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

Euler systems and relative Satake isomorphism

Li Cai Academy for Multidisciplinary Studies, Beijing National Center for Applied Mathematics, Capital Normal University, Beijing, 100048, People’s Republic of China [email protected] Yangyu Fan Key Laboratory of Algebraic Lie Theory and Analysis of Ministry of Education, School of Mathematics and Statistics, Beijing Institute of Technology, Beijing, 100081, People’s Republic of China [email protected]  and  Shilin Lai Department of Mathematics, the University of Texas at Austin, 2515 Speedway, PMA 8.100, Austin, TX 78712, USA. [email protected]
Abstract.

We explain how the tame norm relation in Euler systems is a formal consequence of the unramified Plancherel formula in the relative Langlands program and a mild refinement of the generalized Cartan decomposition. This uniformly recovers many of the known Euler systems and produces a new split anticyclotomic Euler system in a case studied by Cornut.

1. Introduction

The Bloch–Kato conjecture suggests a deep connection between the arithmetic of motives and special values of LL-functions. The method of Euler systems is a method of understanding this connection. An Euler system consists of a family of motivic classes satisfying two types of relations: “wild” ones for relations in a pp-adic tower, and “tame” ones for relations at places away from pp. Experiences suggest that the wild norm relations require quite special conditions on pp, such as some form of ordinarity. On the other hand, tame norm relations seem to exist in much greater generalities, and they are already enough to deduce some cases of the Bloch–Kato conjecture, cf. [JNS24, LS24]. However, in the many known examples of Euler systems, the construction of classes satisfying the tame norm relations usually requires some ad hoc choices followed by extensive case-by-case calculations.

On the automorphic side of this picture, one way to get a handle on special values of LL-functions is through period integrals. Correspondingly, all known constructions of Euler systems are based (perhaps implicitly) on motivic interpretations of period integrals. Recently, Ben-Zvi–Sakellaridis–Venkatesh proposed a relative Langlands program, which is a far-reaching framework for organizing these period integrals centered around spherical varieties and their generalizations [BZSV24]. It is natural to ask if there is an arithmetic analogue of this framework, which should organize the many constructions of motivic classes (or their realizations) in the literature.

In this paper, we examine Euler systems from the relative Langlands point of view. We explain that this gives an automatic way of producing tame norm relations. Our method is computation-free, and we uniformly recover many of the known examples, some of which are summarized in Table 1. More details can be found in §4.4, 4.5.

Group Spherical variety Attribution
U(n)×U(n+1)\mathrm{U}(n)\times\mathrm{U}(n+1) U(n)\U(n)×U(n+1)\mathrm{U}(n)\backslash\mathrm{U}(n)\times\mathrm{U}(n+1) Lai–Skinner [LS24]
U(2n)\mathrm{U}(2n) U(n)×U(n)\U(2n)\mathrm{U}(n)\times\mathrm{U}(n)\backslash\mathrm{U}(2n) Graham–Shah [GS23]
SO(2n+1)\mathrm{SO}(2n+1) U(n)\SO(2n+1)\mathrm{U}(n)\backslash\mathrm{SO}(2n+1) Cornut [Cor18]
GL2×GL2\mathrm{GL}_{2}\times\mathrm{GL}_{2} 𝐆×GL2𝚜𝚝𝚍\mathbf{G}\times^{\mathrm{GL}_{2}}\mathtt{std} Lei–Loeffler–Zerbes [LLZ14]
ResF/GL2\operatorname{Res}_{F/\mathbb{Q}}\mathrm{GL}_{2} \uparrow Grossi [Gro20]
GU(2,1)\mathrm{GU}(2,1) \uparrow Loeffler–Skinner–Zerbes [LSZ22a]
GSp4×𝔾mGL2\mathrm{GSp}_{4}\times_{\mathbb{G}_{m}}\mathrm{GL}_{2} 𝐆×GL2×𝔾mGL2𝚜𝚝𝚍\mathbf{G}\times^{\mathrm{GL}_{2}\times_{\mathbb{G}_{m}}\mathrm{GL}_{2}}\mathtt{std} Hsu–Jin–Sakamoto [HJR20]
Table 1. Examples of spherical varieties giving rise to tame parts of Euler systems

The idea of using spherical varieties in the construction of Euler systems goes back to Cornut [Cor18]. Unfortunately, there is a gap in his construction related to the definition of Hecke operators, as explained in [Sha23]. For this case, we produce a new split anticyclotomic Euler system and provide an independent proof of the main arithmetic application on Bloch-Kato conjecture.

1.1. Simple case of main result

We will state the application of our results for the pushforward of cycles, covering the first three rows of the above table. We construct split anticyclotomic Euler systems in the sense of Jetchev–Nekovář–Skinner [JNS24], which we will simply call JNS Euler systems.

Let E/FE/F be a CM extension. Suppose 𝐇𝐆\mathbf{H}\hookrightarrow\mathbf{G} are reductive groups with Shimura varieties Sh𝐇Sh𝐆\mathrm{Sh}_{\mathbf{H}}\hookrightarrow\mathrm{Sh}_{\mathbf{G}} defined over EE such that the basic numerology

dimSh𝐆=2dimSh𝐇+1\dim\mathrm{Sh}_{\mathbf{G}}=2\dim\mathrm{Sh}_{\mathbf{H}}+1

holds. After fixing level structures for 𝐆\mathbf{G} and 𝐇\mathbf{H}, this defines a special cycle on Sh𝐆\mathrm{Sh}_{\mathbf{G}} in the arithmetic middle dimension, which we expect is related to the central derivative of an LL-function.

We set up some notations for a JNS Euler system. Let \mathscr{L} be the set of places of FF which split in EE, with an explicit finite set of exceptions. For each \ell\in\mathscr{L}, fix a place λ\lambda of EE above \ell, and let Frobλ\mathrm{Frob}_{\lambda} be the arithmetic Frobenius at λ\lambda. Let \mathscr{R} be the set of square-free products of places in \mathscr{L}. For 𝔪\mathfrak{m}\in\mathscr{R}, let E[𝔪]E[\mathfrak{m}] be the ring class field of conductor 𝔪\mathfrak{m}, so it is associated to the order 𝒪F+𝔪𝒪E\mathcal{O}_{F}+\mathfrak{m}\mathcal{O}_{E} by class field theory. Fix a minuscule representation VV of the dual group Gˇ\check{G} of 𝐆\mathbf{G}. For each \ell, let V,(X)\mathcal{H}_{V,\ell}(X) be the Hecke polynomial of VV at the place \ell (cf. Example 3.12).

Theorem 1.1 (Proposition 4.8+Corollary 4.6).

Let d=dimSh𝐆d=\dim\mathrm{Sh}_{\mathbf{G}}. Suppose 𝐗=𝐇\𝐆\mathbf{X}=\mathbf{H}\backslash\mathbf{G} is a spherical 𝐆\mathbf{G}-variety, and there is a character ν:𝐇U(1)\nu:\mathbf{H}\to\mathrm{U}(1) satisfying the “combinatorially trivial” condition of Example 2.3 such that the induced Shimura datum on U(1)\mathrm{U}(1) is non-trivial. Then there exists a collection of classes

{z𝔪Hcontd+1(Sh𝐆/E[𝔪],p(d))|𝔪}\big{\{}z_{\mathfrak{m}}\in\mathrm{H}^{d+1}_{\mathrm{cont}}(\mathrm{Sh}_{\mathbf{G}/E[\mathfrak{m}]},\mathbb{Z}_{p}(d))\,\big{|}\,\mathfrak{m}\in\mathscr{R}\big{\}}

such that whenever 𝔪,𝔪\mathfrak{m},\mathfrak{m}\ell\in\mathscr{R}, we have the tame norm relation

TrE[𝔪]E[𝔪]z𝔪=V,(Frobλ1)z𝔪.\operatorname{Tr}_{E[\mathfrak{m}]}^{E[\mathfrak{m}\ell]}z_{\mathfrak{m}\ell}=\mathcal{H}_{V,\ell}(\mathrm{Frob}_{\lambda}^{-1})\cdot z_{\mathfrak{m}}.

Moreover z1z_{1} is the image of the special cycle under the continuous étale cycle class map.

We can in fact replace p\mathbb{Z}_{p} by more general coefficient systems. Under some automorphic assumptions, a standard Abel–Jacobi map procedure described in §4.3 gives the required Euler system in Galois cohomology.

Corollary 1.2.

Suppose in addition that

  • 𝐆\mathbf{G} is anisotropic modulo centre.

  • Condition (C’) of [MS19] holds for 𝐆\mathbf{G}.

  • Kottwitz’s conjecture [BR94, Conjecture 5.2] holds for the middle degree cohomology of Sh𝐆/E¯\mathrm{Sh}_{\mathbf{G}/\bar{E}}.

Let π\pi be a stable cohomological automorphic representation of 𝐆(𝔸F)\mathbf{G}(\mathbb{A}_{F}) distinguished by 𝐗\mathbf{X}. Let ρπ\rho_{\pi} be the pp-adic Galois representation attached to π\pi and the Shimura cocharacter for 𝐆\mathbf{G}.

Under the above set-up, there is a lattice TπT_{\pi} in ρπ\rho_{\pi} and a collection of Galois cohomology classes

{c𝔪H1(E[𝔪],Tπ)|𝔪}\{c_{\mathfrak{m}}\in\mathrm{H}^{1}(E[\mathfrak{m}],T_{\pi})\,|\,\mathfrak{m}\in\mathscr{R}\}

forming the tame part of a JNS Euler system. In other words, whenever 𝔪,𝔪\mathfrak{m},\mathfrak{m}\ell\in\mathscr{R}, we have the tame norm relation

TrE[𝔪]E[𝔪]c𝔪=Pλ(Frobλ1)c𝔪,\operatorname{Tr}_{E[\mathfrak{m}]}^{E[\mathfrak{m}\ell]}c_{\mathfrak{m}\ell}=P_{\lambda}(\mathrm{Frob}_{\lambda}^{-1})c_{\mathfrak{m}},

where Pλ(X)=det(1XFrobλ|ρπ)P_{\lambda}(X)=\det(1-X\mathrm{Frob}_{\lambda}|\rho_{\pi}) is the characteristic polynomial of Frobλ\mathrm{Frob}_{\lambda}.

Remark 1.3.

We briefly explain the roles of the conditions, which are not too serious thanks to a large body of work in the area.

  • The anisotropic modulo centre is a simplifying condition so that the Shimura variety is compact and we can directly apply the above cited works.

  • Condition (C’) is a collection of statements related to Arthur’s conjecture for 𝐆\mathbf{G}, If 𝐆\mathbf{G} is a unitary or orthogonal group, then condition (C’) is known (cf. the discussion after Remark 1.6 in [MS19]). Their result is only used to modify our classes z𝔪z_{\mathfrak{m}} to be null-homologous. In the unitary case, [LL21, Proposition 6.9] also suffices for this purpose.

  • Kottwitz’s conjecture is used to show that ρπ\rho_{\pi} actually contributes to the cohomology of Sh𝐆/E¯\mathrm{Sh}_{\mathbf{G}/\bar{E}}. In all of our cases, the Shimura variety is of abelian type, and what we need follows from the work of Kisin–Shin–Zhu [KSZ21].

By combining our construction with the results of [JNS24], we obtain the following implication.

Theorem 1.4.

With notations as in the previous corollary, suppose the Galois representation ρπ:GalEGL(Vπ)\rho_{\pi}:\operatorname{Gal}_{E}\to\mathrm{GL}(V_{\pi}) satisfies the following conditions.

  1. (1)

    ρπ\rho_{\pi} is absolutely irreducible.

  2. (2)

    There exists σGalE[1](μp)\sigma\in\operatorname{Gal}_{E[1](\mu_{p^{\infty}})} such that dimVπ/(σ1)Vπ=1\dim V_{\pi}/(\sigma-1)V_{\pi}=1.

  3. (3)

    There exists γGalE[1](μp)\gamma\in\operatorname{Gal}_{E[1](\mu_{p^{\infty}})} such that Vπ/(γ1)Vπ=0V_{\pi}/(\gamma-1)V_{\pi}=0.

Then

c10dimHf1(E,ρπ)=1c_{1}\neq 0\implies\dim\mathrm{H}^{1}_{f}(E,\rho_{\pi})=1

We also have a version of the above results for the pushforward of Eisenstein classes (Example 4.3), which recovers the final four entries in the table. However, they depend on numerical consequences of the results of Sakellaridis–Wang [SW22] which are not yet available in the mixed characteristics setting.

Remark 1.5.

Since we are not using wild norm relations, there is no specific hypothesis on the prime pp in any of the above results.

1.2. Idea of proof

The theorem above is a combination of the following two steps which are completely different in nature:

  1. (1)

    Construction of a “motivic theta series” (Definition 4.1).

  2. (2)

    Local harmonic analysis on spherical varieties (Proposition 3.13).

We now explain each item in turn.

1.2.1. Motivic theta series

By definition, this is a 𝐆(𝔸p)\mathbf{G}(\mathbb{A}^{p\infty})-equivariant map between an adèlic function space and certain “motivic classes”. In the settings considered in this paper, the continuous étale cohomology group plays the role of this space of motivic classes. For arithmetic applications, it is important to have integral coefficients on the target cohomology group.

For the pushforward construction, versions of this map with rational coefficients have appeared, for example in [LSZ22a, Definition 9.2.3] and [GS23, Proposition 9.14]. Our first main idea is that by considering the correct function space, there is a natural integral refinement of this construction.

Observation 1.6 (Definition 4.1, Propositions 4.8, 4.11).

In many cases, including the ones cited above, a more natural statement is that there is an integral, 𝐆(𝔸p)\mathbf{G}(\mathbb{A}^{p\infty})-equivariant map

Cc(𝐗(𝔸p),p){Integral motivic classes},C_{c}^{\infty}(\mathbf{X}(\mathbb{A}^{p\infty}),\mathbb{Z}_{p})\to\{\text{Integral motivic classes}\},

where 𝐗\mathbf{X} is a spherical variety.

