A_0, A) \]

contains $\operatorname {Hom}_{O_F}(T_p(A_0), T_p(A))$ as a product of vertex lattices, where the factor at each $v$ is of the same type as $\Lambda _v$. Since $p\neq 2$ when there are non-split places in $\mathcal {V}_p$, it follows that, if there exists an isometry $\eta _p\colon V_p(A_0, A)\simeq -W \otimes _\mathbb{Q} \mathbb{Q}_p$ at all, then there also exists one that maps these two vertex lattices of identical type into one another, and the class modulo $K_{G, p}$ of such an isometry is then uniquely determined.

Hence we are reduced to showing that there exists an isometry $\eta _p$, that is, the equality of Hasse invariants ${{\mathrm {inv}}}_v(V_v(A_0, A))={{\mathrm {inv}}}_v(-W_v)$ for all $v \in \mathcal {V}_p$. By the Hasse principle and the product formula for hermitian spaces, it suffices to prove that for any $\mathbb{C}$-valued point $(A_0, \iota _0, \lambda _0, A, \iota , \lambda , \overline \eta ^{p})$ of $\mathcal {M}_{K_{\widetilde G}}$,

\[ {{\mathrm{inv}}}_v (V(A_0, A)_v)={{\mathrm{inv}}}_v(-W_v) \quad \text{for all}\ v\neq v_0 , \]

where $v$ runs through all places of $F_0$, including the archimedean ones. Here, as in the proof of Proposition 3.7, $V(A_0, A)=\operatorname {Hom}_F (\mathcal {H}_0, \mathcal {H})$, where $\mathcal {H}_0=H_1(A_0,\mathbb{Q})$ and $\mathcal {H}= H_1(A, \mathbb{Q})$. For the non-archimedean places not lying over $p$, this follows from the existence of the level structure; for the places $v\in \mathcal {V}_p\smallsetminus \{v_0\}$, this follows from the sign condition (4.4) at $v$; and finally, for the archimedean places, this follows from the fact that the signatures of $V(A_0, A)$ and $-W$ at all archimedean places are identical (cf. the proof of Proposition 3.7).

We analogously define the DM stacks $\mathcal {M}_{K_{\widetilde H}}(\widetilde H)$ and $\mathcal {M}_{K_{\widetilde {HG}}}(\widetilde {HG})$ over $\operatorname {Spec} O_{E,(\nu )}$ (cf. the end of § 3.2). Both are again smooth over $\operatorname {Spec} O_{E,(\nu )}$.

Let us now assume that the special vector $u \in W$ has norm $(u, u)\in O_{F_0,(p)}^{\times }$. Then we obtain a finite unramified morphism (respectively, a closed embedding), in analogy with (3.16) (respectively, (3.17)),

(4.13)\begin{equation} \mathcal{M}_{K_{\widetilde H}}(\widetilde H) \longrightarrow \mathcal{M}_{K_{\widetilde G}}(\widetilde G) \quad\text{and}\quad \mathcal{M}_{K_{\widetilde H}}(\widetilde H) {\hookrightarrow} \mathcal{M}_{K_{\widetilde{HG}}}(\widetilde{HG}). \end{equation}

For this we assume that $K_H^{p} \subset H(\mathbb{A}_{F_0,f}^{p}) \cap K_G^{p}$. Furthermore, we assume for each $v\in \mathcal {V}_p$ that the lattices in $W_v$ and $W_v^{\flat }$ satisfy the relation

(4.14)\begin{equation} \Lambda_v = \Lambda_v^{\flat} \oplus O_{F,v}u. \end{equation}

For the lattice in $W_v^{\flat } \oplus W_v$, we take the direct sum $\Lambda _v^{\flat } \oplus \Lambda _v$.

We end this subsection by defining Hecke correspondences attached to adelic elements prime to $p$. We first consider the case of $\mathcal {M}_{K_{\widetilde G}}(\widetilde G)$. Fix $g \in \widetilde G(\mathbb{A}_f^{p})$. Let $K_{\widetilde G}^{p} := Z^{\mathbb{Q}}(\widehat{\mathbb{Z}}^{p}) \times K_G^{p}$ and $K_{\widetilde G,p} := Z^{\mathbb{Q}}(\mathbb{Z}_p) \times K_{G,p}$, and set

\[ K_{\widetilde G}^{\prime p} := K_{\widetilde G}^{p} \cap g K_{\widetilde G}^{p} g{{}^{-1}} \quad\text{and}\quad K_{\widetilde G}' := K_{\widetilde G}^{\prime p} \times K_{\widetilde G,p}. \]

Then we obtain in the standard way a diagram of finite étale morphisms

(4.15)

which we view as a correspondence from $\mathcal {M}_{K_{\widetilde G}}(\widetilde G)$ to itself. Note that for a central element $g=z\in Z(G)(\mathbb{A}_{F_0,f}^{p})=\{z\in (\mathbb{A}^{p}_{F,f})^{\times }\mid \operatorname {Nm}_{F/F_0}(z)=1\}$, the diagram (

4.15

) collapses to a map

(4.16)\begin{equation} \mathcal{M}_{K_{\widetilde G}}(\widetilde G) {\mathop \to \limits^z} \mathcal{M}_{K_{\widetilde G}}(\widetilde G), \end{equation}

and this induces an action of $Z(G)(\mathbb{A}_{F_0,f}^{p})$ on $\mathcal {M}_{K_{\widetilde G}}(\widetilde G)$.

The cases of $\mathcal {M}_{K_{\widetilde H}}(\widetilde H)$ and $\mathcal {M}_{K_{\widetilde {HG}}}(\widetilde {HG})$ are completely analogous, simply replacing $\widetilde G$ everywhere by $\widetilde H$ and $\widetilde {HG}$, respectively. For later use, we record the diagram of finite étale morphisms we obtain for $\widetilde {HG}$:

(4.17)

4.2 Split level at $v_0$

We continue with the setup and assumptions of the previous subsection, except we now allow $v_0$ to be ramified over $p$. In addition, we assume that $v_0$ splits in $F$, say into $w_0$ and another place $\overline w_0$. We will define smooth integral models over $O_{E,(\nu )}$.

We define the moduli functor $\mathcal {M}_{K_{\widetilde G}}(\widetilde G)$ as follows. To each locally noetherian $O_{E,(\nu )}$-scheme $S$, we associate the groupoid of tuples $(A_0,\iota _0,\lambda _0,A,\iota ,\lambda ,\overline \eta ^{p})$ as in the previous subsection, except that we impose the further condition corresponding to $w_0$ that when $p$ is locally nilpotent on $S$, the $p$-divisible group $A[w_0^{\infty }]$ is a Lubin–Tate group of type $r|_{w_0}$, in the sense of [Reference Rapoport and ZinkRZ17, § 8] (note that this involves the Eisenstein condition of [Reference Rapoport and ZinkRZ17, § 8]). Here $r|_{w_0}$ is the restriction of the function $r$ on $\operatorname {Hom}_\mathbb{Q}(F, \overline{\mathbb{Q}})$ to $\operatorname {Hom}_{\mathbb{Q}_p}(F_{w_0}, \overline{\mathbb{Q}}_p)$, in the sense of the identification (4.8). We note that if $v_0$ is unramified over $p$, then this further Eisenstein condition is redundant, and the moduli functor $\mathcal {M}_{K_{\widetilde G}}(\widetilde G)$ is the same as the one defined in the previous subsection (cf. [Reference Rapoport and ZinkRZ17]).

Theorem 4.2 The moduli problem just formulated is representable by a DM stack $\mathcal {M}_{K_{\widetilde G}}(\widetilde G)$ smooth over $\operatorname {Spec} O_{E,(\nu )}$. For $K_G^{p}$ small enough, $\mathcal {M}_{K_{\widetilde G}}(\widetilde G)$ is relatively representable over $\mathcal {M}^{\mathfrak {a},\xi }_0$. Furthermore, the generic fiber $\mathcal {M}_{K_{\widetilde G}} \times _{\operatorname {Spec} O_{E,(\nu )}} \operatorname {Spec} E$ is canonically isomorphic to $M_{K_{\widetilde G}}(\widetilde G)$.

Proof. The proof is the same as that of Theorem 4.1, using in addition that the factor at $w_0$ for the local model for $A$ for the newly introduced Eisenstein condition is smooth (cf. [Reference Rapoport and ZinkRZ17, § 8]).

We analogously define the DM stacks $\mathcal {M}_{K_{\widetilde H}}(\widetilde H)$ and $\mathcal {M}_{K_{\widetilde {HG}}}(\widetilde {HG})$ over $\operatorname {Spec} O_{E,(\nu )}$ (cf. the end of § 3.2). Both are again smooth over $\operatorname {Spec} O_{E,(\nu )}$. We then obtain a finite unramified morphism (respectively, a closed embedding),

(4.18)\begin{equation} \mathcal{M}_{K_{\widetilde H}}(\widetilde H) \longrightarrow \mathcal{M}_{K_{\widetilde G}}(\widetilde G) \quad\text{and}\quad \mathcal{M}_{K_{\widetilde H}}(\widetilde H) {\hookrightarrow} \mathcal{M}_{K_{\widetilde{HG}}}(\widetilde{HG}), \end{equation}

and Hecke correspondences exactly as in the previous subsection.

4.3 Drinfeld level at $v_0$

We continue with the setup and assumptions of § 4.2, with $v_0$ split in $F$ and possibly ramified over $p$. In this subsection we will define integral models over $O_{E,(\nu )}$ where we impose a Drinfeld level structure at $v_0$. To do this, we require that the matching condition between the CM type $\Phi$ and the chosen place $\nu$ of $E$ is satisfied, which demands that

(4.19)\begin{equation} \{\varphi\in \operatorname{Hom}(F, \overline{\mathbb{Q}}) \,|\, w_\varphi=w_0\}\subset \Phi, \end{equation}

where $w_\varphi$ is the place of $F$ induced by $\widetilde \nu \circ \varphi$, as in (4.8). We note that this condition only depends on the place $\nu$ of $E$ induced by $\widetilde \nu$. When condition (4.19) is satisfied, we also say that the CM type $\Phi$ and the place $\nu$ of $E$ are matched. Here are some examples in which the matching condition is guaranteed to hold.

Lemma 4.3 The matching condition for $\Phi$ and $\nu$ is satisfied in each of the following two situations.

  1. (i) $F$ is of the form $K F_0$ for an imaginary quadratic field $K/\mathbb{Q}$, $\Phi$ is the unique CM type induced from $K$ containing $\varphi _0$, and $p$ splits in $K$.

  2. (ii) The place $v_0$ is of degree $1$ over $\mathbb{Q}$.

Proof. The matching condition is obvious in (i), and in (ii) it holds because the left-hand side of (4.19) is the singleton set $\{\varphi _0\}$.

Remark 4.4 We call the case Lemma 4.3(i) the Harris–Taylor case (cf. [Reference Harris and TaylorHT01]).

Now let $m$ be a non-negative integer. We define the level subgroup $K_G^{m} \subset G(\mathbb{A}_{F_0,f})$ in exactly the same way as $K_G$ in § 4.1 (subject to the choice of certain vertex lattices $\Lambda _v$ for $v \in \mathcal {V}_p$), except that in the $v_0$-factor where $G(F_{0, v_0})\simeq \mathrm {GL}_n(F_{0, v_0})$, we take $K^{m}_{G, v_0}$ to be the principal congruence subgroup modulo $\mathfrak {p}_{v_0}^{m}$ inside $K_{G,v_0}$. In particular, $K_G$ coincides with $K^{m}_G$ for $m=0$. We define $K_{\widetilde G}^{m} := K_{Z^{\mathbb{Q}}} \times K_G^{m}$ as in (3.6).

We now define the moduli functor $\mathcal {M}_{K^{m}_{\widetilde G}}(\widetilde G)$ over $\mathcal {M}_{K_{\widetilde G}}(\widetilde G)$. Let $(A_0,\iota _0,\lambda _0,A,\iota ,\lambda , \overline \eta ^{p})\in \mathcal {M}_{K_{\widetilde G}}(\widetilde G)(S)$. Consider the factors occurring in the decomposition (4.5) of the $p$-divisible group $A[p^{\infty }]$,

(4.20)\begin{equation} A[v_0^{\infty}] = A[w_0^{\infty}] \times A[\overline w_0^{\infty}]. \end{equation}

When $p$ is locally nilpotent on $S$, the $p$-divisible group $A[w_0^{\infty }]$ satisfies the Kottwitz condition of type $r|_{w_0}$ for the action of $O_{F,w_0}$ on its Lie algebra, in the sense of the previous subsection. Thus, by the formulation of the Kottwitz condition in Remark 3.6(i) and the matching condition (4.19),

\[ \operatorname{char}(\iota(a)\mid\operatorname{Lie} A[w_0^{\infty}]) = \prod_{\substack{\varphi \colon \! F \rightarrow \mathbb{Q}\\ w_\varphi=w_0}} (T-\varphi(a))^{r_\varphi} = T - \varphi_0(a) \quad\text{for all}\ a \in O_{F,w_0}. \]

Hence $A[w_0^{\infty }]$ is a one-dimensional formal $O_{F,w_0}$-module. Similarly, $A[\overline w_0^{\infty }]$ has dimension $n \cdot [F_{w_0}:\mathbb{Q}_p]-1$. Since $\operatorname {Ros}_\lambda$ induces the conjugation automorphism on $O_F$, $\lambda$ furthermore identifies $A[w_0^{\infty }]$ and $A[\overline w_0^{\infty }]$ with the (absolute) duals of each other, and hence both have (absolute) height $n \cdot [F_{w_0}:\mathbb{Q}_p]$. Analogously,

\[ A_0[v_0^{\infty}] = A_0[w_0^{\infty}] \times A_0[\overline w_0^{\infty}] , \]

where, by (4.19) and the Kottwitz condition on $A_0$, the $p$-divisible group $A_0[w_0^{\infty }]$ with $O_{F,w_0}$-action is étale of height $[F_{w_0}:\mathbb{Q}_p]$, whereas $A_0[\overline w_0^{\infty }]$ is identified with the dual of $A_0[w_0^{\infty }]$.

In analogy with the prime-to-$p$ theory, we introduce the finite flat group scheme over $S$,

\[ T_{w_0}(A_0,A)[w_0^{m}] := \operatorname{\underline{Hom}}_{O_{F,w_0}}(A_0[w_0^{m}],A[w_0^{m}]). \]

Note that as $m$ varies, the right-hand side is naturally an inverse system under restriction of homomorphisms, and making it into a directed system depends on the choice of uniformizer $\pi _{w_0}$ of $F_{0, w_0}$. The colimit $T_{w_0}(A_0,A) := \varinjlim _{m} T_{w_0}(A_0,A)[w_0^{m}]$ is a one-dimensional formal $O_{F,w_0}$-module since $A[w_0^{\infty }]$ is.

For the moduli problem $\mathcal {M}_{K^{m}_{\widetilde G}}(\widetilde G)$, we equip the object $(A_0,\iota _0,\lambda _0,A,\iota ,\lambda , \overline \eta ^{p})\in \mathcal {M}_{K_{\widetilde G}}(\widetilde G)$ with the following additional datum. Let $\Lambda _{v_0} = \Lambda _{w_0} \oplus \Lambda _{\overline w_0}$ denote the natural decomposition of the lattice $\Lambda _{v_0}$ attached to the split place $v_0$. The additional datum is

  • an $O_{F,w_0}$-linear homomorphism of finite flat group schemes,

    (4.21)\begin{equation} \varphi\colon \pi_{w_0}^{-m}\Lambda_{w_0}/\Lambda_{w_0} \longrightarrow \operatorname{\underline{Hom}}_{O_{F,w_0}}(A_0[w_0^{m}],A[w_0^{m}]), \end{equation}
    which is a Drinfeld $w_0^{m}$-structure on the target (cf. [Reference Harris and TaylorHT01, § II.2]).

Theorem 4.5 The moduli problem $\mathcal {M}_{K^{m}_{\widetilde G}}(\widetilde G)$ is relatively representable by a finite flat morphism to $\mathcal {M}_{K_{\widetilde G}}(\widetilde G)$. It is regular and flat over $\operatorname {Spec} O_{E,(\nu )}$. Furthermore, the generic fiber $\mathcal {M}_{K^{m}_{\widetilde G}}(\widetilde G) \times _{\operatorname {Spec} O_{E,(\nu )}} \operatorname {Spec} E$ is canonically isomorphic to $M_{K^{m}_{\widetilde G}}(\widetilde G)$.

Proof. After an étale base change, the subgroup $\mathcal {A}_0[w_0^{m}]$ of the universal abelian scheme $\mathcal {A}_0$ over $\mathcal {M}^{\mathfrak {a},\xi }_0$ becomes constant, and then the proof of [Reference Harris and TaylorHT01, Lemma III.4.1(4)(5)] applies.

We analogously define the DM stack $\mathcal {M}_{K^{m}_{\widetilde H}}(\widetilde H)$ over $\operatorname {Spec} O_{E,(\nu )}$, and obtain for it the analog of Theorem 4.5. We also define $\mathcal {M}_{K^{m}_{\widetilde {HG}}}(\widetilde {HG}) := \mathcal {M}_{K_{\widetilde H}^{m}}(\widetilde H) \times _{\mathcal {M}^{\mathfrak {a},\xi }_0} \mathcal {M}_{K_{\widetilde G}^{m}}(\widetilde G)$, but we note that this stack is not regular for $m > 0$. We then obtain a finite unramified morphism (respectively, a closed embedding)

\[ \mathcal{M}_{K^{m}_{\widetilde H}}(\widetilde H) \longrightarrow \mathcal{M}_{K^{m}_{\widetilde G}}(\widetilde G) \quad\text{and}\quad \mathcal{M}_{K^{m}_{\widetilde H}}(\widetilde H) {\hookrightarrow} \mathcal{M}_{K^{m}_{\widetilde{HG}}}(\widetilde{HG}), \]

provided that $K_H^{p} \subset H(\mathbb{A}_{F_0,f}^{p}) \cap K_G^{p}$ and that the lattices $\Lambda _v$ and $\Lambda _v^{\flat }$ are as in (4.14). Indeed, the Drinfeld level structure $\varphi ^{\flat } \colon \pi _{w_0}^{-m}\Lambda _{w_0}^{\flat }/\Lambda _{w_0}^{\flat } \to \operatorname {\underline {Hom}}_{O_{F,w_0}}(A_0[w_0^{m}],A^{\flat }[w_0^{m}])$ induces a Drinfeld level structure

\[ \varphi\colon \pi_{w_0}^{-m}\Lambda_{w_0}/\Lambda_{w_0} \longrightarrow \operatorname{\underline{Hom}}_{O_{F,w_0}}(A_0[w_0^{m}], (A^{\flat} \times A_0)[w_0^{m}]) \]

by taking the direct sum of $\varphi ^{\flat }$ and the $O_{F,w_0}$-linear homomorphism

(4.22)

with respect to the natural decompositions in the source and target of $\varphi$.

In the present situation there are more Hecke correspondences than those defined for adelic elements prime to $p$ in the previous two subsections (cf. [Reference Harris and TaylorHT01, § III.4]). We again first treat the case for $\widetilde G$. With respect to the decomposition obtained from (2.1),

\[ \widetilde G(\mathbb{Q}_p) = Z^{\mathbb{Q}}(\mathbb{Q}_p) \times \prod_{v \in \mathcal{V}_p} G(F_{0,v}), \]

let $g \in G(F_{0,v_0})$, considered as an element on the left-hand side. In the special case where $g \in K_{G,v_0}$, since $K_{G,v_0}^{m}$ is a normal subgroup of $K_G$, we get a diagram of finite flat morphisms

analogously to (4.15). For general $g \in G(F_{0,v_0})$, choose $\mu$ large enough that

\[ p^{\mu}\Lambda_{v_0} \subset g\Lambda_{v_0} \subset \varpi_{v_0}^{-m} g \Lambda_{v_0} \subset p^{-\mu} \Lambda_{v_0}. \]

Then $K_{G,v_0}^{2\mu e} \subset K_{G,v_0}^{m} \cap gK_{G,v_0}^{m} g{{}^{-1}}$, where $e = e_{v_0}$ is the ramification index of $v_0$ over $p$. Hence we obtain a diagram of finite flat morphisms

(4.23)

as before. In terms of the moduli descriptions above, these morphisms are given as follows. Consider a point $(A_0,\iota _0,\lambda _0,A,\iota ,\lambda , \overline \eta ^{p},\varphi )$ of $\mathcal {M}_{K_{\widetilde G}^{2\mu e}}(\widetilde G)$. Then $\operatorname {nat}_1$ sends this point to the point represented by $(A_0,\iota _0,\lambda _0,A,\iota ,\lambda , \overline \eta ^{p},\varphi |_{(\pi _{w_0}^{-m}\Lambda _{w_0}/\Lambda _{w_0})})$. To describe $\operatorname {nat}_g$, let $C_{w_0} \subset A[w_0^{2\mu e}]$ be the unique closed subgroup scheme for which the set of $\varphi (x)$, $x \in p^{-\mu } g\Lambda _{w_0}/\Lambda _{w_0} \subset p^{-2\mu }\Lambda _{w_0}/\Lambda _{w_0}$, is a complete set of sections; cf. [Reference Katz and MazurKM85, Cor. 1.10.3] (note that $\operatorname {\underline {Hom}}_{O_{F,w_0}}(A_0[w_0^{2\mu e}],A[w_0^{2\mu e}])$ is étale-locally isomorphic to $A[w_0^{2\mu e}]$). Since $\lambda$ induces a principal polarization on the $p$-divisible group $A[v_0^{\infty }]$, the $\lambda$-Weil pairing on $A[p^{2\mu }]$ induces a perfect duality between $A[w_0^{2\mu e}]$ and $A[\overline w_0^{2\mu e}]$; in this way, let $C_{\overline w_0} \subset A[\overline w_0^{2\mu e}]$ be the annihilator of $C_{w_0}$. Let $C$ be the (totally isotropic) subgroup of $A[p^{2\mu }]$,

\[ C := C_{w_0} \times C_{\overline w_0} \times \prod_{v \in \mathcal{V}_p \smallsetminus \{v_0\}} A[v^{\mu e_v}], \]

where $e_v$ denotes the ramification index of $v$ over $p$. Let

\[ A' := A/C. \]

Let $\iota '$ denote the induced $O_F \otimes \mathbb{Z}_{(p)}$-action on $A'$, and let $\alpha \colon A \to A'$ be the corresponding $O_F$-linear isogeny. Then there is a unique polarization $\lambda '$ on $A'$ such that the diagram

commutes, and $\lambda '$ is a polarization of the type in the moduli problem for $\mathcal {M}_{K_{\widetilde G}}(\widetilde G)$. Furthermore, the level structure $\overline \eta ^{p}$ induces, via the quasi-isogeny $p^{-\mu }\alpha$, a level structure $\overline \eta ^{\prime p}$ for the pair $(A_0,A')$. We obtain a Drinfeld level $w_0^{m}$-structure $\varphi '$ on $\operatorname {\underline {Hom}}_{O_{F,w_0}}(A_0[w_0^{m}],A'[w_0^{m}])$ via the following commutative diagram:

Here the dashed arrows are induced by $\varphi$. Then the image of $(A_0,\iota _0,\lambda _0,A,\iota ,\lambda , \overline \eta ^{p},\varphi )$ under $\operatorname {nat}_g$ is $(A_0,\iota _0,\lambda _0,A',\iota ',\lambda ',\overline \eta ^{\prime p}, \varphi ')$. We remark that the construction of the latter tuple is, up to canonical isomorphism in the moduli problem for $\mathcal {M}_{K_{\widetilde G}^{m}}(\widetilde G)$, independent of the choice of $\mu$.

Again, as in (4.16), this defines an action of the center of $G(F_{0, v_0})$ on $\mathcal {M}_{K_{\widetilde G}^{m}}(\widetilde G)$.

The above construction carries over in the obvious way with $\widetilde H$ in place of $\widetilde G$. Taking the product over $\mathcal {M}^{\mathfrak {a},\xi }_0$ of the diagram (4.23) with the one attached to $\widetilde H$ and an element $h \in H(F_{0,v_0})$, we obtain, for $\mu$ sufficiently large, an analogous diagram of finite flat morphisms for $\widetilde {HG}$. We record this as the following diagram, where $g\in (H\times G)(F_{0,v_0})$:

(4.24)

4.4 AT parahoric level at $v_0$

In this subsection we assume that $p \neq 2$ and that the place $v_0$ is unramified over $p$. We again choose a vertex lattice $\Lambda _v \subset W_v$ for each $v \in \mathcal {V}_p$, as in § 4.1. We require that the pair $(v_0,\Lambda _{v_0})$ is of one of the following four types, which we call AT types.

  1. (1) $v_0$ is inert in $F$ and $\Lambda _{v_0}$ is almost self-dual as an $O_{F,v_0}$-lattice.

  2. (2) $v_0$ ramifies in $F$, $n$ is even, and $\Lambda _{v_0}$ is $\pi _{v_0}$-modular.

  3. (3) $v_0$ ramifies in $F$, $n$ is odd, and $\Lambda _{v_0}$ is almost $\pi _{v_0}$-modular.

  4. (4) $v_0$ ramifies in $F$, $n = 2$, and $\Lambda _{v_0}$ is self-dual.

We refer to the end of the Introduction for the terminology on lattice types.

Again recalling the decomposition $K_{\widetilde G} = K_{Z^{\mathbb{Q}}} \times K_G$ from (3.6), we take the subgroup $K_G$ to be of the form

\[ K_G = K_G^{p} \times K_{G,p}, \]

where $K_G^{p} \subset G(\mathbb{A}_{F_0,f}^{p})$ is arbitrary, and $K_{G,p} \subset G(F_0\otimes \mathbb{Q}_p)$ is given by

\[ K_{G,p}=\prod_{v \in \mathcal{V}_p} K_{G, v} \subset \prod_{v \in \mathcal{V}_p} G(F_{0,v}), \]

with $K_{G, v}$ the stabilizer of $\Lambda _v$ in $G(F_{0,v})$.

We define the moduli functor $\mathcal {M}_{K_{\widetilde G}}(\widetilde G)$ over $\operatorname {Spec} O_{E,(\nu )}$ similarly to § 4.1. To be precise, to each $O_{E,(\nu )}$-scheme $S$, we associate the groupoid of tuples $(A_0,\iota _0,\lambda _0,A,\iota ,\lambda ,\overline \eta ^{p})$, where $(A_0,\iota _0,\lambda _0)$ is an object of $\mathcal {M}^{\mathfrak {a},\xi }_0(S)$, $A$ is an abelian scheme over $S$ up to isogeny prime to $p$, $\iota$ is an $O_{F}\otimes \mathbb{Z}_{(p)}$-action on $A$ satisfying the Kottwitz condition (3.11) of signature $((1, n-1)_{\varphi _0}, (0, n)_{\varphi \in \Phi \smallsetminus \{\varphi _0\}})$, and $\lambda$ is a polarization on $A$ whose Rosati involution induces on $O_{F}\otimes \mathbb{Z}_{(p)}$ the non-trivial Galois automorphism of $F/F_0$, subject to the condition that under the decomposition (4.2) of the $p$-divisible group $A[p^{\infty }]$, $\ker \lambda _v$ is contained in $A[\iota (\pi _v)]$ of rank $\#(\Lambda _v^{*}/\Lambda _v)$ for all $v \in \mathcal {V}_p$. Finally, $\overline \eta ^{p}$ is a $K_G^{p}$-equivalence class of $\mathbb{A}_{F,f}^{p}$-linear isometries (4.3).

We also impose for each $v\neq v_0$ over $p$ the sign condition (4.4) at $v$. To ensure flatness of $\mathcal {M}_{K_{\widetilde G}}(\widetilde G)$, we impose the following additional conditions. For each $v\neq v_0$ over $p$ we impose the Eisenstein condition (4.10) at $v$. In addition, when the pair $(v_0,\Lambda _{v_0})$ is of AT type (2), (3), or (4), we impose the following conditions.Footnote 9 As in the case of the Eisenstein condition, it suffices to impose the conditions when the base scheme $S$ lies over $\operatorname {Spf} \overline{\mathbb{Z}}_p$, so that, as in (4.9), there is a decomposition

(4.25)\begin{equation} \operatorname{Lie} A[v_0^{\infty}] \cong \bigoplus_{\psi \in \operatorname{Hom}_{\mathbb{Q}_p}(F_{0,v_0}, \overline{\mathbb{Q}}_p)} \operatorname{Lie}_\psi A[v_0^{\infty}]; \end{equation}

here, in the index set for the direct sum, we have used that $F_{v_0}^{t} = F_{0,v_0}$ ($v_0$ is unramified over $p$ in this section, and in types (2)–(4) $v_0$ ramifies in $F$). For each $\psi \in \operatorname {Hom}_{\mathbb{Q}_p}(F_{0,v_0}, \overline{\mathbb{Q}}_p)$, the extensions of $\psi$ to $F_{v_0}$ form a conjugate pair

(4.26)\begin{equation} \varphi_\psi, \overline\varphi_\psi. \end{equation}

Furthermore, in terms of the identification (4.8), we have $\{r_{\varphi _\psi },r_{\overline \varphi _\psi }\} = \{0,n\}$ unless

\[ \psi = \psi_0 := \varphi_0|_{F_{v_0}}, \]

in which case $\{r_{\varphi _{\psi _0}},r_{\overline \varphi _{\psi _0}}\} = \{r_{\varphi _0},r_{\overline \varphi _0}\} = \{1,n-1\}$. For all $\psi \in \operatorname {Hom}_{\mathbb{Q}_p}(F_{0,v_0}, \overline{\mathbb{Q}}_p) \smallsetminus \{\psi _0\}$, we impose the Eisenstein condition (4.10) on the summand $\operatorname {Lie}_\psi A[v_0^{\infty }]$. This completes the list of conditions in type (4). In types (2) and (3), we impose the following further conditions on $\operatorname {Lie}_{\psi _0} A[v_0^{\infty }]$.

  • If $(v_0,\Lambda _{v_0})$ is of type (2), then we impose the wedge condition

    (4.27)\begin{equation} \bigwedge^{2} (\iota(\pi_{v_0})+\pi_{v_0} \mid \operatorname{Lie}_{ \psi_0} A[v_0^{\infty}]) = 0, \end{equation}
    and the spin condition
    (4.28)\begin{equation} \textit{the endomorphism } \iota(\pi_{v_0})\mid \operatorname{Lie}_{ \psi_0} A[v_0^{\infty}] \textit{ is non-vanishing at each point of } S \end{equation}
    (cf. [Reference Rapoport, Smithling and ZhangRSZ18, § 6]).
  • If $(v_0,\Lambda _{v_0})$ is of type (3), then we impose the refined spin condition (7.9) of [Reference Rapoport, Smithling and ZhangRSZ18] on $\operatorname {Lie}_{ \psi _0} A[v_0^{\infty }]$. More precisely, [Reference Rapoport, Smithling and ZhangRSZ18] applies as written (taking our $F_{v_0}/F_{0,v_0}$ as the local extension $F/F_0$ there) in the case where $v_0$ has degree $1$ over $p$, so that $\operatorname {Lie}_{\psi _0} A[v_0^{\infty }] = \operatorname {Lie} A[v_0^{\infty }]$. For general $v_0$ unramified over $p$, analogously to (4.25), the $\mathcal {O}_S$-module $M(A[v_0^{\infty }])$ (notation of [Reference Rapoport, Smithling and ZhangRSZ18]) decomposes under the action of $O_{F_0,v_0}$ into a direct sum $\bigoplus _{\psi } M_\psi (A[v_0^{\infty }])$, and for each $\psi$ there is an exact sequence

    \[ 0 \longrightarrow \mathrm{Fil}_\psi^{1} \longrightarrow M_\psi(A[v_0^{\infty}]) \longrightarrow \operatorname{Lie}_\psi A[v_0^{\infty}] \longrightarrow 0. \]
    Now taking $\psi = \psi _0$, the condition we impose is given by [Reference Rapoport, Smithling and ZhangRSZ18, condition (7.9)], except with $\mathrm {Fil}_{\psi _0}^{1}$ in place of $\mathrm {Fil}^{1}$, with the almost $\pi _{v_0}$-modular lattice $\Lambda _{v_0}$ in place of $\Lambda _{-m}$, and with the tensor products $- \otimes _{O_{F,v_0}} \mathcal {O}_S$ and $- \otimes _{O_{F_0,v_0}} \mathcal {O}_S$ in terms of the structure morphism
    \[ O_{F_0,v_0} \subset O_{F,v_0} {\mathop \to \limits^{\varphi _0}} \overline{\mathbb{Z}}_p \longrightarrow \mathcal{O}_S. \]
    (The same applies to the definition of $L_{-m,-1}^{n-1,1}(S)$ in [Reference Rapoport, Smithling and ZhangRSZ18, (7.6)].)

Remark 4.6 When $v_0$ is of degree $1$ over $p$, the triple $(A[v_0^{\infty }], \iota [v_0^{\infty }], \lambda [v_0^{\infty }])$ arising from our moduli problem is of the type occurring in the moduli problem for one of the RZ spaces in [Reference Rapoport, Smithling and ZhangRSZ18, § 5–8] (because in the degree-$1$ case, the relative dual of $A[v_0^{\infty }]$ which is used in [Reference Rapoport, Smithling and ZhangRSZ18] is the Serre dual).

Theorem 4.7 The moduli problem just formulated is representable by a DM stack $\mathcal {M}_{K_{\widetilde G}}(\widetilde G)$ flat over $\operatorname {Spec} O_{E,(\nu )}$. For $K_G^{p}$ small enough, $\mathcal {M}_{K_{\widetilde G}}(\widetilde G)$ is relatively representable over $\mathcal {M}^{\mathfrak {a},\xi }_0$. The generic fiber $\mathcal {M}_{K_{\widetilde G}}(\widetilde G) \times _{\operatorname {Spec} O_{E,(\nu )}} \operatorname {Spec} E$ is canonically isomorphic to $M_{K_{\widetilde G}}(\widetilde G)$. Furthermore:

  1. (i) $\mathcal {M}_{K_{\widetilde G}}(\widetilde G)$ is smooth over $\operatorname {Spec} O_{E,(\nu )}$ provided that $(v_0,\Lambda _{v_0})$ is of AT type (2) or (3).

  2. (ii) $\mathcal {M}_{K_{\widetilde G}}(\widetilde G)$ has semi-stable reduction over $\operatorname {Spec} O_{E,(\nu )}$ provided that $(v_0,\Lambda _{v_0})$ is of AT type (1) and $E_\nu$ is unramified over $\mathbb{Q}_p$.