In the cycles case, we take 𝐗=𝐇\𝐆\mathbf{X}=\mathbf{H}\backslash\mathbf{G}, so we are interested in the function space Cc(𝐇\𝐆,p)C_{c}^{\infty}(\mathbf{H}\backslash\mathbf{G},\mathbb{Z}_{p}). In previous works, the source space is often taken to be a space of coinvariants Cc(𝐆,p)𝐇C_{c}^{\infty}(\mathbf{G},\mathbb{Z}_{p})_{\mathbf{H}}. These two spaces are equal after tensoring with p\mathbb{Q}_{p}, but they are different integrally. In other words, our observation is that Cc(𝐗,p)C_{c}^{\infty}(\mathbf{X},\mathbb{Z}_{p}) is the more natural integral structure. The cases involving Eisenstein classes can be treated similarly, taking 𝐗\mathbf{X} to be the vector bundle

𝐗=𝐆×GL2𝚜𝚝𝚍=GL2\(𝐆×𝚜𝚝𝚍),\mathbf{X}=\mathbf{G}\times^{\mathrm{GL}_{2}}\mathtt{std}=\mathrm{GL}_{2}\backslash(\mathbf{G}\times\mathtt{std}),

instead of the previously used coinvariant space Cc(𝐆×𝚜𝚝𝚍,p)GL2C_{c}^{\infty}(\mathbf{G}\times\mathtt{std},\mathbb{Z}_{p})_{\mathrm{GL}_{2}}. Here, 𝚜𝚝𝚍\mathtt{std} is the standard two-dimensional representation of GL2\mathrm{GL}_{2}.

This observation was already used by the third named author in [LS24] to bypass some Iwasawa-theoretic arguments needed previously to deal with torsions.

1.2.2. Local harmonic analysis

Once such a motivic theta series exist, the problem of tame norm relations is entirely reduced to constructing test vectors satisfying norm relations in Cc(X,p)C_{c}^{\infty}(X,\mathbb{Z}_{p}), where X=𝐗(F)X=\mathbf{X}(F_{\ell}).

The first question to understand is the origin of the field extension E[𝔪]E[\mathfrak{m}]. Inspired by [Loe21, GS23], we introduce an augmented group 𝐆~=𝐆×𝐓\widetilde{\mathbf{G}}=\mathbf{G}\times\mathbf{T}, where 𝐓\mathbf{T} is a one-dimensional torus that is supposed to parameterize character twists. By considering the desiderata of such an augmentation, we are led to the definition of a combinatorially trivial 𝐓\mathbf{T}-bundle (Definition 2.2). This ensures that π\pi is 𝐗\mathbf{X}-distinguished implies that π×χ\pi\times\chi is 𝐗~\widetilde{\mathbf{X}}-distinguished for any χ𝐓^\chi\in\hat{\mathbf{T}}, giving rise to a family of character twists.

Using this bundle, we can define two level structures: for i=0,1i=0,1, let Ki=𝐆(𝒪)×𝐓iK^{i}=\mathbf{G}(\mathcal{O})\times\mathbf{T}^{i}, where 𝐓0=𝐓(𝒪)\mathbf{T}^{0}=\mathbf{T}(\mathcal{O}) and 𝐓1\mathbf{T}^{1} is the subset which is congruent to 1 modulo the uniformizer. The existence of tame norm relations is reduced to the following question: given a Hecke operator \mathcal{H}_{\ell} and a “basic element” Φ0\Phi_{0},

(†) is it true that Φ0TrK0K1Cc(X,p)K1?\text{is it true that }\mathcal{H}_{\ell}\cdot\Phi_{0}\in\operatorname{Tr}_{K^{0}}^{K^{1}}C_{c}^{\infty}(X,\mathbb{Z}_{p})^{K^{1}}?

This is stronger than just requiring the function Φ0\mathcal{H}_{\ell}\cdot\Phi_{0} to take values in p\mathbb{Z}_{p}. Indeed, in the extreme case where the K0K^{0}-orbit of a point xXx\in X coincides with its K1K^{1}-orbit, we need the value at xx to be divisible by the index [K0:K1]=1[K^{0}:K^{1}]=\ell-1. We give a necessary condition for when this kind of additional divisibility requirements can occur, in terms of the geometry of 𝐗\mathbf{X} (Proposition 3.4). This is done by extending the proof of the generalized Cartan decomposition due to Gaitsgory–Nadler [GN10, Theorem 8.2.9] and Sakellaridis [Sak12, Theorem 2.3.8],

To verify these divisibility conditions, we need to compute Φ0\mathcal{H}_{\ell}\cdot\Phi_{0}. Since \mathcal{H}_{\ell} is described using its Satake image, we use works on the relative Satake isomorphism by Sakellaridis [Sak13] and Sakellaridis–Wang [SW22]. Our second main idea is the following.

Observation 1.7 (Proposition 3.9).

The structure of the inverse relative Satake transform implies additional divisibility properties.

In the cases considered in this paper, the formula for the inverse relative Satake isomorphism gives the values of the function Φ0\mathcal{H}_{\ell}\cdot\Phi_{0} as polynomials in \ell. Now view \ell as a formal variable. Instead of computing this polynomial, we specialize at =1\ell=1 and show by an anti-symmetry argument that the result is 0 at certain points in XX (along “walls of type T”). This proves its divisibility by 1\ell-1 at those points. This additional divisibility is a new phenomenon in the relative setting and not present in the classical Satake isomorphism (Remark 3.10).

Combining these two computations, we note that this automatic divisibility is stronger than the geometric requirements, so the answer to (1.2.2) is yes, and we obtain the tame norm relation!

1.3. Related works

In our language, the gap in Cornut’s construction [Cor18] (see [Sha23]) is a difference between functions on 𝐗\mathbf{X} and distributions on 𝐗\mathbf{X}. In this paper, we establish tame norm relations at split primes. We believe our framework also works in the original non-split setting studied by Cornut and we plan to pursue such a generalization in a future paper (see §1.4.1 below).

In the other works cited in Table 1, the Euler system is constructed using the zeta integral method first developed by Loeffler–Skinner–Zerbes [LSZ22b]. In this method, one writes down a carefully chosen candidate for the test vector in Cc(X,p)K1C_{c}^{\infty}(X,\mathbb{Z}_{p})^{K^{1}} and verifies that its trace is equal to Φ0\mathcal{H}_{\ell}\cdot\Phi_{0} by an explicit, often intricate, zeta integral computation. It also relies crucially on local multiplicity one. Our method avoids both the computations and the local multiplicity one hypothesis.

In certain spherical settings where classes are obtained by a pushforward construction, Loeffler gave a construction of wild norm relations [Loe21]. He also considered more general cases where a mirabolic subgroup has an open orbit (Definition 4.1.1 of op. cit.). In many such cases, we reinterpret the mirabolic subgroup as the point stabilizer of a naturally occurring affine inhomogeneous spherical variety. However, in some examples such as (GSp4,GL2×𝔾mGL2)(\mathrm{GSp}_{4},\mathrm{GL}_{2}\times_{\mathbb{G}_{m}}\mathrm{GL}_{2}), only a parabolic subgroup of 𝐆\mathbf{G} has an open orbit. Our method does not handle this case. This is an Eisenstein degeneration of the spherical pair (SO4×SO5,SO4)(\mathrm{SO}_{4}\times\mathrm{SO}_{5},\mathrm{SO}_{4}), in a precise sense explained in [LR24], so it would be interesting to understand how this can be interpreted as an operation on spherical varieties.

During the preparation of this work, the preprint [Sha24] was posted. Shah considered a similar local question as (1.2.2), phrased using double cosets on the group 𝐆\mathbf{G}. At this point, the classical Satake isomorphism is used in op. cit., leading to complicated expressions involving Kazhdan–Lusztig polynomials, which need to be computed on a case-by-case basis. However, his method can treat certain non-spherical cases, which at the present falls outside the conjectural framework of [BZSV24].

1.4. Further works

1.4.1. Removing conditions

To make our results unconditional, we would need to have the function-level results of [SW22] for mixed characteristic local fields. Such a statement should follow from motivic integration techniques, along the lines of [CHL11]. The third named author plans to write a short note on this matter in the future, though it would certainly be more desirable to have a sheaf-level statement and proof.

It would also be useful to have a non-split version of our results. Indeed, in certain twisted settings, even the construction of a split anticyclotomic Euler system requires non-split groups. The necessary local harmonic analysis results should be within reach of current methods. For example, [SW22] only assumes quasi-split, and [CZ23] handles many symmetric cases.

1.4.2. Speculations

One byproduct of our observation of automatic divisibility (Proposition 3.9) is that there should be some modification to the local unramified conjecture [BZSV24, Conjecture 7.5.1] when the coefficient ring is not a field, cf. §9.4 of op. cit. We hope this provides a useful piece of phenomenon in the future refinement of this framework.

Perhaps the deepest question raised by our work is to understand a space of the form

Hom𝐆(𝔸)(Fun(𝐗(𝔸),),{Integral motivic classes})\operatorname{Hom}_{\mathbf{G}(\mathbb{A})}(\mathrm{Fun}(\mathbf{X}(\mathbb{A}),\mathbb{Z}),\{\text{Integral motivic classes}\})

in larger generality. Indeed, replacing “integral motivic classes” with “automorphic functions”, then this space (though not its two constituent pieces) has a conjectural dual description in the relative Langlands program. In addition to the pushforward of cycles or motivic classes described above, the arithmetic theta lifting (cf. [Liu11, LL21, Dis24]) should also be part of this framework. It would also be interesting to understand its relation with the recent boundary class construction of Skinner–Sangiovanni-Vincentelli [SV24].

Acknowledgments

It is clear that most of this paper is built on the fundamental works of Yiannis Sakellaridis, and we would like to thank him in particular for some very insightful conversations.

We would also like to thank David Ben-Zvi, Ashay Burungale, Christopher Skinner, Ye Tian and Wei Zhang for their interests and helpful discussions.

The first and second named authors are supported by the National Key R&\&D Program of China No. 2023YFA1009702 and the National Natural Science Foundation of China No. 12371012. The second named author is also supported by National Natural Science Foundation of Beijing, China No. 24A10020.

2. Spherical varieties

Let FF be a field of characteristic 0, not necessarily algebraically closed. Unless otherwise specified, everything in this section will be defined over FF.

Let 𝐆\mathbf{G} be a split reductive group with a Borel subgroup 𝐁\mathbf{B}. Let 𝐗\mathbf{X} be a variety with a right 𝐆\mathbf{G}-action. Recall that 𝐗\mathbf{X} is spherical if 𝐁𝐆\mathbf{B}\subseteq\mathbf{G} acting on 𝐗\mathbf{X} has an open orbit. We make the following assumptions

  1. (1)

    𝐗\mathbf{X} is smooth, affine, connected.

  2. (2)

    𝐗\mathbf{X} has no root of type N (cf. §2.1.2).

  3. (3)

    Every 𝐁\mathbf{B}-orbit on 𝐗/F¯\mathbf{X}_{/\overline{F}} contains an FF-point.

  4. (4)

    𝐗\mathbf{X} has an invariant 𝐆\mathbf{G}-measure.

The assumptions imply that there is a unique open 𝐁\mathbf{B}-orbit even on the level of FF-points. We denote this orbit by 𝐗̊\mathring{\mathbf{X}}. Fix once and for all a point x0𝐗̊(F)x_{0}\in\mathring{\mathbf{X}}(F). Let 𝐇\mathbf{H} be the stabilizer of x0x_{0}. Let 𝐗\mathbf{X}^{\bullet} be the open 𝐆\mathbf{G}-orbit in 𝐗\mathbf{X}, so 𝐗𝐇\𝐆\mathbf{X}^{\bullet}\simeq\mathbf{H}\backslash\mathbf{G}, and it contains 𝐗̊\mathring{\mathbf{X}} as an open dense subset.

Remark 2.1.

Assumptions (1) and (2) roughly correspond to the assumptions imposed in [BZSV24] in the case of polarized Hamiltonian varieties. There have been progresses towards the unramified Plancherel formula without some of these hypothesis, for example [SW22] for certain singular varieties, and [CZ23] for certain varieties with roots of type N. It would be interesting to see if the behaviour observed in this paper still holds in these settings.

Assumption (3) is purely a matter of convenience for the present paper to rule out spherical roots of type T non-split, which requires a separate analysis. In a future work, we plan to remove it and moreover consider the case when 𝐆\mathbf{G} is not split.

Assumption (4) is included also for simplicity of notation. It will hold if 𝐇\mathbf{H} is reductive or if 𝐗\mathbf{X} is of the form 𝐆×𝐇V\mathbf{G}\times^{\mathbf{H}^{\prime}}V, where 𝐇\mathbf{H}^{\prime} is reductive and VV is a linear representation of 𝐇\mathbf{H}^{\prime}. These are the two cases needed for the Euler system constructions in the paper. In general, one may assume there is an 𝐆\mathbf{G}-eigenmeasure after a trivial modification [Sak08, §3.8], and our formulae in §3.3 need to be modified by the corresponding character in a well-understood way.

2.1. Structure theory

We now recall some general results from the theory of spherical varieties, stating only what is needed for the present work. We refer to [Sak13, SV17] and references found therein for a more systematic development with proofs.

2.1.1. Notations

Let 𝐏(𝐗)\mathbf{P}(\mathbf{X}) be the subgroup of 𝐆\mathbf{G} fixing the open orbit. It contains the Borel subgroup, so it is a parabolic subgroup. Choose a good Levi subgroup 𝐋(𝐗)\mathbf{L}(\mathbf{X}) as in [Sak13, §2.1], and let 𝐀\mathbf{A} be a maximal torus of 𝐆\mathbf{G} contained in 𝐁𝐋(𝐗)\mathbf{B}\cap\mathbf{L}(\mathbf{X}). Define the torus

𝐀X=𝐀/(𝐀𝐇)\mathbf{A}_{X}=\mathbf{A}/(\mathbf{A}\cap\mathbf{H})

Write ΛX\Lambda_{X} for its cocharacter lattice, and let 𝔞X=ΛX\mathfrak{a}_{X}=\Lambda_{X}\otimes_{\mathbb{Z}}\mathbb{Q}. Similarly define 𝔞\mathfrak{a}, then we have a quotient map 𝔞𝔞X\mathfrak{a}\twoheadrightarrow\mathfrak{a}_{X}. There is a natural cone 𝒱𝔞X\mathcal{V}\subseteq\mathfrak{a}_{X} containing the image of the negative Weyl chamber. Let ΛX+=𝒱ΛX\Lambda_{X}^{+}=\mathcal{V}\cap\Lambda_{X}. Its elements will be called 𝐗\mathbf{X}-anti-dominant.

Let 𝒱\mathcal{V}^{\perp} be the negative-dual cone to 𝒱\mathcal{V} in X(𝐀X)X^{*}(\mathbf{A}_{X})\otimes_{\mathbb{Z}}\mathbb{Q}, then 𝒱\mathcal{V}^{\perp} is strictly convex. In [Sak13, §6.1], Sakellaridis defined a based root system ΦX\Phi_{X} whose set of simple roots ΔX\Delta_{X} lie on extremal rays of 𝒱\mathcal{V}^{\perp} intersected with X(𝐀X)X^{*}(\mathbf{A}_{X}). Elements of ΔX\Delta_{X} are called (normalized) spherical roots of 𝐗\mathbf{X}. There is a canonical embedding X(𝐀X)X(𝐀)X^{*}(\mathbf{A}_{X})\hookrightarrow X^{*}(\mathbf{A}), allowing us to view spherical roots as characters of 𝐀\mathbf{A}. The Weyl group of ΦX\Phi_{X} is the little Weyl group, denoted by WXW_{X}. It is canonically contained in the Weyl group of 𝐆\mathbf{G}, which we will denote by WW.

Knop and Schalke defined a dual group GˇX\check{G}_{X} whose coroot system is (X(𝐀X),ΦX)(X^{*}(\mathbf{A}_{X}),\Phi_{X}) [KS17]. It is a subgroup of the Langlands dual group Gˇ\check{G} of 𝐆\mathbf{G}. We have an equality GˇX=Gˇ\check{G}_{X}=\check{G} only if for all simple roots α\alpha, the pair (𝐗̊,α)(\mathring{\mathbf{X}},\alpha) is of type T in the classification below. We call 𝐗\mathbf{X} with this property strongly tempered.

2.1.2. Classification of roots

Let α\alpha be a simple root of 𝐆\mathbf{G}. Let 𝐏α\mathbf{P}_{\alpha} be its associated standard parabolic subgroup, with radical (𝐏α)\mathcal{R}(\mathbf{P}_{\alpha}). Let 𝐘\mathbf{Y} be a 𝐁\mathbf{B}-orbit in 𝐗\mathbf{X}. The geometric quotient

𝐘𝐏α/(𝐏α)\mathbf{Y}\mathbf{P}_{\alpha}/\mathcal{R}(\mathbf{P}_{\alpha})

is a homogeneous spherical variety of PGL2\mathrm{PGL}_{2}. There are four cases.

  • Type G: =PGL2\PGL2\ast=\mathrm{PGL}_{2}\backslash\mathrm{PGL}_{2}.

  • Type U: 𝐒\PGL2\mathbf{S}\backslash\mathrm{PGL}_{2}, where 𝐒\mathbf{S} is the subgroup of a Borel subgroup which contains the unipotent radical.

  • Type T: 𝐓\PGL2\mathbf{T}\backslash\mathrm{PGL}_{2}, where 𝐓\mathbf{T} is a maximal torus.

  • Type N: 𝐍(𝐓)\PGL2\mathbf{N}(\mathbf{T})\backslash\mathrm{PGL}_{2}, where 𝐍(𝐓)\mathbf{N}(\mathbf{T}) is the normalizer of the torus. This case is excluded by our assumption.

We say the pair (𝐘,α)(\mathbf{Y},\alpha) is of type G, U, or T according to this classification.

Somewhat confusingly, there is a related but separate classification of spherical roots. Note that it is also standard terminology that the spherical roots only refer to the simple roots of the spherical root system ΦX\Phi_{X}. Let γ\gamma be a spherical root. Under our standing assumptions, it has one of the two types.

  • Type T: This happens exactly if γ\gamma is a root of 𝐆\mathbf{G}.

  • Type G: In all other cases, γ=α+β\gamma=\alpha+\beta, where α,β\alpha,\beta are orthogonal roots of 𝐆\mathbf{G}, and they are simple roots in some choice of basis.

More details, as well as a proof of the above dichotomy, can be found in [Sak13, §6.2]. The ample examples there should illustrate the classification. We simply note that by the description in loc. cit., if there is a pair (𝐘,α)(\mathbf{Y},\alpha) of maximal rank of type T, then there is a spherical root of type T.

2.2. Equivariant bundles

For our intended application, it is necessary to consider certain torus bundles over 𝐗\mathbf{X}.

Definition 2.2.

Let 𝐗~𝐗\widetilde{\mathbf{X}}\to\mathbf{X} be an 𝐆\mathbf{G}-equivariant 𝐓\mathbf{T}-bundle, where 𝐓\mathbf{T} is a torus. It is combinatorially trivial if its restriction to 𝐗̊\mathring{\mathbf{X}} is trivial as a 𝐁\mathbf{B}-equivariant bundle.

In particular, 𝐗~\widetilde{\mathbf{X}} is a spherical variety for the group 𝐆~:=𝐆×𝐓\widetilde{\mathbf{G}}:=\mathbf{G}\times\mathbf{T}, and the stabilizer of any lift of x0x_{0} is isomorphic to 𝐇\mathbf{H} by projection to the first factor. It is immediate that

𝐀X~=𝐀X×𝐓\mathbf{A}_{\widetilde{X}}=\mathbf{A}_{X}\times\mathbf{T}

as quotients of the maximal torus 𝐀×𝐓\mathbf{A}\times\mathbf{T}. As a result, all of the combinatorial data introduced in the previous subsection are either unchanged or change in a trivial way. This explains our terminology. In particular, we note that the classification of spherical roots is unchanged.

Example 2.3.

If 𝐗=𝐇\𝐆\mathbf{X}=\mathbf{H}\backslash\mathbf{G} is homogeneous, then the datum of an equivariant 𝐓\mathbf{T}-bundle over 𝐗\mathbf{X} is equivalent to a character ν:𝐇𝐓\nu:\mathbf{H}\to\mathbf{T} by the recipe

𝐗~=𝐇~\(𝐆×𝐓),𝐇~={(h,ν(h))|h𝐇}\widetilde{\mathbf{X}}=\widetilde{\mathbf{H}}\backslash(\mathbf{G}\times\mathbf{T}),\quad\widetilde{\mathbf{H}}=\{(h,\nu(h))\,|\,h\in\mathbf{H}\}

It is combinatorially trivial if and only if ν|𝐇𝐀=1\nu|_{\mathbf{H}\cap\mathbf{A}}=1, where recall that 𝐀\mathbf{A} is a maximal torus in the Borel subgroup 𝐁\mathbf{B} such that 𝐇𝐁\mathbf{H}\mathbf{B} is open in 𝐆\mathbf{G}. This is the condition imposed by Loeffler for wild norm relations [Loe21, §4.6].

Observe that if 𝐇\mathbf{H} contains a maximal unipotent subgroup of 𝐆\mathbf{G} (so 𝐗\mathbf{X} is horospherical), then any combinatorially trivial bundle is trivial, since ν\nu has to be trivial on the unipotent part.

Remark 2.4.

Combinatorially trivial in particular implies that

GˇX~=GˇX×𝐓ˇ.\check{G}_{\widetilde{X}}=\check{G}_{X}\times\check{\mathbf{T}}.

On the representation theory side, this means if π\pi is 𝐗\mathbf{X}-distinguished, then π×χ\pi\times\chi is 𝐗~\widetilde{\mathbf{X}}-distinguished for any character χ\chi of the torus. So we are looking a family of character twists for π\pi, exactly what is needed for Euler system constructions. If 𝐓=𝔾m\mathbf{T}=\mathbb{G}_{m}, this is related to having an “ss-variable” in the LL-function.

The existence of such a bundle is unfortunately quite restrictive, and it rules out very interesting cases, including the triple product case 𝐆=SO3×SO4\mathbf{G}=\mathrm{SO}_{3}\times\mathrm{SO}_{4}, 𝐇=SO3\mathbf{H}=\mathrm{SO}_{3}.

Proposition 2.5.

Suppose 𝐗\mathbf{X} is homogeneous and has only spherical roots of type G, then every combinatorially trivial 𝐓\mathbf{T}-bundle over 𝐗\mathbf{X} is trivial.

Proof.

Let 𝐘\mathbf{Y} be a 𝐁\mathbf{B}-orbit in 𝐗\mathbf{X} of maximal rank. Let α\alpha be a simple root of 𝐆\mathbf{G} with associated parabolic 𝐏α\mathbf{P}_{\alpha}. The pair (𝐘,α)(\mathbf{Y},\alpha) cannot have type T, since otherwise there would be a spherical root of type T. It follows that the pair is of type G or U. In particular, 𝐘𝐏α\mathbf{Y}\mathbf{P}_{\alpha} is either 𝐘\mathbf{Y} or the disjoint union of an open 𝐁\mathbf{B}-orbit and a closed 𝐁\mathbf{B}-orbit. Moreover, all 𝐁\mathbf{B}-orbits are of maximal rank, so all pairs (𝐘,α)(\mathbf{Y},\alpha) are not of type T.

Let 𝐗~\widetilde{\mathbf{X}} be any equivariant 𝔾m\mathbb{G}_{m}-bundle. Suppose it trivializes as a 𝐁\mathbf{B}-equivariant bundle over 𝐘\mathbf{Y} and α\alpha is a root such that 𝐘\mathbf{Y} is the open orbit in 𝐘𝐏α\mathbf{Y}\mathbf{P}_{\alpha}. Let 𝐇\mathbf{H} be the stabilizer of a point in 𝐘\mathbf{Y}, so the bundle structure becomes a map ν:𝐇𝐓\nu:\mathbf{H}\to\mathbf{T}. Consider the short exact sequence of groups

1𝐇(𝐏α)𝐇𝐏α((𝐇𝐏α)(𝐏α))/(𝐏α)11\to\mathbf{H}\cap\mathcal{R}(\mathbf{P}_{\alpha})\to\mathbf{H}\cap\mathbf{P}_{\alpha}\to\big{(}(\mathbf{H}\cap\mathbf{P}_{\alpha})\mathcal{R}(\mathbf{P}_{\alpha})\big{)}/\mathcal{R}(\mathbf{P}_{\alpha})\to 1

Since 𝐏α𝐁\mathbf{P}_{\alpha}\subseteq\mathbf{B}, the restriction of ν\nu to the first term is trivial, so it descends to a character ν¯\bar{\nu} on the quotient. On the other hand, this quotient is the stabilizer group in the PGL2\mathrm{PGL}_{2}-spherical variety 𝐘𝐏α/(𝐏α)\mathbf{Y}\mathbf{P}_{\alpha}/\mathcal{R}(\mathbf{P}_{\alpha}), which is of type U. But then ν¯\bar{\nu} is trivial by Example 2.3. Therefore, ν|𝐇𝐏α=1\nu|_{\mathbf{H}\cap\mathbf{P}_{\alpha}}=1. In other words, 𝐗~\widetilde{\mathbf{X}} trivializes as a 𝐁α\mathbf{B}_{\alpha}-bundle over 𝐘𝐏α\mathbf{Y}\mathbf{P}_{\alpha}.

Suppose that 𝐗~\widetilde{\mathbf{X}} is combinatorially trivial, then we can construct a 𝐁\mathbf{B}-equivariant section over 𝐗̊\mathring{\mathbf{X}}. By the above discussion, this section extends to 𝐗̊𝐏α\mathring{\mathbf{X}}\mathbf{P}_{\alpha} for any simple root α\alpha. We may continue this process starting from the closed orbits in 𝐗̊𝐏α\mathring{\mathbf{X}}\mathbf{P}_{\alpha}. Since 𝐗\mathbf{X} is homogeneous, all 𝐁\mathbf{B}-orbits are reached this way. Moreover, all of the sections glue since the 𝐁\mathbf{B}-equivariant sections on a 𝐁\mathbf{B}-orbit are unique up to a constant multiple. Therefore, we have shown that 𝐗~\widetilde{\mathbf{X}} is trivial as a 𝐁\mathbf{B}-equivariant bundle.

In particular, 𝐗~\widetilde{\mathbf{X}} is trivial as a line bundle, so the only 𝐆\mathbf{G}-equivariance structure on it comes from a character χ:𝐆𝐓\chi:\mathbf{G}\to\mathbf{T}. Its restriction to 𝐁\mathbf{B} is trivial by the discussion above, so χ\chi itself must be trivial. ∎

3. Local computations

We now specialize to the case where FF is a local field, with ring of integers 𝒪\mathcal{O} and residue field 𝔽\mathbb{F}. Let q=#𝔽q=\#\mathbb{F}. Let ϖ\varpi be a uniformizer. For any of the varieties denoted by bold letters, we will use the normal font to denote its FF-points, so for example X=𝐗(F)X=\mathbf{X}(F).

The group GG has a natural smooth left action on the function space C(X,)C^{\infty}(X,\mathbb{C}). In this section, we recall its spectral decomposition, following the works of Sakellaridis and his collaborators. The goal is to observe certain automatic divisibility properties and match them with a corresponding phenomenon in geometry.

3.1. Assumptions

For all results in this section, we need to impose some “good reduction” hypotheses, which we will specify. In the global setting, they hold for all but finitely many places.

Assumption 3.1.

Both 𝐆\mathbf{G} and 𝐗\mathbf{X} extend to smooth schemes over Spec𝒪\operatorname{Spec}\mathcal{O}, which we denote by the same letter. Moreover, all statements of [Sak12, Proposition 2.3.5] hold. In particular,

  1. (1)

    𝐆\mathbf{G} is reductive and 𝐗\mathbf{X} is affine.

  2. (2)

    The chosen base point x0x_{0} belongs to 𝐗(𝒪)\mathbf{X}(\mathcal{O}).