Proof. The representability assertion and the assertion for the generic fiber are proved in the same way as in Theorem 4.1. The assertions concerning the local structure of $\mathcal {M}_{K_{\widetilde G}}(\widetilde G)$ all reduce to statements about the local model. As in the proof of Theorem 4.1, the local model is a product of local models, with one factor for $A_0$ and one for $A$, and the factor for $A_0$ is trivial. The factor for $A$ furthermore decomposes as in (4.12) into a product of local models $M_v \times _{\operatorname {Spec} O_{E_{r|_v}}} \operatorname {Spec} O_{E_\nu }$ indexed by the places $v \in \mathcal {V}_p$, and the factor for each $v \neq v_0$ is again trivial. At the place $v_0$, let $E_{r|_{v_0}}^{\mathrm{un}}$ denote the maximal unramified extension of $E_{r|_{v_0}}$ in $\overline{\mathbb{Q}}_p$, and let $O_{E_{r|_{v_0}}^{\mathrm{un}}}$ denote its ring of integers. After extending scalars $O_{E_{r|_{v_0}}} \longrightarrow O_{E_{r|_{v_0}}^{\mathrm{un}}}$, the canonical isomorphism $O_{F_0,v_0} \otimes _{\mathbb{Z}_p} O_{E_{r|_{v_0}}^{\mathrm{un}}} \cong \prod _{\psi \in \operatorname {Hom}_{\mathbb{Q}_p}(F_{0,v_0},\overline{\mathbb{Q}}_p)} O_{E_{r|_{v_0}}^{\mathrm{un}}}$ (recall that $v_0$ is unramified over $p$) induces a further decomposition of the local model

\[ M_{v_0} \times_{\operatorname{Spec} O_{E_{r|_{v_0}}}} \operatorname{Spec} O_{E_{r|_{v_0}}^{\mathrm{un}}} = \prod_{\psi \in \operatorname{Hom}_{\mathbb{Q}_p}(F_{0,v_0},\overline{\mathbb{Q}}_p)} M_{v_0,\psi} \times_{\operatorname{Spec} O_{E_{r|_{v_0,\psi}}}} \operatorname{Spec} O_{E_{r|_{v_0}}^{\mathrm{un}}}. \]

Here, in terms of the identifications (4.7) and (4.8) and the notation (4.26), $r|_{v_0,\psi }$ denotes the restriction of $r$ to the set

\[ \{\varphi \in \operatorname{Hom}_{\mathbb{Q}_p}(F_{v_0}, \overline{\mathbb{Q}}_p) \,|\, \varphi|_{F_{0,v_0}} = \psi \} = \{\varphi_\psi,\overline\varphi_\psi\}, \]

and $M_{v_0,\psi } = M(F_v/F_{0,v},r|_{v_0,\psi },\Lambda _v)$ is a local model attached to the algebraic $F_{0,v_0}$-group $\mathrm {GU}(W_{v_0})$, the signature function $r|_{v_0,\psi }$, and the $O_{F,v_0}$-lattice $\Lambda _{v_0}$. For $\psi \neq \psi _0$, the signature function $r|_{v_0,\psi }$ is banal, and $M_{v_0,\psi }$ is again trivial. For $\psi = \psi _0$, the local model $M_{v_0,\psi _0}$ is smooth over $\operatorname {Spec} O_{E_{r|_{v_0,\psi _0}}}$ in types (2) [Reference Rapoport, Smithling and ZhangRSZ17, Proposition 3.10] and (3) [Reference SmithlingSmi15, Theorem 1.4], and is flat of semi-stable reduction $\operatorname {Spec} O_{E_{r|_{v_0,\psi _0}}}$ in types (1) [Reference Rapoport, Smithling and ZhangRSZ18, proof of Theorem 5.1] and (4) [Reference Rapoport, Smithling and ZhangRSZ18, § 8, pp. 1119–20]. This completes the proof, noting in assertion (ii) that semi-stable reduction is preserved under an unramified base extension.

Remark 4.8 The unramifiedness condition on $E_\nu$ in part (ii) of Theorem 4.7 is always satisfied when $F$ contains an imaginary quadratic field $K$ and $\Phi$ is induced from $K$, since $\varphi _0\colon F {\tilde{\to}} E$ in this case (cf. Remark 3.1) and $F_{v_0}$ is unramified over $\mathbb{Q}_p$ in AT type (1).

Remark 4.9 In the case of AT type (4), in the notation of the proof of Theorem 4.7, we have $E_{r|_{v_0,\psi _0}} = F_{0,v_0}$, and the local model $M_{v_0,\psi _0}$ has semi-stable reduction over $\operatorname {Spec} O_{F_0,v_0}$ (cf. [Reference Rapoport, Smithling and ZhangRSZ18]). However, the extension $E_\nu /F_{0,v_0}$ is always ramified (because $F_{v_0}$ maps into $E_\nu$), and therefore $\mathcal {M}_{K_{\widetilde G}}(\widetilde G)$ is not regular over the place $\nu$ in this case. For this reason, we will exclude AT type (4) when considering arithmetic intersections.

We analogously define the DM stack $\mathcal {M}_{K_{\widetilde H}}(\widetilde H)$ over $\operatorname {Spec} O_{E,(\nu )}$ when the pair $(v_0,\Lambda _{v_0}^{\flat })$ is of AT type. To define $\mathcal {M}_{K_{\widetilde {HG}}}(\widetilde {HG})$, we take as before the lattice $\Lambda _v^{\flat } \oplus \Lambda _v$ in $W_v^{\flat } \oplus W$ for each $v \in \mathcal {V}_p$, but the relation between $\Lambda _v^{\flat }$ and $\Lambda _v$ that we allow can be more complicated. Furthermore, the definition of the analog of the morphisms (4.13) requires more care.

Let us first suppose that $(v_0,\Lambda _{v_0})$ is not of AT type (2). In this case, we assume that the lattices $\Lambda _v$ and $\Lambda _v^{\flat }$ satisfy $\Lambda _v = \Lambda _v^{\flat } \oplus O_{F,v}u$ for all $v$. We also assume that $(u,u) \in O_{F_0,v}^{\times }$ for all $v$, with the single exception of $v = v_0$ when $(v_0,\Lambda _{v_0})$ is of AT type (1), in which case we impose that $\Lambda _{v_0}^{\flat }$ is self-dual and $\operatorname {ord}_{v_0} (u,u) = 1$ (and of course we use the definition of $\mathcal {M}_{K_{\widetilde H}}(\widetilde H)$ from § 4.1). Provided that $K_H^{p} \subset H(\mathbb{A}_{F_0,f}^{p}) \cap K_G^{p}$, we then obtain a finite unramified morphism (respectively a closed embedding)

(4.29)\begin{equation} \mathcal{M}_{K_{\widetilde H}}(\widetilde H) \longrightarrow \mathcal{M}_{K_{\widetilde G}}(\widetilde G) \quad\text{and}\quad \mathcal{M}_{K_{\widetilde H}}(\widetilde H) {\hookrightarrow} \mathcal{M}_{K_{\widetilde{HG}}}(\widetilde{HG}), \end{equation}

as in (4.13).

Remark 4.10 The unit $(u,u)$ was chosen very carefully in [Reference Rapoport, Smithling and ZhangRSZ18] because there we made a definite choice between the two isomorphic RZ spaces $\mathcal {N}_n^{(0)}$ and $\mathcal {N}_n^{(1)}$ in the odd ramified case (more precisely, a definite choice of the framing object). Here we make no such choice, and therefore $(u,u)$ can be an arbitrary unit at ramified places.

Now suppose that $(v_0,\Lambda _{v_0})$ is of AT type (2). Then we cannot define such simple embeddings, and it is necessary to consider more complicated diagrams involving additional spaces (cf. [Reference Rapoport, Smithling and ZhangRSZ18, § 12]). In fact we will consider two variants. For both variants, we assume that $(u,u) \in O_{F_0,(p)}^{\times }$, and that $\Lambda _{v_0}^{\flat }$ and $\Lambda _{v_0}$ are related by a chain of inclusions

(4.30)\begin{equation} \pi_{v_0} (\Lambda_{v_0}^{\flat} \oplus O_{F, v_0} u)^{*} \subset^{1} \Lambda_{v_0} \subset^{1} \Lambda_{v_0}^{\flat} \oplus O_{F, v_0} u. \end{equation}

Note that (4.30) is equivalent to the condition that $\Lambda _{v_0}^{\flat }$ is almost $\pi _{v_0}$-modular in $W_{v_0}^{\flat }$, and that $\Lambda _{v_0}$ is one of the two $\pi _{v_0}$-modular lattices contained in $\Lambda _{v_0}^{\flat } \oplus O_{F, v_0} u$. For simplicity, we also assume in type (2) that $v_0$ is of degree $1$ over $p$, as we will do later in § 8 for all places of AT type in the context of arithmetic intersections.

Variant 1. In the first variant, we continue to assume that for all $v \neq v_0$, we have $\Lambda _v = \Lambda _v^{\flat } \oplus O_{F,v}u$. We define $\mathcal {P}_{\widetilde G}$ to be the moduli stack defined in the same way as $\mathcal {M}_{K_{\widetilde G}}(\widetilde G)$, except that at the place $v_0$,

  • $\ker \lambda _{v_0}$ has rank $p^{n-2}$; and

  • when $p$ is locally nilpotent on the base scheme, $\operatorname {Lie} A[v_0^{\infty }]$ satisfies condition (9.2) of [Reference Rapoport, Smithling and ZhangRSZ18].

Now let $\mathcal {L}$ denote the self-dual multichain of $O_F \otimes \mathbb{Z}_p$-lattices in $W \otimes \mathbb{Q}_p = \bigoplus _{v \in \mathcal {V}_p} W_v$ generated by $\Lambda _{v_0}$ and $\Lambda _{v_0}^{\flat } \oplus O_{F, v_0} u$ (and its dual), and by $\Lambda _v$ for all $v \neq v_0$. We define $\mathcal {P}_{\widetilde G}'$ to be the moduli stack of tuples $(A_0,\iota _0,\lambda _0,\mathbf {A},\boldsymbol {\lambda },\overline \eta ^{p})$, where $(A_0,\iota _0,\lambda _0)$ is an object of $\mathcal {M}^{\mathfrak {a},\xi }_0$, $\mathbf {A} = \{A_\Lambda \}$ is an $\mathcal {L}$-set of abelian varieties, $\boldsymbol \lambda$ is a $\mathbb{Q}$-homogeneous principal polarization of $\mathbf {A}$, and $\overline \eta ^{p}$ is a $K_G^{p}$-equivalence class of $\mathbb{A}_{F,f}^{p}$-linear isometries $\eta ^{p}\colon \widehat V^{p}(A_0,\mathbf {A}) \simeq -W \otimes _F \mathbb{A}_{F,f}^{p}$ (cf. [Reference Rapoport and ZinkRZ96, Definition 6.9]). We require that $A_\Lambda$ satisfies the Kottwitz condition (3.11) for all $\Lambda$. We further require that over a base on which $p$ is locally nilpotent, when the $v_0$-summand of $\Lambda$ is $\Lambda _{v_0}$, $\operatorname {Lie} A_\Lambda [v_0^{\infty }]$ satisfies the wedge condition (4.27) and the spin condition (4.28) above; and when the $v_0$-component of $\Lambda$ is $\Lambda _{v_0}^{\flat } \oplus O_{F, v_0} u$, $\operatorname {Lie} A_\Lambda [v_0^{\infty }]$ satisfies condition (9.2) of [Reference Rapoport, Smithling and ZhangRSZ18]. We obtain a diagram

(4.31)

Here the lower left morphism is defined in the usual way, analogously to (3.16), provided that $K_H^{p} \subset H(\mathbb{A}_{F_0,f}^{p}) \cap K_G^{p}$. It is again finite and unramified. The arrows $\pi _1$ and $\pi _2$ are induced by $\mathbf {A} \mapsto A_{(\Lambda _{v_0}^{\flat } \oplus O_{F, v_0}u) \oplus \bigoplus _{v \neq v_0} \Lambda _v}$ and $\mathbf {A} \mapsto A_{\Lambda _{v_0} \oplus \bigoplus _{v \neq v_0} \Lambda _v}$, respectively.

Lemma 4.11 The morphism $\pi _1$ is finite étale of degree $2$.

Proof. The morphism is obviously proper. It is also finite because each geometric fiber has precisely two points. This last assertion follows over the complex numbers by looking at the homology of the abelian varieties in play, and in positive characteristic by looking at their Dieudonné modules. The question of étaleness reduces to the local models, which are isomorphic by [Reference Rapoport, Smithling and ZhangRSZ18, Proposition 9.12(ii)].

The morphism $\pi _2$ is proper, and it is finite étale over the generic fiber of degree $(p^{n}-1)/(p-1)$. However, $\pi _2$ is not finite when $n \geq 4$ (cf. [Reference Rapoport, Smithling and ZhangRSZ18, Remark 9.5]).

Now define

\[ \mathcal{P}_{\widetilde{HG}} := \mathcal{M}_{K_{\widetilde H}}(\widetilde H) \times_{\mathcal{M}^{\mathfrak{a},\xi}_0} \mathcal{P}_{\widetilde G} \quad\text{and}\quad \mathcal{P}_{\widetilde{HG}}' := \mathcal{M}_{K_{\widetilde H}}(\widetilde H) \times_{\mathcal{M}^{\mathfrak{a},\xi}_0} \mathcal{P}_{\widetilde G}'. \]

Applying the functor $\mathcal {M}_{K_{\widetilde H}}(\widetilde H) \times _{\mathcal {M}^{\mathfrak {a},\xi }_0}$ to the rightmost three spaces in (4.31), we obtain a diagram

(4.32)

Here the lower left embedding is the graph of the one in (4.31), again provided that $K_H^{p} \subset H(\mathbb{A}_{F_0,f}^{p}) \cap K_G^{p}$. Of course, the oblique arrows inherit the properties of the corresponding ones in (4.31) under base change. Set

\[ \mathcal{M}_{K_{\widetilde H}'}(\widetilde H) := \mathcal{M}_{K_{\widetilde H}}(\widetilde H) \times_{\mathcal{P}_{\widetilde{HG}}} \mathcal{P}_{\widetilde{HG}}'. \]

Note that the generic fiber of $\mathcal {M}_{K_{\widetilde H}'}(\widetilde H)$ is equal to $M_{K_{\widetilde H}'}(\widetilde H)$, where $K_{\widetilde H}' = K_{Z^{\mathbb{Q}}} \times K_H^{p} \times K_{H,p}'$, with $K_{H,v}' = K_{H,v}$ at all places $v \neq v_0$, and $K_{H,v_0}'$ the simultaneous stabilizer of $\Lambda _{v_0}^{\flat }$ and $\Lambda _{v_0}$.

Lemma 4.12 The morphism

\[ \mathcal{M}_{K_{\widetilde H}'}(\widetilde H) \longrightarrow \mathcal{M}_{K_{\widetilde{HG}}}(\widetilde{HG}) \]

induced by (4.32) is a closed embedding.

Proof. The proof of [Reference Rapoport, Smithling and ZhangRSZ18, Proposition 12.1] applies.

Variant 2. For the second variant, in addition to the place $v_0$, we allow there to be places $v_1,\ldots ,v_{m-1} \in \mathcal {V}_p$ for which the lattice $\Lambda _{v_i}$ is $\pi _{v_i}$-modular. For each $i = 0,\ldots ,m-1$, we then assume that the relation (4.30) holds with $v_i$ in place of $v_0$. For all $v \neq v_0,\ldots ,v_{m-1}$, we again assume that $\Lambda _v = \Lambda _v^{\flat } \oplus O_{F,v} u$. We then define the stack $\mathcal {P}_{\widetilde G}$ exactly as above. We also define $\mathcal {P}_{\widetilde G}'$ exactly as above, except we now take $\mathcal {L}$ to be the self-dual multichain of $O_F \otimes \mathbb{Z}_p$-lattices in $W \otimes \mathbb{Q}_p = \bigoplus _{v \in \mathcal {V}_p} W_v$ generated by the lattices $\Lambda _{v_i}$ and $\Lambda _{v_i}^{\flat } \oplus O_{F_{v_i}}u$ for each $i = 0,\ldots ,m-1$, and by $\Lambda _v$ for all $v \neq v_0,\ldots ,v_{m-1}$. (To be clear, the conditions above on the Lie algebra of the $p$-divisible group when $p$ is locally nilpotent on the base still only involve the place $v_0$.)

In complete analogy with (4.31), there is a diagram

(4.33)

where the lower left morphism is defined provided that $K_H^{p} \subset H(\mathbb{A}_{F_0,f}^{p}) \cap K_G^{p}$, and where the arrows $\pi _1$ and $\pi _2$ are induced by

\[ \mathbf{A} \longmapsto A_{\Lambda_{v_1}'\oplus \cdots \oplus \Lambda_{v_m}' \oplus \bigoplus_{v \neq v_1,\ldots,v_m} \Lambda_v} \quad\text{and}\quad \mathbf{A} \longmapsto A_{\Lambda_{v_1}\oplus \cdots \oplus \Lambda_{v_m} \oplus \bigoplus_{v \neq v_1,\ldots,v_m} \Lambda_v}, \]

respectively. The proof of Lemma 4.11 transposes to yield the following lemma.

Lemma 4.13 The morphism $\pi _1$ is finite étale of degree $2^{m}$.

In analogy with Variant 1, the morphism $\pi _2$ is proper, and finite étale over the generic fiber of degree $((p^{n}-1)/(p-1))^{m}$. However, it is again not finite when $n \geq 4$.

Finally, we again define

\[ \mathcal{P}_{\widetilde{HG}} := \mathcal{M}_{K_{\widetilde H}}(\widetilde H) \times_{\mathcal{M}^{\mathfrak{a},\xi}_0} \mathcal{P}_{\widetilde G} \quad\text{and}\quad \mathcal{P}_{\widetilde{HG}}' := \mathcal{M}_{K_{\widetilde H}}(\widetilde H) \times_{\mathcal{M}^{\mathfrak{a},\xi}_0} \mathcal{P}_{\widetilde G}', \]

and we apply the functor $\mathcal {M}_{K_{\widetilde H}}(\widetilde H) \times _{\mathcal {M}^{\mathfrak {a},\xi }_0} \!-$ to obtain

(4.34)

Here for the lower left embedding we assume, as always, that $K_H^{p} \subset H(\mathbb{A}_{F_0,f}^{p}) \cap K_G^{p}$. Set

\[ \mathcal{M}_{K_{\widetilde H}'}(\widetilde H) := \mathcal{M}_{K_{\widetilde H}}(\widetilde H) \times_{\mathcal{P}_{\widetilde{HG}}} \mathcal{P}_{\widetilde{HG}}'. \]

Note that the generic fiber of $\mathcal {M}_{K_{\widetilde H}'}(\widetilde H)$ is equal to $M_{K_{\widetilde H}'}(\widetilde H)$, where $K_{\widetilde H}' = K_{Z^{\mathbb{Q}}} \times K_H^{p} \times K_{H,p}'$, with $K_{H,v}' = K_{H,v}$ at all places $v \neq v_0,\ldots ,v_{m-1}$, and $K_{H,v_i}'$ the simultaneous stabilizer of $\Lambda _{v_i}^{\flat }$ and $\Lambda _{v_i}$ at all places $v_i$ for $i = 0,\ldots ,m-1$. As in the case of Lemma 4.12, we obtain the following result.

Lemma 4.14 The morphism

\[ \mathcal{M}_{K_{\widetilde H}'}(\widetilde H) \longrightarrow \mathcal{M}_{K_{\widetilde{HG}}}(\widetilde{HG}) \]

induced by (4.34) is a closed embedding.

We note that if $(v_0,\Lambda _{v_0})$ is of type (2) or (3), then the spaces $\mathcal {M}_{K_{\widetilde H}}(\widetilde H)$ and $\mathcal {M}_{K_{\widetilde {HG}}}(\widetilde {HG})$ are smooth. If $(v_0,\Lambda _{v_0})$ is of type (1), then $\mathcal {M}_{K_{\widetilde H}}(\widetilde H)$ is smooth, and $\mathcal {M}_{K_{\widetilde {HG}}}(\widetilde {HG})$ has semi-stable reduction provided that $E_\nu$ is unramified over $\mathbb{Q}_p$.

In these cases, one can define Hecke correspondences prime to $p$, as at the end of § 4.1.

Remark 4.15 In analogy with the above cases in which $(v_0,\Lambda _{v_0})$ is of AT type (2) and $(v_0,\Lambda _{v_0}^{\flat })$ is of AT type (3), one may also consider a situation in which $(v_0,\Lambda _{v_0})$ is hyperspecial and $(v_0,\Lambda _{v_0}^{\flat })$ is of AT type (1). One obtains a closed embedding analogous to the one in Lemma 4.12, where the source is a finite covering of $\mathcal {M}_{K_{\widetilde H}}(\widetilde H)$.

5. Global integral models

In this section we define integral models of the above moduli spaces over $\operatorname {Spec} O_E$. We will take $\mathfrak {a}=O_{F_0}$, that is, we will assume that $\mathcal {M}_0=\mathcal {M}_0^{O_{F_0}}$ is non-empty. Recall from Remark 3.5(ii) that this hypothesis is satisfied whenever $F/F_0$ is ramified at some finite place, a condition which we will eventually impose below in the context of arithmetic intersections (cf. Remark 5.3). We fix $\xi \in \mathcal {L}_\Phi ^{\mathfrak{a}}/{\sim }$ and set $\mathcal {M}_0^{\xi } := \mathcal {M}_0^{O_{F_0},\xi }$.

5.1 Trivial level structure

In this subsection we will define integral models over $\operatorname {Spec} O_E$ of the previously defined moduli spaces in the case where the open compact subgroup is the stabilizer of a lattice of a certain form. Let us start with the case of $\widetilde G$. We consider the following finite set of finite places of $F_0$,

(5.1)\begin{equation} \mathcal{V}_\mathrm{AT}^{W} := \{v \mid v \text{ is either inert in } F \text{ and } W_v \text{ is non-split, or } v \text{ ramifies in } F\}. \end{equation}

Let

\[ \mathfrak{d}_\mathrm{AT}^{W} := \prod_{v\in \mathcal{V}_\mathrm{AT}^{W}} \mathfrak{q}_v \subset O_F, \]

where $\mathfrak {q}_v$ denotes the (unique) prime in $O_F$ determined by $v \in \mathcal {V}_\mathrm {AT}^{W}$. We fix an $O_F$-lattice $\Lambda$ in $W$ with

(5.2)\begin{equation} \Lambda\subset \Lambda^{*}\subset (\mathfrak{d}_\mathrm{AT}^{W}){{}^{-1}}\Lambda. \end{equation}

We assume that the triple $(F/F_0, W, \Lambda )$ satisfies the following conditions.

  1. (1) All finite places $v$ of $F_0$ ramified over $\mathbb{Q}$ or dividing $2$ are split in $F$.Footnote 10

  2. (2) All places $v\in \mathcal {V}_\mathrm {AT}^{W}$ are unramified over $\mathbb{Q}$, and the pair $(v, \Lambda _v)$ is isomorphic to one of the AT types (1)–(4) in § 4.4.

As a consequence, for any finite place $\nu$ of $E$, denoting by $v_0$ the place of $F_0$ induced by $\nu$ via $\varphi _0$, the pair $(v_0,\Lambda _{v_0})$ is of the type considered in one of the four subsections of § 4. Associated to these data is the open compact subgroup

\[ K_G^{\circ} := \{g\in G(\mathbb{A}_{F_0, f})\,|\, g(\Lambda\otimes_{O_F}\widehat O_F)=\Lambda\otimes_{O_F}\widehat O_F \}, \]

and as usual we define $K_{\widetilde G}^{\circ } := K_{Z^{\mathbb{Q}}} \times K_G^{\circ }$.

We formulate a moduli problem over $\operatorname {Spec} O_{E}$ as follows. To each $O_{E}$-scheme $S$, we associate the groupoid of tuples $(A_0,\iota _0,\lambda _0,A,\iota ,\lambda )$, where $(A_0,\iota _0,\lambda _0)$ is an object of $\mathcal {M}_0^{\xi }(S)$. Furthermore,

  • $(A,\iota )$ is an abelian scheme over $S$, with $O_{F}$-action $\iota$ satisfying the Kottwitz condition (3.11) of signature $((1, n-1)_{\varphi _0}, (0, n)_{\varphi \in \Phi \smallsetminus \{\varphi _0\}})$; and

  • $\lambda$ is a polarization whose Rosati involution induces on $O_{F}$ the non-trivial Galois automorphism of $F/F_0$.

We impose the sign condition that at every point $s$ of $S$,

(5.3)\begin{equation} {{\mathrm{inv}}}^{r}_v(A_{0,s}, \iota_{0,s}, \lambda_{0,s}, A_s, \iota_s, \lambda_s) = {{\mathrm{inv}}}_v(-W_v) , \end{equation}

for every finite place $v$ of $F_0$ which is non-split in $F$. Furthermore, we impose that for any finite place $\nu$ of $E$, denoting by $p$ its residue characteristic, the triple up to isogeny prime to $p$ over $S \times _{\operatorname {Spec} O_E} \operatorname {Spec} O_{E,(\nu )}$,

\[ (A\otimes \mathbb{Z}_{(p)}, \iota \otimes \mathbb{Z}_{(p)}, \lambda \otimes \mathbb{Z}_{(p)}), \]

satisfies the conditions in the semi-global moduli problem for $\nu$ defined in § 4.

The morphisms in this category are the isomorphisms.

Remark 5.1 (i) Suppose that $n$ is even. Then when $v$ is a finite place of $F_0$ inert (respectively, ramified) in $F$, the two isometry types of the $n$-dimensional $F_v/F_{0,v}$-hermitian spaces are distinguished by whether they contain a self-dual (respectively, $\pi _v$-modular) lattice. This implies that when there are no places $v$ such that the pair $(v,\Lambda _v)$ above is of AT type (4), the sign condition (5.3) is automatically satisfied.

(ii) Suppose that $F_0 = \mathbb{Q}$. Then there is no need for the sign condition. Indeed, by the Hasse principle for hermitian forms, it is equivalent to impose the condition that for every geometric point $\overline s$ of $S$, there exists an isomorphism of hermitian $O_{F,\ell }$-lattices

\[ \operatorname{Hom}_{O_F}(T_\ell(A_{0,\overline s}),T_\ell(A_{\overline s})) \simeq -\Lambda_\ell \]

for every prime number $\ell \neq \operatorname {char} \kappa (\overline s)$. Hence we recover in this case the definition of the integral moduli problem of [Reference Bruinier, Howard, Kudla, Rapoport and YangBHKR+20, § 2.3] (where the principal polarization in the moduli problem of [Reference Bruinier, Howard, Kudla, Rapoport and YangBHKR+20] is replaced by the polarization type we have specified above).

The only point requiring proof in the next theorem is the representability of $\mathcal {M}_{K^{\circ }_{\widetilde G}}(\widetilde G)$, which is, however, routine.

Theorem 5.2 The moduli problem just formulated is representable by a DM stack $\mathcal {M}_{K^{\circ }_{\widetilde G}}(\widetilde G)$ flat over $\operatorname {Spec} O_{E}$. For every place $\nu$ of $E$, the base change $\mathcal {M}_{K^{\circ }_{\widetilde G}} \times _{\operatorname {Spec} O_{E}} \operatorname {Spec} O_{E, (\nu )}$ is canonically isomorphic to the semi-global moduli space defined in one of §§ 4.1, 4.2, or 4.4 above. Hence:

  1. (i) $\mathcal {M}_{K^{\circ }_{\widetilde G}}(\widetilde G)$ is smooth of relative dimension $n-1$ over the open subscheme of $\operatorname {Spec} O_E$ obtained by removing all places $\nu$ for which the induced pair $(v_0,\Lambda _{v_0})$ is of AT type (1) or (4) in § 4.4.

  2. (ii) $\mathcal {M}_{K^{\circ }_{\widetilde G}}(\widetilde G)$ has semi-stable reduction over the open subscheme of $\operatorname {Spec} O_E$ obtained by removing all places $\nu$ for which $(v_0,\Lambda _{v_0})$ is either of AT type (4) or of AT type (1) and for which $E_\nu$ is ramified over $\mathbb{Q}_p$.

Replacing $W$ by $W^{\flat }$ and choosing a lattice $\Lambda ^{\flat } \subset W^{\flat }$ analogously to above, we define the DM stack $\mathcal {M}_{K_{\widetilde H}^{\circ }}(\widetilde H)$, where $K_{\widetilde H}^{\circ } := K_{Z^{\mathbb{Q}}} \times K_H^{\circ }$ with $K_H^{\circ } \subset H(\mathbb{A}_{F_0,f})$ the stabilizer of $\Lambda ^{\flat } \otimes _{O_F} \widehat O_F$. Similarly, we define $\mathcal {M}_{K_{\widetilde {HG}}^{\circ }}(\widetilde {HG})$, where $K_{\widetilde {HG}}^{\circ } := K_{Z^{\mathbb{Q}}} \times K_H^{\circ } \times K_G^{\circ }$, and where we impose the following additional conditions on $\Lambda ^{\flat }$ and $\Lambda$. We first require that $\Lambda _v^{\flat }$ is self-dual for all $v$ which are split or inert in $F$ (i.e. we require that $\mathcal {V}_\mathrm {AT}^{W^{\flat }}$ consists of exactly the finite places of $F_0$ which ramify in $F$), and that $\Lambda _v^{\flat }$ and $\Lambda _v$ are $\pi _v$-modular or almost $\pi _v$-modular for all $v$ which ramify in $F$ (i.e. $\mathcal {V}_\mathrm {AT}^{W^{\flat }}$ and $\mathcal {V}_\mathrm {AT}^{W}$ contain no $v$ for which $(v,\Lambda _v^{\flat })$ or $(v,\Lambda _v)$ is of AT type (4)).

Remark 5.3 The conditions that we have just imposed on $\Lambda ^{\flat }$ place non-trivial constraints on the extension $F/F_0$ and on the hermitian space $W^{\flat }$. Let $d := [F_0 : \mathbb{Q}]$.

First consider the case when $n=2m+1$ is odd. Then our assumptions on $\Lambda ^{\flat }$ imply that $W^{\flat }$ is split at all finite places of $F_0$. On the other hand, at each archimedean place $\varphi$, the Hasse invariant ${{\mathrm {inv}}}_\varphi (W_\varphi ^{\flat })$ is equal to $(-1)^{m-1}$ if $\varphi =\varphi _0$, and to $(-1)^{m}$ if $\varphi \neq \varphi _0$. Hence the product formula (1.5) imposes the congruence

(5.4)\begin{equation} dm\equiv 1 \bmod 2. \end{equation}

In particular, since this requires $d$ to be odd, the extension $F/F_0$ is forced to be ramified at at least one finite place (otherwise the product formula for the norm residue symbol $(-1,F/F_0)$ would fail). On the other hand, if the congruence (5.4) is satisfied, then the hermitian space $W^{\flat }$ will exist and admit a lattice $\Lambda ^{\flat }$ as above.

Now consider the case where $n=2m$ is even. If $F/F_0$ is ramified at some finite place, then a hermitian space $W^{\flat }$ admitting a lattice $\Lambda ^{\flat }$ as above will exist for any $d$ and $m$. On the other hand, we claim that $F/F_0$ being everywhere unramified is again disallowed. Indeed, the same argument applies: our assumptions on $\Lambda ^{\flat }$ would again imply that $W^{\flat }$ is split at all finite places of $F_0$, and the Hasse invariants at the infinite places are again given by $(-1)^{m-1}$ at $\varphi _0$ and by $(-1)^{m}$ at each $\varphi \neq \varphi _0$. Hence we again obtain the congruence $dm \equiv 1 \bmod 2$, forcing $d$ to be odd.

In future work, we plan to handle more cases of an AT conjecture which will allow us to weaken the constraints we have placed above on the extension $F/F_0$ and on the lattices $\Lambda$ and $\Lambda ^{\flat }$; this will allow for more general hermitian spaces $W$ and $W^{\flat }$. More precisely, we expect to allow places $v$ which are inert in $F$ and unramified over $\mathbb{Q}$ such that $\Lambda _v$ is self-dual and $\Lambda _v^{\flat }$ is almost self-dual (cf. Remark 4.15).

Let us continue with the conditions we impose on $\Lambda ^{\flat }$ and $\Lambda$. When $n$ is odd, we require that

\[ \Lambda = \Lambda^{\flat} \oplus O_Fu, \]

and that $(u,u)$ is a unit at each finite place $v$ unless $v$ is inert in $F$ and $W_v$ is non-split, in which case $\operatorname {ord}_v(u,u) = 1$ (and hence $(v,\Lambda _v)$ is of AT type (1)). In this case, we have closed embeddings

(5.5)\begin{equation} \mathcal{M}_{K_{\widetilde H}^{\circ}}(\widetilde H) {\hookrightarrow} \mathcal{M}_{K_{\widetilde G}^{\circ}}(\widetilde G) \quad\text{and}\quad \mathcal{M}_{K_{\widetilde H}^{\circ}}(\widetilde H) {\hookrightarrow} \mathcal{M}_{K_{\widetilde{HG}}^{\circ}}(\widetilde{HG}) \end{equation}

completely analogous to those we have considered before (e.g. (4.13), (4.18), and (4.29)).