  3. (3)

    A local structure theorem for 𝐗\mathbf{X} and its compactifications hold.

We will write K=𝐆(𝒪)K=\mathbf{G}(\mathcal{O}). By hypothesis, this is a hyperspecial maximal compact subgroup of GG.

3.2. Generalized Cartan decomposition

Under the assumptions made above, we will state a generalized Cartan decomposition due to Gaitsgory–Nadler in the equal characteristics case [GN10, Theorem 8.2.9] and adpoted to the mixed characteristics case by Sakellaridis [Sak12, Theorem 2.3.8]. Recall that ΛX𝐀𝐗(F)/𝐀𝐗(𝒪)\Lambda_{X}\simeq\mathbf{A}_{\mathbf{X}}(F)/\mathbf{A}_{\mathbf{X}}(\mathcal{O}) contains an 𝐗\mathbf{X}-anti-dominant monoid ΛX+\Lambda_{X}^{+}.

Proposition 3.2 (Genralized Cartan decomposition).

For each λˇΛX+\check{\lambda}\in\Lambda_{X}^{+}, fix a representation xλˇAXx_{\check{\lambda}}\in A_{X}, viewed as an element of XX by the orbit map through x0x_{0}. Then there is a disjoint union decomposition

X=λˇΛX+xλˇK.X^{\bullet}=\bigsqcup_{\check{\lambda}\in\Lambda_{X}^{+}}x_{\check{\lambda}}K.

3.2.1. Representation of functions

Since 𝐗𝐗\mathbf{X}-\mathbf{X}^{\bullet} is Zariski closed in 𝐗\mathbf{X}, a function ϕCc(X,)\phi\in C_{c}^{\infty}(X,\mathbb{C}) is uniquely determined by its restriction to XX^{\bullet}, where it no longer has to be compactly supported. By the generalized Cartan decomposition, we have a canonical isomorphism C(X,)K[[ΛX+]]C^{\infty}(X^{\bullet},\mathbb{C})^{K}\xrightarrow{\,\sim\,}\mathbb{C}[[\Lambda_{X}^{+}]] defined by

(1) ϕλˇΛX+ϕ(xλˇ)eλˇ[[ΛX+]]\phi\mapsto\sum_{\check{\lambda}\in\Lambda_{X}^{+}}\phi(x_{\check{\lambda}})e^{\check{\lambda}}\in\mathbb{C}[[\Lambda_{X}^{+}]]

where eλˇe^{\check{\lambda}} is a formal symbol representing the monoid element λˇ\check{\lambda}. We will use such a formal power series to represent functions in Cc(X,)C_{c}^{\infty}(X,\mathbb{C}) in the future.

3.2.2. Interaction with bundle

Let 𝐗~𝐗\widetilde{\mathbf{X}}\to\mathbf{X} be a combinatorially trivial 𝔾m\mathbb{G}_{m}-bundle defined over Spec𝒪\operatorname{Spec}\mathcal{O}. Let

J0=𝒪×,J1={xF×|x1(modϖ)}.J_{0}=\mathcal{O}^{\times},\quad J_{1}=\{x\in F^{\times}\,|\,x\equiv 1\pmod{\varpi}\}.

The maximal compact subgroup K0:=𝐆~(𝒪)K^{0}:=\widetilde{\mathbf{G}}(\mathcal{O}) is equal to 𝐆(𝒪)×J0\mathbf{G(\mathcal{O})}\times J_{0}. For global reasons, we are also interested in the subgroup

K1:=𝐆(𝒪)×J1K0.K^{1}:=\mathbf{G}(\mathcal{O})\times J_{1}\subseteq K^{0}.

Clearly, K1K^{1} is a normal subgroup of K0K^{0} whose quotient is 𝔽×\mathbb{F}^{\times}, which has size q1q-1. The main result of this subsection is to describe the space X~/K1\widetilde{X}/K^{1}.

Definition 3.3.

A coroot λˇ\check{\lambda} is said to lie on a wall of type T if there is a spherical root γ\gamma of type T such that λˇ,γ=0\langle\check{\lambda},\gamma\rangle=0.

A point xXx\in X lies on a wall of type T if it is in the orbit xλˇK0x_{\check{\lambda}}K^{0}, where λˇ\check{\lambda} lies on a wall of type T.

Proposition 3.4.

Let xX~x\in{\widetilde{X}}. If the action of K1K^{1} on xK0xK^{0} has fewer than q1q-1 orbits, then xx lies on a wall of type T.

Proof.

In this proof, we will only work over the generic fibre, so the standard theory of spherical embeddings apply without comment. Further note that we are only working over the open 𝐆\mathbf{G}-orbit 𝐗\mathbf{X}^{\bullet}, so we will assume that 𝐗\mathbf{X} is homogeneous.

Over the open orbit 𝐗̊\mathring{\mathbf{X}}, fix a trivialization of 𝐗~\widetilde{\mathbf{X}}. Identify AXA_{X} with a subset of X̊\mathring{X} using the orbit map. Let xAX+x\in A_{X}^{+} be an element such that K1K^{1} splits xK0xK^{0} into less than q1q-1 pieces. Then there exists gKg\in K such that (x,1)g=(x,α)(x,1)g=(x,\alpha), where α1(modϖ)\alpha\not\equiv 1\pmod{\varpi}. For any n1n\geq 1, we have

(2) (xn,1)g=(xn1,1)(x,1)g=(xn,α),(x^{n},1)g=(x^{n-1},1)(x,1)g=(x^{n},\alpha),

where we are using the identification of AXA_{X} (as a group) with its orbit through the base point x0x_{0} (a subset of XX). It follows from this equation that the orbit xnK0x^{n}K^{0} also does not split completely.

Let λˇ\check{\lambda} be the image of xx in ΛX+\Lambda_{X}^{+}. Suppose for contradiction that λˇ\check{\lambda} does not lie on a wall of type T, then the set Θ:={γΔX|λˇ,γ=0}\Theta:=\{\gamma\in\Delta_{X}\,|\,\langle\check{\lambda},\gamma\rangle=0\} consists only of spherical roots of type G. The idea of the proof is to show that the bundle trivializes at Θ\Theta-infinity. This notion is introduced in [SV17, §2.3], and we recall some aspects of their construction.

Let Θ\mathcal{F}_{\Theta} be the face in 𝒱\mathcal{V} consisting of vectors orthogonal to all spherical roots in Θ\Theta, then it contains λˇ\check{\lambda} in its relative interior. Choose a fan decomposition of Θ\mathcal{F}_{\Theta} so that λˇ\check{\lambda} belongs to a face \mathcal{F}. Let 𝐗¯\overline{\mathbf{X}} be the spherical embedding defined by this fan decomposition, following Luna–Vust theory [Kno91, Theorem 3.3]. The face \mathcal{F} corresponds to a 𝐆\mathbf{G}-orbit 𝐙\mathbf{Z} satisfying the condition that xnx^{n} converges to a point x𝐙x^{\infty}\in\mathbf{Z}.

The combinatorially trivial 𝔾m\mathbb{G}_{m}-bundle 𝐗~\widetilde{\mathbf{X}} also extends to the boundary using the compactification given by the same fan decomposition. By [SV17, Proposition 2.3.8(3)], the spherical roots of 𝐙\mathbf{Z} are exactly Θ\Theta, so 𝐙\mathbf{Z} is a spherical variety with only spherical roots of type G. Proposition 2.5 implies that the bundle over 𝐙\mathbf{Z} must split. On the other hand, taking limit as nn\to\infty in equation (2), we see that gg translates between the q1q-1 disks above the point xx^{\infty}. This is a contradiction, proving that λˇ\check{\lambda} must lie on a wall of type T. ∎

3.3. Relative Satake isomorphism

In this subsection, we will introduce the relative Satake isomorphism. In the homogeneous case, the results are unconditional and due to Sakellaridis [Sak13, Sak18]. In a second case, which we call “totally type T”, the analogous result holds in the equal-characteristic case [SW22], and we conjecture that it carries over to the mixed characteristic case.

3.3.1. Hecke algebra and Satake isomorphism

Fix the Haar measure on GG so that KK has volume 1, then we can define the Hecke algebra (G):=Cc(K\G/K,)\mathcal{H}(G):=C_{c}^{\infty}(K\backslash G/K,\mathbb{C}) whose algebra structure is given by convolution

(f1f2)(x)=Gf1(xg1)f2(g)𝑑g(f_{1}\ast f_{2})(x)=\int_{G}f_{1}(xg^{-1})f_{2}(g)\,dg

This algebra acts on Cc(X,)C_{c}^{\infty}(X,\mathbb{C}) on the left in the natural way

(fϕ)(x)=Gϕ(xg)f(g)𝑑g.(f\cdot\phi)(x)=\int_{G}\phi(xg)f(g)\,dg.

Note that when X=GX=G, this action is different from the convolution.

Recall the classical Satake isomorphism, cf. [Gro98]

(3) (G)(A)W[Λ]W,ff^(t):=δ(t)12Nf(tn)𝑑n\mathcal{H}(G)\xrightarrow{\,\sim\,}\mathcal{H}(A)^{W}\simeq\mathbb{C}[\Lambda]^{W},\quad f\mapsto\hat{f}(t):=\delta(t)^{\frac{1}{2}}\int_{N}f(tn)dn

where Λ=X(𝐀)=𝐀(F)/𝐀(𝒪)\Lambda=X_{*}(\mathbf{A})=\mathbf{A}(F)/\mathbf{A}(\mathcal{O}). By restriction, we have a natural ring homomorphism [Λ]W[ΛX+]\mathbb{C}[\Lambda]^{W}\to\mathbb{C}[\Lambda_{X}^{+}]. Let AA^{*} be the set of unramified characters of AA, so A=X(𝐀)×A^{*}=X^{*}(\mathbf{A})\otimes_{\mathbb{Z}}\mathbb{C}^{\times} is an algebraic torus. There is a Fourier transform

(4) [Λ][A],eλˇ(χχ,λˇ×)\mathbb{C}[\Lambda]\xrightarrow{\,\sim\,}\mathbb{C}[A^{*}],\quad e^{\check{\lambda}}\mapsto(\chi\mapsto\langle\chi,\check{\lambda}\rangle\in\mathbb{C}^{\times})

where as before, eλˇe^{\check{\lambda}} is a formal symbol representing the monoid element λˇ\check{\lambda}. We will use the Fourier transform to identify the two spaces.

3.3.2. Homogeneous case

In this subsection, suppose in addition that 𝐗\mathbf{X} is a homogeneous spherical variety. We will also assume Statements 6.3.1 and 7.1.5 from [Sak13]. These are easy to verify for each given 𝐗\mathbf{X}, and it is expected they always hold under the assumptions we have made already.

Let AXA_{X}^{*} (resp. AA^{*}) be the set of unramified characters of AXA_{X} (resp. AA). Since we are assuming all 𝐁\mathbf{B} orbits on 𝐗/F¯\mathbf{X}_{/\bar{F}} are defined over FF, the tours 𝐀X\mathbf{A}_{X} is split, and we have the canonical injection AXAA_{X}^{*}\hookrightarrow A^{*}. We can alternatively describe AXA_{X}^{*} as the \mathbb{C}-points of the dual torus 𝐀ˇX\check{\mathbf{A}}_{X}.

Let δ(X)\delta_{(X)} (resp. δP(X)\delta_{P(X)}) be the modulus character for the good Levi subgroup 𝐋(𝐗)\mathbf{L}(\mathbf{X}) (resp. parabolic subgroup 𝐏(X)\mathbf{P}(X)), cf. §2.1. The translate δ(X)12AX\delta_{(X)}^{\frac{1}{2}}A_{X}^{*} defines a subvariety of AA^{*}. Its ring of polynomial functions will be denoted by [δ(X)12AX]\mathbb{C}[\delta_{(X)}^{\frac{1}{2}}A_{X}^{*}]. We make the following canonical identification

(5) [ΛX][AX][δ(X)12AX]\mathbb{C}[\Lambda_{X}]\xrightarrow{\,\sim\,}\mathbb{C}[A_{X}^{*}]\xrightarrow{\,\sim\,}\mathbb{C}[\delta_{(X)}^{\frac{1}{2}}A_{X}^{*}]

where the first map is the Fourier transform as defined in (4), and the second map is induced by translation.

Theorem 3.5 ([Sak13, Theorem 8.0.2]).

There exists an isomorphism

ϕϕ^:Cc(X,)K[δ(X)12AX]WX\phi\mapsto\hat{\phi}:C_{c}^{\infty}(X,\mathbb{C})^{K}\simeq\mathbb{C}[\delta_{(X)}^{\frac{1}{2}}A_{X}^{*}]^{W_{X}}

such that

  1. (1)

    The image of the basic function

    Φ0:=𝟏[𝐗(𝒪)]\Phi_{0}:=\mathbf{1}[\mathbf{X}(\mathcal{O})]

    is the constant function 1.

  2. (2)

    Let Cc(K\G/K,)[A]WC_{c}^{\infty}(K\backslash G/K,\mathbb{C})\simeq\mathbb{C}[A^{*}]^{W} be the classical Satake isomorphism, then the above isomorphism is equivariant with respect to the natural action on the left hand side and multiplication on the right hand side.

Let Cc(K\G/K,)\mathcal{L}\in C_{c}^{\infty}(K\backslash G/K,\mathbb{C}) be a Hecke operator, with Satake transform ^[A]W\hat{\mathcal{L}}\in\mathbb{C}[A^{*}]^{W}. By restriction and using identification (5), this defines an element

(6) ^(X)[δ(X)12AX][ΛX]\hat{\mathcal{L}}_{(X)}\in\mathbb{C}[\delta_{(X)}^{\frac{1}{2}}A_{X}^{*}]\xrightarrow{\,\sim\,}\mathbb{C}[\Lambda_{X}]

We would like to compute the function Φ0\mathcal{L}\cdot\Phi_{0} in terms of ^\hat{\mathcal{L}}, which can be done using the inverse Satake isomorphism [Sak18]. To state it, recall that Sakellaridis defined a multiset ΘX+\Theta_{X}^{+} in [Sak13, §7.1]. By our split hypothesis, all the signs σθˇ=1\sigma_{\check{\theta}}=1 in the notation of loc. cit., and the remaining two pieces of data can be combined as a representation of 𝐀ˇ×𝔾m\check{\mathbf{A}}\times\mathbb{G}_{m}, cf. [BZSV24, §9.3.5].