Now suppose that $n$ is even, and let $v_1,\ldots ,v_m$ be the finite places of $F_0$ which ramify in $F$ (cf. Remark 5.3). As in the discussion in § 4.4 after Remark 4.10, we assume for simplicity that each $v_i$ is of degree $1$ over $\mathbb{Q}$. By our assumptions already made, each $(v_i,\Lambda _{v_i}^{\flat })$ is of AT type (3). For each $i$, we further require that $(u,u)$ is a unit at $v_i$, and that $\Lambda _{v_i}$ is one of the two lattices for which the relation (4.30) holds with $v_i$ in place of $v_0$. Then $(v_i,\Lambda _{v_i})$ is indeed of AT type (2). At the split and inert places $v$, we again require that $\Lambda _v = \Lambda _v^{\flat } \oplus O_{F,v}u$, where $(u,u)$ is a unit at $v$ unless $v$ is inert in $F$ and $W_v$ is non-split, in which case $\operatorname {ord}_v(u,u) = 1$. Then we obtain natural global analogs of the diagrams (4.33) and (4.34),

and

Here the global analogs $\mathcal {R}_{\widetilde G}$, $\mathcal {R}_{\widetilde G}'$, $\mathcal {R}_{\widetilde {HG}}$, and $\mathcal {R}_{\widetilde {HG}}'$ of the auxiliary spaces appearing in (4.33) and (4.34) are defined in the obvious way. Similarly, we obtain a closed embedding

(5.6)\begin{equation} \mathcal{M}_{K_{\widetilde H}^{\circ\prime}}(\widetilde H) := \mathcal{M}_{K_{\widetilde H}^{\circ}}(\widetilde H) \times_{\mathcal{R}_{\widetilde{HG}}} \mathcal{R}_{\widetilde{HG}}' {\hookrightarrow} \mathcal{M}_{K_{\widetilde{HG}}^{\circ}}(\widetilde{HG}). \end{equation}

We note that $\mathcal {M}_{K_{\widetilde H}}(\widetilde H)$ and $\mathcal {M}_{K_{\widetilde {HG}}}(\widetilde {HG})$ are smooth over the open subscheme of $\operatorname {Spec} O_E$ obtained by removing all places $\nu$ for which the induced pair $(v_0,\Lambda _{v_0})$ is of AT type (1). Furthermore, $\mathcal {M}_{K^{\circ }_{\widetilde {HG}}}(\widetilde {HG})$ has semi-stable reduction over the open subscheme of $\operatorname {Spec} O_E$ obtained by removing all places $\nu$ for which $E_\nu$ is ramified over $\mathbb{Q}_p$.

5.2 Drinfeld level structure

We continue with the setup at the beginning of the previous subsection. In particular, we have the lattice $\Lambda \subset W$ satisfying relation (5.2) and the assumptions that follow it. Let

(5.7)\begin{equation} \begin{aligned} \Sigma^{\mathrm{spl}} & := \{\text{places } v \text{ of } F_0\mid v \text{ splits in } F\},\\ \Sigma^{\mathrm{spl},\Phi} & := \{v\in \Sigma^{\mathrm{spl}}\mid \text{every place } \nu \hbox{ of } E \hbox{ above } v \hbox{ matches } \Phi\} \end{aligned} \end{equation}

(cf. (4.19)). In addition, we fix a function

(5.8)\begin{equation} \mathbf{m}\colon \Sigma^{\mathrm{spl}}\longrightarrow \mathbb{Z}_{\geq 0} \end{equation}

with finite support contained in $\Sigma ^{\mathrm {spl},\Phi }$.

Associated to these data is the open compact subgroup

\[ K^{\mathbf{m}}_G := \{g\in G(\mathbb{A}_{F_0, f})\,|\, g(\Lambda\otimes_{O_F}\widehat O_F)=\Lambda\otimes_{O_F}\widehat O_F \text{ and } g\equiv \mathrm{id} \bmod N(\mathbf{m}) \}, \]

where

\[ N(\mathbf{m}) := \prod_{v \in\Sigma^{\mathrm{spl}}}\mathfrak{p}_v^{\mathbf{m} (v)}. \]

As usual, we define $K_{\widetilde G}^{\mathbf{m}} := K_{Z^{\mathbb{Q}}} \times K_G^{\mathbf{m}}$ as in (3.6). Note that if $\mathbf {m}=0$, then $K^{\mathbf{m}}_G=K^{\circ }_{G}$ and $K_{\widetilde G}^{\mathbf{m}} =K^{\circ }_{\widetilde G}$.

The subgroup $K_{\widetilde G}^{\mathbf{m}}$ defines a moduli stack $M_{K_{\widetilde G}^{\mathbf{m}}}(\widetilde G)$ as in § 3.2, which maps via a finite flat morphism to $M_{K_{\widetilde G}^{\circ }}(\widetilde G)$. We then define $\mathcal {M}_{K_{\widetilde G}^{\mathbf{m}}}(\widetilde G)$ to be the normalization of $\mathcal {M}_{K_{\widetilde G}^{\circ }}(\widetilde G)$ in $M_{K_{\widetilde G}^{\mathbf{m}}}(\widetilde G)$.

As with Theorem 5.2, the proof of the following theorem is routine.

Theorem 5.4 $\mathcal {M}_{K^{\mathbf{m}}_{\widetilde G}}(\widetilde G)$ is a regular DM stack finite and flat over $\mathcal {M}_{K^{\circ }_{\widetilde G}}(\widetilde G)$. For $N(\mathbf {m})$ big enough, $\mathcal {M}_{K^{\mathbf{m}}_{\widetilde G}}(\widetilde G)$ is relatively representable over $\mathcal {M}_0^{\xi }$. For every finite place $\nu$ of $E$, the base change $\mathcal {M}_{K^{\mathbf{m}}_{\widetilde G}} \times _{\operatorname {Spec} O_{E}} \operatorname {Spec} O_{E, (\nu )}$ is canonically isomorphic to (variants of) the semi-global moduli space defined in one of §§ 4.1, 4.3, or 4.4 above.

Note that in the last assertion of this theorem, in the case where $\mathbf {m}$ is not identically zero, we are implicitly allowing smaller open compact subgroups of $K_p$ in these subsections (this is what is meant by ‘variants’). We analogously define the DM stacks $\mathcal {M}_{K_{\widetilde H}^{\mathbf{m}}}(\widetilde H)$ and $\mathcal {M}_{K_{\widetilde {HG}}^{\mathbf{m}}}(\widetilde {HG})$ over $\operatorname {Spec} O_E$.

To define embeddings between the stacks we have introduced, we make the same assumptions on the lattices, on $(u,u)$, and on the ramified places of $F_0$ when $n$ is even as after Theorem 5.2 in § 5.1. When $n$ is odd, we analogously obtain closed embeddings

(5.9)\begin{equation} \mathcal{M}_{K_{\widetilde H}^{\mathbf{m}}}(\widetilde H) {\hookrightarrow} \widetilde{\mathcal{M}}_{K_{\widetilde G}^{\mathbf{m}}}(\widetilde G) \quad\text{and}\quad \mathcal{M}_{K_{\widetilde H}^{\mathbf{m}}}(\widetilde H) {\hookrightarrow} \widetilde{\mathcal{M}}_{K_{\widetilde{HG}}^{\mathbf{m}}}(\widetilde{HG}). \end{equation}

When $n$ is even, we analogously obtain a closed embedding

(5.10)\begin{equation} \mathcal{M}_{K_{\widetilde H}^{\prime\mathbf{m}}}(\widetilde H) {\hookrightarrow} \mathcal{M}_{K_{\widetilde{HG}}^{\mathbf{m}}}(\widetilde{HG}). \end{equation}

Also, in all cases we obtain Hecke correspondences for elements $g\in \prod _{v\in \Sigma ^{\mathrm {spl},\Phi }}'(H\times G)(F_{0,v})$ (restricted direct product),

(5.11)

6. The arithmetic Gan–Gross–Prasad conjecture

In this section we state a version of the arithmetic Gan–Gross–Prasad conjecture [Reference Gan, Gross and PrasadGGP12]. It is based on some wide open standard conjectures about algebraic cycles.

6.1 Standard conjectures on height pairing

Consider the category $\mathscr {V}$ of smooth proper varieties over a number field $E$, and let $H^{*}\colon \mathscr {V}\to \mathrm {grVec}_K$ be a Weil cohomology theory with coefficients in a field $K$ of characteristic zero. Let $X$ be an object in $\mathscr {V}$, and let ${{\mathrm {Ch}}}^{i}(X)$ be the group of codimension-$i$ algebraic cycles in $X$ modulo rational equivalence. We have a cycle class map

\[ {{\mathrm{cl}}}_i\colon {{\mathrm{Ch}}}^{i}(X)_\mathbb{Q}\longrightarrow H^{2i}(X). \]

Its kernel is the group of cohomologically trivial cycles, denoted by ${{\mathrm {Ch}}}^{i}(X)_{\mathbb{Q},0}$. We take the Weil cohomology theory $H^{*}$ as either the Betti cohomology $H^{*}(X(\mathbb{C}),\mathbb{Q})$, or, for a prime $\ell$, the $\ell$-adic cohomology $H^{*}(X\otimes _E \overline E,\mathbb{Q}_\ell )$, endowed with its continuous $\operatorname {Gal}(\overline E/E)$-action. Comparison theorems between Betti cohomology and étale cohomology show that the subspace ${{\mathrm {Ch}}}^{i}(X)_{\mathbb{Q},0}$ is independent of the choice of these two.

We will base ourselves on the following conjectures of Beilinson and Bloch (cf. [Reference JannsenJan90, § 2]).

Conjecture 6.1 There exists a regular proper flat model $\mathcal {X}$ of $X$ over $\operatorname {Spec} O_E$.

Let $\mathcal {X}$ be such a model, and consider its $i$th Chow group ${{\mathrm {Ch}}}^{i}(\mathcal {X})_\mathbb{Q}$. Restriction to the generic fiber defines a map

\[ {{\mathrm{Ch}}}^{i}(\mathcal{X})_{\mathbb{Q}}\longrightarrow {{\mathrm{Ch}}}^{i}(X)_{\mathbb{Q}}. \]

Let ${{\mathrm {Ch}}}^{i}_\mathrm {fin}(\mathcal {X})_\mathbb{Q}$ be the kernel of this map (cycles supported on ‘finite fibers’), and let ${{\mathrm {Ch}}}^{i}(\mathcal {X})_{\mathbb{Q},0}$ be the pre-image of ${{\mathrm {Ch}}}^{i}(X)_{\mathbb{Q},0}$. We will use the Arakelov pairing defined by Gillet and Soulé [Reference Gillet and SouléGS90, § 4.2.10],

(6.1)\begin{equation} {(\text{,})}_{\mathrm{GS}}\colon {{\mathrm{Ch}}}^{i}(\mathcal{X})_{\mathbb{Q},0}\times {{\mathrm{Ch}}}^{d+1-i}(\mathcal{X})_{\mathbb{Q},0}\longrightarrow \mathbb{R},\quad d:=\dim X. \end{equation}

Let ${{\mathrm {Ch}}}^{d+1-i}_\mathrm {fin}(\mathcal {X})_\mathbb{Q}^{\perp }\subset {{\mathrm {Ch}}}^{i}(\mathcal {X})_{\mathbb{Q},0}$ be the orthogonal complement of ${{\mathrm {Ch}}}^{d+1-i}_\mathrm {fin}(\mathcal {X})_\mathbb{Q}$ under the pairing (6.1).

Conjecture 6.2 The natural map ${{\mathrm {Ch}}}^{d+1-i}_\mathrm {fin}(\mathcal {X})_\mathbb{Q}^{\perp }\to {{\mathrm {Ch}}}^{i}(X)_{\mathbb{Q},0}$ is surjective.

Assuming Conjectures 6.1 and 6.2, the height pairing of Beilinson and Bloch

(6.2)\begin{equation} {(\text{,})}_\mathrm{BB}\colon {{\mathrm{Ch}}}^{i}(X)_{\mathbb{Q},0}\times {{\mathrm{Ch}}}^{d+1-i}(X)_{\mathbb{Q},0}\longrightarrow \mathbb{R} \end{equation}

can be defined as follows. Lift the elements $c_1\in {{\mathrm {Ch}}}^{i}(X)_{\mathbb{Q},0}$ and $c_2\in {{\mathrm {Ch}}}^{d+1-i}(X)_{\mathbb{Q},0}$ to $\widetilde c_1\in {{\mathrm {Ch}}}^{d+1-i}_\mathrm {fin}(\mathcal {X})_\mathbb{Q}^{\perp }$ and $\widetilde c_2\in {{\mathrm {Ch}}}^{i}_\mathrm {fin}(\mathcal {X})_\mathbb{Q}^{\perp }$, respectively. Define

(6.3)\begin{equation} (c_1,c_2)_{\mathrm{BB}} :=(\widetilde c_1,\widetilde c_2)_{\mathrm{GS}}. \end{equation}

It is easy to see that this is independent of the choices of the liftings.

Remark 6.3 Assuming Conjectures 6.1 and 6.2, the pairing (6.2) is independent of the choice of the (regular proper flat) integral model $\mathcal {X}$ (cf. [Reference KünnemannKün01, Lemma 1.5]).

Remark 6.4 Assume that there exists a smooth proper model $\mathcal {X}$ of $X$ over $O_E$. Then by [Reference KünnemannKün98, Theorem 6.11], Conjecture 6.2 holds for $\mathcal {X}$ and therefore the intersection product (6.2) is defined; again, by [Reference KünnemannKün01, Lemma 1.5], the intersection product is independent of the choice of $\mathcal {X}$ (assumed to be smooth and proper).

6.2 Cohomology and Hecke–Künneth projectors

We apply these considerations to the Shimura varieties defined in § 3.Footnote 11 In order to simplify notation, we write $K$ for $K_{\widetilde {HG}}$ in $\mathrm {Sh}_{K_{\widetilde {HG}}}(\widetilde {HG})$ throughout the rest of this section.

Denote by $\mathscr {H}_K$ the Hecke algebra of bi-$K$-invariant $\mathbb{Q}$-valued functions with compact support, with multiplication given by the convolution product,

(6.4)\begin{equation} \mathscr{H}_K := C_c^{\infty}(\widetilde{HG}(\mathbb{A}_f)/\!\!/K, \mathbb{Q}). \end{equation}

The variety $\mathrm {Sh}_K(\widetilde {HG})$ is equipped with a collection of algebraic correspondences ($\mathbb{Q}$-linear combinations of algebraic cycles on the self-product of $\mathrm {Sh}(\widetilde {HG})_K$ with itself), the Hecke correspondences associated to $f\in \mathscr {H}_K$.

The Hodge conjecture implies that the Künneth projector from $\bigoplus _{i\in \mathbb{Z}}H^{i}(\mathrm {Sh}_K(\widetilde {HG}),\mathbb{Q})$ to each summand $H^{i}(\mathrm {Sh}_K(\widetilde {HG}),\mathbb{Q})$, or to the primitive cohomology, is induced by algebraic cycles (with $\mathbb{Q}$-linear combinations). Morel and Suh [Reference Morel and SuhMS19] prove a partial result on the algebraicity of Künneth projectors for Shimura varieties (the so-called ‘standard sign conjecture’), conditional on Arthur's conjecture. We now recall their theorem.

Remark 6.5 The hypotheses on which [Reference Morel and SuhMS19] is based are in fact known in our setting. (1) The stabilization of the twisted trace formula is known from the work of Waldspurger. (2) The Arthur conjecture on the expression of the discrete spectrum in terms of discrete Arthur parameters is known (cf. Mok [Reference MokMok15] for quasi-split unitary groups, and Kaletha, Minguez, Shin, and White [Reference Kaletha, Minguez, Shin and WhiteKMSW14] (and its sequels) for inner forms of unitary groups). Here we note that our group $\widetilde {HG}$ is a product of a unitary group and a torus by (2.1). (3) The comparison between the Adams–Johnson classification and the Arthur classification of cohomological Arthur parameters of real groups is known from Arancibia, Moeglin, and Renard [Reference Arancibia, Moeglin and RenardAMR18].

Theorem 6.6 (Morel–Suh) Let $\varepsilon \in \mathbb{Z}/2\mathbb{Z}$. Then there exists $f^{\varepsilon }$ in $\mathscr {H}_{K}$ such that the associated Hecke correspondence induces the projector to the even (respectively, odd) degree cohomology,

\[ \bigoplus_{i\in\mathbb{Z}}H^{i}(\mathrm{Sh}_K(\widetilde{HG}),\mathbb{Q})\longrightarrow \bigoplus_{i\equiv\varepsilon \bmod 2}H^{i}(\mathrm{Sh}_K(\widetilde{HG}),\mathbb{Q}). \]

Definition 6.7 We set $f^{+}=f^{0}$ and $f^-=f^{1}$ as in Theorem 6.6, and call them the even (respectively, odd) Hecke–Künneth projectors in $\mathscr {H}_{K}$.

Proof Proof of Theorem 6.6 This follows from [Reference Morel and SuhMS19, Theorem 1.4, Lemma 2.2]. We outline the proof given in [Reference Morel and SuhMS19]. It suffices to prove the assertions after tensoring with $\mathbb{C}$. We henceforth consider $H^{i}(\mathrm {Sh}_K(\widetilde {HG}),\mathbb{C})$ with the action by $\mathscr {H}_{K,\mathbb{C}}=\mathscr {H}_K\otimes \mathbb{C}$.

By Matsushima's formula [Reference Borel and WallachBW00, VII], we have a decomposition of the Betti cohomologyFootnote 12 into a finite direct sum,

(6.5)\begin{equation} H^{*}(\mathrm{Sh}_K(\widetilde{HG}),\mathbb{C}) \simeq \bigoplus_{\pi=\pi_f\otimes \pi_\infty} V_{\pi, K}, \end{equation}

where we set

\[ V_{\pi, K}:= m_{\operatorname{disc}}(\pi)\,(\pi_f^{K}\otimes H^{*}(\widetilde{\mathfrak{h}\mathfrak{g}},K_\infty;\pi_\infty)). \]

Here

  • $\pi$ runs through the set $\Pi _\textrm {disc}(\widetilde {HG})$ of irreducible representations of $\widetilde {HG}(\mathbb{A})$ in the discrete automorphic spectrum $L^{2}_{\operatorname {disc}}(Z^{\mathbb{Q}}(\mathbb{R}) \widetilde {HG}(\mathbb{Q})\backslash \widetilde {HG}(\mathbb{A}))$, and $m_{\operatorname {disc}}(\pi )$ is the multiplicity of $\pi$;Footnote 13

  • $\widetilde {\mathfrak {h}\mathfrak {g}}$ is the complex Lie algebra of $\widetilde {HG}(\mathbb{R})$;

  • $K_\infty$ is the centralizer of $h_{\widetilde {HG}}$ in $\widetilde {HG}(\mathbb{R})$;Footnote 14 and

  • $\pi _f^{K}$ denotes the invariants of $\pi _f$ under $K$.

The isomorphism (6.5) is $\mathscr {H}_K$-equivariant, where $\mathscr {H}_K$ acts on the right-hand side through the space $\pi ^{K}_f$.

We have a decomposition of the discrete spectrum as $\widetilde {HG}(\mathbb{A})$-modules (cf. [Reference ArthurArt89], [Reference KottwitzKot90, § 8]),

\[ L^{2}_{\operatorname{disc}} (Z^{\mathbb{Q}}(\mathbb{R}) \widetilde{HG}(\mathbb{Q})\backslash \widetilde{HG}(\mathbb{A}))\simeq \bigoplus_{\psi} \bigoplus_{\pi\in \Pi_\psi} m(\psi,\pi)\cdot \pi. \]

Here the sum is taken over all equivalence classes of global Arthur parameters $\psi$, with trivial associated quasi-character of $Z^{\mathbb{Q}}(\mathbb{R})$. Furthermore, $\Pi _\psi$ denotes the Arthur packet attached to $\psi$, and $m(\psi ,\pi )$ denotes the Arthur multiplicity. Hence we may rewrite the decomposition (6.5) according to global Arthur parameters as

(6.6)\begin{equation} H^{*}(\mathrm{Sh}_K(\widetilde{HG}),\mathbb{C}) \simeq \bigoplus_{\psi} V_{\psi,K} , \end{equation}

where we set

(6.7)\begin{equation} V_{\psi,K}:= \bigoplus_{\pi=\pi_f\otimes\pi_\infty\in \Pi_\psi} m(\psi,\pi) (\pi_f^{K}\otimes H^{*}(\widetilde{\mathfrak{h}\mathfrak{g}},K_\infty;\pi_\infty)). \end{equation}

The isomorphism (6.6) is $\mathscr {H}_{K}$-equivariant. Moreover, there is a canonical Lefschetz class (coming from the cup product with the Killing form) which induces an $\mathrm {SL}_2(\mathbb{C})$-action on the graded vector space $H^{*}(\mathrm {Sh}_K(\widetilde {HG}),\mathbb{C})$ (cf. [Reference ArthurArt89, Proposition 9.1]). Correspondingly, there is an $\mathrm {SL}_2(\mathbb{C})$-action on the graded vector space $H^{*}(\widetilde {\mathfrak {h}\mathfrak {g}},K_\infty ;\pi _\infty )$. The decomposition respects the $\mathrm {SL}_2(\mathbb{C})$-action and the grading on both sides of (6.6). We refer to [Reference ArthurArt89] for details.

In general, the definition of Arthur parameter involves the hypothetical automorphic Langlands group. In the case of classical groups, Arthur avoids the Langlands group by using cuspidal (or isobaric) automorphic representations of general linear groups as substitute parameters [Reference ArthurArt13]. In our setting, our group is a product $\widetilde {HG}=Z^{\mathbb{Q}}\times {HG}$, where

(6.8)\begin{equation} {HG} := \operatorname{Res}_{F_0/\mathbb{Q}}(H\times G) \end{equation}

(cf. (2.1)). Accordingly, we will write an Arthur parameter $\psi$ as a pair $(\psi _0,\psi _1)$ where $\psi _0$ and $\psi _1$ are Arthur parameters for the two factors $Z^{\mathbb{Q}}$ and ${HG}$ and those are defined in terms of cuspidal (or isobaric) automorphic representations of general linear groups.

By the product decomposition (3.4) we have

\[ \mathrm{Sh}_K(\widetilde{HG})_\mathbb{C}\simeq \mathrm{Sh}_{K_{Z^{\mathbb{Q}}}}(Z^{\mathbb{Q}})_\mathbb{C}\times_{\operatorname{Spec}\mathbb{C}} \mathrm{Sh}_{K_{{HG}}}({HG})_\mathbb{C} \]

and an induced isomorphism (note that the first factor above is zero-dimensional)

(6.9)\begin{equation} H^{*}(\mathrm{Sh}_K(\widetilde{HG}),\mathbb{C}) \simeq H^{0}(\mathrm{Sh}_{K_{Z^{\mathbb{Q}}}}(Z^{\mathbb{Q}}),\mathbb{C}) \otimes H^{*}(\mathrm{Sh}_{K_{{HG}}}({HG}),\mathbb{C}). \end{equation}

We may therefore replace $\mathrm {Sh}_K(\widetilde {HG})$ by $\mathrm {Sh}_{K_{{HG}}}({HG})$ in the conclusion of Theorem 6.6. We record the decomposition similar to (6.5):

(6.10)\begin{equation} H^{*}(\mathrm{Sh}_{K_{{HG}}}({HG}),\mathbb{C}) \simeq \bigoplus_{\pi_1} V_{\pi_1,K_{{HG}}}. \end{equation}

For a fixed $\pi _{1,f}$, let

\[ \Pi_\infty(\pi_ {1,f}) := \{\pi_{1,\infty}\in {\Pi}({HG}(\mathbb{R})) \,|\, \pi_{1,f}\otimes\pi_{1,\infty}\in \Pi_\textrm{disc}({HG})\}. \]

Here ${\Pi }({HG}(\mathbb{R}))$ denotes the set of equivalence classes of irreducible admissible representations of ${HG}(\mathbb{R})$. Then we may rewrite (6.10) as

(6.11)\begin{equation} H^{*}(\mathrm{Sh}_{K_{{HG}}}({HG}),\mathbb{C}) \simeq \bigoplus_{\pi_{1,f}} V_{\pi_{1,f},K_{{HG}}}, \end{equation}

where

\[ V_{\pi_{1,f},K_{{HG}}}:= \bigoplus_{ \pi_{1,\infty} \in \Pi_\infty(\pi_ {1,f}) } V_{\pi_{1,\infty}\otimes\pi_{1,f},K_{{HG}}}. \]

Now by [Reference Morel and SuhMS19, Theorem 1.4], the degree $i$ modulo $2$ such that $H^{i}(\mathfrak {h}\mathfrak {g},K_\infty ;\pi _{1,\infty })\neq 0$ is constant (in $\mathbb{Z}/2\mathbb{Z}$) as $\pi _{1,\infty }$ varies through $\Pi _\infty (\pi _{1,f})$ and $i$ varies through $\mathbb{Z}$. We denote the constant $i$ mod $2$ above by $\varepsilon (\pi _{1,f})\in \mathbb{Z}/2\mathbb{Z}$.

The (finitely many) non-zero direct summands $V_{\pi _{1,f},K_{{HG}}}$ are distinct $\mathscr {H}_{K,\mathbb{C}}$-modules. Therefore, for each $\pi _{1,f}$, there exists $f_{\pi _{1,f}}\in \mathscr {H}_{K,\mathbb{C}}$ that induces the projector to $V_{\pi _{1,f},K_{{HG}}}$. Now set

\[ f^{+}:=\sum_{\substack{\pi_{1,f}\\ \varepsilon(\pi_{1,f})=0}}f_{\pi_{1,f}} \quad\text{and}\quad f^-:=\sum_{\substack{\pi_{1,f}\\ \varepsilon(\pi_{1,f})=1}}f_{\pi_{1,f}}. \]

This completes the proof of Theorem 6.6.

Remark 6.8 Let $H^{\varepsilon }_K:=\bigoplus _{i\equiv \varepsilon \bmod 2}H^{i}(\mathrm {Sh}_K(\widetilde {HG}),\mathbb{Q})$. Then [Reference Morel and SuhMS19, Theorem 1.4] asserts that $H^{+}_K$ and $H^-_K$, both being semisimple $\mathscr {H}_{K}$-modules, do not share any common irreducible $\mathscr {H}_K$-submodule.

6.3 Arithmetic diagonal cycles

We now apply the considerations of §§ 6.1 and 6.2 to the canonical model $M_K(\widetilde {HG})$ of $\mathrm {Sh}_K(\widetilde {HG})$ over $E$. When $K=K_{\widetilde {HG}}$ is as in (3.15), we have given in § 3.2 a moduli interpretation of $M_K(\widetilde {HG})$.

We consider the cycle class map in degree $n-1$ (for the Betti cohomology),

\[ {{\mathrm{cl}}}_{n-1}\colon {{\mathrm{Ch}}}^{n-1}(M_K(\widetilde{HG}))_\mathbb{Q} \longrightarrow H^{2(n-1)}(\mathrm{Sh}_K(\widetilde{HG}), \mathbb{Q}). \]

Note that $\dim M_K(\widetilde {HG}))=2n-3$ is odd. In particular, the above cohomology group is not in the middle degree, but just above.

Let $K_{\widetilde H}$ be a compact open subgroup of $\widetilde H(\mathbb{A}_f)$ contained in $K\cap \widetilde H(\mathbb{A}_f)$. We have a finite and unramified morphism

\[ M_{K_{\widetilde H}}(\widetilde H)\longrightarrow M_{K}(\widetilde{HG}). \]

The proper push-forward defines a cycle class $[M_{K_{\widetilde H}}(\widetilde H)]\in {{\mathrm {Ch}}}^{n-1}(M_K(\widetilde {HG}))_\mathbb{Q}$. We now fix a Haar measure on $\widetilde H(\mathbb{A})$ such that the volume $\operatorname {vol}(K_{\widetilde H})\in \mathbb{Q}$. We then define the normalized class

(6.12)\begin{equation} z_K=\operatorname{vol}(K_{\widetilde H})[M_{K_{\widetilde H}}(\widetilde H)]\in {{\mathrm{Ch}}}^{n-1}(M_K(\widetilde{HG}))_\mathbb{Q}, \end{equation}

which is independent of the choice of the group $K_{\widetilde H}$. We call $z_K$ the arithmetic diagonal cycle since it lies in the arithmetic middle dimension, that is, $2\dim z_K= \dim M_K(\widetilde {HG})+1$. Let $\mathcal {Z}_{K}$ be the cyclic Hecke submodule of ${{\mathrm {Ch}}}^{n-1}(M_K(\widetilde {HG}))_\mathbb{Q}$ generated by $z_K$.

Let $f^-\in \mathscr {H}_{K}$ be an odd Hecke–Künneth projector (cf. Definition 6.7). We obtain a map to the Chow group of cohomologically trivial cycles,

(6.13)\begin{equation} R(f^-)\colon {{\mathrm{Ch}}}^{n-1}(M_K(\widetilde{HG}))_\mathbb{Q}\longrightarrow {{\mathrm{Ch}}}^{n-1}(M_K(\widetilde{HG}))_{\mathbb{Q},0}. \end{equation}

We obtain a cohomologically trivial cycle in the Chow group,

(6.14)\begin{equation} z_{K,0}=R(f^-)( z_K)\in {{\mathrm{Ch}}}^{n-1}(M_K(\widetilde{HG}))_{\mathbb{Q},0}. \end{equation}

Accordingly we call $z_{K,0}$ the cohomologically trivial arithmetic diagonal cycle.

Remark 6.9 (i) Since the Hecke–Künneth projector $f^-\in \mathscr {H}_{K}$ is not unique, we comment on the (in)dependence of the induced cohomologically trivial cycle in the Chow group $z_{K,0}=R(f^-)( z_K)\in {{\mathrm {Ch}}}^{n-1}(M_K(\widetilde {HG}))_{\mathbb{Q},0}$. Conjecturally, for any smooth projective variety $X$ over a number field, it should be true that, given $i$, ${{\mathrm {Ch}}}^{i}(X)_{\mathbb{Q}}$ is finite-dimensional, and if a correspondence induces the zero endomorphism on $H^{2i-1}(X)$, then the induced endomorphism on ${{\mathrm {Ch}}}^{i}(X)_{\mathbb{Q}, 0}$ is zero (cf. [Reference BeilinsonBei87, Lemma 5.6] and [Reference JannsenJan90, Conjecture 9.12]). Moreover, one expects that ${{\mathrm {Ch}}}^{i}(M_K(\widetilde {HG}))_{\mathbb{Q}}$ is a semisimple $\mathscr {H}_K$-module. If these were true for $X=M_K(\widetilde {HG})$ and $i=n-1$, then by Remark 6.8 it would follow that the simple $\mathscr {H}_K$-modules appearing in ${{\mathrm {Ch}}}^{n-1}(M_K(\widetilde {HG}))_{\mathbb{Q}, 0}$ and ${\rm Im} ({{\mathrm {cl}}}_{n-1})$ are disjoint. Furthermore, one expects that there is a unique decomposition of (finite-dimensional) semisimple $\mathscr {H}_K$-modules

(6.15)\begin{equation} {{\mathrm{Ch}}}^{n-1}(M_K(\widetilde{HG}))_{\mathbb{Q}}= {{\mathrm{Ch}}}^{n-1}(M_K(\widetilde{HG}))_{\mathbb{Q}, 0}\oplus {\rm Im}({{\mathrm{cl}}}_{n-1}). \end{equation}

The endomorphism $R(f^-)$ would then define the projector onto the summand ${{\mathrm {Ch}}}^{n-1} (M_K(\widetilde {HG}))_{\mathbb{Q}, 0}$. In particular, the induced endomorphism of ${{\mathrm {Ch}}}^{n-1}(M_K(\widetilde {HG}))_{\mathbb{Q}}$, and hence the element $z_{K,0}=R(f^-)( z_K)\in {{\mathrm {Ch}}}^{n-1}(M_K(\widetilde {HG}))_{\mathbb{Q},0}$, would be independent of the choice of $f^-\in \mathscr {H}_{K}$. The summands in (6.15) would decompose as $\mathscr {H}_K$-modules,

(6.16)\begin{equation} \begin{aligned} {{\mathrm{Ch}}}^{n-1}(M_K(\widetilde{HG}))_{\mathbb{C},0} & =\bigoplus_{\pi}{{\mathrm{Ch}}}^{n-1}(M_K(\widetilde{HG}))_{\mathbb{C}, 0}[\pi^{K}_f] , \\ {\rm Im}({{\mathrm{cl}}}_{n-1})_\mathbb{C} & =\bigoplus_{\sigma}{\rm Im}({{\mathrm{cl}}}_{n-1})_{\mathbb{C}}[\sigma^{K}_f]. \end{aligned} \end{equation}

Here $\pi$ runs through all automorphic representations contributing to $H^{2n-3}(\mathrm {Sh}_K(\widetilde {HG}), \mathbb{C})$, and $\sigma$ runs through all automorphic representations contributing to $H^{2n-2}(\mathrm {Sh}_K(\widetilde {HG}), \mathbb{C})$. Also, ${{\mathrm {Ch}}}^{n-1}(M_K(\widetilde {HG}))_{\mathbb{C},0}[\pi _f^{K}]$ denotes the $\pi ^{K}_f$-isotypic component of the $\mathscr {H}_K$-module, that is, the image under the evaluation map, which is injective,

\[ \pi_f^{K}\otimes \operatorname{Hom}_{\mathscr{H}_K}(\pi_f^{K},{{\mathrm{Ch}}}^{n-1}(M_K(\widetilde{HG}))_{\mathbb{C},0}) \longrightarrow {{\mathrm{Ch}}}^{n-1}(M_K(\widetilde{HG}))_{\mathbb{C},0}. \]

Then

(6.17)\begin{equation} {{\mathrm{Ch}}}^{n-1}(M_K(\widetilde{HG}))_{\mathbb{C},0}[\pi_f^{K}]\simeq \pi_f^{K}\otimes \operatorname{Hom}_{\mathscr{H}_K}(\pi_f^{K},{{\mathrm{Ch}}}^{n-1}(M_K(\widetilde{HG}))_{\mathbb{C},0}).\end{equation}

Similar definitions apply to ${\rm Im} ({{\mathrm {cl}}}_{n-1})_{\mathbb{C}}[\sigma ^{K}_f]$.

(ii) In [Reference ZhangZha09] the Chow–Künneth decomposition was used to modify the arithmetic diagonal cycle to make its cohomology class trivial. However, it is difficult to show the existence of a Chow–Künneth decomposition except in some special cases. Therefore, the procedure above is preferable.

6.4 The arithmetic Gan–Gross–Prasad conjecture, for a fixed level $K\subset \widetilde {HG}(\mathbb{A}_f)$

Let $\mathcal {Z}_{K,0}$ denote the cyclic Hecke submodule of ${{\mathrm {Ch}}}^{n-1}(M_K(\widetilde {HG}))_{\mathbb{Q},0}$ generated by $z_{K,0}$ or, equivalently, the image of $\mathcal {Z}_K$ under the map $R(f^-)$. We would like to decompose as $\mathscr {H}_K$-modules

\[ \mathcal{Z}_{K,0}\subset {{\mathrm{Ch}}}^{n-1}(M_K(\widetilde{HG}))_{\mathbb{C},0}. \]

In the conjectural decomposition (6.16), we only consider the tempered part of the spectrum. The non-tempered part is also interesting but will be postponed to the future.