Theorem 3.6.

Let ΘX+\Theta_{X}^{+} be the multiset of weights of 𝐀ˇ×𝔾m\check{\mathbf{A}}\times\mathbb{G}_{m} described in [Sak13, §7.1], then

δP(X)12(Φ0):=^(X)γˇΦˇX+(1eγˇ)(θˇ,dθˇ)ΘX+(1q12dθˇeθˇ)|ΛX+\delta_{P(X)}^{\frac{1}{2}}\cdot(\mathcal{L}\cdot\Phi_{0}):=\hat{\mathcal{L}}_{(X)}\cdot\frac{\prod_{\check{\gamma}\in\check{\Phi}_{X}^{+}}(1-e^{\check{\gamma}})}{\prod_{(\check{\theta},d_{\check{\theta}})\in\Theta_{X}^{+}}(1-q^{-\frac{1}{2}d_{\check{\theta}}}e^{\check{\theta}})}\bigg{|}_{\Lambda_{X}^{+}}

under the identification (1).

Proof.

This is proven in [Sak18] under the additional assumption that 𝐗\mathbf{X} is wavefront. However, the only time the wavefront hypothesis is used in the derivation is to compute the normalization factor in the Plancherel measure. This exact normalization is not needed for the above result since the expressions do not depend on the choice of the GG-invariant measure on XX. ∎

3.3.3. Strongly tempered case

We no longer suppose XX is homogeneous. Instead, we work in the strongly tempered case, namely GˇX=Gˇ\check{G}_{X}=\check{G}. We state a conjectural analogue of Theorem 3.6. In the case when FF is an equal-characteristic local field, the conjecture minus the final part is known by the main theorem of [SW22].

In this case, we have 𝐀=𝐀X\mathbf{A}=\mathbf{A}_{X}, so we can drop the subscript XX from all combinatorial data. Moreover, δ(X)=1\delta_{(X)}=1 and δP(X)=δ\delta_{P(X)}=\delta. Finally, all elements of ΘX+\Theta_{X}^{+} are in degree 1 by the recipe constructing them. Therefore, the following is the exact analogue of Theorem 3.6, with the final statement playing the role of [Sak13, Statement 7.1.5].

Conjecture 3.7.

Let ΘX+\Theta_{X}^{+} be the multiset of weights of 𝐀ˇ\check{\mathbf{A}} denoted by 𝔅+\mathfrak{B}^{+} in [SW22, Theorem 1.1.2], then

δ12(Φ0)=^γˇΦˇ+(1eγˇ)θˇΘX+(1q12eθˇ)|Λ+\delta^{\frac{1}{2}}\cdot(\mathcal{L}\cdot\Phi_{0})=\hat{\mathcal{L}}\cdot\frac{\prod_{\check{\gamma}\in\check{\Phi}^{+}}(1-e^{\check{\gamma}})}{\prod_{\check{\theta}\in\Theta_{X}^{+}}(1-q^{-\frac{1}{2}}e^{\check{\theta}})}\bigg{|}_{\Lambda^{+}}

Moreover, all weights of ΘX+\Theta_{X}^{+} are minuscule as coweights of 𝐆\mathbf{G}.

Proof in the equal-characteristic case.

All references in this proof are to [SW22]. Define the Radon transform π!\pi_{!} by the expression

π!ϕ(t)=Nϕ(tn)𝑑n.\pi_{!}\phi(t)=\int_{N}\phi(tn)\,dn.

By unwinding definitions, it is easy to check that π!(ϕ)=^π!ϕ\pi_{!}(\mathcal{L}\cdot\phi)=\hat{\mathcal{L}}\ast\pi_{!}\phi, where the convolution takes place in the space of smooth (not necessarily compactly supported) functions on 𝐀\mathbf{A}. In terms of the ring [[Λ]]\mathbb{C}[[\Lambda]], this convolution is just usual multiplication. Since π!Φ0\pi_{!}\Phi_{0} is computed in Theorem 1.1.2, the conjecture follows by applying the theory of asymptotics, exactly as in the proof of Corollary 1.2.1. ∎

3.4. Integral structure

Let the assumptions be as in the previous subsection. We give an ad hoc definition of an integral structure on [δ(X)12AX]\mathbb{C}[\delta_{(X)}^{\frac{1}{2}}A_{X}^{*}], whose only purpose is to specify when a half power of qq can show up. This is abstractly explained by the analytic shearing construction in [BZSV24, §6.8]. Let (q)=[q1]\mathbb{Z}_{(q)}=\mathbb{Z}[q^{-1}].

Definition 3.8.

Let ρ\rho (resp. ρ(X)\rho_{(X)}, ρP(X)\rho_{P(X)}) be the half sum of positive roots for 𝐆\mathbf{G} (resp. 𝐋(𝐗)\mathbf{L}(\mathbf{X}), 𝐏(𝐗)\mathbf{P}(\mathbf{X})). Define (q)[[A]]\fatslash\mathbb{Z}_{(q)}[[A^{*}]]^{{\mathbin{\mkern-6.0mu\fatslash}}} to be the functions in [[A]]\mathbb{C}[[A^{*}]] of the form

λˇΛXaλˇqλˇ,ρeλˇ\sum_{\check{\lambda}\in\Lambda_{X}}a_{\check{\lambda}}q^{\langle\check{\lambda},\rho\rangle}e^{\check{\lambda}}

where aλˇ(q)a_{\check{\lambda}}\in\mathbb{Z}_{(q)}. We are again identifying [[A]]\mathbb{C}[[A^{*}]] with [[Λ]]\mathbb{C}[[\Lambda]] by (1).

The ring (q)[[δ(X)12AX]]\fatslash\mathbb{Z}_{(q)}[[\delta_{(X)}^{\frac{1}{2}}A_{X}^{*}]]^{{\mathbin{\mkern-6.0mu\fatslash}}} consists of restrictions of functions in [[A]]\fatslash\mathbb{Z}[[A^{*}]]^{{\mathbin{\mkern-6.0mu\fatslash}}}.

Proposition 3.9.

Let \mathcal{L} be a Hecke operator such that ^(X)(q)[δ(X)12AX]\fatslash\hat{\mathcal{L}}_{(X)}\in\mathbb{Z}_{(q)}[\delta_{(X)}^{\frac{1}{2}}A_{X}^{*}]^{{\mathbin{\mkern-6.0mu\fatslash}}}, in the notation of (6). Let ϕ=Φ0\phi=\mathcal{L}\cdot\Phi_{0}, then ϕ(x)[q1]\phi(x)\in\mathbb{Z}[q^{-1}] for all xXx\in X. Moreover, if xx lies on a wall of type T, then ϕ(x)(q1)[q1]\phi(x)\in(q-1)\mathbb{Z}[q^{-1}].

Proof.

Under the various hypotheses discussed in the previous section, the second term in Theorem 3.6 and Conjecture 3.7 lies in the ring (q)[[δ(X)12AX]]\fatslash\mathbb{Z}_{(q)}[[\delta^{\frac{1}{2}}_{(X)}A_{X}^{*}]]^{\mathbin{\mkern-6.0mu\fatslash}}. It follows that

ϕ(xλˇ)\displaystyle\phi(x_{\check{\lambda}}) δP(X)12(xλˇ)δ(X)12(xλˇ)qλˇ,ρ(q)\displaystyle\in\delta_{P(X)}^{-\frac{1}{2}}(x_{\check{\lambda}})\delta_{(X)}^{\frac{1}{2}}(x_{\check{\lambda}})q^{\langle\check{\lambda},\rho\rangle}\mathbb{Z}_{(q)}
=qλˇ,ρP(X)+ρ(X)+ρ(q)\displaystyle=q^{\langle\check{\lambda},-\rho_{P(X)}+\rho_{(X)}+\rho\rangle}\mathbb{Z}_{(q)}

Since ρ=ρP(X)+ρ(X)\rho=\rho_{P(X)}+\rho_{(X)} and 2ρP(X)2\rho_{P(X)} is integral, this is an integral power of qq.

From this computation, we see that ϕ(xλˇ)\phi(x_{\check{\lambda}}) is a polynomial in [q±1]\mathbb{Z}[q^{\pm 1}], where we now treat qq as a formal variable. To prove the required divisibility, we specialize it to q=1q=1, so ϕ(xλˇ)|q=1\phi(x_{\check{\lambda}})|_{q=1} is the coefficient of eλˇe^{\check{\lambda}} in the power series expansion of

^(X)|q=1𝚊𝚍𝙻,where 𝚊𝚍=γˇΦˇX+(1eγˇ),𝙻=θˇΘ+(1eθˇ).\hat{\mathcal{L}}_{(X)}|_{q=1}\cdot\frac{\mathtt{ad}}{\mathtt{L}},\quad\text{where }\mathtt{ad}=\prod_{\check{\gamma}\in\check{\Phi}_{X}^{+}}(1-e^{\check{\gamma}}),\quad\mathtt{L}=\prod_{\check{\theta}\in\Theta^{+}}(1-e^{\check{\theta}}).

Suppose γ\gamma is a spherical root of type T, with corresponding coroot γˇ\check{\gamma}. Let wγWXw_{\gamma}\in W_{X} be its associated simple reflection. We consider the behaviour of the above expression under wγw_{\gamma}. There are three terms.

  1. (1)

    The term ^(X)|q=1\hat{\mathcal{L}}_{(X)}|_{q=1} is wγw_{\gamma}-invariant by construction.

  2. (2)

    Since ΦX+\Phi_{X}^{+} is the positive part of a root system, we see that

    wγ𝚊𝚍𝚊𝚍=1eγˇ1eγˇ=eγˇ\frac{w_{\gamma}\mathtt{ad}}{\mathtt{ad}}=\frac{1-e^{-\check{\gamma}}}{1-e^{\check{\gamma}}}=-e^{-\check{\gamma}}
  3. (3)

    We need to understand the difference between wγΘ+w_{\gamma}\Theta^{+} and Θ+\Theta^{+}. In the homogeneous case, this is described in [Sak13, Statement 7.1.5]. Since γ\gamma has type T, there are two virtual colors DD and DD^{\prime} belonging to γ\gamma. Their associated valuations vDv_{D} and vDv_{D^{\prime}} sum to γˇ\check{\gamma}, so

    wγ𝙻𝙻\displaystyle\frac{w_{\gamma}\mathtt{L}}{\mathtt{L}} =1evD1evD1evD1evD\displaystyle=\frac{1-e^{-v_{D}}}{1-e^{v_{D}}}\frac{1-e^{-v_{D^{\prime}}}}{1-e^{v_{D^{\prime}}}}
    =e(vD+vD)=eγˇ\displaystyle=e^{-(v_{D}+v_{D^{\prime}})}=e^{-\check{\gamma}}

    In the strongly tempered case, the above discussion still holds due to our minuscule hypothesis combined with [SW22, Theorem 7.1.9(iii)].

It follows that

(7) wγ(ϕ^|q=1𝚊𝚍𝙻)=ϕ^|q=1𝚊𝚍𝙻w_{\gamma}\left(\hat{\phi}|_{q=1}\cdot\frac{\mathtt{ad}}{\mathtt{L}}\right)=-\hat{\phi}|_{q=1}\cdot\frac{\mathtt{ad}}{\mathtt{L}}

If wγλˇ=λˇw_{\gamma}\check{\lambda}=\check{\lambda}, or equivalently if λˇ\check{\lambda} lies on the wall defined by γ\gamma, then we see that the coefficient of eλˇe^{\check{\lambda}} must be 0, as required. ∎

Remark 3.10.

Performing the above computation for a spherical root of type G recovers equation (7), except with a plus sign. Therefore, we gain no information. In particular, all roots are of type G in the group case 𝐇\𝐇×𝐇\mathbf{H}\backslash\mathbf{H}\times\mathbf{H}, so we do not see any additional divisibility in the classical Satake isomorphism.

3.5. Abstract Euler system I

To state the abstract Euler system relation, we need to specify the Hecke operator. Our approach depends very little on this choice.

Definition 3.11.

A Hecke operator Cc(K\G/K,)\mathcal{L}\in C_{c}^{\infty}(K\backslash G/K,\mathbb{C}) is 𝐗\mathbf{X}-integral if ^(X)(q)[δ(X)12AX]\fatslash\hat{\mathcal{L}}_{(X)}\in\mathbb{Z}_{(q)}[\delta_{(X)}^{\frac{1}{2}}A_{X}^{*}]^{{\mathbin{\mkern-6.0mu\fatslash}}}.

Example 3.12.

Let VV be a minuscule representation of Gˇ\check{G}, then the function tdet(1q12t|V)t\mapsto\det(1-q^{-\frac{1}{2}}t|V) belongs to [A]W\mathbb{C}[A^{*}]^{W}. The Hecke operator associated to it by the Satake transform lies in (q)[A]\mathbb{Z}_{(q)}[A^{*}], so it is 𝐗\mathbf{X}-integral for all 𝐗\mathbf{X}. More generally, replacing tt by tXtX gives a polynomial V(X)\mathcal{H}_{V}(X), all of whose coefficients are 𝐗\mathbf{X}-integral. This is the Hecke polynomial considered in [BR94, §6].

Proposition 3.13.

Let \mathcal{L} be an 𝐗~\widetilde{\mathbf{X}}-integral Hecke operator, then there exists Φ1Cc(X~,[q1])K1\Phi_{1}\in C_{c}^{\infty}({\widetilde{X}},\mathbb{Z}[q^{-1}])^{K^{1}} such that

TrK0K1Φ1=𝟏[𝐗(𝒪)],\operatorname{Tr}_{K^{0}}^{K^{1}}\Phi_{1}=\mathcal{L}\cdot\mathbf{1}[\mathbf{X}(\mathcal{O})],

where TrK0K1:=gK0/K1g\operatorname{Tr}_{K^{0}}^{K^{1}}:=\sum_{g\in K^{0}/K_{1}}g is the trace operator.