Let $\pi$ be an automorphic representation of $\widetilde {HG}(\mathbb{A})$ with trivial restriction to the central subgroup $Z^{\mathbb{Q}}(\mathbb{A})$. By (2.1) we may consider $\pi$ as an automorphic representation of $(H\times G)(\mathbb{A}_{F_0})$. Let $R$ be the tensor product representation of the $L$-group of $H\times G$ defined in [Reference Gan, Gross and PrasadGGP12, § 22]. The $L$-function $L(s,\pi ,R)$ depends only on the Arthur parameter $\psi$ of $\pi$.

To explain this $L$-function, we write a tempered Arthur parameter $\psi =\psi ^{(n-1)}\boxtimes \psi ^{(n)}$ formally as

(6.18)\begin{equation} \psi^{(n-1)}=\bigoplus_{i}\psi^{(n-1)}_{i},\quad \psi^{(n)}=\bigoplus_{j}\psi^{(n)}_{j} , \end{equation}

where $\psi _i^{(n-1)}$ correspond to distinct cuspidal automorphic representations of $\mathrm {GL}_{N^{(n-1)}_i}$ such that $\sum N^{(n-1)}_i=n-1$, and similarly for $\psi ^{(n)}$. Then the $L$-function in question equals

(6.19)\begin{equation} L(s,\pi,R):=\prod_{i,j} L(s,\psi^{(n-1)}_{i}\boxtimes \psi^{(n)}_{j}), \end{equation}

where each factor is a Rankin–Selberg convolution.

We assume that $\pi$ lies in the packet $\Pi _\psi$ of a cohomological tempered packet $\psi$. Here ‘cohomological’ is a condition on the archimedean component $\psi _\infty$ and refers to the trivial coefficient system.

The first version of the AGGP conjecture can now be stated as follows.

Conjecture 6.10 Let $K\subset \widetilde {HG}(\mathbb{A}_f)$ be an open compact subgroup. Let $\pi$ be as above, that is, with trivial restriction to the central subgroup $Z^{\mathbb{Q}}(\mathbb{A})$ and lying in a cohomological tempered Arthur packet. Consider the following conditions on $\pi$.

  1. (a) $\dim \operatorname {Hom}_{\mathscr {H}_K}(\pi _f^{K}, \mathcal {Z}_{K,0})=1$.Footnote 15

  2. (b) The order of vanishing $\operatorname {ord}_{s=1/2}L(s,\pi ,R)$ equals $1$, the space $\operatorname {Hom}_{\widetilde H(\mathbb{A}_f)}(\pi _f,\mathbb{C})$ is one-dimensional, and its generator does not vanish on the subspace $\pi _f^{K}\subset \pi _f$.

  3. (c) $\operatorname {Hom}_{\mathscr {H}_K}(\pi _f^{K}, {{\mathrm {Ch}}}^{n-1}(M_K(\widetilde {HG}))_{\mathbb{C},0})\neq 0$.

Then (a) and (b) are equivalent and imply (c). If $E=F$, then (a), (b) and (c$'$) are equivalent, where

  1. (c′) $\dim \operatorname {Hom}_{\mathscr {H}_K}(\pi _f^{K}, {{\mathrm {Ch}}}^{n-1}(M_K(\widetilde {HG}))_{\mathbb{C},0})=1$, the space $\operatorname {Hom}_{\widetilde H(\mathbb{A}_f)}(\pi _f,\mathbb{C})$ is one-dimensional, and its generator does not vanish on the subspace $\pi _f^{K}\subset \pi _f$.

Remark 6.11 If $E=F$, the equivalence between (b) and (c$'$) is part of the Beilinson–Bloch conjecture [Reference BlochBlo84, Reference BeilinsonBei87] that generalizes the Birch–Swinnerton-Dyer conjecture.

We would like to test the conjecture quantitatively through height pairings. Now we assume that Conjectures 6.1 and 6.2 hold for $M_K(\widetilde {HG})$. For instance, this is the case when $M_K(\widetilde {HG})$ has everywhere good reductionFootnote 16 (cf. Remark 6.4). In particular, we have the Beilinson–Bloch height pairing (6.2) between cohomologically trivial cycles for $i=n-1$. We extend it to a hermitian form on ${{\mathrm {Ch}}}^{n-1}(M_K(\widetilde {HG}))_{\mathbb{C},0}$. Pairing against the distinguished element $z_{K,0}\in {{\mathrm {Ch}}}^{n-1}(M_K(\widetilde {HG}))_{\mathbb{Q},0}$ then defines a linear functional

Let $\mathcal {Z}_{K,0}[\pi _f^{K}]$ be the $\pi ^{K}_f$-isotypic component of $\mathcal {Z}_{K,0}$ as an $\mathscr {H}_K$-module, so that

\[ \mathcal{Z}_{K,0}[\pi_f^{K}]\simeq \pi_f^{K}\otimes \operatorname{Hom}_{\mathscr{H}_K}(\pi_f^{K},\mathcal{Z}_{K,0}). \]

The second version of the AGGP conjecture in terms of the height pairing can be stated as follows.

Conjecture 6.12 Let $K\subset \widetilde {HG}(\mathbb{A}_f)$ be an open compact subgroup. Let $\pi$ be as above. Then the following conditions on $\pi$ are equivalent.

  1. (a) $\ell _{K}|_{\mathcal {Z}_{K,0}[\pi _f^{K}]}\neq 0$.

  2. (b) The order of vanishing $\operatorname {ord}_{s=1/2}L(s,\pi ,R)$ equals $1$, the space $\operatorname {Hom}_{\widetilde H(\mathbb{A}_f)}(\pi _f,\mathbb{C})$ is one-dimensional, and its generator does not vanish on the subspace $\pi _f^{K}\subset \pi _f$.

  3. (c) $\ell _{K}|_{{{\mathrm {Ch}}}^{n-1}(M_K(\widetilde {HG}))_{\mathbb{C},0}[\pi _f^{K}]}\neq 0$.

Remark 6.13 Our formulation differs in several aspects from [Reference Gan, Gross and PrasadGGP12, Conjecture 27.1]. First, in [Reference Gan, Gross and PrasadGGP12], the Shimura varieties are associated to unitary groups, whereas here we consider Shimura varieties associated to groups which differ from those in [Reference Gan, Gross and PrasadGGP12] by a central subgroup (cf. Remark 3.2(iii)). Correspondingly, the varieties in [Reference Gan, Gross and PrasadGGP12] are defined over $F$, whereas our varieties are defined over the extension $E$ of $F$. As a consequence, we cannot predict the dimension of $\operatorname {Hom}_{\mathscr {H}_K}(\pi _f^{K},{{\mathrm {Ch}}}^{n-1}(M_K(\widetilde {HG}))_{\mathbb{C},0})$ in Conjecture 6.10(b) when $F\neq E$ (in the version of [Reference Gan, Gross and PrasadGGP12], this space is one-dimensional if it is non-zero). Note, however, that if $F= K F_0$ for an imaginary quadratic field $K$ and the CM type is induced from $K$ (as in [Reference Harris and TaylorHT01]), then $F=E$. Second, we exploit that the standard sign conjecture is known in our case, and we use it to construct the cohomologically trivial diagonal cycle $z_{K,0}$ and the corresponding linear functional $\ell _K$ that occur in our version of the conjecture. Third, we work with a fixed level $K$ and specify the compact open subgroup $K$ over which the linear functional $\ell _K$ should be non-zero. Finally, we note that, in the terminology of [Reference Gan, Gross and PrasadGGP12], we are only considering the case of a trivial local system $\mathcal {F}$.

Remark 6.14 The space $\operatorname {Hom}_{\widetilde H(\mathbb{A}_f)}(\pi _f,\mathbb{C})$ is at most one-dimensional. It is one-dimensional if and only if the member $\pi$ in the packet $\Pi _\psi$ is as prescribed by the local Gan–Gross–Prasad conjecture [Reference Gan, Gross and PrasadGGP12, Conjecture 17.3]. The local Gan–Gross–Prasad conjecture [Reference Gan, Gross and PrasadGGP12, Conjecture 17.1] predicts that there is a unique $\pi$ in the packet $\Pi _\psi$ such that $\operatorname {Hom}_{\widetilde H(\mathbb{A}_f)}(\pi _f,\mathbb{C})\neq 0$ (in which case $\dim \operatorname {Hom}_{\widetilde H(\mathbb{A}_f)}(\pi _f,\mathbb{C})=1$).

Remark 6.15 Let us restrict our attention to the tempered part in the decomposition in the first line of (6.16),

\[ {{\mathrm{Ch}}}^{n-1}(M_K(\widetilde{HG}))_{\mathbb{C}, 0, \mathrm{temp}} =\bigoplus_{\pi \text{ tempered}}{{\mathrm{Ch}}}^{n-1}(M_K(\widetilde{HG}))_{\mathbb{C}, 0}[\pi^{K}_f]. \]

Let $\mathcal {Z}_{K, 0,\mathrm {temp}}$ be the Hecke submodule of ${{\mathrm {Ch}}}^{n-1}(M_K(\widetilde {HG}))_{\mathbb{C}, 0, \text { temp}}$ generated by (the projection of) $z_{K, 0}$. Then Conjecture 6.10 (together with the expectations in Remark 6.9(i)) implies that when $E=F$,

\[ \mathcal{Z}_{K, 0, \mathrm{temp}}=\bigoplus_{\pi}{{\mathrm{Ch}}}^{n-1}(M_K(\widetilde{HG}))_{\mathbb{C}, 0}[\pi^{K}_f] , \]

where the sum runs over all tempered automorphic representations $\pi$ such that

\[ \operatorname{ord}_{s=1/2}L(\pi, s, R)=1 \]

and such that the space $\operatorname {Hom}_{\widetilde H(\mathbb{A}_f)}(\pi _f,\mathbb{C})$ is one-dimensional, with generator not vanishing on the subspace $\pi _f^{K}\subset \pi _f$.

Remark 6.16 A parallel question is to investigate the structure of the $\mathscr {H}_K$-submodule in $H^{2(n-2)}(M_K(\widetilde {HG}),\mathbb{C})$ generated by the cohomology class ${{\mathrm {cl}}}_{n-1}(z_{K})$. However, since every automorphic representation contributing to the cohomology in off-middle degree must be non-tempered, the answer to such a question must involve the non-tempered version of the Gan–Gross–Prasad conjecture. We hope to return to this in the future.

Remark 6.17 As remarked above, the height pairing is defined unconditionally if $M_K(\widetilde {HG})$ has everywhere good reduction. To construct such cases, let us assume now that $K=K_{\widetilde {HG}}=K_{Z^{\mathbb{Q}}} \times K_H \times K_G$, where $K_{Z^{\mathbb{Q}}}$ is the usual maximal compact subgroup (3.7), $K_G$ is the stabilizer of a lattice $\Lambda$ in $W$, and $K_H$ is the stabilizer of a lattice $\Lambda ^{\flat }$ in $W^{\flat }$. We make the following assumptions on the field extensions $F/F_0/\mathbb{Q}$.

  • Each finite place $v$ of $F_0$ which is ramified over $\mathbb{Q}$ or of residue characteristic $2$ is split in $F$.

  • Each finite place $v$ of $F_0$ which ramifies in $F$ is of degree $1$ over $\mathbb{Q}$.

We also make the following assumptions on the hermitian spaces $W$ and $W^{\flat }$. We distinguish the case where $n$ is odd from the case where $n$ is even (cf. Remark 5.3).

When $n=2m+1$ is odd, we impose that

  • $W$ is split at all finite places of $F_0$ which are inert in $F$; and

  • $W^{\flat }$ is split at all finite places, which forces $m$ and $[F_0 : \mathbb{Q}]$ to be odd (cf. Remark 5.3).

Then we choose $\Lambda _v$ to be self-dual when $v$ is split or inert in $F$, and almost $\pi _v$-modular when $v$ is ramified in $F$. Furthermore, we choose $\Lambda _v^{\flat }$ to be self-dual when $v$ is inert in $F$ and $\pi _v$-modular when $v$ is ramified in $F$. Such lattices exist, even when we impose that $\Lambda =\Lambda ^{\flat }\oplus O_F u$ with $(u,u) \in O_{F_0}^{\times }$. With these definitions, $M_K(\widetilde {HG})$ has everywhere good reduction.

When $n=2m$ is even, we impose that

  • $W$ is split at all finite places of $F_0$, which again forces $m$ and $[F_0 : \mathbb{Q}]$ to be odd (cf. Remark 5.3); and

  • $W^{\flat }$ is split at all finite places of $F_0$ which are inert in $F$.

Now we choose $\Lambda _v$ to be self-dual when $v$ is split or inert in $F$, and $\pi _v$-modular when $v$ is ramified in $F$. Furthermore, we choose $\Lambda _v^{\flat }$ to be self-dual when $v$ is inert in $F$ and almost $\pi _v$-modular when $v$ is ramified in $F$. Such lattices exist, even when we impose that there exists $u\in W$ with $(u, u) \in O_{F_0}^{\times }$ such that, for all inert finite places $\Lambda _v=\Lambda _v^{\flat }\oplus O_{F,v} u$, and such that, for all ramified places, $\Lambda _v^{\flat }$ and $\Lambda _v$ are related by a chain of inclusions

\[ \pi_{v} (\Lambda_{v}^{\flat} \oplus O_{F,v} u)^{*} \subset^{1} \Lambda_{v} \subset^{1} \Lambda_{v}^{\flat} \oplus O_{F,v} u \]

(cf. (4.30)). With these definitions, $M_K(\widetilde {HG})$ has everywhere good reduction.

One may ask in this connection whether Conjecture 6.10 is non-empty for $M_K(\widetilde {HG})$ (with everywhere good reduction). By our expectations in Remark 6.9(i), this comes down to asking whether there are representations $\pi \in \Pi _\textrm {disc}(\widetilde {HG})$ with $\pi _f^{K}\neq 0$ which contribute to the cohomology $H^{2n-3}(\mathrm {Sh}_K(\widetilde {HG}), \mathbb{C})$. Chenevier [Reference ChenevierChe16] has indicated to us a method of producing such $\pi$ for low values of $n$, when $F/F_0$ is everywhere unramified. The method should also apply when $F/F_0$ is ramified once one has a better understanding of the local Langlands correspondence for unitary groups in ramified cases.

Remark 6.18 Assume $n=2$. In this case, the Beilinson–Bloch pairing is defined unconditionally and coincides with the Néron–Tate height. Conjecture 6.12 is closely related to the Gross–Zagier formula in [Reference Yuan, Zhang and ZhangYZZ13]. It would be interesting to clarify this relation.

7. $L$-functions and the relative trace formula

In this section we recall certain distributions on the group $G'= \operatorname {Res}_{F/F_0}(\mathrm {GL}_{n-1} \times \mathrm {GL}_n)$ that appear in the context of the relative trace formula. For test functions with some local hypotheses, we follow [Reference ZhangZha12a, § 3.1] and [Reference ZhangZha14a, §§ 2.1–2.4]. In general, our definition relies on the truncation of relative trace formulas of Zydor [Reference ZydorZyd20].

On the one hand, these distributions are related to $L$-functions via the Rankin–Selberg theory (for $\mathrm {GL}_{n-1}\times \mathrm {GL}_n$) of Jacquet, Piatetskii-Shapiro, and Shalika [Reference Jacquet, Piatetskii-Shapiro and ShalikaJPS83]. On the other hand, they will serve as the analytic side in our conjectures on arithmetic intersection numbers formulated in the next section.

7.1 The $L$-function

Let $\Pi =\Pi _1\boxtimes \Pi _2$ be a cuspidal automorphic representation of $G'(\mathbb{A}_{F_0})$, where $\Pi _1, \Pi _2$ are automorphic representations of $\mathrm {GL}_{n-1}(\mathbb{A}_F)$ and $\mathrm {GL}_{n}(\mathbb{A}_F)$, respectively. Let $L(s,\Pi _1\boxtimes \Pi _2)$ be the Rankin–Selberg convolution $L$-function. This is an entire function in $s\in \mathbb{C}$ and it satisfies a functional equation of the form

\[ L(s,\Pi_1\boxtimes\Pi_2)=\epsilon(s,\Pi_1\boxtimes\Pi_2)L(1-s,\Pi_1^{\vee}\boxtimes\Pi_2^{\vee}) \]

(cf. [Reference Jacquet, Piatetskii-Shapiro and ShalikaJPS83]). Here $\Pi _i^{\vee }$ denotes the contragredient representation of $\Pi _i$.

The $L$-function $L(s,\Pi _1\boxtimes \Pi _2)$ is represented by an integral. Let $\varphi =\bigotimes _v\varphi _v\in \Pi =\bigotimes _v\Pi _v$ be a decomposable vector. Consider the integral

(7.1)\begin{equation} \lambda(\varphi,s):= \int_{H_1'(F_0)\backslash H_1'(\mathbb{A}_{F_0})}\varphi(h)\lvert\det(h)\rvert_F^{s}\,dh,\quad s\in\mathbb{C}. \end{equation}

Then by [Reference Jacquet, Piatetskii-Shapiro and ShalikaJPS83], with the Haar measures specified in [Reference ZhangZha14b, § 2], we have a decomposition (cf. [Reference ZhangZha14b, § 3.3])

\[ \lambda(\varphi,s) = L\Bigl(s+\frac{1}{2},\Pi_1\boxtimes\Pi_2\Bigr)\prod_{v} \lambda_v(\varphi_v,s). \]

Here the left-hand side is an entire function in $s\in \mathbb{C}$, and the local factors $\lambda (\varphi _v,s)$ have the following properties.

  1. (1) For every $\varphi _v\in \Pi _v$, the function $s\mapsto \lambda _v(\varphi _v,s)$ is entire, and there exists $\varphi _v^{\circ }$ such that $\lambda _v(\varphi _v^{\circ },s) \equiv 1$.

  2. (2) For any decomposable $\varphi =\bigotimes _v\varphi _v\in \Pi =\bigotimes _v\Pi _v$, we have $\lambda _v(\varphi _v,s)\equiv 1$ for all but finitely many $v$.

It follows that if $L(\frac {1}{2},\Pi _1\boxtimes \Pi _2)=0$ (e.g. $\Pi _1$ and $\Pi _2$ are self-dual, and $\epsilon (1/2,\Pi _1\boxtimes \Pi _2)=-1$), then we have

\[ \left.\frac{d}{ds}\right|_{s=0}\int_{H_1'(F_0)\backslash H_1'(\mathbb{A}_{F_0})}\varphi(h)\lvert\det(h)\rvert_F^{s}\,dh=L'\Bigl(\frac{1}{2},\Pi_1\boxtimes\Pi_2\Bigr)\prod_{v} \lambda(\varphi_v,0). \]

We note that, if $\Pi _1$ and $\Pi _2$ correspond to the Arthur parameters $\psi ^{(n-1)}$ and $\psi ^{(n)}$ in (6.18), then we may write the $L$-function in (6.19) as

\[ L(s,\pi,R)=L(s,\Pi_1\boxtimes\Pi_2). \]

7.2 The global distribution on $G'$

We briefly recall the global distribution on $G'$ from [Reference ZhangZha12a, § 3.1] and [Reference ZhangZha14a, § 2] (the notation in these papers is slightly different). We denote by $A_{G'}$ and $A_{H_2'}$ the maximal $F_0$-split tori in the centers of $G'$ and $H_2'$, respectively. We have (via the embedding $H_2'\hookrightarrow G'$) an equality $A_{G'}=A_{H_2'}$ and both are isomorphic to $\mathbb{G}_{m,F_0}\times \mathbb{G}_{m,F_0}$.

Let $f'=\bigotimes _v f'_v\in \mathscr {H}(G'(\mathbb{A}_{F_0}))=C_c^{\infty } (G'(\mathbb{A}_{F_0}))$ be a pure tensor. We consider the associated automorphic kernel function

(7.2)\begin{equation} K_{f'}(x,y) := \int_{A_{G'}(F_0)\backslash A_{G'}(\mathbb{A}_{F_0})}\sum_{\gamma\in G'(F_0)}f'(x^{-1}z\gamma y)\,dz,\quad x,y\in G'(\mathbb{A}_{F_0}). \end{equation}

We define

(7.3)\begin{equation} J(f',s) := \int_{H'_1(F_0)\backslash H_1'(\mathbb{A}_{F_0}) }\int_{A_{H'_2}(\mathbb{A}_{F_0})H_2'(F_0)\backslash H_2'(\mathbb{A}_{F_0})} K_{f'}(h_1,h_2)\eta(h_2)\lvert\det(h_1)\rvert_F^{s}\,dh_1\,dh_2, \end{equation}

where the quadratic character $\eta \colon H_2'(\mathbb{A}_{F_0})\to \{\pm 1\}$ is defined as follows: for $h_2=(x_{n-1},x_{n})\in H_2'(\mathbb{A}_{F_0})$, with $x_i\in \mathrm {GL}_i(\mathbb{A}_{F_0})$,

\[ \eta(h_2)=\eta_{F/F_0}(\det(x_{n-1})^{n}\det(x_{n})^{n-1})). \]

Here $\eta _{F/F_0}\colon \mathbb{A}_{F_0}^{\times }\to \{\pm 1\}$ is the idele class character associated to the extension of global fields $F/F_0$.

The kernel function (7.2) has a spectral decomposition. The contribution of a cuspidal automorphic representation $\Pi$ to the kernel function is given by

\[ K_{\Pi,f'}(x,y)=\sum_{\varphi\in \mathrm{OB}(\Pi)}(\Pi(f')\varphi)(x)\,\overline{ \varphi(y)}, \]

where the sum runs over an orthonormal basis $\mathrm {OB}(\Pi )$ of $\Pi$. Correspondingly, a cuspidal automorphic representation $\Pi$ contributes to the global distribution $J(f',s)$,

\[ J_{\Pi}(f',s) := \sum_{\varphi\in \mathrm{OB}(\Pi)}\lambda(\Pi(f')\varphi ,s)\,\beta(\overline\varphi), \]

where $\beta$ is the Flicker–Rallis period integral with respect to the subgroup $H_{2}'$ (cf. [Reference ZhangZha14b, § 3.2]). It follows from the endoscopic classification for unitary groups that the period $\beta$ does not vanish identically if and only if $\Pi$ is in the image of base change from (quasi-split) unitary groups associated to the quadratic extension $F/F_0$ (cf. [Reference MokMok15, Theorem 2.5.4, Remark 2.5.5], and note that the corresponding Asai $L$-function has a pole if and only if the period $\beta$ does not vanish identically (cf. [Reference ZhangZha14a, Remark 8, p. 976] and references therein)).

We again use the Haar measures specified in [Reference ZhangZha14b, § 2].

Proposition 7.1 Let $\Pi =\Pi _1\boxtimes \Pi _2$ be cuspidal, and assume that it is the base change of an automorphic representation $\pi$ on (quasi-split) unitary groups. If $L(1/2,\Pi _1\boxtimes \Pi _2)=0$, then

\[ \frac{d}{ds}\Big|_{s=0} J_{\Pi}(f',s)=L(1,\eta)^{2}\frac{L'(1/2,\Pi_1\boxtimes\Pi_2)}{L(1,\pi,{\rm Ad})}\prod_{v} J_{\Pi_v}(f'_v), \]

where $J_{\Pi _v}(f'_v)$ is the local distribution defined in [Reference ZhangZha14b, § 3, (3.31)], and $L(1,\pi ,{{\rm Ad}})$ is the adjoint $L$-function (cf. [Reference ZhangZha14b, § 3.4], [Reference MokMok15], [Reference Kaletha, Minguez, Shin and WhiteKMSW14]).

Proof. It follows from [Reference ZhangZha14b, Proposition 3.6] that

(7.4)\begin{equation} J_{\Pi}(f',s)=L(1,\eta)^{2}\frac{L(s+1/2,\Pi_1\boxtimes\Pi_2)}{L(1,\pi,{\rm Ad})}\prod_{v} J_{\Pi_v}(f'_v,s). \end{equation}

Here the local distribution $J_{\Pi _v}(f'_v,s)$ is defined in an analogous way to that of its value at $s=0$ given by [Reference ZhangZha14b, § 3, (3.31)],

(7.5)\begin{equation} J_{\Pi_v}(f'_v,s) = \sum_{\{\varphi_v\}} \frac{\lambda_v^{{\natural}}(\Pi_v(f_v')\varphi_v,s)\overline{\beta^{{\natural}}}(\varphi_v)} {\theta^{{\natural}}(\varphi_v,\varphi_v)}, \end{equation}

where the linear functional $\varphi _v\mapsto \lambda _v^{{\natural }}(\varphi _v,s)$ on $\Pi _v$ is defined by [Reference ZhangZha14b, § 3, (3.24)], using the Whittaker model. We refer to [Reference ZhangZha14b, § 3] for the precise normalization of measures and the linear functionals $\beta ^{{\natural }}$ and $\theta ^{{\natural }}$. In particular, if $L(1/2,\Pi _1\boxtimes \Pi _2)=0$, then we have

\[ \left.\frac{d}{ds}\right|_{s=0} J_{\Pi}(f',s)=L(1,\eta)^{2}\frac{L'(1/2,\Pi_1\boxtimes\Pi_2)}{L(1,\pi,{\rm Ad})}\prod_{v} J_{\Pi_v}(f'_v,0). \]

Since $J_{\Pi _v}(f'_v,0)=J_{\Pi _v}(f'_v)$ by definition, the proof is complete.

Proposition 7.2 Let $f'=\bigotimes _v f'_v\in \mathscr {H}(G'(\mathbb{A}_{F_0}))=C_c^{\infty } (G'(\mathbb{A}_{F_0}))$ be a pure tensor. Suppose that for a split place $v$ the function $f_v'$ has the property that, for every character $\chi _v$ of the center $Z_{G'}(F_{0,v})$ of $G'(F_{0,v})$, the function $f'_{v,\chi _v}:g\mapsto \int _{Z_{G'}(F_{0,v})}f'_v(zg)\chi ^{-1}_v(z)\,dz$ is the sum of matrix coefficients of supercuspidal representations. Then the integral (7.3) converges absolutely and it decomposes as

\[ J(f',s)=\sum_{\Pi} J_{\Pi}(f',s) =\sum_{\Pi}L(1,\eta)^{2}\frac{L(s+1/2,\Pi_1\boxtimes\Pi_2)}{L(1,\pi,{\rm Ad})}\prod_{v} J_{\Pi_v}(f'_v,s) , \]

where the sum runs through the set of cuspidal automorphic representations $\Pi =\Pi _1\boxtimes \Pi _2$ of $G'(\mathbb{A}_{F_0})$ coming by base change from automorphic representations $\pi$ of (quasi-split) unitary groups. Here the distribution $J_{\Pi _v}(f'_v,s)$ is defined by (7.5).

Proof. The spectral decomposition (i.e. the first equality) follows from the simple version of the relative trace formula in [Reference ZhangZha14a, Theorem 2.3]. Note that there the test function $f'$ is required to be ‘nice’ with respect to a character of the center $\chi$ of $G'(\mathbb{A}_{F_0})$. However, the spectral decomposition only requires the existence of a place $v$ where the function $f'_{v,\chi _v}$ is a matrix coefficient of a super-cuspidal representation. Hence the result holds by the linearity of $J$ under the current assumption. Moreover, though only the case $s=0$ is stated in [Reference ZhangZha14a, Theorem 2.3], the same proof works for all $s\in \mathbb{C}$. The second equality follows from Proposition 7.1.

Remark 7.3 There are many test functions $f_v'$ with the property in the above statement. It suffices to construct such functions for $\mathrm {GL}_m(F)$ where $F$ is a $p$-adic local field. Let $\pi$ be a supercuspidal representation of $\mathrm {GL}_m(F)$ and $\widetilde f$ a matrix coefficient. Let $\textbf {1}_{[\textrm {val}=0]}$ be the characteristic function of the set consisting of $g\in \mathrm {GL}_m(F)$ such that $\det (g)\in O_F^{\times }$. Then we claim that $f=\widetilde f\cdot \textbf {1}_{[\textrm {val}=0]}$ has the desired property. To see this, let $\chi _0$ be the central character of $\pi$, and $\nu$ an unramified character of $F^{\times }$ of order (exactly) $m$. Then $f_{\chi _0}= ({1}/{m})\sum _{i=0}^{m-1} \widetilde f\cdot (\nu ^{i}\circ \det )$ (we normalize the measure such that $\operatorname {vol}(O_F^{\times })=1$). Now note that $f(zg)=\chi _{0}(z)f(g)$ for $z\in O_F^{\times }$. Hence $f_{\chi }=0$ unless $\chi \chi _0^{-1}$ is unramified. If $\chi \chi _0^{-1}$ is unramified, we may assume that $\chi \chi _0^{-1}=\xi ^{m}$ for some unramified character $\xi$. Then $f_{\chi }=({1}/{m})\sum _{i=0}^{m-1}\widetilde f\cdot ((\xi \nu ^{i})\circ \det )$, hence a sum of matrix coefficients of $\pi \otimes \xi \nu ^{i}$, $0\leq i\leq m-1$.

It follows that for $f'$ as in Proposition 7.2, we have an expansion for the first derivative

(7.6)\begin{align} \frac{d}{ds}\Big|_{s=0}J(f', s)&=\sum_{\substack{\Pi\\ \epsilon(\Pi)=-1}}L(1,\eta)^{2}\frac{L'(1/2,\Pi_1\boxtimes\Pi_2)}{L(1,\pi,{\rm Ad})}\prod_{v} J_{\Pi_v}(f'_v,0)\nonumber\\ &\quad+\sum_{\substack{\Pi\\ \epsilon(\Pi)=1}}L(1,\eta)^{2}\frac{L(1/2,\Pi_1\boxtimes\Pi_2)}{L(1,\pi,{\rm Ad})}\cdot \frac{d}{ds}\Big(\prod_{v} J_{\Pi_v}(f'_v,s) \Big)\Big|_{s=0}. \end{align}

Here $\epsilon (\Pi )=\epsilon (1/2,\Pi _1\boxtimes \Pi _2)$ is the global root number for the Rankin–Selberg convolution.

Remark 7.4 We note that the contribution to the spectral decomposition from non-cuspidal automorphic representations is more complicated and we will not touch on this topic in this paper. A full spectral decomposition to remove the local restriction on $f'_v$ in the last proposition is work in progress by Chaudouard and Zydor, and a coarse spectral decomposition has been obtained by Zydor in [Reference ZydorZyd20].

Definition 7.5 Let $\lambda$ be a place of $F_0$. A function $f'_\lambda \in C_c^{\infty } (G'(F_{0, \lambda }))$ has regular support if $\operatorname {supp}(f_\lambda ')\subset G'(F_{0,\lambda })_\mathrm {rs}$. A pure tensor $f'=\bigotimes _v f'_v\in \mathscr {H}(G'(\mathbb{A}_{F_0}))$ has regular support at $\lambda$ if $f'_\lambda$ has regular support.

Let us assume that $f'$ has regular support at $\lambda$. Later we will assume that $\lambda$ is non-archimedean. Then, by [Reference ZhangZha12a, Lemma 3.2], the integral (7.3) is absolutely convergent for all $s\in \mathbb{C}$, and admits a decomposition into a finite sum for a given $f'$ (see the proof of [Reference ZhangZha12a, Lemma 3.2]),

(7.7)\begin{equation} J(f',s)=\sum_{\gamma\in G'(F_0)_\mathrm{rs}/H_{1,2}'(F_0)} \operatorname{Orb}(\gamma,f',s), \end{equation}

where each term is a product of local orbital integrals [Reference ZhangZha12a, (3.2)],

(7.8)\begin{equation} \operatorname{Orb}(\gamma,f',s)=\prod_{v}\operatorname{Orb}(\gamma,f'_v,s), \end{equation}

where in turn

\[ \operatorname{Orb}(\gamma, f_v', s) := \int_{H_{1,2}'(F_{0,v})} f_v'(h_1^{-1}\gamma h_2) \lvert\det h_1\rvert^{s} \eta(h_2)\, dh_1\, dh_2. \]

We set

(7.9)\begin{equation} J(f'):=J(f', 0). \end{equation}

Then the decomposition (7.7) specializes to

\[ J(f')=\sum_{\gamma\in G'(F_0)_\mathrm{rs}/H_{1,2}'(F_0)} \operatorname{Orb}(\gamma,f'), \]

where

\[ \operatorname{Orb}(\gamma,f') := \operatorname{Orb}(\gamma,f',0). \]

We introduce

\[ J_v(f',s) := \sum_{\gamma\in G'(F_0)_\mathrm{rs}/H_{1,2}'(F_0)} \operatorname{Orb}(\gamma,f'_v,s)\cdot \prod_{u\neq v} \operatorname{Orb}(\gamma,f'_u). \]

We set

(7.10)\begin{equation} \begin{aligned} \partial J(f') & :=\left.\frac{d}{ds}\right|_{s=0}J(f',s),\\ \partial J_v(f') & := \left.\frac{d}{ds}\right|_{s=0}J_v(f',s),\\ \operatorname{\partial Orb}(\gamma,f'_v) & := \left.\frac{d}{ds}\right|_{s=0} \operatorname{Orb}(\gamma,f_v',s). \end{aligned} \end{equation}

Note that

(7.11)\begin{equation} \partial J_v(f')= \sum_{\gamma\in G'(F_0)_\mathrm{rs}/H_{1,2}'(F_0)} \operatorname{\partial Orb}(\gamma,f'_v)\cdot \prod_{u\neq v} \operatorname{Orb}(\gamma,f'_u). \end{equation}

Then we may decompose

(7.12)\begin{equation} \partial J(f')=\sum_{v} \partial J_v(f'). \end{equation}

Without the regularity assumption on $f'$, the integral (7.3) may diverge in general. For all $f'\in \mathscr {H}(G'(\mathbb{A}_{F_0}))$, the truncation process of Zydor [Reference ZydorZyd20] allows us to define a meromorphic distribution $J(\cdot ,s)$ [Reference ZydorZyd20, Theorem 0.1] which is holomorphic away from $s=\pm 1$. This allows us to define (7.9) and (7.10). We will use these distributions to formulate our conjectures in the next section. Zydor also obtains a coarse decomposition of $J(\cdot ,s)$ into a sum of global orbital integrals, although for a non-regular-semisimple orbit there is no natural decomposition into a product of local orbital integrals.

7.3 Smooth transfer

The notion of smooth transfer between functions on unitary groups and on linear groups or their symmetric spaces is based on the concept of matching (cf. [Reference Rapoport, Smithling and ZhangRSZ18, § 2] and [Reference ZhangZha12a, Reference ZhangZha14a]). Using the results of § 2.2, we can transpose this concept to our situation at hand.

Our definitions below depend on the choice of the transfer factor $\omega$ and on the choices of Haar measures. For definiteness, we will always take the transfer factor from [Reference Rapoport, Smithling and ZhangRSZ18, § 2.4] (this is a slight variant of [Reference ZhangZha14a, § 2.4]), which works for all places $v$,

(7.13)\begin{equation} \omega(\gamma)=\prod_v\omega_v(\gamma_v),\quad \gamma=(\gamma_v)\in G'(\mathbb{A}_{F_0}). \end{equation}

The transfer factor has the following properties.

  1. (1) ($\eta$-invariance) For $h_1\in H_1'(\mathbb{A}_{F_0})$ and $h_2\in H_2'(\mathbb{A}_{F_0})$, we have $\omega (h_1^{-1}\gamma h_2)=\eta (h_2)\omega (\gamma )$.

  2. (2) (product formula) For $\gamma \in G'(F_0)$, we have

    (7.14)\begin{equation} \prod_v\omega_v(\gamma)=1. \end{equation}

Now let $p$ be a rational prime. We fix Haar measures on $Z^{\mathbb{Q}}(\mathbb{Q}_p)$, $H(F_{0,v})$, and $G_W(F_{0,v})$ for $v \mid p$. We choose the Haar measures on $\widetilde H(\mathbb{Q}_p)$ and $\widetilde {HG}(\mathbb{Q}_p)$ compatible with the product decompositions $\widetilde {HG}(\mathbb{Q}_p)=Z^{\mathbb{Q}}(\mathbb{Q}_p)\times \prod _{v \mid p}H(F_{0,v})$ and $\widetilde {HG}(\mathbb{Q}_p)=Z^{\mathbb{Q}}(\mathbb{Q}_p)\times \prod _{v \mid p}G_W(F_{0,v})$, induced by (2.1). We choose the quotient measure on $(\widetilde H(\mathbb{Q}_p)\times \widetilde H(\mathbb{Q}_p)) /\Delta (Z^{\mathbb{Q}}(\mathbb{Q}_p))$. For the archimedean places, we also choose Haar measures satisfying similar compatibilities.

We define the orbital integral for $f_p \in C_c^{\infty }(\widetilde {HG}(\mathbb{Q}_p))$ and $g\in \widetilde H(\mathbb{Q}_p) \backslash \widetilde {HG}(\mathbb{Q}_p)_\mathrm {rs} / \widetilde H(\mathbb{Q}_p)$,

(7.15)\begin{equation} \operatorname{Orb}(g,f_p) :=\int_{(\widetilde H(\mathbb{Q}_p)\times\widetilde H(\mathbb{Q}_p)) /\Delta(Z^{\mathbb{Q}}(\mathbb{Q}_p)) } f_p(h_1^{-1}g h_2)\, dh_1\, dh_2. \end{equation}

Definition 7.6 A function $f_p \in C_c^{\infty }(\widetilde {HG}(\mathbb{Q}_p))$ and a collection $(f_v') \in \prod _{v\mid p} C_c^{\infty }(G'(F_{0,v}))$ of functions are transfers of each other if for any element $\gamma = (\gamma _v) \in \prod _{v\mid p} G'(F_{0, v})_\mathrm {rs}$, the following identity holds:

\[ \omega(\gamma) \prod_{v\mid p} \operatorname{Orb}(\gamma_v,f_v') = \begin{cases} \operatorname{Orb}(g,f_p), & \text{whenever } g \text{ matches } \gamma,\\ 0, & \text{no } g\in \widetilde{HG}(\mathbb{Q}_p) \text{ matches } \gamma. \end{cases} \]

We make the same definition for a function $f_\infty \in C_c^{\infty } (\widetilde {HG}(\mathbb{R}))$ and a collection of functions $(f_v') \in \prod _{v\mid \infty } C_c^{\infty }(G'(F_{0,v}))$.

Let us explain the relation to smooth transfer between functions in $C^{\infty }_c(G'(F_{0, v}))$ and functions in $C^{\infty }_c(G_W(F_{0, v}))$, as $W$ varies through the isomorphism classes of hermitian spaces of dimension $n$ over $F_v$. This definition is based on the concept of matching between elements of $G'(F_{0, v})_\mathrm {rs}$ and elements of $G_W(F_{0, v})_\mathrm {rs}$ (see [Reference Rapoport, Smithling and ZhangRSZ17, § 2] for non-archimedean places $v$ of $F_0$ that are non-split in $F$; the definition extends in an obvious way to the archimedean places and to the split non-archimedean places).

Definition 7.7 A function $f_p\in C_c^{\infty }(\widetilde {HG}(\mathbb{Q}_p))$ is completely decomposed if it is of the form

(7.16)\begin{equation} f_p = \phi_{p}\otimes\bigotimes_{v\mid p}f_v, \end{equation}

where $\phi _p\in C_c^{\infty }(Z^{\mathbb{Q}}(\mathbb{Q}_p))$ and $f_v\in C_c^{\infty }(G_W(F_{0,v}))$. A pure tensor $f=\bigotimes _p f_p\in \mathscr {H}(\widetilde {HG}(\mathbb{A}_f))$ is completely decomposed if $f_p$ is completely decomposed for all $p$.

Note that an arbitrary element in $\mathscr {H}(\widetilde {HG}(\mathbb{A}_f))$ is a linear combination of completely decomposed pure tensors.

Remark 7.8 Let $f_p = \phi _{p}\otimes \bigotimes _{v\mid p}f_v$ be completely decomposed. Set

\[ c(\phi_p):=\int_{Z^{\mathbb{Q}}(\mathbb{Q}_p)}\phi_p(z) \, dz. \]

By Lemma 2.1, we have, for $g\in \widetilde {HG}(\mathbb{Q}_p)$ corresponding to the collection $g_v\in (H\times G)(F_{0,v})$,

\[ \operatorname{Orb}(g,f_p) =c(\phi_p)\prod_{v\mid p} \operatorname{Orb}(g_v,f_v), \]

where

\[ \operatorname{Orb}(g_v,f_v)=\int_{H(F_{0,v})\times H(F_{0,v})} f_v(h_1^{-1}g_v h_2)\,dh_1\,dh_2 \]

is the orbital integral in [Reference Rapoport, Smithling and ZhangRSZ18, § 2]. If the orbital integrals of $f_p$ do not vanish identically, then $f_p$ and $(f_v')_{v\mid p}$ are transfers of each other in the sense of Definition 7.6 if and only if for some non-zero constants $c_v$ such that $c(\phi _p)=\prod _{v\mid p}c_v$, the functions $f_v$ and $c_v f_v'$ are transfers of each other for each $v$ in the sense of [Reference Rapoport, Smithling and ZhangRSZ18, § 2].

Definition 7.9 Let $v$ be an archimedean place of $F_0$. A function $f'_v\in C^{\infty }_c(G'(F_{0, v}))$ is a Gaussian test function if it transfers to the constant function $\mathbf {1}$ on $G_{W_0}(F_{0, v})$, where $W_0$ denotes the negative-definite hermitian space, and transfers to the zero function on $G_W(F_{0, v})$ for any other hermitian space $W$ (i.e. in the terminologyFootnote 17 of [Reference ZhangZha12a, Definition 3.5], $f'_v$ is pure of type $W_0$ and a transfer of $\mathbf {1}_{G_{W_0}(F_{0, v})}$).

A pure tensor $f'=\bigotimes _v f'_v\in \mathscr {H}(G'(\mathbb{A}_{F_0}))$ is a Gaussian test function if the archimedean components $f'_v$ for $v\mid \infty$ are all (up to scalar factor) Gaussian test functions.

We have fixed a Haar measure on each (compact) $H_{W_0}(F_{0, v})$. We will assume that the volume of $H_{W_0}(F_{0, v})$ is $1$. Then $f_v'$ is a Gaussian test function if and only if for all $\gamma \in G'(F_{0,v})_\mathrm {rs}$,

(7.17)\begin{align} \omega_v(\gamma)\operatorname{Orb}(\gamma, f'_v)=\begin{cases} 1, & \text{there exists } g \in G_{W_0}(F_{0,v}) \text{ matching } \gamma ,\\ 0, & \text{no } g\in G_{W_0}(F_{0,v}) \text{ matches } \gamma. \end{cases} \end{align}

The existence of Gaussian test functions is still conjectural. A Gaussian test function does not have regular support, in the sense of Definition 7.5.

Definition 7.10 A pure tensor $f=\bigotimes _p f_p\in \mathscr {H}(\widetilde {HG}(\mathbb{A}_f))$ and a pure tensor $f'=\bigotimes _v f'_v\in \mathscr {H}(G'(\mathbb{A}_{F_0, f}))$ are smooth transfers of each other if they are expressible in such a way that $f_p$ and $(f_v')_{v\mid p}$ are transfers of each other for each prime $p$. Here we choose the product measures on $Z^{\mathbb{Q}}(\mathbb{A}_f), \widetilde {HG}(\mathbb{A}_f)$, and $G'(\mathbb{A}_{F_0, f})$ (implicitly we require that our local choices are made such that the product is convergent); also, the adelic transfer factor (7.13) is simply the product of the local transfer factors.

Remark 7.11 The existence of local smooth transfer is known for non-archimedean places [Reference ZhangZha14a]; hence for any $f\in \mathscr {H}(\widetilde {HG}(\mathbb{A}_f))$ as above, there exists a smooth transfer $f'\in \mathscr {H}(G'(\mathbb{A}_{F_0, f}))$ as above, and conversely.

Lemma 7.12 Let $f'=\bigotimes _v f'_v\in \mathscr {H}(G'(\mathbb{A}_{F_0}))$ be a Gaussian test function. Assume that $f'$ has regular support at some place $\lambda$ of $F_0$. Then for any place $v_0$ of $F_0$ split in $F$,

\[ \partial J_{v_0}(f')=0. \]

Proof. This is [Reference ZhangZha12a, Proposition 3.6(ii)]. Note that implicitly our test function $f'$ is pure of an incoherent type [Reference ZhangZha12a, § 3.1], so that for every regular semisimple $\gamma$, the local orbital integral $\operatorname {Orb}(\gamma ,f'_v,s)$ vanishes at $s=0$ for at least one non-split place $v$.

8. The conjectures for the arithmetic intersection pairing

In this section we formulate a conjectural formula for the Gillet–Soulé arithmetic intersection pairing for cycles on the integral models of $M_K(\widetilde {HG})$ we introduced earlier. This formula uses the distributions introduced in § 7.

Throughout this section we assume that the extension $F/F_0$ and the hermitian space $W$ are such that all places $v \in \mathcal {V}_\mathrm {AT}^{W}$ (cf. (5.1)) are of degree $1$ over $\mathbb{Q}$.

8.1 The global conjecture, trivial level structure

Let $\Lambda ^{\flat }\subset W^{\flat }$ and $\Lambda \subset W$ be a pair of $O_F$-lattices related as in § 5.1, and recall the models $\mathcal {M}_{K_{\widetilde H}^{\circ }}(\widetilde H)$, $\mathcal {M}_{K_{\widetilde G}^{\circ }}(\widetilde G)$, and $\mathcal {M}_{K_{\widetilde {HG}}^{\circ }}(\widetilde {HG})$ over $\operatorname {Spec} O_E$ defined in § 5.1 for the Shimura varieties of § 3. In the case where $F_0=\mathbb{Q}$ and $\mathcal {M}_{{K_{\widetilde {HG}}^{\circ }}}(\widetilde {HG})$ is non-compact, we are implicitly replacing $\mathcal {M}_{{K_{\widetilde {HG}}^{\circ }}}(\widetilde {HG})$ by its toroidal compactification. Then the model $\mathcal {M}_{K_{\widetilde {HG}}^{\circ }}(\widetilde {HG})$ is proper and flat over $\operatorname {Spec} O_E$. Furthermore, it is regular provided that there are no places $\nu$ of $E$ for which $E_\nu$ is ramified over $\mathbb{Q}_p$ and $(v_0,\Lambda _{v_0})$ is of AT type (1), where $v_0$ denotes the place of $F_0$ induced by $\nu$ via $\varphi _0$. Throughout this section we assume that there are no places $v_0$ for which $(v_0,\Lambda _{v_0})$ is of AT type (4) (the justification for this assumption is given by Remark 4.9).

The compact open subgroup $K^{\circ }_{\widetilde H} \subset \widetilde H(\mathbb{A}_f)$ contains ${K_{\widetilde {HG}}^{\circ }}\cap \widetilde H(\mathbb{A}_f)$, with equality when there are no places $v$ of $F_0$ for which $(v,\Lambda _v)$ is of AT type (2). In this case, there is a closed embedding

\[ \mathcal{M}_{K^{\circ}_{\widetilde H}}(\widetilde H) {\hookrightarrow} \mathcal{M}_{{K_{\widetilde{HG}}^{\circ}}}(\widetilde{HG}). \]

As in (6.12), we obtain a cycle (with $\mathbb{Q}$-coefficients) $z_{K_{\widetilde {HG}}^{\circ }}=\operatorname {vol}(K^{\circ }_{\widetilde H})[\mathcal {M}_{K^{\circ }_{\widetilde H}}(\widetilde H)]$ on $\mathcal {M}_{K_{\widetilde {HG}}^{\circ }}(\widetilde {HG})$. We denote by the same symbol its class in the rational Chow group,

(8.1)\begin{equation} z_{K_{\widetilde{HG}}^{\circ}}\in {{\mathrm{Ch}}}^{n-1}(\mathcal{M}_{K_{\widetilde{HG}}^{\circ}}(\widetilde{HG}))_\mathbb{Q}. \end{equation}

In general, let $v_1,\ldots ,v_m \in \mathcal {V}_\mathrm {AT}^{W}$ be the places for which $(v_i,\Lambda _{v_i})$ is of AT type (2). We use the closed embedding (5.6) to define the cycle $z_{K_{\widetilde {HG}}^{\circ }} = \operatorname {vol}(K_{\widetilde H}^{\circ \prime })[\mathcal {M}_{K_{\widetilde H}^{\circ \prime }}(\widetilde H)]$ and its class in the Chow group.

Remark 8.1 The definitions above of the cycle class $z_{K_{\widetilde {HG}}^{\circ }}$ use a Haar measure on $\widetilde H(\mathbb{A}_f)$. We will always choose the product of the measures used to define the local orbital integrals (7.15).

We denote by $\widehat {{\mathrm {Ch}}}{}^{n-1}(\mathcal {M}_{K_{\widetilde {HG}}^{\circ }}(\widetilde {HG}))$ the arithmetic Chow group. Elements are represented by pairs $(Z, g_Z)$, where $Z$ is a cycle and $g_Z$ is a Green's current (cf. [Reference Gillet and SouléGS90, § 3.3]). We will use the Gillet–Soulé arithmetic intersection pairing (cf. [Reference Gillet and SouléGS90]),

\[ {(\text{,})}_{\mathrm{GS}}\colon \widehat{{\mathrm{Ch}}}{}^{n-1}(\mathcal{M}_{K_{\widetilde{HG}}^{\circ}}(\widetilde{HG})) \times \widehat{{\mathrm{Ch}}}{}^{n-1}(\mathcal{M}_{K_{\widetilde{HG}}^{\circ}}(\widetilde{HG}))\longrightarrow \mathbb{R}. \]

We extend this from a symmetric pairing to a hermitian pairing on the corresponding $\mathbb{C}$-vector space ($\mathbb{C}$-linear combinations of $(Z, g_Z)$),

(8.2)\begin{equation} {(\text{,})}_{\mathrm{GS}}\colon \widehat{{\mathrm{Ch}}}{}^{n-1}(\mathcal{M}_{K_{\widetilde{HG}}^{\circ}}(\widetilde{HG}))_{\mathbb{C}}\times \widehat{{\mathrm{Ch}}}{}^{n-1}(\mathcal{M}_{K_{\widetilde{HG}}^{\circ}}(\widetilde{HG}))_{\mathbb{C}}\longrightarrow \mathbb{C}. \end{equation}

We choose a Green's current $g_{z_{K_{\widetilde {HG}}^{\circ }} }$ of the cycle (with $\mathbb{Q}$-coefficients) $z_{K_{\widetilde {HG}}^{\circ }}$ to get an element in the rational arithmetic Chow group,

(8.3)\begin{equation} \widehat{z}_{K_{\widetilde{HG}}^{\circ}}= (z_{K_{\widetilde{HG}}^{\circ}},g_{z_{K_{\widetilde{HG}}^{\circ}} })\in \widehat{{\mathrm{Ch}}}{}^{n-1}(\mathcal{M}_{K_{\widetilde{HG}}^{\circ}}(\widetilde{HG}))_\mathbb{Q}. \end{equation}

The Green's current is not unique. We shall work in the following with an arbitrary but fixed choice.

Let

(8.4)\begin{equation} \mathscr{H}_{K_{\widetilde{HG}}^{\circ}}^{\mathrm{spl},\Phi} \subset \mathscr{H}_{K_{\widetilde{HG}}^{\circ}}= \mathscr{H}(\widetilde{HG}(\mathbb{A}_f),K_{\widetilde{HG}}^{\circ}) \end{equation}

be the partial Hecke algebra spanned by completely decomposed pure tensors of the form $f=\bigotimes _p f_p\in \mathscr {H}_{K_{\widetilde {HG}}^{\circ }}$, where $f_p=\phi _p\otimes \bigotimes _{v|p}f_v$, as in Definition 7.7, with $\phi _p=\mathbf {1}_{K_{Z^{\mathbb{Q}}, p}}$ for all $p$ and where $f_v=\mathbf {1}_{{K^{\circ }_{H\times G,v}}}$ unless $v\in \Sigma ^{\mathrm {spl}, \Phi }$. Here $\Sigma ^{\mathrm {spl}, \Phi }$ is as in (5.7). We have

\[ \mathscr{H}_{K_{\widetilde{HG}}^{\circ}}^{\mathrm{spl},\Phi}\simeq \bigotimes_{v\in \Sigma^{\mathrm{spl}, \Phi}}\mathscr{H}_{{HG},K^{\circ}_{{HG},v}}. \]

By Lemma 4.3(ii), we have $\Sigma ^{\mathrm {spl}, \Phi }\supset \Sigma ^{\deg =1}:=\{v\in \Sigma ^{\mathrm {spl}} \mid \deg _\mathbb{Q} v=1\}$ and hence the algebra $\mathscr {H}_{K_{\widetilde {HG}}^{\circ }}^{\mathrm {spl},\Phi }$ is not too small.

For $f\in \mathscr {H}_{K_{\widetilde {HG}}^{\circ }}^{\mathrm {spl},\Phi }$, we introduce via (5.11) a Hecke correspondence, hence an induced endomorphism $\widehat R(f)$ on the arithmetic Chow group $\widehat {{\mathrm {Ch}}}{}^{n-1}(\mathcal {M}_{K_{\widetilde {HG}}^{\circ }}(\widetilde {HG}))_\mathbb{C}$ (cf. [Reference Gillet and SouléGS90, 5.2.1]). Using the arithmetic intersection pairing (8.2), we define

(8.5)\begin{equation} \begin{aligned} \operatorname{\mathrm{Int}}^{\natural}(f) & := (\widehat R(f)\widehat z_{K_{\widetilde{HG}}^{\circ}}, \widehat z_{K_{\widetilde{HG}}^{\circ}})_{ \mathrm{GS}} ,\\ \operatorname{\mathrm{Int}}(f) & := \frac{1}{\tau(Z^{\mathbb{Q}})[E:F]}\cdot\operatorname{\mathrm{Int}}^{\natural}(f). \end{aligned} \end{equation}

Here

\[ \tau(Z^{\mathbb{Q}}) := \operatorname{vol}(Z^{\mathbb{Q}}(\mathbb{A}_f)/Z^{\mathbb{Q}}(\mathbb{Q}) ). \]

Conjecture 8.2 (Global conjecture, trivial level structure) Let $f=\bigotimes _p f_p\in \mathscr {H}_{K_{\widetilde {HG}}^{\circ }}^{\mathrm {spl},\Phi }$, and let $f'=\bigotimes _{v}f'_v\in \mathscr {H}(G'(\mathbb{A}_{F_0}))$ be a Gaussian test function such that $\bigotimes _{v<\infty }f'_v$ is a smooth transfer of $f$. Then

\[ \operatorname{\mathrm{Int}}(f)=-\partial J(f')-J(f'_{\mathrm{corr}}), \]

where $f'_{\mathrm {corr}}\in C_c^{\infty }(G'(\mathbb{A}_{F_0}))$ is a correction function. Furthermore, we may choose $f'$ such that $f'_{\mathrm {corr}}=0.$

Remark 8.3 The notion of smooth transfer at each individual place $v$ depends on the choice of transfer factors, and of Haar measures on various groups. However, the validity of the conjecture does not depend on these choices (use the product formula (7.14)).

This conjecture has the following drawback. Since we cannot impose any regular support assumptions on functions in $\mathscr {H}_{K_{\widetilde {HG}}^{\circ }}^{\mathrm {spl},\Phi }$, the left-hand side of the asserted equality may involve self-intersection numbers, and these are difficult to calculate explicitly. Analogously, on the right-hand side, the terms in Zydor's truncation that are not regular-semisimple orbital integrals are more delicate. Nevertheless, assuming a spectral decomposition of $J(f',s)$ that generalizes the case of special test functions in Propositions 7.2 and (7.6), Conjecture 8.2 relates the intersection number $\operatorname {\mathrm {Int}}(f)$ to the first derivative of $L$-functions in (the AGGP) Conjectures 6.10 and 6.12.

8.2 The global conjecture, non-trivial level structure

In this subsection we use the integral models of the Shimura varieties with deeper level structures depending on the choice of a function $\mathbf {m}$ as in (5.8). Note that the models $\mathcal {M}_{K_{\widetilde {HG}}^{\mathbf{m}}}(\widetilde {HG})$ are not regular in the fibers over places lying above the support of $\mathbf {m}$ (in effect, we are taking here the product of two copies of the Drinfeld moduli scheme). Therefore, the Gillet–Soulé pairing (8.2) is not defined for them. However, under certain hypotheses that ensure that our physical cycle $z_K$ and its physical transform under a Hecke correspondence do not intersect in the generic fiber, we can define a naive intersection number for them, by the usual derived tensor product formula.

Similarly to the case with trivial level structure, we obtain a cycle (with $\mathbb{Q}$-coefficients) $z_{K_{\widetilde {HG}}^{\mathbf{m}}} =\operatorname {vol}(K_{\widetilde H}^{\mathbf {m} \prime } )[\mathcal {M}_{K_{\widetilde H}^{\mathbf {m} \prime }}(\widetilde H)]$ on $\mathcal {M}_{K_{\widetilde {HG}}^{\mathbf {m}}}(\widetilde {HG})$ (cf. (8.1)) and, again, we denote by the same symbol its class in the Chow group. We choose a Green's current $g_{z_{K_{\widetilde {HG}}^{\mathbf{m}}} }$ of the cycle (with $\mathbb{Q}$-coefficients) $z_{K_{\widetilde {HG}}^{\mathbf{m}}}$ to get an element in the arithmetic Chow group,

(8.6)\begin{equation} \widehat{z}_{K_{\widetilde{HG}}^{\mathbf{m}}} = (z_{K_{\widetilde{HG}}^{\mathbf{m}}},g_{z_{K_{\widetilde{HG}}^{\mathbf{m}}}})\in \widehat{{\mathrm{Ch}}}{}^{n-1}(\mathcal{M}_{K_{\widetilde{HG}}^{\mathbf{m}}}(\widetilde{HG}))_\mathbb{Q}. \end{equation}

Let

(8.7)\begin{equation} \mathscr{H}_{{K_{\widetilde{HG}}^{\mathbf{m}}}}^{\mathrm{spl},\Phi} := \mathscr{H}(\widetilde{HG}(\mathbb{A}_f),{{K_{\widetilde{HG}}^{\mathbf{m}}}})^{\mathrm{spl},\Phi} \end{equation}

be the partial Hecke algebra spanned by completely decomposed pure tensors of the form $f=\bigotimes _p f_p\in \mathscr {H}_{K_{\widetilde {HG}}^{\mathbf{m}}}$, where $f_p=\phi _p\otimes \bigotimes _{v|p}f_v$, as in Definition 7.7, with $\phi _p=\mathbf {1}_{K_{Z^{\mathbb{Q}}, p}}$ for all $p$, and where $f_v=\mathbf {1}_{K^{\circ }_{H\times G,v}}$ unless $v\in \Sigma ^{\mathrm {spl}, \Phi }$. We have

\[ \mathscr{H}_{K_{\widetilde{HG}}^{\mathbf{m}}}^{\mathrm{spl},\Phi}\simeq \bigotimes_{ v\in \Sigma^{\mathrm{spl}, \Phi}}\mathscr{H}_{{HG},K^{\mathbf{m}}_{{HG},v}}. \]

By Lemma 4.3(ii), we have $\Sigma ^{\mathrm {spl}, \Phi }\supset \Sigma ^{S,\deg =1}:=\{v\in \Sigma ^{\mathrm {spl}} \mid v\nmid S\ \text {and}\ \deg _\mathbb{Q} v=1\}$, for any finite set $S$ of places of $\mathbb{Q}$ such that the places of $F_0$ above $S$ contain the support of $\mathbf {m}$.

To any $f \in \mathscr {H}_{{K_{\widetilde {HG}}^{\mathbf{m}}}}^{\mathrm {spl},\Phi }$ we associate via (5.11) a Hecke correspondence on $\mathcal {M}_{K_{\widetilde {HG}}^{\mathbf{m}}}(\widetilde {HG})$.

Definition 8.4 Let $\lambda$ be a non-archimedean place of $F_0$, of residue characteristic $\ell$. Let $f_\ell \in C_c^{\infty }(\widetilde {HG}(\mathbb{Q}_\ell ))$ be completely decomposed, that is, $f_\ell = \phi _{\ell }\otimes \bigotimes _{v\mid \ell }f_v$ (cf. Definition 7.7). Then $f_\ell$ is said to have regular support at $\lambda$ if $\operatorname {supp} f_\lambda \subset G_W(F_{0,\lambda })_\mathrm {rs}$. If $f=\bigotimes _p f_p\in \mathscr {H}_{{K_{\widetilde {HG}}^{\mathbf{m}}}}^{\mathrm {spl},\Phi }$ is a completely decomposed pure tensor, then $f$ has regular support at $\lambda$ if $f_\ell$ has regular support at $\lambda$.

Theorem 8.5 Let $f=\bigotimes _p f_p\in \mathscr {H}_{{K_{\widetilde {HG}}^{\mathbf{m}}}}^{\mathrm {spl},\Phi }$ be a completely decomposed pure tensor. Assume that $f$ has regular support at some place $\lambda$ of $F_0$. Then the following statements on the support of the intersection of the cycles $z_{K_{\widetilde {HG}}^{\mathbf{m}}}$ and $R(f) z_{K_{\widetilde {HG}}^{\mathbf{m}}}$ of $\mathcal {M}_{K_{\widetilde {HG}}^{\mathbf{m}}}(\widetilde {HG})$ hold.

  1. (i) The support does not meet the generic fiber.

  2. (ii) Let $\nu$ be a place of $E$ lying over a place of $F_0$ which splits in $F$. Then the support does not meet the special fiber $\mathcal {M}_{K_{\widetilde {HG}}^{\mathbf{m}}}(\widetilde {HG})\otimes _{O_E} \kappa _{\nu }$.

  3. (iii) Let $\nu$ be a place of $E$ lying over a place of $F_0$ which does not split in $F$. Then the support meets the special fiber $\mathcal {M}_{K_{\widetilde {HG}}^{\mathbf{m}}}(\widetilde {HG})\otimes _{O_E} \kappa _{\nu }$ only in its basic locus.Footnote 18

Remark 8.6 Let $F_0 = \mathbb{Q}$ and assume that $\mathcal {M}_{K_{\widetilde {HG}}^{\mathbf{m}}}(\widetilde {HG})$ is non-compact. Then the closure in the toroidal compactification [Reference HowardHow15, § 2] of $\mathcal {M}_{K_{\widetilde {HG}}^{\mathbf{m}}}(\widetilde {HG})$ of the support of the intersection of the cycles $z_{K_{\widetilde {HG}}^{\mathbf{m}}}$ and $R(f) z_{K_{\widetilde {HG}}^{\mathbf{m}}}$ does not meet the boundary. Indeed, this follows from Theorem 8.5(iii) because the basic locus of $\mathcal {M}_{K_{\widetilde {HG}}^{\mathbf{m}}}(\widetilde {HG})\otimes _{O_E} \kappa _{\nu }$ does not meet the boundary.

The proof of Theorem 8.5 will be based on the following lemma.

Lemma 8.7 Let $k$ be an algebraically closed field which is an $O_E$-algebra. Let $(A_0, \iota _0)$ be an abelian variety with $O_F$-action with Kottwitz condition of signature $((0, 1)_{\varphi \in \Phi })$ (cf. (3.8)), and let $(A, \iota )$ be an abelian variety with $O_F$-action with Kottwitz condition of type $r$ as in Remark 3.6(i). Assume there exists an $F$-linear isogeny

\[ A_0^{n}\longrightarrow A. \]

Then $k$ is of positive characteristic $p$. Let $\nu$ be the corresponding place of $E$. The place $v_0$ of $F_0$ induced by $\nu$ is non-split in $F$, and the isogeny classes of $A_0$ and $A$ only depend on the CM type $\Phi$. If $v_0$ is the only place of $F_0$ above $p$, then $A_0$ and $A$ are supersingular.

Proof. We prove the first statement by contradiction. Assume that $k$ is of characteristic zero. Then an isogeny as above induces an $F\otimes k$-linear isomorphism

\[ \operatorname{Lie} (A_0)^{n}{\tilde{\to}} \operatorname{Lie} A. \]

Such an isomorphism cannot exist due to the different Kottwitz conditions on $A_0^{n}$ and $A$.

Now suppose that the place $v_0$ of $F_0$ induced by $\nu$ is split in $F$. Then there is a splitting of the $p$-divisible group of $A_0^{n}$ (respectively, $A$) according to the two places of $F$ above $F_0$,

\[ A_0^{n}[p^{\infty}]=X_0^{(1)}\times X_0^{(2)}, \quad A[p^{\infty}]=X^{(1)}\times X^{(2)}, \]

An $F$-linear isogeny as above induces isogenies of $p$-divisible groups,

\[ X_0^{(1)}\longrightarrow X^{(1)}, \quad X_0^{(2)} \longrightarrow X^{(2)}. \]

However, by the Kottwitz conditions, the dimension of $X_0^{(1)}$ (respectively, $X_0^{(2)}$) is divisible by $n$, whereas the dimension of $X^{(1)}$ (respectively, $X^{(2)}$) is $\equiv \pm 1\bmod n$. Hence such an isogeny cannot exist, hence $v_0$ is non-split in $F$.

The rational Dieudonné module $M_0$ of $A_0[p^{\infty }]$ is a free $F\otimes W(k)$-module of rank $1$. Consider its decomposition according to the places $w$ of $F$ above $p$,

\[ M_0=\bigoplus_w M_{0, w}. \]

Then each summand is isoclinic. More precisely, if $w$ lies above a non-split place of $F_0$ (such as the place $w_0$ induced by $\nu$), then $M_{0, w}$ is isoclinic of slope $1/2$; and if $\overline w\neq w$, then the slope of $M_{0, w}$ is equal to $a_w/d_w$, where

\[ a_w := \#\{\,\varphi\in \Phi \mid w_\varphi=w \,\}, \quad d_w := [F_w:\mathbb{Q}_p]. \]

Here we chose an embedding of $\overline{\mathbb{Q}}$ into $\overline{\mathbb{Q}}_p$ so that any $\varphi \in \Phi$ induces a place $w_\varphi$ of $F$.

Proof Proof of Theorem 8.5 (i) Suppose that

\[ (A_0,\iota_0,\lambda_0,A^{\flat},\iota^{\flat},\lambda^{\flat},\overline\eta^{\flat}, \varphi^{\flat}, A,\iota,\lambda,\overline\eta,\varphi)\in \mathcal{M}_{K^\textbf{m}_{\widetilde{HG}}}(\widetilde{HG})(k) \]

is a point in the support, where $k$ is an algebraically closed field of characteristic zero. Then $(A, \iota ,\lambda )=(A^{\flat } \times A_0,\iota ^{\flat } \times \iota _0,\lambda ^{\flat } \times \lambda _0(u))$, and there exists $g\in (H\times G)(\mathbb{A}_{F_0,f})_\mathrm {rs}$ such that there is an isogeny

\[ \phi\colon A= A^{\flat}\times A_0\longrightarrow A \]

which makes the diagram

(8.8)

commute. From the splitting $A= A^{\flat }\times A_0$, we also have

\[ u\colon A_0\longrightarrow A, \]

which makes the diagram

(8.9)

commute, where $\eta _0$ is defined in the obvious way. We consider the homomorphism

\[ (\phi^{i} u)_{0\leq i\leq n-1}: A_0^{n}\longrightarrow A \]

whose $i$th component is $\phi ^{i} u: A_0\to A$. We claim that this is an isogeny. It suffices to show that its induced map on any rational Tate module is an isomorphism. This follows by the commutativity of the diagram (

8.8

) from the regular semisimplicity of $g$. This conclusion contradicts Lemma 8.7.

(ii), (iii) Now let $(A_0,\iota _0,\lambda _0,A^{\flat },\iota ^{\flat },\lambda ^{\flat },\overline \eta ^{\flat }, \varphi ^{\flat }, A,\iota ,\lambda ,\overline \eta , \varphi )\in \mathcal {M}_{K^\textbf {m}_{\widetilde {HG}}}(\widetilde {HG})(k)$ be a point in the support, where $k$ is an algebraically closed field of positive characteristic $p$. When there exists $g\in (H\times G)(\mathbb{A}_{F_0,f}^{p})_\mathrm {rs}$ such that there exists a commutative diagram (8.8) (with an upper index $^{p}$ added everywhere), the argument is as before, by reduction to Lemma 8.7.