Proof.

Let ϕCc(X~,)K1\phi\in C_{c}^{\infty}({\widetilde{X}},\mathbb{C})^{K^{1}}. Let xXx\in X. Write xK0=iIxiK1xK^{0}=\bigsqcup_{i\in I}x_{i}K^{1}, then

(TrK0K1ϕ)(x)=iI[K0:K1][xK0:xiK1]ϕ(xi).(\operatorname{Tr}_{K^{0}}^{K^{1}}\phi)(x)=\sum_{i\in I}\frac{[K^{0}:K^{1}]}{[xK^{0}:x_{i}K^{1}]}\phi(x_{i}).

Since K1K^{1} is normal in K0K^{0}, the multiplier in front of each term is the same. In other words, the image of Cc(X~,[q1])K1C_{c}^{\infty}({\widetilde{X}},\mathbb{Z}[q^{-1}])^{K^{1}} under the trace map is characterized by the divisibility condition

ϕ(x)q1[xK0:xK1][q1] for every x\phi(x)\in\frac{q-1}{[xK^{0}:xK^{1}]}\mathbb{Z}[q^{-1}]\text{ for every }x

By Proposition 3.4, if this coefficient is not 1, then xx lies on a wall of type T. Proposition 3.9 applied to 𝐗~\widetilde{\mathbf{X}} shows that this divisibility requirement is automatic. ∎

4. Global picture

Now let FF be a global field. The goal of this section is to apply the above theorems and conjectures to construct the tame part of Euler systems over FF.

4.1. Motivic theta series

Let 𝔸\mathbb{A} be the ring of adèles of FF. Let pp be a rational prime. Let SS be a finite set of places of FF such that

  1. (1)

    SS contains all places above pp and \infty.

  2. (2)

    Away from SS, the assumptions made in §3.1 hold [Sak12, Proposition 2.3.5].

In particular, we may extend 𝐗\mathbf{X} and 𝐆\mathbf{G} to smooth varieties over Spec𝒪F,S\operatorname{Spec}\mathcal{O}_{F,S}, which allows us to talk about their adèlic points. Again denote these extensions by the same letters.

Let KS=vS𝐆(𝒪Fv)K^{S}=\prod_{v\notin S}\mathbf{G}(\mathcal{O}_{F_{v}}). Define the basic vector to be

δ:=vS𝟏[𝐗(𝒪Fv)]Cc(𝐗(𝔸S),p)\delta:=\prod_{v\notin S}\mathbf{1}[\mathbf{X}(\mathcal{O}_{F_{v}})]\in C_{c}^{\infty}(\mathbf{X}(\mathbb{A}^{S}),\mathbb{Z}_{p})

By construction, it is KSK^{S}-invariant.

Definition 4.1.

Let \mathcal{M} be a p\mathbb{Z}_{p}-module with a smooth 𝐆(𝔸S)\mathbf{G}(\mathbb{A}^{S})-action. An 𝐗\mathbf{X}-theta series for \mathcal{M} is a 𝐆(𝔸S)\mathbf{G}(\mathbb{A}^{S})-equivariant map

Θ:Cc(𝐗(𝔸S),p)\Theta:C_{c}^{\infty}(\mathbf{X}(\mathbb{A}^{S}),\mathbb{Z}_{p})\to\mathcal{M}

The basic element zz\in\mathcal{M} is defined to be Θ(δ)\Theta(\delta).

Remark 4.2.

In general, the discussions of [Sak12, §3] suggests that the correct function space should be a restricted tensor product of smooth functions on 𝐗\mathbf{X}^{\bullet} with bounded growth near the boundary 𝐗𝐗\mathbf{X}-\mathbf{X}^{\bullet}. However, this subspace is sufficient for our applications.

If \mathcal{M} is a space of automorphic functions, then the classical theta series is an example of this definition. In the relative Langlands program, such objects are expected to play an important role, cf. [BZSV24, §15.1]. Their motivic analogues should also exist in great generality, though they are harder to construct. We will later specify \mathcal{M} to be certain p\mathbb{Z}_{p}-coefficient continuous étale cohomology groups, which we consider as pp-adic realizations of motives.

Example 4.3.

Consider the case 𝐆=GL2\mathbf{G}=\mathrm{GL}_{2}, F=F=\mathbb{Q}, and 𝐗=𝚜𝚝𝚍\mathbf{X}=\mathtt{std}, the 2-dimensional affine space with the standard action of 𝐆\mathbf{G}. The theory of Siegel units (cf. [Kat04, §1.4], [Col04, Théoréme 1.8]) gives a motivic theta series

Cc(𝐗(𝔸S),)limU𝒪(Y(U))×,C_{c}^{\infty}(\mathbf{X}(\mathbb{A}^{S}),\mathbb{Z})\to\varinjlim_{U}\mathcal{O}(Y(U))^{\times},

where UU runs over all open compact subgroups of 𝐆(𝔸)\mathbf{G}(\mathbb{A}^{\infty}), and Y(U)Y(U) is the open modular curve of level UU. By taking pp-adic étale realization, this produces a motivic theta series valued in Hcont1(Y,p(1))\mathrm{H}^{1}_{\mathrm{cont}}(Y,\mathbb{Z}_{p}(1)). This is exactly the construction of GL2\mathrm{GL}_{2}-Eisenstein classes in weight 2. For higher weights, the coefficient p\mathbb{Z}_{p} is replaced by an integral étale local system, cf. [LSZ22b, Proposition 7.2.4].

We are hiding one subtlety in this definition. In the above references, one need to restrict to functions which vanish at (0,0)(0,0) to ensure convergence. To get this vanishing automatically, we fix an auxiliary place vSv\in S and fix the test vector there to vanish at (0,0)(0,0). This was used in the choices of [LSZ22b, §8.4.4].

Example 4.4.

By taking cup product, we obtain the following motivic theta series for 𝐗=𝚜𝚝𝚍2\mathbf{X}=\mathtt{std}^{2} with the diagonal GL2\mathrm{GL}_{2}-action

Cc(𝐗(𝔸S),)Hcont2(Y,p(2)).C_{c}^{\infty}(\mathbf{X}(\mathbb{A}^{S}),\mathbb{Z})\to\mathrm{H}^{2}_{\mathrm{cont}}(Y,\mathbb{Z}_{p}(2)).

This is Kato’s construction of his zeta element, reformulated in [Col04]. As observed by Colmez, this makes verifying their norm relations “almost automatic”, since it reduces to some elementary manipulation of indicator functions on 𝐗\mathbf{X}.

4.2. Abstract Euler system II

Let 𝐓\mathbf{T} be a 1-dimensional torus defined over 𝒪F,S\mathcal{O}_{F,S}, the SS-integers of FF. Let 𝐗~𝐗\widetilde{\mathbf{X}}\to\mathbf{X} be a combinatorially trivial 𝐓\mathbf{T}-bundle. Let 𝐆~=𝐆×𝐓\widetilde{\mathbf{G}}=\mathbf{G}\times\mathbf{T} be the augmented group. For each square-free ideal 𝔪\mathfrak{m} not divisible by any place in SS, we may form the level structure

(8) J[𝔪]:=v𝔪vS𝐓(𝒪Fv)v|𝔪𝐓(𝒪Fv)1𝐓(𝔸S),K[𝔪]:=KS×J[𝔪]𝐆~(𝔸S),J[\mathfrak{m}]:=\prod_{\begin{subarray}{c}v\nmid\mathfrak{m}\\ v\notin S\end{subarray}}\mathbf{T}(\mathcal{O}_{F_{v}})\prod_{v|\mathfrak{m}}\mathbf{T}(\mathcal{O}_{F_{v}})^{1}\subseteq\mathbf{T}(\mathbb{A}^{S}),\quad K[\mathfrak{m}]:=K^{S}\times J[\mathfrak{m}]\subseteq\widetilde{\mathbf{G}}(\mathbb{A}^{S}),

where 𝐓(𝒪Fv)1\mathbf{T}(\mathcal{O}_{F_{v}})^{1} consists of the elements of 𝐓(𝒪Fv)\mathbf{T}(\mathcal{O}_{F_{v}}) whose reduction is the identity.

Let \mathscr{L} be the set of places of FF away from SS where the hypotheses of Section 2 are satisfied for 𝐗~\widetilde{\mathbf{X}}. This in particular requires 𝐓\mathbf{T} to be split at those places. For each \ell\in\mathscr{L}, let \mathcal{H}_{\ell} be an 𝐗~\widetilde{\mathbf{X}}-integral Hecke operator in the group 𝐆~(F)\widetilde{\mathbf{G}}(F_{\ell}) (Definition 3.11). They give rise to the following abstract Euler system.

Theorem 4.5.

Let Θ\Theta be an 𝐗~\widetilde{\mathbf{X}}-theta element for \mathcal{M}. Let \mathscr{R} be the collection of square-free products of places in \mathscr{L}. There exists a collection of elements {z𝔪K[𝔪]|𝔪}\{z_{\mathfrak{m}}\in\mathcal{M}^{K[\mathfrak{m}]}\,|\,\mathfrak{m}\in\mathscr{R}\} such that z1z_{1} is the basic element, and for any 𝔪\mathfrak{m}\in\mathscr{R} and any \ell\in\mathscr{L} such that 𝔪\ell\nmid\mathfrak{m}, we have the norm relation

TrK[𝔪]K[𝔪]z𝔪=z𝔪\operatorname{Tr}_{K[\mathfrak{m}]}^{K[\mathfrak{m}\ell]}z_{\mathfrak{m}\ell}=\mathcal{H}_{\ell}\cdot z_{\mathfrak{m}}
Proof.

For each \ell\in\mathscr{L}, let δ()\delta^{(\ell)} be the element Φ1\Phi_{1} from Proposition 3.13. Define δ[𝔪]\delta[\mathfrak{m}] to be δ\delta, except the local component at each |𝔪\ell|\mathfrak{m} are replaced by δ()\delta^{(\ell)}. The theorem follows from the 𝐆~(𝔸)\widetilde{\mathbf{G}}(\mathbb{A}^{\infty})-equivariance of Θ\Theta. ∎

4.3. Étale classes and Galois action

We now specialize further and suppose that 𝐆\mathbf{G} has Shimura varieties defined over a number field EE containing FF. We can then form Shimura varieties for 𝐆~\widetilde{\mathbf{G}}. Let 𝕃\mathbb{L} be a p\mathbb{Z}_{p}-local system defined using an algebraic representation of 𝐆\mathbf{G}, as in for example [HT01, §III.2]. Define

(9) =limKSHcontd(Sh𝐆~(KSKS),𝕃),\mathcal{M}=\varinjlim_{K^{S}}\mathrm{H}^{d}_{\mathrm{cont}}(\mathrm{Sh}_{\widetilde{\mathbf{G}}}(K_{S}K^{S}),\mathbb{L}),

where KSK_{S} is a fixed open compact subgroup of 𝐆~(𝔸S)\widetilde{\mathbf{G}}(\mathbb{A}_{S}), and KSK^{S} runs over all open compact subgroups of 𝐆~(𝔸S)\widetilde{\mathbf{G}}(\mathbb{A}^{S}). The cohomology theory used is Jannsen’s continuous étale cohomology [Jan88]. Note that all the transition maps are well-defined since we are not changing the level structure at the pp-adic places.

Suppose K𝐆(𝔸)K\subseteq\mathbf{G}(\mathbb{A}^{\infty}) and U𝐓(𝔸)U\subseteq\mathbf{T}(\mathbb{A}^{\infty}) are both open compact, then we have

(10) Sh𝐆~(K×U)Sh𝐆(K)×Sh𝐓(U)\mathrm{Sh}_{\widetilde{\mathbf{G}}}(K\times U)\simeq\mathrm{Sh}_{\mathbf{G}}(K)\times\mathrm{Sh}_{\mathbf{T}}(U)

as algebraic varieties over EE. Take U=J[𝔪]×JSU=J[\mathfrak{m}]\times J_{S} for some appropriate choice of the ramified level structure JSJ_{S}, then it is often possible to interpret the above product as a base change from EE to an extension E[𝔪]E[\mathfrak{m}]. We single out two cases.

  • If 𝐓=𝔾m\mathbf{T}=\mathbb{G}_{m} with Shimura cocharacter z1z^{-1}, then E[𝔪]E[\mathfrak{m}] is the cyclotomic extension of conductor 𝔪\mathfrak{m}.

  • If 𝐓=U(1)\mathbf{T}=\mathrm{U}(1) attached to a quadratic extension E/FE/F, and the Shimura cocharacter is zz¯/zz\mapsto\bar{z}/z, then E[𝔪]E[\mathfrak{m}] is the anticyclotomic extension of EE of conductor 𝔪\mathfrak{m}.

In these cases, the trace from Theorem 4.5 can be identified with a trace map between fields.

Locally at \ell\in\mathscr{L}, we have the following description of the Hecke algebra for 𝐆~\widetilde{\mathbf{G}} as a ring of Laurent polynomials

(𝐆~(F),)(𝐆(F),)[𝚃±1].\mathcal{H}(\widetilde{\mathbf{G}}(F_{\ell}),\mathbb{C})\simeq\mathcal{H}(\mathbf{G}(F_{\ell}),\mathbb{C})[\mathtt{T}^{\pm 1}].

Here, 𝚃\mathtt{T} is the indicator function of ϖ𝒪F×𝐓(F)\varpi\mathcal{O}_{F_{\ell}}^{\times}\subseteq\mathbf{T}(F_{\ell}). Moreover, a Hecke operator is 𝐗~\widetilde{\mathbf{X}}-integral if and only if all of its coefficients on the right hand side are 𝐗\mathbf{X}-integral. In particular, if VV is a minuscule representation of Gˇ\check{G}, then we may use the Hecke polynomial V(𝚃)\mathcal{H}_{V}(\mathtt{T}) defined in Example 3.12.