Now assume that the place $v_0$ of $F_0$ induced by $\nu$ is split in $F$ and that there exists $g\in G(F_{0, v_0})_\mathrm {rs}$ such that there is an isogeny

\[ \phi\colon A= A^{\flat}\times A_0\longrightarrow A \]

which makes a diagram analogous to (8.8) commute. To explain this diagram, we use a compatible system of Drinfeld level structures,

\[ \widetilde \varphi\colon \Lambda_{w_0}[\pi_{w_0}^{-1}]/\Lambda_{w_0}\longrightarrow T_{w_0}(A_0, A) , \]

which induces for every $m$ a Drinfeld level ${w^{m}_0}$-structure on the $w_0^{m}$-torsion subgroup,

\[ \varphi\colon \pi^{-m}_{w_0}\Lambda_{w_0}/\Lambda_{w_0}\longrightarrow T_{w_0}(A_0, A) [w_0^{m}]. \]

Then the analog of (8.8) is the following commutative diagram for sufficiently large $m'$:

(8.10)

From the splitting $A= A^{\flat }\times A_0$, we also have

\[ u\colon A_0\longrightarrow A, \]

which makes the diagram

(8.11)

commute, where $\widetilde \varphi _0$ is the limit over $m$ of the homomorphisms (4.22). We consider the homomorphism

\[ (\phi^{i} u)_{0\leq i\leq n-1}: A_0^{n}\longrightarrow A \]

whose $i$th component is $\phi ^{i} u: A_0\to A$. We again claim that this is an isogeny, which would contradict Lemma 8.7. It suffices to show that its induced map on the $p$-divisible group is an isogeny. This follows by the commutativity of diagram (

8.10

) from the regular semisimplicity of $g$.

Let us assume that $f\in \mathscr {H}_{{K_{\widetilde {HG}}^{\mathbf{m}}}}^{\mathrm {spl},\Phi }$ is a completely decomposed pure tensor which has regular support at some place $\lambda$ of $F_0$. Then, by Theorem 8.5(i), the generic fibers of the cycles $z_{K_{\widetilde {HG}}^{\mathbf{m}}}$ and $R(f) z_{K_{\widetilde {HG}}^{\mathbf{m}}}$ do not intersect, and we may define

(8.12)\begin{equation} \begin{aligned} \operatorname{\mathrm{Int}}_\nu^{\natural} (f) & :=\langle\widehat R(f)\widehat z_{K_{\widetilde{HG}}^{\mathbf{m}}}, \widehat z_{K_{\widetilde{HG}}^{\mathbf{m}}}\rangle_{{\nu}} \log q_\nu ,\\ \operatorname{\mathrm{Int}}(f) & :=\frac{1}{\tau(Z^{\mathbb{Q}}) [E:F]}\sum_\nu \operatorname{\mathrm{Int}}_\nu^{\natural} (f). \end{aligned} \end{equation}

Here the first quantity is defined for a non-archimedean place $\nu$ through the Euler–Poincaré characteristic of a derived tensor product on $\mathcal {M}_{K_{\widetilde {HG}}^{\mathbf{m}}}(\widetilde {HG})\otimes _{O_E} O_{E, (\nu )}$ (cf. [Reference Gillet and SouléGS90, 4.3.8(iv)]). Note that the intersection numbers are indeed defined for a non-archimedean place because the intersection of the cycles $z_{K_{\widetilde {HG}}^{\mathbf{m}}}$ and $R(f)z_{K_{\widetilde {HG}}^{\mathbf{m}}}$ avoids all fibers of $\mathcal {M}_{K_{\widetilde {HG}}^{\mathbf{m}}}(\widetilde {HG})$ over places $\nu$ lying over the support of $\mathbf {m}$, as follows from Theorem 8.5(ii). Therefore the intersection of these cycles takes place in the regular locus of $\mathcal {M}_{K_{\widetilde {HG}}^{\mathbf{m}}}(\widetilde {HG})$ (cf. Theorem 8.5), and hence the Euler–Poincaré characteristic is finite. For an archimedean place $\nu$, the last quantity is defined by the archimedean component of the arithmetic intersection theory and we have set $\log q_\nu :=[E_\nu :\mathbb{R}]=2$ (cf. [Reference Gillet and SouléGS90]).

Conjecture 8.8 (Global conjecture, non-trivial level structure) Let $f=\bigotimes _p f_p\in \mathscr {H}_{{K_{\widetilde {HG}}^{\mathbf{m}}}}^{\mathrm {spl},\Phi }$ be a completely decomposed pure tensor and let $f'=\bigotimes _{v}f'_v\in \mathscr {H}(G'(\mathbb{A}))$ be a Gaussian test function such that $\bigotimes _{v<\infty }f'_v$ is a smooth transfer of $f$. Assume that $f$ has regular support at some place $\lambda$ of $F_0$. Then

\[ \operatorname{\mathrm{Int}}(f)=-\partial J(f')-J(f'_{\mathrm{corr}}), \]

where $f'_{\mathrm {corr}}\in C_c^{\infty }(G'(\mathbb{A}))$ is a correction function. Furthermore, $f'=\bigotimes _{v}f'_v$ may be chosen such that $f'$ has regular support at $\lambda$ and that $f'_{\mathrm {corr}}=0.$

Remark 8.9 Part of the conjecture asserts that a change of the Green's current is compensated by a change of the correction function $f'_{\mathrm {corr}}$.

Remark 8.10 Similar to the case of trivial level structure, when there exists a split place $v$ such that $f_v'$ satisfies the assumption in Proposition 7.2, by (7.6), Conjecture 8.8 relates the intersection number $\operatorname {\mathrm {Int}}(f)$ to the first derivative of $L$-functions in (the AGGP) Conjectures 6.10 and 6.12. The hypothesis on the existence of such a split place $v$ could be dropped once a full spectral decomposition of $J(f',s)$ for all test functions is available.

Note that here the right-hand side is well defined (cf. § 7.1). We also point out that if $f\neq 0$ has regular support at some place $\lambda$ of $F_0$, then $\lambda$ must be in $\operatorname {supp} \mathbf {m}$ (in particular, $\lambda$ is split in $F$). Since we will not need this statement, we omit the proof.

Lemma 8.11

  1. (i) Let $u \in \operatorname {supp} \mathbf {m}$ be a place above $p$ (in particular, $u$ is split in $F$). There exists a non-zero function $f_p \in \mathscr {H}_{{K_{\widetilde {HG}}^{\mathbf{m}},p}}$ that has regular support at $u$.

  2. (ii) For any $f_p\in \mathscr {H}_{{K_{\widetilde {HG}}^{\mathbf{m}},p}}$ with regular support at the place $u$ above $p$, there exists a transfer $(f'_v)_{v\mid p}$ such that $f'_{u}$ has regular support at $u$.

Proof. Let $m$ be a positive integer. At a place $u$ split in $F$, by choosing a basis of a self-dual lattice in $W^{\flat }\otimes F_{0,u}$ (respectively, $W\otimes F_{0,u}$), we may identify $H(F_{0,u})=\mathrm {GL}_{n-1}(F_{0,u})$ and $G(F_{0,u})=\mathrm {GL}_{n}(F_{0,u})$, and $K^{m}_{H,u}$ (respectively, $K^{m}_{G,u}$) with the principal congruence subgroups of level $m$. We may choose the basis of the lattice in $W\otimes F_{0,u}$ by adding the special vector $u$ to the basis of the lattice in $W^{\flat }\otimes F_{0,u}$. Therefore we may further assume that the embedding $H(F_{0,u})\hookrightarrow G(F_{0,u})$ has the property that $K^{m}_{H,u}=K^{m}_{G,u}\cap H(F_{0,u})$. Moreover, the stabilizers of the lattices are identified with $\mathrm {GL}_{n-1}(O_{F_0,u})$ and $\mathrm {GL}_{n}(O_{F_0,u})$ respectively, and $\mathrm {GL}_{n-1}\hookrightarrow \mathrm {GL}_{n}$ sends $h$ to $\operatorname {diag} (h,1)$.

Now let $f_p = \phi _{p}\otimes \bigotimes _{v\mid p}f_v$ be completely decomposed. It suffices to show that there exists a non-zero $f_u=f_{n-1,u}\otimes f_{n, u}\in \mathscr {H}((H\times G)(F_{0, u}), K^{m}_{H, u}\times K^{m}_{G, u})$ with regular support. We construct such a function by first setting $f_{n-1,u}=\mathbf {1}_{K^{m}_{H,u}}$. Note that the pair of functions $( f_{n-1,u}, f_{n, u})$ defines the function $\widetilde f_u\in \mathscr {H}(G(F_{0,u}), K^{m}_{G,u})$ by ‘contraction’ under the map

(8.13)

namely,

\[ \widetilde f_u(g)=\int_{H(F_{0,u})} f_{n-1,u}(h) f_{n,u}(hg)\,dh. \]

We then have $\widetilde f_u=\operatorname {vol}(K^{m}_{H,u}) f_{n,u}.$ Then the function $f_u=f_{n-1,u}\otimes f_{n, u}$ has regular support (with respect to the $H\times H$-action) if (and only if) $\widetilde f_u$ or, equivalently, $f_{n,u}$ has regular support (with respect to the conjugation action of $H$). This holds because the inverse image of $G(F_{0,u})_{\mathrm {rs}}$ under the contraction map (

8.13

) is exactly $(H\times G)(F_{0,u})_{\mathrm {rs}}$.

We now choose $f_{n,u}$ supported in $\mathrm {GL}_{n}(O_{F_0,u})$. Recall that $G(F_{0,u})_{\mathrm {rs}}$ is defined by the equation $\Delta \neq 0$ where $\Delta$ is a polynomial in the entries of $\mathrm {GL}_n$ with coefficients in $\mathbb{Z}$, and it is easy to exhibit an element $g\in \mathrm {GL}_{n}(O_{F_0,u})$ such that $\Delta (g)\in \mathbb{Z}^{\times }=\{\pm 1\}$.Footnote 19 Then the function $f_{n,u}=\mathbf {1}_{K^{m}_{G,u}g K^{m}_{G,u}}$ has regular support. Indeed, consider the reduction map $\mathrm {GL}_{n}(O_{F_0,u})\to \mathrm {GL}_{n}(O_{F_0,u}/\varpi _u^{m})$ where $\varpi _u$ is a uniformizer. It is easy to see that for $k,k'\in 1+\varpi _u^{m} M_{n}(O_{F_0,u})$ we have $\Delta (kgk')\equiv \Delta (g)\equiv \pm 1\mod \varpi _u^{m}$. In particular, we have $K^{m}_{G,u}g K^{m}_{G,u}\subset \mathrm {GL}_{n}(O_{F_0,u})_\mathrm {rs}$. This completes the proof of the first part.

To show the second part, by a reduction process similar to the first part (cf. [Reference ZhangZha14a, Proposition 2.5]), it suffices to work with the inhomogeneous version, that is, to show that there exists a smooth transfer $f'_u$ with support in $S_n(F_{0,u})_\mathrm {rs}$. We may identify $S_n(F_{0,u})=\mathrm {GL}_n(F_{0,u})$ and then the notions of regular semisimplicity on $S_n(F_{0, u})$ and on $G(F_{0,u})$ coincide. This completes the proof.

The left-hand side of (8.12) can be localized, that is, we can write it as a sum over all non-archimedean places,

(8.14)\begin{equation} \operatorname{\mathrm{Int}}(f) = \sum_v \operatorname{\mathrm{Int}}_v(f) , \end{equation}

where

(8.15)\begin{equation} \operatorname{\mathrm{Int}}_v(f) :=\frac{1}{\tau(Z^{\mathbb{Q}})\cdot [E:F]} \sum_{\nu\mid v} \operatorname{\mathrm{Int}}^{\natural}_\nu(f). \end{equation}

By Lemma 8.11, the smooth transfer $f'$ of $f$ can be chosen such that $f'$ has regular support at $\lambda$, which we assume from now on. Then also the right-hand side of Conjecture 8.8 can be written as a sum of local contributions of each place of $F_0$ (cf. (7.12) for $\partial J(f')$ and (7.9) for $J(f'_{\mathrm {corr}}))$.

Proposition 8.12 In the situation of Conjecture 8.8, let $v_0$ be a place of $F_0$ that is split in $F$. Then

\[ \operatorname{\mathrm{Int}}_{v_0}(f)=\partial J_{v_0}(f')=0. \]

Proof. In Lemma 7.12 we have proved $\partial J_{v_0}(f')=0$. Now $\operatorname {\mathrm {Int}}_{v_0}(f)=0$ follows from Theorem 8.5(ii).

In the next subsection we will formulate a semi-global conjecture for each non-split place $v$ (including archimedean ones) which refines Conjecture 8.8.

8.3 The semi-global conjecture

Let $v_0$ be a place of $F_0$ above the place $p\leq \infty$ of $\mathbb{Q}$. By Proposition 8.12, from now on we may and do assume that $v_0$ is non-split in $F$.

Now assume that $v_0$ is non-archimedean. We assume that $v_0$ is either of hyperspecial level type or of AT parahoric level type, in the sense of § 4. We also take up the notation of §§ 4.1 and 4.4 and denote, for any place $\nu$ of $E$ lying over $v_0$, by $\mathcal {M}_{K_{\widetilde G}}(\widetilde G)_{(\nu )}$, $\mathcal {M}_{K_{\widetilde H}}(\widetilde H)_{(\nu )}$, and $\mathcal {M}_{K_{\widetilde {HG}}}(\widetilde {HG})_{(\nu )}$ the corresponding semi-global moduli stacks over $\operatorname {Spec} O_{E,(\nu )}$. Let $K_{\widetilde {HG}} = K_{\widetilde {HG}}^{p} \times K_{\widetilde {HG},p} \subset \widetilde {HG}(\mathbb{A}_f)$ be as in § 4.1 (respectively, § 4.4). Let

(8.16)\begin{equation} \mathscr{H}_{K_{\widetilde{HG}}}^{p} \subset \mathscr{H}_{{K_{\widetilde{HG}}}} = \mathscr{H}(\widetilde{HG}(\mathbb{A}_f),{{K_{\widetilde{HG}}}}) \end{equation}

be the partial Hecke algebra spanned by completely decomposed pure tensors of the form $f=\bigotimes _\ell f_\ell \in \mathscr {H}_{K_{\widetilde {HG}}}$, where $f_\ell =\phi _\ell \otimes \bigotimes _{v|\ell }f_v$, as in Definition 7.7, with $\phi _\ell =\mathbf {1}_{K_{Z^{\mathbb{Q}}, \ell }}$ for all $\ell$, and where

\[ f_p=\mathbf{1}_{K_{\widetilde{HG}, p}}. \]

We note that this defines a bigger Hecke algebra than (8.7) when ${K_{\widetilde {HG}}}={K_{\widetilde {HG}}^{\mathbf{m}}}$,

(8.17)\begin{equation} \mathscr{H}_{{K_{\widetilde{HG}}}}^{p} \supset \mathscr{H}_{{K_{\widetilde{HG}}^{\mathbf{m}}}}^{\mathrm{spl},\Phi}. \end{equation}

Let $f=\bigotimes _{\ell } f_\ell \in \mathscr {H}_{{K_{\widetilde {HG}}}}^{p}$ be completely decomposed with regular support at some place $\lambda$. We define, as before in (8.15),

(8.18)\begin{equation} \begin{aligned} \operatorname{\mathrm{Int}}_\nu^{\natural} (f) & :=\langle\widehat R(f)\widehat z_{K_{\widetilde{HG}}}, \widehat z_{K_{\widetilde{HG}}}\rangle_{{\nu}} \log q_\nu ,\\ \operatorname{\mathrm{Int}}_{v_0}(f) & :=\frac{1}{\tau(Z^{\mathbb{Q}}) [E:F]}\sum_{\nu|v_0} \operatorname{\mathrm{Int}}_\nu^{\natural} (f) , \end{aligned} \end{equation}

where again the contribution of the place $\nu$ is defined through the Euler–Poincaré characteristic of a derived tensor product on $\mathcal {M}_{K_{\widetilde {HG}}}(\widetilde {HG})_{(\nu )}$. This extends definition (8.15) to the bigger Hecke algebra $\mathscr {H}_{{K_{\widetilde {HG}}}}^{p}$.

We proceed similarly for an archimedean place $v_0\in \operatorname {Hom}(F_0,\mathbb{R})$. Denote, for any place $\nu$ of $E$ lying over $v_0$, by $\mathcal {M}_{K_{\widetilde G}}(\widetilde G)_{(\nu )}$, $\mathcal {M}_{K_{\widetilde H}}(\widetilde H)_{(\nu )}$, and $\mathcal {M}_{K_{\widetilde {HG}}}(\widetilde {HG})_{(\nu )}$ the corresponding complex analytic spaces (in fact, orbifolds). Note that the Green's current $g_{z_{K_{\widetilde {HG}} }}$ is the multi-set $g_{z_{K_{\widetilde {HG}} },\nu }$ indexed by $\nu \in \operatorname {Hom}(E,\mathbb{C})$. We define

(8.19)\begin{equation} \begin{aligned} \operatorname{\mathrm{Int}}_\nu^{\natural} (f) & :=\langle\widehat R(f)\widehat z_{K_{\widetilde{HG}}}, \widehat z_{K_{\widetilde{HG}}}\rangle_{{\nu}} \log q_\nu ,\\ \operatorname{\mathrm{Int}}_{v_0}(f) & :=\frac{1}{\tau(Z^{\mathbb{Q}}) [E:F]}\sum_{\nu|v_0} \operatorname{\mathrm{Int}}_\nu^{\natural} (f), \end{aligned} \end{equation}

where the first quantity is defined before Conjecture 8.8.

A refinement of Conjecture 8.8 is now given by the following statement.

Conjecture 8.13 (Semi-global conjecture) Fix a place $v_0$ of $F_0$ above a place $p\leq \infty$ of $\mathbb{Q}$. Let $f=\bigotimes _{\ell } f_\ell \in \mathscr {H}_{{K_{\widetilde {HG}}}}^{p}$ ($\mathscr {H}_{{K_{\widetilde {HG}}}}$ if $p$ is archimedean) be completely decomposed, and let $f'=\bigotimes _{v}f'_v\in \mathscr {H}(G'(\mathbb{A}_{F_0}))$ be a Gaussian test function such that $\bigotimes _{v < \infty }f'_v$ is a smooth transfer of $f$. Assume that for some $\ell$ prime to $v_0$ and some place $\lambda$ above $\ell$, the function $f$ has regular support at $\lambda$ in the sense of Definition 8.4 and that $f'$ has regular support at $\lambda$ in the sense of Definition 7.5.

  1. (i) Assume that $v_0$ is non-archimedean of hyperspecial type (cf. § 4.1), and that $f_{v_0}'=\mathbf {1}_{G'(O_{F_0,v_0})}$. Then

    \[ \operatorname{\mathrm{Int}}_{v_0}(f)=-\partial J_{v_0}(f'). \]
  2. (ii) Assume that $v_0$ is archimedean, or non-archimedean of AT type (cf. § 4.4). Then

    \[ \operatorname{\mathrm{Int}}_{v_0}(f)=-\partial J_{v_0}(f')-J(f'_{\mathrm{corr}}[v_0]), \]
    where $f'_{\mathrm {corr}}[v_0]=\bigotimes _{v}f'_{\mathrm {corr}, v}$, with $f'_{\mathrm {corr}, v}=f'_v$ for $v\neq v_0$, is a correction function. Furthermore, $f'$ may be chosen such that $f'_{\mathrm {corr}}[v_0]$ is zero.

Theorem 8.14 The semi-global Conjecture 8.13 for all places $v_0$ implies the global Conjecture 8.8.

Proof. By (7.12), and (8.14), it follows from Proposition 8.12 for split places $v_0$ of $F_0$ and the semi-global Conjecture 8.13 for non-split places $v_0$ that

\[ \operatorname{\mathrm{Int}}(f)= - \partial J(f')-\sum_{v_0\text{ bad}}J(f'_{\mathrm{corr}}[v_0]) , \]

where the sum runs over a finite set of places $v_0$ in Conjecture 8.13(ii). Here we note that by (8.17) we may apply the semi-global Conjecture 8.13 for the given test function $f$ in the global Conjecture 8.8. This proves Conjecture 8.8 by taking

\[ f'_\mathrm{corr}=\sum_{v_0\text{ bad}}f'_{\mathrm{corr}}[v_0].\]

Working towards the semi-global conjecture, there are the following results.

Theorem 8.15 Conjecture 8.13(i) holds when $n\leq 3$.

Proof. Let $p$ denote the residue characteristic of $v_0$, and as previously in the paper, let $\mathcal {V}_p$ denote the set of places of $F_0$ lying above $p$. It suffices to show that the result holds if we assume the AFL conjecture, which is known under these circumstances (see [Reference ZhangZha12a, Theorem 5.5]; cf. also [Reference MihatschMih17]).

We imitate the proof of [Reference ZhangZha12a, Theorem 3.11]. More precisely, we consider the non-archimedean uniformization along the basic locus,

(8.20)\begin{equation} (\mathcal{M}_{(\nu)} \otimes_{O_{E,(\nu)}} O_{\breve E_\nu})^\hat \, = \widetilde{HG}\vphantom{G}'(\mathbb{Q}) \left\backslash\, \big[\mathcal{N}'\times \widetilde{HG} (\mathbb{A}_f^{p}) /K_{\widetilde{HG}}^{p} \big]\right.. \end{equation}

Here the hat on the left-hand side denotes the completion along the basic locus in the special fiber of $\mathcal {M}_{(\nu )}:=\mathcal {M}_{K_{\widetilde {HG}}}(\widetilde {HG})_{(\nu )}$. The group $\widetilde {HG}\vphantom {G}'$ is an inner twist of $\widetilde {HG}$. More precisely, the group $\widetilde {HG}\vphantom {G}'$ is associated to the pair of hermitian spaces $(W^{\prime \flat },W')$, where $W^{\prime \flat }$ and $W'$ are negative definite at all archimedean places, and isomorphic to $W^{\flat }$ (respectively, $W$) locally at all non-archimedean places except at $v_0$. Furthermore, $\mathcal {N}'$ is the RZ space relevant in this situation. Using Lemma 8.16 below, can write

(8.21)\begin{equation} \mathcal{N}'\simeq (Z^{\mathbb{Q}}(\mathbb{Q}_p)/K_{Z^{\mathbb{Q}}, p})\times\mathcal{N}_{O_{\breve E_\nu}} \times \prod_{v\in \mathcal{V}_p\smallsetminus\{v_0\}}(H\times G)(F_{0,v})/(K_{H,v}\times K_{G,v}). \end{equation}

Here $\mathcal {N}_{O_{\breve E_\nu }}=\mathcal {N} \mathbin {\widehat \otimes }_{O_{\breve F_{w_0}}}O_{\breve E_\nu }$, where $\mathcal {N} = \mathcal {N}_{n-1} \times _{O_{\breve F_{w_0}}} \mathcal {N}_n$ is the relative RZ space of [Reference ZhangZha12a]. More precisely, the formal scheme in the uniformization theorem is the RZ space of polarized $p$-divisible groups with action by $O_{F,w_0}$ satisfying the Kottwitz condition (3.11) of signature $((1, n-1)_{\varphi _0}, (0, n)_{\varphi \in \Phi _{v_0}\smallsetminus \{\varphi _0\}})$. That this coincides with the relative RZ space of [Reference ZhangZha12a, § 2.2] follows from [Reference MihatschMih16, Theorem 4.1], where in the notation of this paper we take $E_0=F_{0,v_0}$, $E=F_{w_0}$, $(r, s)=(1, n-1)$, and $K=\mathbb{Q}_p$.

Therefore we may rewrite (8.20) as

(8.22)\begin{equation} (\mathcal{M}_{(\nu)} \otimes_{O_{E,(\nu)}} O_{\breve E_\nu})^\hat \, = \widetilde{HG}\vphantom{G}'(\mathbb{Q}) \Big\backslash \big[ \mathcal{N}_{O_{\breve E_\nu}} \times \widetilde{HG} (\mathbb{A}_f^{v_0}) /K_{\widetilde{HG}}^{v_0} \big]. \end{equation}

Here we denote for simplicity (even though $\widetilde {HG}$ is not an algebraic group over $F_0$)

\[ \widetilde{HG} (\mathbb{A}_f^{v_0})/K_{\widetilde{HG}}^{v_0} = \widetilde{HG} (\mathbb{A}_f^{p})/K_{\widetilde{HG}}^{p}\times (Z^{\mathbb{Q}}(\mathbb{Q}_p)/K_{Z^{\mathbb{Q}}, p})\times \prod_{v\in \mathcal{V}_p\smallsetminus\{v_0\}}(H\times G)(F_{0,v})/(K_{H,v}\times K_{G,v}). \]

There is also a similar uniformization of the basic locus of $\mathcal {M}_{K_{\widetilde H}}(\widetilde H)_{(\nu )}$ involving the twisted form $\widetilde H'$ of $H$.

By Theorem 8.5(iii), the intersection of the cycles is supported in the basic locus, and hence we can imitate the proof of [Reference ZhangZha12a, Theorem 3.9]. The difference is that here we have an extra central subgroup $Z^{\mathbb{Q}}$. By the same procedure as in that paper, we see that (8.19) can be written as a sum

(8.23)\begin{equation} \operatorname{\mathrm{Int}}_{v_0}(f) = \frac 1{[E:F]} \sum_{g \in \mathbb{O}(\widetilde{HG}(\mathbb{Q}))_\mathrm{rs}} \operatorname{Orb}(g, f^{p})\cdot \bigg(c(\phi_p) \prod_{v\in \mathcal{V}_p\smallsetminus\{v_0\}} \operatorname{Orb}(g,f_v)\bigg) \cdot \operatorname{\mathrm{Int}}_{v_0}^{\natural} (g) \end{equation}

(cf. Remark 7.8). Here the volume factor $\tau (Z^{\mathbb{Q}})=\operatorname {vol}(Z^{\mathbb{Q}}(\mathbb{A}_f)/Z^{\mathbb{Q}}(\mathbb{Q}))$ is canceled with the one in the definition of $\operatorname {\mathrm {Int}}_{v_0}(f)$, and

\[ \operatorname{Orb}(g, f^{p})=\prod_{\ell \nmid p} \operatorname{Orb}(g, f_\ell). \]

Also, we have set

\[ \mathbb{O}(\widetilde{HG}(\mathbb{Q}))_\mathrm{rs} := \widetilde H'(\mathbb{Q}) \backslash \widetilde{HG}\vphantom{G}'(\mathbb{Q})_\mathrm{rs} / \widetilde H'(\mathbb{Q}) , \]

and, for $g\in (H\times G)(F_{0, v_0})$, in analogy with (8.18),

(8.24)\begin{equation} \begin{aligned} \operatorname{\mathrm{Int}}^{\natural}_{v_0}(g) & :=\sum_{\nu|v_0}\operatorname{\mathrm{Int}}^{\natural}_\nu(g),\\ \operatorname{\mathrm{Int}}^{\natural}_\nu(g) & :=\langle \Delta(\mathcal{N}_{n-1,O_{\breve E_\nu}}), g\Delta(\mathcal{N}_{n-1,O_{\breve E_\nu}}) \rangle_{\mathcal{N}_{O_{\breve E_\nu}}} \log q_\nu. \end{aligned} \end{equation}

Now note that

(8.25)\begin{align} \sum_{\nu \mid v_0} \operatorname{\mathrm{Int}}^{\natural}_\nu(g)&= \sum_{\nu \mid v_0} \langle \Delta(\mathcal{N}_{n-1,O_{\breve E_\nu}}), g\Delta(\mathcal{N}_{n-1,O_{\breve E_\nu}}) \rangle_{\mathcal{N}_{O_{\breve E_\nu}}} \log q_\nu \nonumber\\ &=[E:F]\langle \Delta(\mathcal{N}_{n-1}), g\Delta(\mathcal{N}_{n-1}) \rangle_\mathcal{N} \log q_{w_0}. \end{align}

Here we use the equality $\sum _{\nu \mid w_0} e_{\nu /w_0} f_{\nu /w_0}=\sum _{\nu \mid w_0} d_{\nu /w_0}=[E:F]$. By the AFL identity (cf. [Reference Rapoport, Smithling and ZhangRSZ18, § 4]), we have

(8.26)\begin{equation} 2\langle \Delta(\mathcal{N}_{n-1}), g\Delta(\mathcal{N}_{n-1}) \rangle_\mathcal{N} \log q_{v_0}=-\omega_{v_0}(\gamma)\operatorname{\partial Orb}(\gamma,f_{v_0}'), \end{equation}

for any $\gamma \in G'(F_{0,v_0})_\mathrm {rs}$ matching $g$. Since $v_0$ is inert in $F$, we have $\log q_{w_0}=2 \log q_{v_0}$, and hence

(8.27)\begin{equation} \operatorname{\mathrm{Int}}^{\natural}_{v_0}(g)=- [E:F]\omega_{v_0}(\gamma)\operatorname{\partial Orb}(\gamma,f_{v_0}'). \end{equation}

Since $f$ and $f'$ are smooth transfers of each other, we have, for $\gamma \in G'(F_{0})$ matching $g\in \mathbb{O}(\widetilde {HG}(\mathbb{Q}))_\mathrm {rs}$,

(8.28)\begin{equation} \operatorname{Orb}(g, f^{p})=\prod_{v<\infty,\, v\nmid p }\omega_v(\gamma)\operatorname{Orb}(\gamma,f_v'). \end{equation}

By Remark 7.8, since the orbital integrals of $f_{v_0}$ and $f'_{v_0}$ do not vanish identically, one of the following holds.

  1. (1) $c(\phi _p)=0$ or one of $f_v$, for $v\in \mathcal {V}_p\smallsetminus \{v_0\}$, has identically vanishing orbital integrals.

  2. (2) The orbital integrals of $f_p$ do not vanish identically.

In the first case, one of $f'_v$, for $v\in \mathcal {V}_p\smallsetminus \{v_0\}$, has identically vanishing orbital integrals, and we have

\[ \operatorname{\mathrm{Int}}_{v_0}(f)=0=\partial J_{v_0}(f'), \]

where the second equality follows from (7.11).

Therefore we only need to consider the second case. By Remark 7.8, there are non-zero constants $c_v$ such that $\prod _{v \in \mathcal {V}_p}c_v=c(\phi _p)$, and for every $v \in \mathcal {V}_p$, $c_v f_v'$ is a transfer of $f_v$. By computing orbital integrals at some special $g$ and $\gamma$ (e.g. [Reference YunYun11]) we conclude that $c_{v_0}=1$ (this follows in any case if the fundamental lemma (FL) conjecture of Jacquet–Rallis holds, which is known when the residue characteristic is big enough). It follows that

(8.29)\begin{equation} c(\phi_p) \prod_{v\in \mathcal{V}_p\smallsetminus\{v_0\}}\operatorname{Orb}(g,f_v)=\prod_{v\in \mathcal{V}_p\smallsetminus\{v_0\}}\omega_v(\gamma)\operatorname{Orb}(\gamma,f'_v). \end{equation}

Note that for $v$ archimedean, $f_v'$ is a Gaussian test function so that we have, for regular semisimple $\gamma \in G'(F_{0,v})_\mathrm {rs}$ (cf. (7.17)),

\[ \omega_v(\gamma)\operatorname{Orb}(\gamma, f'_v)=\begin{cases} 1, & \text{there exists } g\in (H\times G)(F_{0,v})\text{ matching } \gamma,\\ 0, & \text{no } g\in (H\times G)(F_{0,v}) \text{ matches } \gamma. \end{cases} \]

By the last equality for archimedean $v$, by (8.28) for non-archimedean $v$ with $v\nmid p$, by (8.29) for $v\in \mathcal {V}_p\smallsetminus \{v_0\}$, and by (8.27) for $v=v_0$, we have (cf. (8.23))

\[ \operatorname{\mathrm{Int}}_{v_0}(f)=-\sum_{\gamma\in G'(F_0)_\mathrm{rs}/H_{1,2}'(F_0)} \omega_{v_0}(\gamma)\operatorname{\partial Orb}(\gamma,f'_{v_0})\cdot \prod_{v\neq v_0} \omega_{v}(\gamma)\operatorname{Orb}(\gamma,f'_v). \]

Here the sum runs a priori only over those $\gamma$ which match some $g\in G_W(F_0)$. However, those $\gamma$ which do not match any $g\in G_W(F_0)$ have vanishing orbital integrals away from $v_0$ (cf. [Reference ZhangZha12a, Propositon 3.6, (3.4)]).

By the product formula for transfer factors (7.14), we have $\prod _{v}\omega _{v}(\gamma ) =1$, and hence

\[ \operatorname{\mathrm{Int}}_{v_0}(f)=-\sum_{\gamma\in G'(F_0)_\mathrm{rs}/H_{1,2}'(F_0)} \operatorname{\partial Orb}(\gamma,f'_{v_0})\cdot \prod_{v\neq v_0}\operatorname{Orb}(\gamma,f'_v). \]

By (7.11), the right-hand side equals $-\partial J_{v_0}(f'),$ and this completes the proof.

In the preceding proof, we used the following lemma.

Lemma 8.16 Let $L_0/\mathbb{Q}_p$ be a finite extension and $L/L_0$ an étale $L_0$-algebra of rank $2$. Assume $p\neq 2$ if $L$ is a field. Let $\Phi _L \subset \operatorname {Hom}_{\mathbb{Q}_p}( L, \overline{\mathbb{Q}}_p)$ be a local CM type for $L/L_0$, and let $r_L\colon \operatorname {Hom}_{\mathbb{Q}_p}(L, \overline{\mathbb{Q}}_p)\longrightarrow \{0, n\}$ be a banal generalized CM type of rank $n$. Let $E'$ be the join of the reflex fields of $\Phi _L$ and $r_L$. Let $\overline k$ be an algebraic closure of the residue field of $O_{E'}$. Let $(\mathbb{X}_0, \iota _{\mathbb{X}_0}, \lambda _{\mathbb{X}_0})$ be a local CM triple of type $\overline \Phi _L$ over $\overline k$, where $\lambda _{\mathbb{X}_0}$ is principal. Let $(\mathbb{X}, \iota _{\mathbb{X}}, \lambda _{\mathbb{X}})$ be a local CM triple of type $r$ over $\overline k$ which satisfies the Eisenstein condition (cf. (B.5)). Assume that $\ker \lambda _{\mathbb{X}}\subset \mathbb{X}[\pi ]$, where $\pi$ denotes a uniformizerFootnote 20 of $L$. Let $\mathcal {N}_{\overline \Phi _L, r_L}$ be the formal scheme over $\operatorname {Spf} O_{\breve E'}$ that represents the functor which associates to $S\in \textrm {Nilp}_{O_{\breve E'}}$ the set of isomorphism classes of tuples $(X_0, \iota _0, \lambda _0, X, \iota , \lambda , \varrho _0, \varrho )$, where $(X_0, \iota _0, \lambda _0)$ is a local CM triple of type $\overline \Phi _L$, and $(X, \iota , \lambda )$ is a local CM triple of type $r_L$, over $S$, and where

\[ \rho_0\colon (X_0, \iota_0, \lambda_0)\times_S \overline S\longrightarrow \mathbb{X}_0\times_{\operatorname{Spec} \overline k} \overline S, \quad \rho\colon (X, \iota, \lambda)\times_S \overline S\longrightarrow \mathbb{X}\times_{\operatorname{Spec} \overline k} \overline S \]

are quasi-isogenies respecting the $O_L$-actions and the polarizations. Here $(X, \iota , \lambda )$ is supposed to satisfy the Eisenstein condition. Then

\[ \mathcal{N}_{\overline\Phi_L, r_L}\simeq G(L_0)/K , \]

where $G$ is the unitary group of an $L/L_0$-hermitian vector space of invariant ${{\mathrm {inv}}}^{r_L}(\mathbb{X}_0, \mathbb{X})$, and where $K$ is the stabilizer of a vertex lattice of type $t := \log _q \lvert \ker \lambda _\mathbb{X}\rvert$, $q := \# O_L/\pi O_L$.

Here the invariant ${{\mathrm {inv}}}^{r_L}(\mathbb{X}_0, \mathbb{X})$ is defined in Remark A.3.

Proof. By Lemmas B.1 and B.4, $\mathcal {N}_{\overline \Phi _L, r_L}$ is formally étale over $\operatorname {Spf} O_{\breve E'}$. So it remains to determine the point set $\mathcal {N}_{\overline \Phi _L, r_L}(\overline k)$ or, equivalently, the point set $\mathcal {N}_{\overline \Phi _L, r_L}^\textrm {rig}(\breve E')$, where $\mathcal {N}_{\overline \Phi _L, r_L}^\textrm {rig}$ denotes the generic fiber. Consider the crystalline period map $\pi \colon \mathcal {N}_{\overline \Phi _L, r_L}^\textrm {rig}\to { \mathcal {F}}\otimes _{E'}\breve E'$ (cf. [Reference Rapoport and ZinkRZ96]). However, by the banality of $r_L$, the Grassmannian $\mathcal {F}$ consists of a single point. Furthermore, the fiber over this point is identified with $G(L_0)/K$, since $\mathcal {N}_{\overline \Phi _L, r_L}^\textrm {rig}$ corresponds in the RZ tower to the level $K$.

Remark 8.17 We have considered local CM triples of type $\overline \Phi _L$ in the lemma because these are what naturally arise from the Kottwitz condition (3.8) in the moduli problem for $\mathcal {M}_0$. Of course, one could just as well consider the moduli space $\mathcal {N}_{\Phi _L,r_L}$; then one replaces $\Phi _L$ with $\overline \Phi _L$ in the definition of ${{\mathrm {inv}}}^{r_L}(\mathbb{X}_0,\mathbb{X})$ (cf. Remark A.2).

Remark 8.18 Consider the $J$-group in the sense of Kottwitz [Reference Rapoport and ZinkRZ96] associated to the situation of Lemma 8.16, when $L$ is a field. Let $M(\mathbb{X}_0)$ (respectively, $M(\mathbb{X})$) be the rational Dieudonné module of $\mathbb{X}_0$ (respectively, $\mathbb{X}$). Let $M=\operatorname {Hom}_L(M(\mathbb{X}_0), M(\mathbb{X}))$. Then $M$ is a rational Dieudonné module free of rank $n$ over $L\otimes _{\mathbb{Z}_p} W(\overline k)$ which, by our assumption, is isoclinic of slope zero. It is equipped with a hermitian form $h \colon M\times M \to L\otimes _{\mathbb{Z}_p} W(\overline k)$. Let $C$ be the space of Frobenius invariants in $M$. Then the restriction of $h$ makes $C$ into an $L/L_0$-hermitian vector space of dimension $n$. The $J$-group is the unitary group $J=\mathrm {U}(C)$. We claim that $J\simeq G$. Indeed, the difference between ${{\mathrm {inv}}}^{r_L}(\mathbb{X}_0, \mathbb{X})$ and ${{\mathrm {inv}}}(C)$ is equal to $\operatorname {sgn}(r_L)$, where

\[ \operatorname{sgn}(r_L)=(-1)^{\sum_{\varphi \in \Phi_L} (r_L)_\varphi} \]

(cf. (A.7), (A.11)). Since $r_L$ is banal, the exponent is a multiple of $n$. Therefore the assertion follows in the case when $n$ is even. When $n$ is odd, any two unitary groups of $L/L_0$-hermitian vector spaces of dimension $n$ are isomorphic.

Theorem 8.19 Conjecture 8.13(ii) holds when $n\leq 3$ and $v_0$ is non-archimedean.

Proof. This follows in the same way as in the proof of Theorem 8.15, from the AT conjecture proved for $n\leq 3$ in [Reference Rapoport, Smithling and ZhangRSZ17, Reference Rapoport, Smithling and ZhangRSZ18].

We indicate where we need to modify the proof. To simplify, we only prove the ‘Furthermore’ part of Conjecture 8.13(ii). We first consider the ramified case in § 4.4. In (8.21), we have $\mathcal {N}_{O_{\breve E_\nu }}=\mathcal {N} \mathbin {\widehat \otimes }_{O_{\breve F_{w_0}}}O_{\breve E_\nu }$, where $\mathcal {N} = \mathcal {N}_{n-1} \times _{O_{\breve F_{w_0}}} \mathcal {N}_n$ is the relative RZ space in [Reference Rapoport, Smithling and ZhangRSZ17, Reference Rapoport, Smithling and ZhangRSZ18]. Then we replace the AFL identity (8.26) by the AT identity (cf. [Reference Rapoport, Smithling and ZhangRSZ17, § 5] when $n$ is odd, and [Reference Rapoport, Smithling and ZhangRSZ18, § 12] when $n$ is even),

(8.30)\begin{equation} \langle \Delta(\mathcal{N}_{n-1}), g\Delta(\mathcal{N}_{n-1}) \rangle_\mathcal{N} \log q_{v_0}=-\omega_{v_0}(\gamma)\operatorname{\partial Orb}(\gamma,f_{v_0}'), \end{equation}

where $f_{v_0}'$ is a function in the AT conjecture. Now since $v_0$ is ramified in $F$, we have $\log q_{w_0}=\log q_{v_0}$ and hence equation (8.27) remains true by (8.25).

In the case of unramified AT type, we replace the corresponding space $\mathcal {N}$ after (8.21) by the RZ space in [Reference Rapoport, Smithling and ZhangRSZ18].

In both cases, the rest of the proof is the same.

Remark 8.20 The proof in the ramified case explains the discrepancy of the factor $2$ in the ramified AT conjecture [Reference Rapoport, Smithling and ZhangRSZ17, Reference Rapoport, Smithling and ZhangRSZ18] and the AFL. In these identities, it would be most natural to normalize the intersection number by the factor $\log q_{w_0}$.

Acknowledgements

We are grateful to G. Chenevier, B. Gross, B. Howard, S. Kudla, Y. Liu, P. Scholze and Y. Tian for helpful remarks. We also thank the referees for their work. We acknowledge the hospitality of the ESI (Vienna) and the MFO (Oberwolfach), where part of this work was carried out. M.R. is supported by a grant from the Deutsche Forschungsgemeinschaft through the grant SFB/TR 45. B.S. is supported by Simons Foundation grant #359425 and NSA grant H98230-16-1-0024. W.Z. is supported by NSF grants DMS #1601144, #1901642 and a Simons Fellowship.

Sign invariants

In this appendix we adapt the sign invariants of [Reference Kudla and RapoportKR15] to the setting of the moduli problems introduced in §§ 4 and 5 (see also [Reference Kudla, Rapoport and ZinkKRZ20]). We continue with the notation in the main body of the paper, with $F/\mathbb{Q}$ a CM field, $F_0$ its maximal totally real subfield of index $2$, and $\Phi$ a CM type for $F$. However, we will allow for more general functions $r$ than in (3.14). Set $d := [F_0 : \mathbb{Q}]$, and let $v$ be a finite place of $F_0$ which is non-split in $F$. (In the case where $v$ splits in $F$, the analog of the theory in this appendix is trivial, in so far as the value group $F_{0,v}^{\times }/ \operatorname {Nm} F_v^{\times }$ is trivial.) Let $k$ be an arbitrary field. We will first define an invariant

(A.1)\begin{equation} {{\mathrm{inv}}}_v(A_0,\iota_0,\lambda_0,A,\iota,\lambda)^{{\natural}} \in F_{0,v}^{\times} / \operatorname{Nm} F_v^{\times} \end{equation}

attached to the following objects over $k$:

  • abelian varieties $A_0$ and $A$ over $k$ of respective dimensions $d$ and $nd$;

  • rational actions $\iota _0\colon F \to \operatorname {End}^{\circ }(A_0)$ and $\iota \colon F \to \operatorname {End}^{\circ }(A)$; and

  • quasi-polarizations $\lambda _0 \in \operatorname {Hom}^{\circ }(A_0,A_0^{\vee })$ and $\lambda \in \operatorname {Hom}^{\circ }(A,A^{\vee })$ whose corresponding Rosati involutions induce the non-trivial automorphism on $F/F_0$.

We will then write

(A.2)\begin{equation} {{\mathrm{inv}}}_v(A_0,\iota_0,\lambda_0,A,\iota,\lambda) \in \{\pm 1\} \end{equation}

for the image of ${{\mathrm {inv}}}_v(A_0,\iota _0,\lambda _0,A,\iota ,\lambda )^{{\natural }}$ under the identification $F_{0,v}^{\times } / \operatorname {Nm} F_v^{\times } \cong \{\pm 1\}$. We note in advance that this invariant will depend on $A_0$ and $A$ only up to isogeny. We give the definition separately in the cases where $\operatorname {reschar} v \neq \operatorname {char} k$ and $\operatorname {reschar} v = \operatorname {char} k$, where $\operatorname {reschar} v$ denotes the residue characteristic of $v$.

(a) $\operatorname {reschar} v \neq \operatorname {char} k$. Let $\ell := \operatorname {reschar} v$, and let $V_\ell (A_0)$ and $V_\ell (A)$ denote the respective rational $\ell$-adic Tate modules of $A_0$ and $A$. Set

\[ V_\ell(A_0, A) := \operatorname{Hom}_F(V_\ell(A_0), V_\ell(A)). \]

Then $V_\ell (A_0, A)$ is a free $F \otimes _\mathbb{Q} \mathbb{Q}_\ell$-module of rank $n$. The polarizations $\lambda _0$ and $\lambda$ endow $V_\ell (A_0, A)$ with a non-degenerate hermitian form $h$, defined by

\[ h(\phi_1,\phi_2) := \lambda_0{{}^{-1}} \circ \phi_2^{\vee} \circ \lambda \circ \phi_1 \in \operatorname{End}_{F \otimes \mathbb{Q}_\ell}( V_\ell(A_0, A) ) \cong F \otimes \mathbb{Q}_\ell. \]

The decomposition $F_0 \otimes \mathbb{Q}_\ell = \prod _{v'\mid \ell } F_{0,v'}$, where $v'$ runs through the places of $F_0$ over $\ell$, induces a decomposition

\[ V_\ell(A_0, A) = \bigoplus_{v'\mid \ell} V_{v'}(A_0, A), \]

and each $V_{v'}(A_0, A)$ is a free $F_{v'}$-module of rank $n$. By assumption $F_v$ is a field, and we define the invariant at $v$ as for any $n$-dimensional $F_v/F_{0,v}$-hermitian space in the main body of the paper (1.4),

\[ {{\mathrm{inv}}}_v(A_0,\iota_0,\lambda_0,A,\iota,\lambda)^{{\natural}} := (-1)^{n(n-1)/2} \det V_v(A_0, A) \in F_{0,v}^{\times} / \operatorname{Nm} F_v^{\times}, \]

where $\det V_v(A_0, A)$ is the class mod $\operatorname {Nm} F_v^{\times }$ of any hermitian matrix representing the component $h_v$. We note that ${{\mathrm {inv}}}_v(A_0,\iota _0,\lambda _0,A,\iota ,\lambda )^{{\natural }}$ is unchanged after any base change $k \to k'$.

(b) $\operatorname {reschar} v = \operatorname {char} k = p$. Let $\overline k$ be an algebraic closure of $k$. Let $W = W(\overline k)$ denote the ring of Witt vectors of $\overline k$, and let $\sigma$ denote the Frobenius operator on $W$. The decomposition $O_{F_0} \otimes \mathbb{Z}_{p} = \prod _{v'\mid p} O_{F_0,v'}$ induces decompositions of $p$-divisible groups $A_0[p^{\infty }] = \prod _{v'\mid p}A_0[v^{\prime \infty }]$ and $A[p^{\infty }] = \prod _{v'\mid p}A[v^{\prime \infty }]$, and we denote by $M(A_0)_{\mathbb{Q},v}$ and $M(A)_{\mathbb{Q},v}$ the respective rational Dieudonné modules of $A_0[v^{\infty }]$ and $A[v^{\infty }]$ over $W_\mathbb{Q}$. Let $\underline {F}_0$ and $\underline {F}$ denote the respective Frobenius operators of $M(A_0)_{\mathbb{Q},v}$ and $M(A)_{\mathbb{Q},v}$. Then $M(A_0)_{\mathbb{Q},v}$ and $M(A)_{\mathbb{Q},v}$ carry actions of $F_v$ which commute with the Frobenius operators and make $M(A_0)_{\mathbb{Q},v}$ and $M(A)_{\mathbb{Q},v}$ into free $F_v \otimes _{\mathbb{Q}_p} W_\mathbb{Q}$-modules of respective ranks $1$ and $n$. Furthermore, $M(A_0)_{\mathbb{Q},v}$ is isoclinic of slope $1/2$. We consider the internal Hom in the category of isocrystals with $F_v$-action,

\[ M_{\mathbb{Q},v} := \operatorname{Hom}_{F_v \otimes W_\mathbb{Q}}(M(A_0)_{\mathbb{Q},v}, M(A)_{\mathbb{Q},v}), \]

which is a free $F_v \otimes _{\mathbb{Q}_p}W_\mathbb{Q}$-module of rank $n$. Here, as for any internal Hom object, the Frobenius operator $\underline {F}_{M_{\mathbb{Q},v}}$ on $M_{\mathbb{Q},v}$ sends $\phi \mapsto \underline {F} \circ \phi \circ \underline {F}_0{{}^{-1}}$. The polarizations $\lambda _0$ and $\lambda$ endow $M_{\mathbb{Q},v}$ with an $F_v \otimes _{\mathbb{Q}_p} W_\mathbb{Q}/F_{0,v} \otimes _{\mathbb{Q}_p} W_\mathbb{Q}$-hermitian form $h$, defined by

\[ h(\phi_1,\phi_2) := \lambda_0{{}^{-1}} \circ \phi_2^{\vee} \circ \lambda \circ \phi_1 \in \operatorname{End}_{F_v \otimes W_\mathbb{Q}}(M(A_0)_{\mathbb{Q},v}) \cong F_v \otimes_{\mathbb{Q}_p} W_\mathbb{Q}. \]

Let

\[ N_{\mathbb{Q},v} := \bigwedge_{F_v \otimes W_\mathbb{Q}}^{n} M_{\mathbb{Q},v}. \]

Then $N_{\mathbb{Q},v}$ is isoclinic of slope zero, and $h$ induces a hermitian form ${(\text {,})}$ on it satisfying

\[ (\underline{F}_{N_{\mathbb{Q},v}} x, \underline{F}_{N_{\mathbb{Q},v}} y) = (x,y)^{\sigma}, \quad x,y \in N_{\mathbb{Q},v}. \]

For any element $x_0 \in N_{\mathbb{Q},v}$ fixed by $\underline {F}_{N_{\mathbb{Q},v}}$, the class $(x_0,x_0) \in F_{0,v}^{\times }/\operatorname {Nm} F_v^{\times }$ is independent of the choice of $x_0$, and we define

\[ {{\mathrm{inv}}}_v(A_0,\iota_0,\lambda_0,A,\iota,\lambda)^{{\natural}} := (-1)^{n(n-1)/2}(x_0,x_0) \in F_{0,v}^{\times}/\operatorname{Nm} F_v^{\times} \]

for such $x_0$.

Now let $r\colon \operatorname {Hom}_\mathbb{Q}(F,\overline{\mathbb{Q}}) \to \mathbb{Z}_{\geq 0}$ be a generalized CM type for $F$ of rank $n$, that is, a function $\varphi \mapsto r_\varphi$ satisfying $r_\varphi + r_{\overline \varphi } = n$ for all $\varphi \in \operatorname {Hom}_\mathbb{Q}(F,\overline{\mathbb{Q}})$ (cf. [Reference Kudla and RapoportKR15, Definition 2.1]). Also, let $r_0$ be the opposite of the canonical generalized CM type for $F$ of rank $1$ attached to the CM type $\Phi$,

(A.3)\begin{equation} r_{0,\varphi} = \begin{cases} 0, & \varphi\in \Phi,\\ 1, & \varphi\notin\Phi, \end{cases}\quad \varphi \in \operatorname{Hom}_\mathbb{Q}(F,\overline{\mathbb{Q}}). \end{equation}

Let $E$ be the subfield of $\overline{\mathbb{Q}}$ characterized by

(A.4)\begin{equation} \operatorname{Gal}(\overline{\mathbb{Q}}/E) = \{\sigma\in\operatorname{Gal}(\overline{\mathbb{Q}}/\mathbb{Q}) \,|\, \sigma \circ\Phi=\Phi \text{ and } r_{\sigma\varphi}= r_\varphi \text{ for all } \varphi\in\Phi \}. \end{equation}

Thus $E$ is the join of the reflex fields of $r$ and of $r_0$, in the sense of [Reference Kudla and RapoportKR15, § 2]. Note that, in the situation of the main body of the paper, this definition of $E$ agrees with (3.1); but in contrast to the main body, in general $F$ need not admit an embedding into $E$.

Recall from [Reference Kudla and RapoportKR15, § 2] that a triple of CM type $r$ over an $O_E$-scheme $S$ is a triple $(A,\iota ,\lambda )$ consisting of an abelian scheme $A$ over $S$, an action $\iota \colon O_F \to \operatorname {End}_S(A)$ satisfying the Kottwitz condition of type $r$, and a polarization $\lambda \colon A \to A^{\vee }$ such that $\operatorname {Ros}_\lambda$ induces on $O_F$, via $\iota$, the non-trivial Galois automorphism of $F/F_0$. We denote by $\mathcal {M}_{r_0,r}$ the stack over $\operatorname {Spec} O_E$ of tuples $(A_0,\iota _0,\lambda _0,A,\iota ,\lambda )$, where $(A_0,\iota _0,\lambda _0)$ is a CM triple of type $r_0$ and $(A,\iota ,\lambda )$ is a CM triple of type $r$.

Let $k$ be a field which is an $O_E$-algebra, and let $(A_0,\iota _0,\lambda _0,A,\iota ,\lambda ) \in \mathcal {M}_{r_0,r}(k)$. Again let $v$ be a finite place of $F_0$ which is non-split in $F$. We will define the $r$-adjusted invariant ${{\mathrm {inv}}}_v^{r}(A_0,\iota _0,\lambda _0,A,\iota ,\lambda )$ (it depends on both $r_0$ and $r$). If the residue characteristic of $v$ is different from the characteristic of $k$, then we set

(A.5)\begin{equation} {{\mathrm{inv}}}_v^{r}(A_0,\iota_0,\lambda_0,A,\iota,\lambda) := {{\mathrm{inv}}}_v(A_0,\iota_0,\lambda_0,A,\iota,\lambda). \end{equation}

Now suppose that the residue characteristic of $v$ is equal to the characteristic $p$ of $k$. Let $\nu$ be the place of $E$ determined by the structure map $O_E \to k$, and let $\widetilde \nu \colon \overline{\mathbb{Q}} \to \overline{\mathbb{Q}}_p$ be an embedding which induces $\nu$. Let

\[ \Phi_{\nu,v} := \{\varphi \in \Phi \,|\, \widetilde\nu \circ \varphi|_{F_0} \text{ induces } v\, \}. \]

Then the set $\Phi _{\nu ,v}$ is independent of the choice of $\widetilde \nu$ inducing $\nu$, and, using $\widetilde \nu$, we may identify

(A.6)\begin{equation} \operatorname{Hom}_{\mathbb{Q}_p}(F_v,\overline{\mathbb{Q}}_p) \simeq \Phi_{\nu,v} \sqcup \overline\Phi_{\nu,v}. \end{equation}

Let

\[ r_{\nu,v} := r|_{\Phi_{\nu,v} \sqcup \overline\Phi_{\nu,v}}. \]

Then we define

(A.7)\begin{equation} \operatorname{sgn}(r_{\nu,v}) := (-1)^{\sum_{\varphi\in \Phi_{\nu,v}} r_\varphi} \end{equation}

and

(A.8)\begin{equation} {{\mathrm{inv}}}_v^{r}(A_0,\iota_0,\lambda_0,A,\iota,\lambda) := \operatorname{sgn}(r_{\nu,v}) \cdot {{\mathrm{inv}}}_v(A_0,\iota_0,\lambda_0,A,\iota,\lambda). \end{equation}

The analog of [Reference Kudla, Rapoport and ZinkKRZ20, Appendix] (which corrects [Reference Kudla and RapoportKR15, Proposition 3.2]) is now as follows.

Proposition A.1 Let $(A_0,\iota _0,\lambda _0,A,\iota ,\lambda ) \in \mathcal {M}_{r_0,r}(S)$, where $S$ is a connected scheme over $\operatorname {Spec} O_E$. Then, for every non-archimedean place $v$ of $F_0$ which is non-split in $F$, the function

\[ s \longmapsto {{\mathrm{inv}}}_v^{r}(A_{0,s},\iota_{0,s},\lambda_{0,s},A_s,\iota_s,\lambda_s) \]

is constant on $S$.

Proof. The proof is easy when the residue characteristic $\ell$ of $v$ is invertible in $\mathcal {O}_S$, in which case $\bigwedge ^\textrm {max} V_\ell (A_0, A)$ is a lisse étale sheaf on $S$ (cf. the proof of [Reference Kudla and RapoportKR15, Proposition 3.2]). A similar argument proves the assertion when $S$ is a scheme over $\mathbb{F}_p$ and $\ell = p$ (cf. [Reference Kudla, Rapoport and ZinkKRZ20, Lemma 8.2.2]).Footnote 21 The remaining cases are reduced to the case $\ell = p$ and $S = \operatorname {Spec} O_L$, where $L$ is the completion of a subfield of $\overline{\mathbb{Q}}_p$ which contains $E$ and such that its ring of integers $O_L$ is a discrete valuation ring with residue field $k = \overline{\mathbb{F}}_p$. To compare the invariants at the generic and closed points of $S$, as in [Reference Kudla and RapoportKR15, proof of Proposition 3.2], we will use $p$-adic Hodge theory. Let $A_L$ and $A_{0, L}$ denote the respective generic fibers of $A$ and $A_0$, and let $A_k$ and $A_{0,k}$ denote the respective special fibers. Let $\nu$ denote the induced place of $E$.

We decompose the homomorphism module of the rational $p$-adic Tate modules of $A_{0,L}$ and $A_L$ (respectively, the homomorphism module of the rational Dieudonné modules of $A_{0, k}$ and $A_k$) with respect to the actions of $F \otimes _\mathbb{Q} \mathbb{Q}_p \cong \prod _{w \mid p} F_w$,

\begin{align*} V_p &:= \operatorname{Hom}_{F \otimes \mathbb{Q}_p}(V_p(A_{0, L}),V_p(A_L))= \bigoplus_{w\mid p} V_{w} ,\\ M_\mathbb{Q} &:= \operatorname{Hom}_{F \otimes \breve{\mathbb{Q}}_p}(M(A_{0,k})_\mathbb{Q},M(A_k)_\mathbb{Q}) = \bigoplus_{w\mid p} M_{\mathbb{Q},w}. \end{align*}

Here $w$ runs through the places of $F$, and we recall the notation $\breve{\mathbb{Q}}_p = W(k)_\mathbb{Q}$. Furthermore, for each place $w \mid p$, the summand $V_{w}$ is a free $F_{w}$-vector space of rank $n$, and $M_{\mathbb{Q},w}$ is a free $F_{w} \otimes _{\mathbb{Q}_p} \breve{\mathbb{Q}}_p$-module of rank $n$. At our distinguished place $v$, we set

\[ S_v := \bigwedge_{F_v}^{n} V_v \quad\text{and}\quad N_{\mathbb{Q},v} := \bigwedge_{F_v \otimes \breve{\mathbb{Q}}_p}^{n} M_{\mathbb{Q},v}. \]

Then $S_v$ and $N_{\mathbb{Q},v}^{\text {Frob}=1}$ are one-dimensional $F_v$-vector spaces (the latter because $N_{\mathbb{Q},v}$ is isoclinic of slope zero) equipped with natural $F_v/F_{0,v}$-hermitian forms. Our problem is to compare these hermitian spaces, which we will do via [Reference Rapoport and ZinkRZ96, Proposition 1.20].

To explain the group-theoretic setup in our application of [Reference Rapoport and ZinkRZ96, Proposition 1.20], let $T$ be the torus over $\mathbb{Q}_p$ which is the kernel in the exact sequence

\[ 1 \longrightarrow T \longrightarrow \operatorname{Res}_{F_v /\mathbb{Q}_p} \mathbb{G}_{m,F_v} {\mathop \to \limits^{\operatorname{Nm}}} \operatorname{Res}_{F_{0,v} /\mathbb{Q}_p} \mathbb{G}_{m,F_{0,v}} \longrightarrow 1. \]

Then $H^{1}(\mathbb{Q}_p,T) = F_{0,v}^{\times }/\operatorname {Nm} F_v^{\times }$. The spaces $S_v$ and $N_{\mathbb{Q},v}^{\text {Frob}=1}$ are natural $\mathbb{Q}_p$-rational representations of $T$, and we may regard the isomorphisms of hermitian vector spaces $\operatorname {Isom}(N_{\mathbb{Q},v}^{\text {Frob}=1},S_v)$ as an étale sheaf on $\operatorname {Spec} \mathbb{Q}_p$. This is a $T$-torsor. To calculate its class ${{\mathrm {cl}}}(N_{\mathbb{Q},v}^{\text {Frob}=1},S_v)$ via [Reference Rapoport and ZinkRZ96, Proposition 1.20], we seek to express the filtered isocrystal $N_{\mathbb{Q},v}$ in the form $\mathcal {I}(N_{\mathbb{Q},v}^{\text {Frob}=1})$ for an admissible pair $(\mu ,b)$ in $T$, in the notation of [Reference Rapoport and ZinkRZ96, 1.17].

Since $N_{\mathbb{Q},v}$ is isoclinic of slope zero, we take $b \in T(\breve{\mathbb{Q}}_p)$ to be the identity. To determine the cocharacter $\mu$, we need to identify the filtration on $N_{\mathbb{Q},v} \otimes _{\breve{\mathbb{Q}}_p} \overline {\breve{\mathbb{Q}}}_p$. Choose any embedding $\overline{\mathbb{Q}}_p \to \overline {\breve{\mathbb{Q}}}_p$, and identify $\operatorname {Hom}_{\mathbb{Q}_p}(F_v,\overline{\mathbb{Q}}_p) \simeq \Phi _{\nu ,v} \sqcup \overline \Phi _{\nu ,v}$ as in (A.6). By the Kottwitz condition, the filtration on $M(A_k)_{\mathbb{Q},v} \otimes _{\breve{\mathbb{Q}}_p} \overline {\breve{\mathbb{Q}}}_p \cong \bigoplus _{\varphi \in \Phi _{\nu , v} \sqcup \overline \Phi _{\nu ,v}} M(A_k)_{\mathbb{Q},\varphi }$ is given by, for each $\varphi$,

\[ M(A_k)_{\mathbb{Q}, \varphi} \supset^{r_\varphi} \mathrm{Fil}_\varphi^{1} \supset 0, \]

where the displayed terms are in respective degrees $0$, $1$, and $2$, and the upper index on the first containment means that the cokernel is of dimension $r_\varphi$. The filtration on $M(A_{0,k})_{\mathbb{Q},v} \otimes _{\breve{\mathbb{Q}}_p} \overline {\breve{\mathbb{Q}}}_p \cong \bigoplus _{\varphi \in \Phi _{\nu , v} \sqcup \overline \Phi _{\nu , v}} M(A_{0,k})_{\mathbb{Q}, \varphi }$ is similarly given by, for each $\varphi$,

\[ M(A_{0,k})_{\mathbb{Q},\varphi} \supset^{r_{0,\varphi}} \mathrm{Fil}_{0,\varphi}^{1} \supset 0, \]

where $r_{0,\varphi }$ is given in (A.3). The unique jump in this filtration occurs in degree $1-r_{0, \varphi }$. The filtration on the dual space $M(A_{0,k})_{\mathbb{Q},\varphi }^{\vee }$ therefore has unique jump in degree $r_{0, \varphi }-1$. Therefore in the filtration on the one-dimensional space

\[ N_{\mathbb{Q},\varphi} = \bigwedge_{\overline{\breve{\mathbb{Q}}}_p}^{n} M_{\mathbb{Q},\varphi} \cong (M(A_{0,k})_{\mathbb{Q},\varphi}^{\vee})^{\otimes n} \otimes \bigwedge_{\overline{\breve{\mathbb{Q}}}_p}^{n} M(A_k)_{\mathbb{Q},\varphi}, \]

the unique jump occurs in degree $n - r_\varphi + n(r_{0,\varphi } -1) = nr_{0,\varphi } - r_\varphi$.

Now consider the natural identification

\[ X_*(T) \cong \ker[{{\mathrm{Ind}}}_{F_{0,v}}^{F_v} ({{\mathrm{Ind}}}_{\mathbb{Q}_p}^{F_{0,v}}\mathbb{Z}) \longrightarrow {{\mathrm{Ind}}}_{\mathbb{Q}_p}^{F_{0,v}}\mathbb{Z} ]. \]

The corresponding filtration on $N_{\mathbb{Q},\varphi }$ is then given by the cocharacter $\mu \in X_*(T)$ with

\[ \mu_\varphi = n r_{0,\varphi} - r_\varphi, \quad \varphi \in \Phi_{\nu,v} \sqcup \overline\Phi_{\nu,v}. \]

Then $N_{\mathbb{Q},\varphi } = \mathcal {I}(N_{\mathbb{Q},v}^{\text {Frob}=1})$ for the above choices of $\mu$ and $b$.

Now, by $p$-adic Hodge theory, in the case of the abelian scheme $A$, there is a canonical isomorphism

(A.9)\begin{equation} V_p(A_L) \otimes_{\mathbb{Q}_p} B_\mathrm{crys} \cong M(A_k)_{\mathbb{Q}} \otimes_{\breve{\mathbb{Q}}_p} B_\mathrm{crys} \end{equation}

compatible with all structures on both sides (e.g. the Frobenii, the $F$-actions, and the polarization forms); cf. [Reference TsujiTsu99, Reference FaltingsFal02]. Here $B_\mathrm {crys}$ is the crystalline period ring of Fontaine [Reference FontaineFon94]. Moreover, after extension of scalars under the inclusion $B_\mathrm {crys} \subset B_\text {dR}$, this isomorphism is compatible with the filtrations on both sides. Furthermore, there is an analogous isomorphism with $A_0$ in place of $A$. Taking homomorphism modules on both sides between the $v$-components of (A.9) and its analog for $A_0$, and then passing to top exterior powers, we obtain an isomorphism between free $F_v \otimes _{\mathbb{Q}_p} B_\mathrm {crys}$-modules of rank $1$,

\[ S_v \otimes_{\mathbb{Q}_p} B_\mathrm{crys} \cong N_{\mathbb{Q},v} \otimes_{\breve{\mathbb{Q}}_p} B_\mathrm{crys}, \]

again compatible with all structures on both sides, and in particular with the hermitian forms. Taking Frobenius-fixed elements in the zeroth filtration modules, we obtain an isometry of $F_v/F_{0,v}$-hermitian spaces,

\[ S_v \cong \mathcal{F}( \mathcal{I}(N_{\mathbb{Q},v}^{\text{Frob}=1}) ), \]

where $\mathcal {F}$ denotes Fontaine's functor from admissible filtered isocrystals to Galois representations.

We conclude that the class ${{\mathrm {cl}}}(N_{\mathbb{Q},v}^{\text {Frob}=1},S_v)$ is computed by the formula $\kappa (b) - \mu ^{\sharp }$ in [Reference Rapoport and ZinkRZ96, Proposition 1.20]. Here $\mu ^{\sharp }$ denotes the image of $\mu$ in the coinvariants $X_*(T)_\Gamma$, where $\Gamma := \operatorname {Gal}(\overline{\mathbb{Q}}_p/\mathbb{Q}_p)$. Since $b$ is trivial, under the identification $X_*(T)_\Gamma \cong H^{1}(\mathbb{Q}_p,T) \cong \mathbb{Z}/2\mathbb{Z}$, we obtain

(A.10)\begin{equation} {{\mathrm{cl}}}(N_{\mathbb{Q},v}^{\text{Frob}=1},S_v) = \mu^{\sharp} = \sum_{\varphi \in \Xi_v} \mu_\varphi , \end{equation}

where $\Xi _v$ is any half-system, that is, $\Xi _v \sqcup \overline \Xi _v \cong \operatorname {Hom}_{\mathbb{Q}_p}(F_v,\overline{\mathbb{Q}}_{p})$. In the formula for $\operatorname {sgn} (r_{\nu ,v})$, we took $\Xi _v=\Phi _{\nu ,v}$.

Remark A.2 In the definition of the $r$-adjusted invariant above, we took the function $r_0$ in (A.3) to be the opposite of the canonical rank-$1$ function for $\Phi$ because this is what occurs in the moduli problem for $\mathcal {M}_0^{\mathfrak{a}}$ in the main body of the paper (cf. (3.8)). Of course, we could instead have worked with respect to the canonical function (sending $\varphi \mapsto 1$ for $\varphi \in \Phi$ and $\varphi \mapsto 0$ for $\varphi \notin \Phi$), which is tantamount to replacing $\Phi$ by $\overline \Phi$. In this case one defines the $r$-adjusted invariant at a place $v$ dividing $\operatorname {char} k$ to be $(-1)^{\sum _{\varphi \in \overline \Phi _v} r_\varphi }\cdot {{\mathrm {inv}}}_v(A_0,\iota _0,\lambda _0,A,\iota ,\lambda )$, and the statement and proof of Proposition A.1 for this version go through virtually without change.