Under the identification (10), the action of 𝚃\mathtt{T} matches with the action of the geometric Frobenius element Frob1\mathrm{Frob}_{\ell}^{-1}. In the 𝐓=U(1)\mathbf{T}=\mathrm{U}(1) case, this requires further comments: the identification of 𝐓/F\mathbf{T}_{/F_{\ell}} with 𝔾m\mathbb{G}_{m} depends on the choice of a place λ\lambda of EE above \ell, and Frob\mathrm{Frob}_{\ell} really means the arithmetic Frobenius at λ\lambda. Using this, Theorem 4.5 leads to the following Corollary.

Corollary 4.6.

Let the setting be as in Theorem 4.5, with \mathcal{M} given by equation (9). Let VV be a minuscule representation of Gˇ\check{G}. Then there exists a collection of elements

{z𝔪Hcontd(Sh𝐆/E[𝔪],𝕃)|𝔪}\big{\{}z_{\mathfrak{m}}\in\mathrm{H}^{d}_{\mathrm{cont}}(\mathrm{Sh}_{\mathbf{G}/E[\mathfrak{m}]},\mathbb{L})\,\big{|}\,\mathfrak{m}\in\mathscr{R}\big{\}}

satisfying the norm relations

TrE[𝔪]E[𝔪]z𝔪=V(Frob1)z𝔪\operatorname{Tr}_{E[\mathfrak{m}]}^{E[\mathfrak{m}\ell]}z_{\mathfrak{m}\ell}=\mathcal{H}_{V}(\mathrm{Frob}_{\ell}^{-1})\cdot z_{\mathfrak{m}}

whenever 𝔪,𝔪\mathfrak{m},\mathfrak{m}\ell\in\mathscr{R}.

Remark 4.7.

It is crucial that the Shimura cocharacter on 𝐓\mathbf{T} is non-trivial, otherwise we do not recover a Galois action. In particular, choosing the trivial equivariant bundle 𝐗~=𝐗×𝐓\widetilde{\mathbf{X}}=\mathbf{X}\times\mathbf{T} does not lead to interesting arithmetic in any of the examples we consider, as expected.

This result of Corollary 4.6 may be called a “motivic Euler system”, cf. [LSZ22a, §9.4]. We now explain the steps required to convert this to an Euler system in Galois cohomology, proving Corollary 1.2 from the introduction. These steps are all standard in the literature.

  1. (1)

    Null-homologous modification: the main result of [MS19] constructs an element 𝚝±\mathtt{t}^{\pm} in the Hecke algebra of 𝐆\mathbf{G} whose action on the cohomology of Sh𝐆\mathrm{Sh}_{\mathbf{G}} are the two sign projectors. In particular, 𝚝:=𝚝(1)d+1\mathtt{t}:=\mathtt{t}^{(-1)^{d+1}} annihilates the cohomology group Hd(Sh𝐆/E¯,𝕃)\mathrm{H}^{d}(\mathrm{Sh}_{\mathbf{G}/\bar{E}},\mathbb{L}). Therefore, for each 𝔪\mathfrak{m} as above, 𝚝z𝔪\mathtt{t}\cdot z_{\mathfrak{m}} is cohomologically trivial, and the spectral sequence in continuous étale cohomology [Jan88, Remark 3.5(b)] gives an element

    c~𝔪H1(E[𝔪],Hd1(Sh𝐆/E¯,𝕃))\tilde{c}_{\mathfrak{m}}\in\mathrm{H}^{1}(E[\mathfrak{m}],\mathrm{H}^{d-1}(\mathrm{Sh}_{\mathbf{G}/\bar{E}},\mathbb{L}))

    A simpler argument is given in [LL21, Proposition 6.9(1)] which also suffices for our purpose.

  2. (2)

    Projection to Galois representations: Kottwitz’s conjecture gives a concrete description of the cohomology group Hd1(Sh𝐆/E¯,𝕃pp)\mathrm{H}^{d-1}(\mathrm{Sh}_{\mathbf{G}/\bar{E}},\mathbb{L}\otimes_{\mathbb{Z}_{p}}\mathbb{Q}_{p}). In the case considered in Corollary 1.2, it gives a projection from this group to the Galois representation ρπ\rho_{\pi}. Moreover, the Hecke polynomial specialized to the Satake parameters of π\pi agrees with the characteristic polynomial of ρπ\rho_{\pi}.

    The integral structure given by the lattice 𝕃\mathbb{L} is mapped to a lattice TπT_{\pi} in ρπ\rho_{\pi}. The image of c~𝔪\tilde{c}_{\mathfrak{m}} under this projection is the class c𝔪H1(E[𝔪],Tπ)c_{\mathfrak{m}}\in\mathrm{H}^{1}(E[\mathfrak{m}],T_{\pi}).

In the following two subsections, we will explain how pushforwards of cycles and Eisenstein classes both give instances of such a motivic theta series. In these settings, the above corollary produces many of the known motivic Euler systems in the literature.

4.4. Special cycles

Suppose that 𝐗\mathbf{X} is homogeneous, then 𝐗=𝐇\𝐆\mathbf{X}=\mathbf{H}\backslash\mathbf{G}, where we recall that 𝐇\mathbf{H} is the stabilizer of a point in the open orbit. In this case, 𝐗\mathbf{X} is automatically smooth, and it is affine if and only if 𝐇\mathbf{H} is reductive. Assume the following conditions.

  1. (1)

    Both 𝐆\mathbf{G} and 𝐇\mathbf{H} have Shimura varieties, and dimSh𝐆=2dimSh𝐇+1\dim\mathrm{Sh}_{\mathbf{G}}=2\dim\mathrm{Sh}_{\mathbf{H}}+1.

  2. (2)

    ξ\xi is an algebraic representation of 𝐆\mathbf{G}_{\infty} whose restriction to 𝐇\mathbf{H}_{\infty} contains the trivial representation.

Under the above assumptions, the discussions of [LS24, §2.4] carries over verbatim, and Proposition 2.7 of op. cit. gives a motivic theta series.

Proposition 4.8.

Let d=dimSh𝐇d=\dim\mathrm{Sh}_{\mathbf{H}}. Let 𝕃\mathbb{L} be the p\mathbb{Z}_{p}-local system on Sh𝐆\mathrm{Sh}_{\mathbf{G}} attached to ξ\xi. The diagonal cycle construction defines an 𝐗\mathbf{X}-theta series

Θ:Cc(𝐗(𝔸p),p)Hcont2d(Sh𝐆,𝕃(d))\Theta:C_{c}^{\infty}(\mathbf{X}(\mathbb{A}^{p\infty}),\mathbb{Z}_{p})\to\mathrm{H}_{\mathrm{cont}}^{2d}(\mathrm{Sh}_{\mathbf{G}},\mathbb{L}(d))
Remark 4.9.

The algebraic representation ξ\xi determines the archimedean part of the representations distinguished by Θ\Theta. The above construction depends on the choice of an invariant vector in ξ|𝐇\xi|_{\mathbf{H}}. It is likely that including the archimedean place in the space 𝐗(𝔸p)\mathbf{X}(\mathbb{A}^{p\infty}) would give a more canonical construction.

Since we are constructing motivic classes using cycles, we should limit our deformations to the self-dual setting. Therefore, it seems necessary to take 𝐓=U(1)\mathbf{T}=\mathrm{U}(1), defined with respect to a CM extension E/FE/F. To satisfy the conditions in §2, we must restrict our attention to split primes. The result is a split anticyclotomic Euler system, in the sense of [JNS24].

We now explain the content of Corollary 4.6 in a few cases. In each case, we need to specify the data of an embedding of reductive groups ι:𝐇𝐆\iota:\mathbf{H}\hookrightarrow\mathbf{G} and a character ν:𝐇U(1)\nu:\mathbf{H}\to\mathrm{U}(1). The latter is used to specify the U(1)\mathrm{U}(1)-bundle over 𝐗\mathbf{X}. In all cases, the local spherical variety appears in the table [Sak13, Appendix A], so they have been verified to satisfy the combinatorial conditions alluded to in §3.3.2.

4.4.1. Gan–Gross–Prasad [LS24]

Let 𝚅n𝚅n+1\mathtt{V}_{n}\subseteq\mathtt{V}_{n+1} be Hermitian spaces of dimensions nn, n+1n+1 respectively which are nearly definite. Consider the setting

𝐇=U(𝚅n),𝐆=U(𝚅n)×U(𝚅n+1)\displaystyle\mathbf{H}=\mathrm{U}(\mathtt{V}_{n}),\quad\mathbf{G}=\mathrm{U}(\mathtt{V}_{n})\times\mathrm{U}(\mathtt{V}_{n+1})
ι:h(h,diag(h,1)),ν:hdeth\displaystyle\iota:h\mapsto(h,\operatorname{diag}(h,1)),\quad\nu:h\mapsto\det h

The intersection of 𝐇\mathbf{H} with a good Borel 𝐁\mathbf{B} is trivial, so the bundle is automatically combinatorially trivial. Take VV to be the standard tensor product representation of degree n(n+1)n(n+1), then Corollary 4.6 recovers [LS24, Proposition 5.4], bypassing the explicit matrix calculations in §4 of op. cit.

4.4.2. Friedberg–Jacquet [GS23]

Let 𝚅2n\mathtt{V}_{2n} be a Hermitian space with signature (1,2n1)(1,2n-1) at one archimedean place and (0,2n)(0,2n) at the other archimedean places. Let 𝚆\mathtt{W} be a totally definite subspace of dimension nn, and let 𝚆\mathtt{W}^{\perp} be its dual. Consider the setting

𝐇=U(𝚆)×U(𝚆),𝐆=U(𝚅)\displaystyle\mathbf{H}=\mathrm{U}(\mathtt{W}^{\perp})\times\mathrm{U}(\mathtt{W}),\quad\mathbf{G}=\mathrm{U}(\mathtt{V})
ι:(a,b)diag(a,b),ν:(a,b)detadetb\displaystyle\iota:(a,b)\mapsto\operatorname{diag}(a,b),\quad\nu:(a,b)\mapsto\frac{\det a}{\det b}

At a split place, the local picture is GLn×GLn\GL2n\mathrm{GL}_{n}\times\mathrm{GL}_{n}\backslash\mathrm{GL}_{2n}. For an appropriate choice of the Borel subgroup 𝐁\mathbf{B}, the intersection 𝐀𝐇\mathbf{A}\cap\mathbf{H} has the form

(diag(z1,,zn),diag(zn,,z1)),(\operatorname{diag}(z_{1},\cdots,z_{n}),\operatorname{diag}(z_{n},\cdots,z_{1})),

so the bundle defined by ν\nu is combinatorially trivial (which explains why (a,b)detadetb(a,b)\mapsto\det a\det b is not the correct choice for ν\nu). In this case, the equivariant bundle is non-trivial even as a line bundle, which explains why it is essential to include the additional U(1)\mathrm{U}(1)-factor.

Take VV to be the standard representation of degree 2n2n, then Corollary 4.6 recovers the tame part of the split anticyclotomic Euler system for symplectic representations constructed by Graham and Shah [GS23, Proposition 9.27] without assumptions on the prime pp.

4.4.3. Cornut’s Euler system [Cor18]

Let FF be a totally real field. Let 𝚅\mathtt{V} be a quadratic space over FF of signature (2,2n1)(2,2n-1) at one archimedean place and istotropic at the other archimedean places. Suppose 𝚅\mathtt{V} splits over EE, then there is an EE-Hermitian FF-hyperplane 𝚆𝚅\mathtt{W}\subseteq\mathtt{V}, unique up to 𝐆\mathbf{G}-translation. Let

𝐇=U(𝚆),𝐆=SO(𝚅)\mathbf{H}=\mathrm{U}(\mathtt{W}),\quad\mathbf{G}=\mathrm{SO}(\mathtt{V})

The definition gives an inclusion ι:𝐇𝐆\iota:\mathbf{H}\hookrightarrow\mathbf{G}. The character ν\nu is again the determinant.

At a split place \ell, the local picture is GLn\SO2n+1\mathrm{GL}_{n}\backslash\mathrm{SO}_{2n+1}. The combinatorially trivial condition is automatic. Take VV to be the standard representation of degree 2n2n. Corollary 4.6 then recovers a split anticyclotomic Euler system for representations of SO(𝚅)\mathrm{SO}(\mathtt{V}), where the “Euler factor” is the one corresponding to the standard LL-function. This is the split analogue of the one constructed by Cornut [Cor18].

Remark 4.10.

In this case, 𝐀=𝐀𝐗\mathbf{A}=\mathbf{A}_{\mathbf{X}}, but the little Weyl group WXW_{X} is smaller, so we do not have multiplicity one [Sak08, Theorem 1.2.1]. This observation also shows that 𝐗\mathbf{X} is not wavefront, which suggests that a representation π\pi which is distinguished by 𝐗\mathbf{X} is necessarily not stable. If this is the case, then Kottwitz’s conjecture predicts that the Galois representation appearing in the orthogonal Shimura variety is not the standard one. We hope further automorphic studies of the setting will shed light on this question.

4.5. Eisenstein classes

We now move to a family of non-homogeneous spherical varieties. Suppose VV is an affine space with a spherical action of a group 𝐇\mathbf{H}. Let

𝐗=𝐆×𝐇V\mathbf{X}=\mathbf{G}\times^{\mathbf{H}}V

In a change of notation from §2, the stabilizer of a generic point is no longer 𝐇\mathbf{H}. In the Eisenstein class applications below, the stabilizer is a mirabolic subgroup of 𝐇\mathbf{H}, which gives an explanation for their prominence in [Loe21].

We expect there to be a general pushforward construction sending motivic theta series for VV to ones for 𝐗\mathbf{X}. If both 𝐇\mathbf{H} and 𝐆\mathbf{G} have Shimura varieties, then this should just be the usual pushforward construction, as described for example in [Loe21]. The following proposition describes that construction in our framework.

Proposition 4.11.