Remark A.3 There is an obvious variant of the sign invariant for $p$-divisible groups. More precisely, let $L_0$ be a finite extension of $\mathbb{Q}_p$, and let $L/L_0$ be a quadratic extension. Fix a local CM type $\Phi _L\subset \operatorname {Hom}_{\mathbb{Q}_p}(L, \overline{\mathbb{Q}}_p)$ for $L/L_0$, and let $r_L\colon \operatorname {Hom}_{\mathbb{Q}_p}(L, \overline{\mathbb{Q}}_p)\longrightarrow \mathbb{Z}_{\geq 0}$ be a local generalized CM type of rank $n$ for $L/L_0$ (cf. [Reference Kudla and RapoportKR15, § 5]). Let $k$ be a field of characteristic $p$ which is an $O_{E_{\Phi _L,r_L}}$-algebra, where $E_{\Phi _L,r_L} \subset \overline{\mathbb{Q}}_p$ denotes the join of the reflex fields for $\Phi _L$ and for $r_L$. Let $(X_0, \iota _0, \lambda _0)$ and $(X, \iota , \lambda )$ be $p$-divisible groups over $k$ with actions by $O_L$ and quasi-polarizations whose associated Rosati involutions induce on $F$ the Galois conjugation over $L_0$. Assume that $(X_0, \iota _0)$ is of CM type $\overline \Phi _L$ and that $(X, \iota )$ is of generalized CM type $r_L$ (cf. [Reference Kudla and RapoportKR15, § 5]). Then there is a sign invariant

(A.11)\begin{equation} {{\mathrm{inv}}}^{r_L}(X_0,\iota_0,\lambda_0,X,\iota,\lambda) := \operatorname{sgn}(r_L) \cdot {{\mathrm{inv}}}(X_0,\iota_0,\lambda_0,X,\iota,\lambda) , \end{equation}

with properties analogous to the case of abelian varieties. In fact, returning to all of our notation from the global setting, suppose that $k$ is an $O_E$-algebra, and let $(A_0,\iota _0,\lambda _0,A,\iota ,\lambda ) \in \mathcal {M}_{r_0,r}(k)$. Consider the decomposition of the corresponding $p$-divisible groups induced by the decomposition $O_{F_0}\otimes \mathbb{Z}_p=\prod _{v\mid p} O_{F_{0}, v}$,

\[ A_0[p^{\infty}]=\prod_{v\mid p}A_0[v^{\infty}] \quad\text{and}\quad A[p^{\infty}]=\prod_{v\mid p}A[v^{\infty}]. \]

Let $\nu$ denote the place of $E$ determined by $k$, and for a place $v \mid p$ which is non-split in $F$, take $L := F_v$, $\Phi _L := \Phi _{\nu ,v}$, and $r_L := r_{\nu ,v}$ (where we implicitly choose an identification $\operatorname {Hom}_{\mathbb{Q}_p}(F_v,\overline{\mathbb{Q}}_p) \simeq \Phi _{\nu ,v} \sqcup \overline \Phi _{\nu ,v}$ as above). Then

\[ {{\mathrm{inv}}}_v^{r}(A_0,\iota_0,\lambda_0,A,\iota,\lambda) = {{\mathrm{inv}}}^{r_{\nu,v}}(A_0[v^{\infty}],\iota_0[v^{\infty}],\lambda_0[v^{\infty}],A[v^{\infty}],\iota[v^{\infty}],\lambda[v^{\infty}]). \]

Local models in the case of banal signature

In this appendix we prove that local models attached to Weil restrictions of $\mathrm {GL}_n$ and $\mathrm {GU}_n$, defined using an analog of the Eisenstein condition of [Reference Rapoport and ZinkRZ17], are trivial in the case of banal signature. Let $L/K$ be a finite separable extension of discretely valued henselian fields, with respective valuation rings $O_L$ and $O_K$. Let $\pi$ be a uniformizer for $L$, and fix an algebraic closure $\overline K$ of $K$. The material in this appendix applies to the main body of the paper in the case $K = \mathbb{Q}_p$ and $L = F_w$ for $w$ a $p$-adic place of the number field $F$.

B.1 The $\mathrm {GL}_n$ case

Let $n$ be a positive integer, and fix a function

(B.1)

The reflex field attached to $r$ is the fixed field $E_r \subset \overline K$ of the subgroup of the Galois group,

\[ \{\sigma \in \operatorname{Gal}(\overline K/K) \,|\, r_{\sigma \circ \varphi} = r_\varphi \text{ for all } \varphi \in \operatorname{Hom}_K(L,\overline K)\}. \]

Then $E_r$ is a finite extension of $K$, contained in the normal closure of $L$ in $\overline K$. Note that in contrast to the analogous global situation considered in the main body of the paper (with the particular choice of $r$ in (3.14)), $L$ need not admit an embedding into $E_r$. Let $\mathcal {L}$ be a periodic $O_L$-lattice chain in $L^{n}$.

The local model attached to the group $\operatorname {Res}_{L/K} \mathrm {GL}_n$, the function $r$, and the lattice chain $\mathcal {L}$ is the scheme $M = M_{\operatorname {Res}_{L/K} \mathrm {GL}_n,r,\mathcal {L}}$ over $\operatorname {Spec} O_{E_r}$ representing the following functor. To each $O_{E_r}$-scheme $S$, the functor associates the set of isomorphism classes of families $(\Lambda \otimes _{O_K} \mathcal {O}_S \twoheadrightarrow \mathcal {P}_\Lambda )_{\Lambda \in \mathcal {L}}$ such that

  • for each $\Lambda$, $\mathcal {P}_\Lambda$ is an $O_L \otimes _{O_K} \mathcal {O}_S$-linear quotient of $\Lambda \otimes _{O_K} \mathcal {O}_S$, locally free on $S$ as an $\mathcal {O}_S$-module;

  • for each inclusion $\Lambda \subset \Lambda '$ in $\mathcal {L}$, the arrow $\Lambda \otimes _{O_K} \mathcal {O}_S \to \Lambda ' \otimes _{O_K} \mathcal {O}_S$ induces an arrow $\mathcal {P}_\Lambda \to \mathcal {P}_{\Lambda '}$;

  • for each $\Lambda$, the isomorphism $\Lambda \otimes _{O_K} \mathcal {O}_S {\mathop \to \limits_\sim ^{\pi \otimes 1}} (\pi \Lambda ) \otimes _{O_K} \mathcal {O}_S$ identifies $\mathcal {P}_\Lambda {\tilde{\to}} \mathcal {P}_{\pi \Lambda }$; and

  • for each $\Lambda$, $\mathcal {P}_\Lambda$ satisfies the Kottwitz condition

    (B.2)\begin{equation} \operatorname{char}_{\mathcal{O}_S} (a \otimes 1 \mid \mathcal{P}_\Lambda) = \prod_{\varphi \in \operatorname{Hom}_K(L,\overline K)} (T - \varphi(a))^{r_\varphi} \quad\text{for all} \quad a \in O_L. \end{equation}

We further require that the family $(\Lambda \otimes _{O_K} \mathcal {O}_S \twoheadrightarrow \mathcal {P}_\Lambda )_{\Lambda \in \mathcal {L}}$ satisfies the (analog of the) Eisenstein condition of [Reference Rapoport and ZinkRZ17, (8.2)], which in our case takes the following form. Let $L^{t}$ denote the maximal unramified extension of $K$ in $L$. We first formulate the condition when $S$ is an $O_{E_r \widetilde L^{t}}$-scheme, where $E_r \widetilde L^{t}$ is the compositum in $\overline K$ of $E_r$ and the normal closure $\widetilde L^{t}$ of $L^{t}$; by Lemma B.1 below, the condition will descend over $O_{E_r}$ (and yield $M {\tilde{\to}} \operatorname {Spec} O_{E_r}$). For each $\psi \in \operatorname {Hom}_K(L^{t},\overline K)$, set

\begin{align*} A_\psi &:= \{\varphi \in \operatorname{Hom}_K(L,\overline K)\,|\, \varphi|_{F^{t}} = \psi \text{ and } r_\varphi = n \}, \\ B_\psi &:= \{\varphi \in \operatorname{Hom}_K(L,\overline K)\,|\, \varphi|_{F^{t}} = \psi \text{ and } r_\varphi = 0 \}. \end{align*}

Further, set

\[ Q_{A_\psi}(T) := \prod_{\varphi \in A_\psi} (T - \varphi(\pi)) \quad\text{and}\quad Q_{B_\psi}(T) := \prod_{\varphi \in B_\psi} (T - \varphi(\pi)). \]

Then $Q_{A_\psi }$ and $Q_{B_\psi }$ are polynomials with coefficients in $O_{E_r\widetilde L^{t}}$. Since we assume that $S$ is an $O_{E_r \widetilde L^{t}}$-scheme, there is a natural isomorphism

(B.3)\begin{equation} O_{L^{t}} \otimes_{O_K} \mathcal{O}_S {\tilde{\to}} \prod_{\psi \in \operatorname{Hom}_K(L^{t},\overline K)} \mathcal{O}_S, \end{equation}

whose $\psi$-component is $\psi \otimes \mathrm {id}$. This induces a decomposition, for each $\Lambda$,

(B.4)\begin{equation} \mathcal{P}_\Lambda {\tilde{\to}} \bigoplus_{\psi \in \operatorname{Hom}_K(L^{t},\overline K)} (\mathcal{P}_\Lambda)_\psi. \end{equation}

The Eisenstein condition is that, for each $\Lambda$,

(B.5)\begin{equation} Q_{A_\psi}(\pi \otimes 1)|_{(\mathcal{P}_\Lambda)_\psi} = 0 \quad\text{for all} \quad \psi \in \operatorname{Hom}_K(L^{t},\overline K). \end{equation}

To complete the definition of the moduli problem, an isomorphism from $(\Lambda \otimes _{O_K} \mathcal {O}_S \twoheadrightarrow \mathcal {P}_\Lambda )_{\Lambda \in \mathcal {L}}$ to $(\Lambda \otimes _{O_K} \mathcal {O}_S \twoheadrightarrow \mathcal {P}'_\Lambda )_{\Lambda \in \mathcal {L}}$ consists of an isomorphism $\mathcal {P}_\Lambda {\tilde{\to}} \mathcal {P}_\Lambda '$ for each $\Lambda$, compatible with the given epimorphisms in the obvious way. Note that such an isomorphism is unique if it exists.

The main result is that the moduli scheme $M$ we have defined is trivial, in the following sense.

Lemma B.1 Let $S$ be an $O_{E_r \widetilde L^{t}}$-scheme. Then $M(S)$ consists of a single point.

Proof. It suffices to consider the case where $\mathcal {L}$ consists of the homothety class of a single lattice $\Lambda$; the general case then follows immediately. Let $\Lambda \otimes _{O_K} \mathcal {O}_S \cong \bigoplus _{\psi \in \operatorname {Hom}_K(L^{t},\overline K)} (\Lambda \otimes _{O_K} \mathcal {O}_S)_\psi$ denote the decomposition induced by (B.3). Then the Eisenstein condition forces

(B.6)\begin{equation} (\mathcal{P}_\Lambda)_\psi = (\Lambda \otimes_{O_K} \mathcal{O}_S)_\psi / Q_{A_\psi}(\pi \otimes 1) \cdot (\Lambda \otimes_{O_K} \mathcal{O}_S)_\psi, \end{equation}

which completes the proof.

As we have already noted, it follows by descent that $M \cong \operatorname {Spec} O_{E_r}$. This also shows that the Eisenstein condition is independent of the choice of uniformizer $\pi$.

Remark B.2 In the special case $n = 1$, we may take $\Lambda = O_L$ in the proof of Lemma B.1, and then the Kottwitz condition already implies (B.6). Thus the Eisenstein condition is redundant in this case.

We also note that the Eisenstein condition is redundant in the unramified case (cf. [Reference Rapoport and ZinkRZ17, Proposition 2.2]).

Lemma B.3 Suppose that $L/K$ is unramified. Then the Kottwitz condition (B.2) on $\mathcal {P}_\Lambda$ implies the Eisenstein condition (B.5).

Proof. When $L=L^{t}$ is unramified over $K$, all sets $A_\psi$ have at most one element. If $A_\psi$ is empty, then $(\mathcal {P}_\Lambda )_\psi =0$ and condition (B.5) is empty. If $A_\psi$ is non-empty, then condition (B.5) is equivalent to the definition of the $\psi$-eigenspace in the decomposition (B.4).

B.2 The unitary case

In this subsection we assume that the residue characteristic $p$ of $K$ is not $2$. We retain the setup of the previous subsection, and we assume in addition that $L$ is a quadratic extension of a field $L_0/K$. Let $a \mapsto \overline a$ denote the non-trivial automorphism of $L/L_0$, and for each $\varphi \in \operatorname {Hom}_K(L,\overline K)$, define $\overline \varphi (a) := \varphi (\overline a)$. We assume that the function $r$ in (B.1) satisfies $r_\varphi + r_{\overline \varphi } = n$ for all $\varphi \in \operatorname {Hom}_K(L,\overline K)$. Furthermore, we endow $L^{n}$ with a non-degenerate $L/L_0$-hermitian form $h$, and we assume that the lattice chain $\mathcal {L}$ is self-dual for $h$. We define the alternating $K$-bilinear form $\langle \text {,}\rangle \colon L^{n} \times L^{n} \to K$ as follows. Let $\vartheta _{L_0/K}^{-1}$ be a generator of the inverse different $\mathfrak {d}_{L_0/K}^{-1}$. If $L/L_0$ is unramified, then choose an element $\zeta \in O_L^{\times }$ such that $\overline \zeta = -\zeta$ (since $p \neq 2$, such a $\zeta$ always exists), and set

\[ \langle x,y \rangle := \operatorname{tr}_{L/K}(\vartheta_{L_0/K}^{-1}\zeta h(x,y)), \quad x,y \in L^{n}. \]

If $L/L_0$ is ramified, then choose the uniformizer $\pi$ to satisfy $\pi ^{2} \in L_0$ (since $p \neq 2$, such a $\pi$ always exists), and set

\[ \langle x,y \rangle := \operatorname{tr}_{L/K}(\vartheta_{L_0/K}^{-1}\pi^{-1}h(x,y)), \quad x,y \in L^{n}. \]

Then in both cases, the dual $\Lambda ^{*}$ of an $O_L$-lattice $\Lambda$ in $L^{n}$ is the same with respect to $h$ as it is with respect to $\langle \text {,}\rangle$.

The local model attached to the group $\operatorname {Res}_{L_0/K} \mathrm {GU}(h)$, the function $r$, and the lattice chain $\mathcal {L}$ is by definition the closed subscheme $M_{\operatorname {Res}_{L_0/K} \mathrm {GU}(h),r,\mathcal {L}}$ of $M_{\operatorname {Res}_{L/K}\mathrm {GL}_n,r,\mathcal {L}}$ defined by the following additional condition:

  • for each $\Lambda$, the perfect pairing

    \[ (\Lambda \otimes_{O_K} \mathcal{O}_S) \times (\Lambda^{*} \otimes_{O_K} \mathcal{O}_S) {\mathop \to \limits^{\langle {\rm ,}\rangle \otimes {\rm {\cal O}}_S}} \mathcal{O}_S \]
    identifies $\ker [\Lambda \otimes _{O_K} \mathcal {O}_S \twoheadrightarrow \mathcal {P}_\Lambda ]^{\perp }$ with $\ker [\Lambda ^{*} \otimes _{O_K} \mathcal {O}_S \twoheadrightarrow \mathcal {P}_{\Lambda ^{*}}]$.

It is a trivial consequence of Lemma B.1 that this additional condition is redundant, and that we again have the following result.

Lemma B.4 $M_{\operatorname {Res}_{L_0/K} \mathrm {GU}(h),r,\mathcal {L}} \cong \operatorname {Spec} O_{E_r}$.

Footnotes

1 However, we point out that the work of X. Yuan, S. Zhang and W. Zhang (in preparation) proves certain variants of this conjecture in a higher-dimensional case for orthogonal groups of type ${{\mathrm {SO}}}(3)\times {{\mathrm {SO}}}(4)$.

2 This notation differs from the main body of the paper, where $G$ denotes the unitary group of $W$ over $F_0$.

3 Kottwitz [Reference KottwitzKot92] does not need any assumptions on the signature of $W$; neither do we (cf. [Reference Rapoport, Smithling and ZhangRSZ19]).

4 Since the submission of the paper, the last author has made progress in the good reduction case (arXiv:1909.02697).

5 More precisely, in the case where $F_v/F_{0,v}$ is the unramified extension $\mathbb{Q}_{p^{2}}/\mathbb{Q}_p$, we have defined what [Reference Kudla and RapoportKR11] calls a vertex lattice of level $0$ and type $n-r$, where $n$ denotes the dimension of the hermitian space. In the case where $F_v$ is a ramified extension of $F_{0,v} = \mathbb{Q}_p$, our $\Lambda ^{*}$ is what [Reference Rapoport, Terstiege and WilsonRTW14] calls a vertex lattice of type $r$, except that [Reference Rapoport, Terstiege and WilsonRTW14] should also require the containment $\Lambda ^{*} \subset {\pi _v^{-1}} \Lambda$.

6 Strictly speaking, Proposition 3.1.2 in [Reference HowardHow12] is stated only for CM algebras and CM types $\Phi$ which are of a rather special sort, but the proof relies only on the very general Theorem 2.2.1 in [Reference HowardHow12] and applies equally well to our situation.

7 Indeed, let $\mathfrak {d}$ denote the different of $F/\mathbb{Q}$, and set $\mathfrak {b} := \sqrt \Delta \mathfrak {d}{{}^{-1}}\mathfrak {a}$. Let $\mathcal {H}_{\ll 0}^{\mathfrak{b}}$ be the set of isomorphism classes of hermitian $O_F$-modules $(\Lambda _0,{(\text {,})}_0)$ of rank $1$ (with ${(\text {,})}_0$ valued in $F$) such that the $F$-hermitian space $\Lambda _0 \otimes _\mathbb{Z} \mathbb{Q}$ has signature $(0,1)$ at every archimedean place of $F_0$, and such that the dual lattice $\Lambda _0^{*}$ of $\Lambda _0$ inside $\Lambda _0 \otimes _\mathbb{Z} \mathbb{Q}$ with respect to ${(\text {,})}_0$ equals $\mathfrak {b}{{}^{-1}} \Lambda _0$. Then $(\Lambda _0,{(\text {,})}_0) \mapsto (\Lambda _0, \operatorname {tr}_{F/\mathbb{Q}}\sqrt \Delta {{}^{-1}} {(\text {,})}_0)$ defines a bijection $\mathcal {H}_{\ll 0}^{\mathfrak{b}} {\tilde{\to}} \mathcal {L}_\Phi ^{\mathfrak{a}}$. It is easy to deduce from the finiteness of the class group for $F$ and the product formula for hermitian spaces that $\mathcal {H}_{\ll 0}^{\mathfrak{b}}$ is finite. Note that in the particular situation of [Reference Kudla and RapoportKR14], where $F_0 = \mathbb{Q}$ and $\Delta$ is the discriminant of $F$, one has $\mathfrak {a} = \mathfrak {b}$, but $\mathfrak {a}$ and $\mathfrak {b}$ certainly may differ in general.

8 In the construction of semi-global models in [Reference Rapoport, Smithling and ZhangRSZ19, §§ 4–5], this assumption is relaxed to the assumption that, when $p = 2$, all $v \in \mathcal {V}_2$ are unramified in $F$.

9 In particular, no further conditions are required in AT type (1); this was already observed by Cho [Reference ChoCho18], who based himself on a preprint version of the present paper in which we required $v_0$ to be of degree $1$ over $p$ in all AT types.

10 In [Reference Rapoport, Smithling and ZhangRSZ19, § 6.1] this assumption is relaxed to the assumption that all such $v$ are unramified in $F$, with the analog of Theorem 5.2 below continuing to hold true.

11 Note that these spaces can in fact be (and in particular cases of interest to us, are) DM stacks, but we will suppress this point in our discussion throughout the rest of the paper. The extension of the usual intersection theory to DM stacks is supplied by Gillet's paper [Reference GilletGil09].

12 When the Shimura variety $\mathrm {Sh}_K(\widetilde {HG})$ is non-compact, one has to replace $H^{*}(\mathrm {Sh}_K(\widetilde {HG}),\mathbb{C})$ by the image of the Betti cohomology of the toroidal compactification in the Betti cohomology of $\mathrm {Sh}_K(\widetilde {HG})$. This also coincides with the intersection cohomology of the Baily–Borel compactification of $\mathrm {Sh}_K(\widetilde {HG})$.

13 When the group $\widetilde {HG}$ is anisotropic modulo its center, the quotient is compact and then $L^{2}_{\operatorname {disc}}=L^{2}$.

14 In particular, $Z^{\mathbb{Q}}(\mathbb{R})$ is contained in $K_\infty$.

15 Note that we always have $\dim \operatorname {Hom}_{\mathscr {H}_K}(\pi _f^{K}, \mathcal {Z}_{K,0})\leq 1$, because $\mathcal {Z}_{K,0}$ is a cyclic $\mathscr {H}_K$-module.

16 Instances of everywhere good reduction can in fact be constructed (cf. Remark 6.17).

17 But note that here $G_W$ is the product of two unitary groups.

18 In this special case, the basic locus is characterized as follows. Let $( A_0,\iota _0,\lambda _0,A^{\flat },\iota ^{\flat },\lambda ^{\flat },\overline \eta ^{\flat }, \varphi ^{\flat }, A,\iota ,\lambda ,\overline \eta , \varphi )$ correspond to a point of $\mathcal {M}_{K_{\widetilde {HG}}^{\mathbf{m}}}(\widetilde {HG})$ with values in an algebraically closed extension of $\kappa (\nu )$. Consider the decomposition $A[p^{\infty }]=\prod _{w|p}A[w^{\infty }]$ (respectively $A^{\flat }[p^{\infty }]=\prod _{w|p}A^{\flat }[w^{\infty }]$) of the $p$-divisible group of $A$ (respectively, $A^{\flat }$) under the action of $O_F\otimes \mathbb{Z}_p$; then all factors $A[w^{\infty }]$ (respectively, $A^{\flat }[w^{\infty }]$) are isoclinic.

19 For example, $g=\left ( \begin {array}{ccccc} 0 & 1 & \cdots & 0 & 0\\ 0 & 0 & \cdots & 1 & 0\\ 0 & 0 & \cdots & 0 & 1\\ 1 & 1 & \cdots & 1 & 0 \end {array}\right )\in \mathrm {GL}_{n}(\mathbb{Z})$ satisfies $\Delta (g)\in \{\pm 1\}$.

20 If $L=L_0\oplus L_0$, this means, as elsewhere in the paper, an ordered pair of uniformizers in the usual sense.

21 The corresponding passage in the proof of [Reference Kudla and RapoportKR15, Proposition 3.2] is an over-simplification.

References

Arancibia, N., Moeglin, C. and Renard, D., Paquets d'Arthur des groupes classiques et unitaires, Ann. Fac. Sci. Toulouse Math. (6) 27 (2018), 10231105.10.5802/afst.1590CrossRefGoogle Scholar
Arthur, J., Unipotent automorphic representations: conjectures, Astérisque 171–172 (1989), 1371.Google Scholar
Arthur, J., The endoscopic classification of representations. Orthogonal and symplectic groups, American Mathematical Society Colloquium Publications, vol. 61 (American Mathematical Society, Providence, RI, 2013).Google Scholar
Beilinson, A., Height pairing between algebraic cycles, in Current trends in arithmetical algebraic geometry (Arcata, Calif., 1985), Contemporary Mathematics, vol. 67 (American Mathematical Society, Providence, RI, 1987), 124.Google Scholar
Bloch, S., Height pairings for algebraic cycles, J. Pure Appl. Algebra 34 (1984), 119145.10.1016/0022-4049(84)90032-XCrossRefGoogle Scholar
Borel, A. and Wallach, N., Continuous cohomology, discrete subgroups, and representations of reductive groups, second edition, Mathematical Surveys and Monographs, vol. 67 (American Mathematical Society, Providence, RI, 2000).Google Scholar
Bruinier, J., Howard, B., Kudla, S., Rapoport, M. and Yang, T., Modularity of generating series of divisors on unitary Shimura varieties, Astérisque 421 (2020), 7125.Google Scholar
Chai, C.-L., Conrad, B. and Oort, F., Complex multiplication and lifting problems, Mathematical Surveys and Monographs, vol. 195 (American Mathematical Society, Providence, RI, 2014).Google Scholar
Chenevier, G., email to the authors, 1 Oct. 2016.Google Scholar
Cho, S., The basic locus of the unitary Shimura variety with parahoric level structure, and special cycles, Preprint (2018), arXiv:1807.09997.Google Scholar
Deligne, P., Travaux de Shimura, in Séminaire Bourbaki, 23ème année (1970/71), Exp. No. 389, Lecture Notes in Mathematics, vol. 244 (Springer, Berlin, 1971), 123165.10.1007/BFb0058700CrossRefGoogle Scholar
Faltings, G., Almost étale extensions, Astérisque 279 (2002), 185270.Google Scholar
Fontaine, J.-M., Le corps des périodes $p$-adiques, with an appendix by P. Colmez, Astérisque 223 (1994), 59111.Google Scholar
Gan, W. T., Gross, B. and Prasad, D., Symplectic local root numbers, central critical $L$-values, and restriction problems in the representation theory of classical groups, Astérisque 346 (2012), 1109.Google Scholar
Gillet, H., Arithmetic intersection theory on Deligne–Mumford stacks, in Motives and algebraic cycles, Fields Institute Communications, vol. 56 (American Mathematical Society, Providence, RI, 2009), 93109.Google Scholar
Gillet, H. and Soulé, C., Arithmetic intersection theory, Publ. Math. Inst. Hautes Etudes Sci. 72 (1990), 93174.Google Scholar
Görtz, U., On the flatness of models of certain Shimura varieties of PEL-type, Math. Ann. 321 (2001), 689727.CrossRefGoogle Scholar
Gross, B. and Zagier, D., Heegner points and derivatives of $L$-series, Invent. Math. 84 (1986), 225320.10.1007/BF01388809CrossRefGoogle Scholar
Harris, M. and Taylor, R., The geometry and cohomology of some simple Shimura varieties, with an appendix by V. Berkovich, Annals of Mathematics Studies, vol. 151 (Princeton University Press, Princeton, NJ, 2001).Google Scholar
Howard, B., Complex multiplication cycles and Kudla–Rapoport divisors, Ann. of Math. (2) 176 (2012), 10971171.10.4007/annals.2012.176.2.9CrossRefGoogle Scholar
Howard, B., Complex multiplication cycles and Kudla–Rapoport divisors, II, Amer. J. Math. 137 (2015), 639698.10.1353/ajm.2015.0021CrossRefGoogle Scholar
Jacquet, H., Piatetskii-Shapiro, I. I. and Shalika, J. A., Rankin–Selberg convolutions, Amer. J. Math. 105 (1983), 367464.Google Scholar
Jacquet, H. and Rallis, S., On the Gross-Prasad conjecture for unitary groups, in On certain L-functions, Clay Mathematics Proceedings, vol. 13 (American Mathematical Society, Providence, RI, 2011), 205264.Google Scholar
Jannsen, U., Mixed motives and algebraic K-theory, with appendices by S. Bloch and C. Schoen, Lecture Notes in Mathematics, vol. 1400 (Springer, Berlin, 1990).10.1007/BFb0085080CrossRefGoogle Scholar
Kaletha, T., Minguez, A., Shin, S. W. and White, P.-J., Endoscopic classification of representations: inner forms of unitary groups, Preprint (2014), arXiv:1409.3731.Google Scholar
Katz, N. and Mazur, B., Arithmetic moduli of elliptic curves, Annals of Mathematics Studies, vol. 108 (Princeton University Press, Princeton, NJ, 1985).Google Scholar
Kisin, M. and Pappas, G., Integral models of Shimura varieties with parahoric level structure, Publ. Math. Inst. Hautes Études Sci. 128 (2018), 121218.10.1007/s10240-018-0100-0CrossRefGoogle Scholar
Kottwitz, R., Shimura varieties and λ-adic representations, in Automorphic forms, Shimura varieties, and L-functions, vol. I (Ann Arbor, MI, 1988), Perspectives in Mathematics, vol. 10 (Academic Press, Boston, MA, 1990), 161209.Google Scholar
Kottwitz, R., Points on some Shimura varieties over finite fields, J. Amer. Math. Soc. 5 (1992), 373444.Google Scholar
Kudla, S. and Rapoport, M., Special cycles on unitary Shimura varieties I. Unramified local theory, Invent. Math. 184 (2011), 629682.10.1007/s00222-010-0298-zCrossRefGoogle Scholar
Kudla, S. and Rapoport, M., Special cycles on unitary Shimura varieties II. Global theory, J. Reine Angew. Math. 697 (2014), 91157.Google Scholar
Kudla, S. and Rapoport, M., New cases of $p$-adic uniformization, Astérisque 370 (2015), 207241.Google Scholar
Kudla, S., Rapoport, M. and Zink, Th., On the p-adic uniformization of unitary Shimura curves, Preprint (2020), arXiv:2007.05211Google Scholar
Künnemann, K., Projective regular models for abelian varieties, semistable reduction, and the height pairing, Duke Math. J. 95 (1998), 161212.CrossRefGoogle Scholar
Künnemann, K., Height pairings for algebraic cycles on abelian varieties, Ann. Sci. Éc. Norm. Supér. 34 (2001), 503523.Google Scholar
Mihatsch, A., Relative unitary RZ-spaces and the arithmetic fundamental lemma, J. Inst. Math. Jussieu, to appear. PhD thesis, Bonn (2016), arXiv:1611.06520.Google Scholar
Mihatsch, A., On the arithmetic fundamental lemma conjecture through Lie algebras, Math. Z. 287 (2017), 181197.Google Scholar
Mok, C. P., Endoscopic classification of representations of quasi-split unitary groups, Mem. Amer. Math. Soc. 235 (2015).Google Scholar
Morel, S. and Suh, J., The standard sign conjecture on algebraic cycles: The case of Shimura varieties, J. Reine Angew. Math. 748 (2019), 139151.CrossRefGoogle Scholar
Pappas, G. and Rapoport, M., Local models in the ramified case. III. Unitary groups, J. Inst. Math. Jussieu 8 (2009), 507564.CrossRefGoogle Scholar
Pappas, G., Rapoport, M. and Smithling, B., Local models of Shimura varieties, I. Geometry and combinatorics, in Handbook of moduli, vol. III, Advanced Lectures in Mathematics, vol. 26 (International Press, Somerville, MA, 2013), 135217.Google Scholar
Rapoport, M., Smithling, B. and Zhang, W., On the arithmetic transfer conjecture for exotic smooth formal moduli spaces, Duke Math. J. 166 (2017), 21832336.10.1215/00127094-2017-0003CrossRefGoogle Scholar
Rapoport, M., Smithling, B. and Zhang, W., Regular formal moduli spaces and arithmetic transfer conjectures, Math. Ann. 370 (2018), 10791175.10.1007/s00208-017-1526-2CrossRefGoogle Scholar
Rapoport, M., Smithling, B. and Zhang, W., On Shimura varieties for unitary groups, Pure Appl. Math. Q., to appear. Preprint (2019), arXiv:1906.12346.Google Scholar
Rapoport, M., Terstiege, U. and Wilson, S., The supersingular locus of the Shimura variety for $\textrm {GU}(1, n - 1)$ over a ramified prime, Math. Z. 276 (2014), 11651188.10.1007/s00209-013-1240-zCrossRefGoogle Scholar
Rapoport, M., Terstiege, U. and Zhang, W., On the arithmetic fundamental lemma in the minuscule case, Compos. Math. 149 (2013), 16311666.10.1112/S0010437X13007239CrossRefGoogle Scholar
Rapoport, M. and Zink, Th., Period spaces for p-divisible groups, Annals of Mathematics Studies, vol. 141 (Princeton University Press, Princeton, NJ, 1996).10.1515/9781400882601CrossRefGoogle Scholar
Rapoport, M. and Zink, Th., On the Drinfeld moduli problem of $p$-divisible groups, Camb. J. Math. 5 (2017), 229279.10.4310/CJM.2017.v5.n2.a2CrossRefGoogle Scholar
Smithling, B., On the moduli description of local models for ramified unitary groups, Int. Math. Res. Not. IMRN 2015 (2015), 1349313532.Google Scholar
Tsuji, T., $p$-adic étale cohomology and crystalline cohomology in the semi-stable reduction case, Invent. Math. 137 (1999), 233411.CrossRefGoogle Scholar
Yuan, X., Zhang, S.-W. and Zhang, W., The Gross–Zagier formula on Shimura curves, Annals of Mathematics Studies, vol. 184 (Princeton University Press, Princeton, NJ, 2013).10.1515/9781400845644CrossRefGoogle Scholar
Yun, Z., The fundamental lemma of Jacquet–Rallis in positive characteristics, Duke Math. J. 156 (2011), 167228.10.1215/00127094-2010-210CrossRefGoogle Scholar
Zhang, W., Relative trace formula and arithmetic Gross–Prasad conjecture, Unpublished manuscript (2009).Google Scholar
Zhang, W., On arithmetic fundamental lemmas, Invent. Math. 188 (2012), 197252.10.1007/s00222-011-0348-1CrossRefGoogle Scholar
Zhang, W., Gross-Zagier formula and arithmetic fundamental lemma, in Fifth International Congress of Chinese Mathematicians. Part 1, 2, AMS/IP Studies in Advanced Mathematics, vol. 51 (American Mathematical Society, Providence, RI, 2012), 447459.10.1090/amsip/051.1/32CrossRefGoogle Scholar
Zhang, W., Fourier transform and the global Gan–Gross–Prasad conjecture for unitary groups, Ann. of Math. (2) 180 (2014), 9711049.10.4007/annals.2014.180.3.4CrossRefGoogle Scholar
Zhang, W., Automorphic periods and the central value of Rankin-Selberg $L$-function, J. Amer. Math. Soc. 27 (2014), 541612.Google Scholar
Zhang, W., Periods, cycles, and $L$-functions: a relative trace formula approach, in Proceedings of the International Congress of Mathematicians—Rio de Janeiro 2018. Vol. II. Invited lectures (World Scientific, Hackensack, NJ, 2018), 487521.Google Scholar
Zink, T., Über die schlechte Reduktion einiger Shimuramannigfaltigkeiten, Compos. Math. 45 (1982), 15107.Google Scholar
Zydor, M., Les formules des traces relatives de Jacquet-Rallis grossières, J. Reine Angew. Math. 762 (2020), 195259.Google Scholar