Suppose we have a VV-theta series

Θ𝐇:Cc(V(𝔸p),p)Hconti(Sh𝐇,p(j)),\Theta_{\mathbf{H}}:C_{c}^{\infty}(V(\mathbb{A}^{p\infty}),\mathbb{Z}_{p})\to\mathrm{H}^{i}_{\mathrm{cont}}(\mathrm{Sh}_{\mathbf{H}},\mathbb{Z}_{p}(j)),

then its pushforward defines an 𝐗\mathbf{X}-theta series

Θ𝐆:Cc(𝐗(𝔸p),p)Hconti+2d(Sh𝐆,p(j+d))/tors,\Theta_{\mathbf{G}}:C_{c}^{\infty}(\mathbf{X}(\mathbb{A}^{p\infty}),\mathbb{Z}_{p})\to\mathrm{H}_{\mathrm{cont}}^{i+2d}(\mathrm{Sh}_{\mathbf{G}},\mathbb{Z}_{p}(j+d))/\mathrm{tors},

where d=dimSh𝐆dimSh𝐇d=\dim\mathrm{Sh}_{\mathbf{G}}-\dim\mathrm{Sh}_{\mathbf{H}} is the codimension.

Proof.

Exactly as in [LSZ22b, §8.2], one may define a map

Θ𝐆:Cc((𝐆×V)(𝔸p),p)Hconti+2d(Sh𝐆,p(j+d))\Theta_{\mathbf{G}}^{\prime}:C_{c}^{\infty}((\mathbf{G}\times V)(\mathbb{A}^{p\infty}),\mathbb{Z}_{p})\to\mathrm{H}_{\mathrm{cont}}^{i+2d}(\mathrm{Sh}_{\mathbf{G}},\mathbb{Q}_{p}(j+d))

which is left 𝐆\mathbf{G}-equivariant and right 𝐇\mathbf{H}-invariant. Therefore, we have a commutative diagram

Cc((𝐆×V)(𝔸p),p){C_{c}^{\infty}((\mathbf{G}\times V)(\mathbb{A}^{p\infty}),\mathbb{Z}_{p})}Cc(𝐗(𝔸p),p){C_{c}^{\infty}(\mathbf{X}(\mathbb{A}^{p\infty}),\mathbb{Q}_{p})}Hconti+2d(Sh𝐆,p(j+d)){\mathrm{H}_{\mathrm{cont}}^{i+2d}(\mathrm{Sh}_{\mathbf{G}},\mathbb{Q}_{p}(j+d))}Θ𝐆\scriptstyle{\Theta_{\mathbf{G}}}Θ𝐆\scriptstyle{\Theta_{\mathbf{G}}^{\prime}}

where the downward arrow is the 𝐇\mathbf{H}-coinvariant map. One simply observes that the coinvariant map and Θ𝐆\Theta_{\mathbf{G}}^{\prime} both introduce the same volume factors on a basis of functions, so Θ𝐆\Theta_{\mathbf{G}} is integral. ∎

Remark 4.12.

The proof of [LS24, Proposition 2.7] should carry over to this setting, thereby removing the “mod torsion” part of the statement. One can likely axiomatize it using Loeffler’s definition of Cartesian cohomology functors [Loe21, GS23].

If 𝐇=GL2\mathbf{H}=\mathrm{GL}_{2} and VV is its standard two-dimensional representation, then Example 4.3 gives a VV-theta series as above, with i=j=1i=j=1. This gives rise to the tame parts of the following Euler systems, with the caveat that we need the function-level statements of [SW22] in mixed characteristics.

4.5.1. Rankin–Selberg [LLZ14]

Let 𝐆=GL2×GL2\mathbf{G}=\mathrm{GL}_{2}\times\mathrm{GL}_{2} and 𝐇=GL2\mathbf{H}=\mathrm{GL}_{2}, embedded diagonally in 𝐆\mathbf{G}. Let

𝐗=𝐆×𝐇𝚜𝚝𝚍\mathbf{X}=\mathbf{G}\times^{\mathbf{H}}\mathtt{std}

The pushforward construction gives rise to Beilinson–Flach elements considered in [LLZ14]. To obtain an Euler system, we endow the trivial line bundle over 𝐗\mathbf{X} with the 𝐆\mathbf{G}-action through the character (g1,g2)detg1(g_{1},g_{2})\mapsto\det g_{1}. Our main theorem applied to the associated 𝔾m\mathbb{G}_{m}-bundle recovers the tame norm relations of Beilinson–Flach elements [LLZ14, Theorem 3.4.1].

4.5.2. Asai representation [Gro20]

Let FF be real quadratic field. In the previous subsection, take instead

𝐆=ResF/GL2/F,𝐇=GL2\mathbf{G}=\operatorname{Res}_{F/\mathbb{Q}}\mathrm{GL}_{2/F},\quad\mathbf{H}=\mathrm{GL}_{2}

This gives rise to the Asai–Flach classes considered in [Gro20]. The 𝔾m\mathbb{G}_{m}-bundle defined above recovers the tame norm relation at primes \ell which split in FF.

4.5.3. GSp4×GL2\mathrm{GSp}_{4}\times\mathrm{GL}_{2} [HJR20]

Let 𝐆=GSp4×𝔾mGL2\mathbf{G}=\mathrm{GSp}_{4}\times_{\mathbb{G}_{m}}\mathrm{GL}_{2} and 𝐇=GL2×𝔾mGL2\mathbf{H}=\mathrm{GL}_{2}\times_{\mathbb{G}_{m}}\mathrm{GL}_{2}. By 𝚜𝚝𝚍\mathtt{std} we now mean the 2-dimensional affine space where 𝐇\mathbf{H} acts through its first component. The variety 𝐗=𝐆×𝐇𝚜𝚝𝚍\mathbf{X}=\mathbf{G}\times^{\mathbf{H}}\mathtt{std} is spherical. The trivial line bundle on 𝐗\mathbf{X} with the obvious character 𝐆𝔾m\mathbf{G}\to\mathbb{G}_{m} gives rise to the tame norm relation [HJR20, Proposition 8.17].

4.5.4. GU(2,1)\mathrm{GU}(2,1) [LSZ22a]

Let EE be an imaginary quadratic field. Consider the pair of algebraic groups defined over \mathbb{Q}:

(𝐆,𝐇)=(GU(2,1),GL2×ResE/𝔾m).(\mathbf{G},\mathbf{H})=(\mathrm{GU}(2,1),\mathrm{GL}_{2}\times\operatorname{Res}_{E/\mathbb{Q}}\mathbb{G}_{m}).

The variety 𝐗=𝐆×𝐇𝚜𝚝𝚍\mathbf{X}=\mathbf{G}\times^{\mathbf{H}}\mathtt{std} is spherical by [LSZ22a, Lemma 2.5.1]. The authors also defined a character μ:𝐆ResE/𝔾m\mu:\mathbf{G}\to\operatorname{Res}_{E/\mathbb{Q}}\mathbb{G}_{m} in §2.2 of op. cit. and use it to vary the class over field extensions. In our set-up, this corresponds to the trivial ResE/𝔾m\operatorname{Res}_{E/\mathbb{Q}}\mathbb{G}_{m}-bundle over 𝐗\mathbf{X} with the action of 𝐆\mathbf{G} given by μ\mu.

An interesting additional feature is that we have a 2-variable p\mathbb{Z}_{p}-deformation, since this is a 2-dimensional torus bundle. Let =ww¯\ell=w\bar{w} be a split prime, then Theorem A of op. cit. includes the hypothesis that at most one of ww and w¯\bar{w} divides the ideal 𝔪\mathfrak{m}. Our result also implies the tame norm relation at \ell under this hypothesis: allowing ww¯|𝔪w\bar{w}|\mathfrak{m} would require a divisibility by (1)2(\ell-1)^{2}, which appears to be false in this case. On the other hand, at an inert place, the divisibility requirement is by 21=(1)(+1)\ell^{2}-1=(\ell-1)(\ell+1), and this can be detected by specializing at =1\ell=1 and =1\ell=-1 separately. We hope to examine the inert case in a future paper.

References

  • [BR94] Don Blasius and Jonathan D. Rogawski. Zeta functions of Shimura varieties. In Motives (Seattle, WA, 1991), volume 55 of Proc. Sympos. Pure Math., pages 525–571. Amer. Math. Soc., Providence, RI, 1994.
  • [BZSV24] David Ben-Zvi, Yiannis Sakellaridis, and Akshay Venkatesh. Relative langlands duality. Preprint, available at arXiv:2409.04677, 2024.
  • [CHL11] Raf Cluckers, Thomas Hales, and François Loeser. Transfer principle for the fundamental lemma. In On the stabilization of the trace formula, volume 1 of Stab. Trace Formula Shimura Var. Arith. Appl., pages 309–347. Int. Press, Somerville, MA, 2011.
  • [Col04] Pierre Colmez. La conjecture de Birch et Swinnerton-Dyer 𝐩\mathbf{p}-adique. In Séminaire Bourbaki : volume 2002/2003, exposés 909-923, number 294 in Astérisque, pages 251–319. Association des amis de Nicolas Bourbaki, Société mathématique de France, Paris, 2004. talk:919.
  • [Cor18] Christophe Cornut. An Euler system of Heegner type. Preprint, 2018.
  • [CZ23] Murillo Corato-Zanarella. Spherical functions on symmetric spaces of Friedberg–Jacquet type. Preprint, available at arXiv:2311.00148, 2023.
  • [Dis24] Daniel Disegni. Euler systems for conjugate-symplectic motives. Preprint, available at arXiv:2410.08419, 2024.
  • [GN10] Dennis Gaitsgory and David Nadler. Spherical varieties and Langlands duality. Mosc. Math. J., 10(1):65–137, 271, 2010.
  • [Gro98] Benedict H. Gross. On the Satake isomorphism. In Galois representations in arithmetic algebraic geometry (Durham, 1996), volume 254 of London Math. Soc. Lecture Note Ser., pages 223–237. Cambridge Univ. Press, Cambridge, 1998.
  • [Gro20] Giada Grossi. On norm relations for Asai-Flach classes. Int. J. Number Theory, 16(10):2311–2377, 2020.
  • [GS23] Andrew Graham and Syed Waqar Ali Shah. Anticyclotomic euler systems for unitary groups. Proceedings of the London Mathematical Society, 127(6):1577–1680, 2023.
  • [HJR20] Chi-Yun Hsu, Zhaorong Jin, and Sakamoto Ryotaro. Euler systems for GSp4×GL2\mathrm{GSp}_{4}\times\mathrm{GL}_{2}. Preprint, available at arXiv:2011.12894, 2020.
  • [HT01] Michael Harris and Richard Taylor. The geometry and cohomology of some simple Shimura varieties, volume 151 of Annals of Mathematics Studies. Princeton University Press, Princeton, NJ, 2001. With an appendix by Vladimir G. Berkovich.
  • [Jan88] Uwe Jannsen. Continuous étale cohomology. Math. Ann., 280(2):207–245, 1988.
  • [JNS24] Dimitar Jetchev, Jan Nekovář, and Christopher Skinner. Split Euler systems for conjugate-dual Galois representations. Preprint, 2024.
  • [Kat04] Kazuya Kato. pp-adic Hodge theory and values of zeta functions of modular forms. Astérisque, (295):ix, 117–290, 2004. Cohomologies pp-adiques et applications arithmétiques. III.
  • [Kno91] Friedrich Knop. The Luna-Vust theory of spherical embeddings. In Proceedings of the Hyderabad Conference on Algebraic Groups (Hyderabad, 1989), pages 225–249. Manoj Prakashan, Madras, 1991.
  • [KS17] F. Knop and B. Schalke. The dual group of a spherical variety. Trans. Moscow Math. Soc., 78:187–216, 2017.
  • [KSZ21] Mark Kisin, Sug Woo Shin, and Yihang Zhu. The stable trace formula for Shimura varieties of abelian type. Preprint, available at arXiv:2110.05381, 2021.
  • [Liu11] Yifeng Liu. Arithmetic theta lifting and LL-derivatives for unitary groups, I. Algebra Number Theory, 5(7):849–921, 2011.
  • [LL21] Chao Li and Yifeng Liu. Chow groups and LL-derivatives of automorphic motives for unitary groups. Ann. of Math. (2), 194(3):817–901, 2021.
  • [LLZ14] Antonio Lei, David Loeffler, and Sarah Livia Zerbes. Euler systems for Rankin-Selberg convolutions of modular forms. Ann. of Math. (2), 180(2):653–771, 2014.
  • [Loe21] David Loeffler. Spherical varieties and norm relations in Iwasawa theory. Journal de Théorie des Nombres de Bordeaux, 33(3.2):1021–1043, 2021.
  • [LR24] David Loeffler and Oscar Rivero. Eisenstein degeneration of Euler systems. Journal für die reine und angewandte Mathematik (Crelles Journal), 2024(814):241–282, 2024.
  • [LS24] Shilin Lai and Christopher Skinner. Anti-cyclotomic Euler system of diagonal cycles. Preprint, available at arXiv:2408.01219, 2024.
  • [LSZ22a] David Loeffler, Christopher Skinner, and Sarah Livia Zerbes. An Euler system for GU(2,1)\rm GU(2,1). Math. Ann., 382(3-4):1091–1141, 2022.
  • [LSZ22b] David Loeffler, Christopher Skinner, and Sarah Livia Zerbes. Euler systems for GSp(4). J. Eur. Math. Soc. (JEMS), 24(2):669–733, 2022.
  • [MS19] Sophie Morel and Junecue Suh. The standard sign conjecture on algebraic cycles: the case of Shimura varieties. J. Reine Angew. Math., 748:139–151, 2019.
  • [Sak08] Yiannis Sakellaridis. On the unramified spectrum of spherical varieties over pp-adic fields. Compositio Mathematica, 144(4):978–1016, 2008.
  • [Sak12] Yiannis Sakellaridis. Spherical varieties and integral representations of LL-functions. Algebra Number Theory, 6(4):611–667, 2012.
  • [Sak13] Yiannis Sakellaridis. Spherical functions on spherical varieties. American Journal of Mathematics, 135(5):1291–1381, 2013.
  • [Sak18] Yiannis Sakellaridis. Inverse Satake transforms. In Geometric aspects of the trace formula, Simons Symp., pages 321–349. Springer, Cham, 2018.
  • [Sha23] Syed Waqar Ali Shah. Explicity Hecke descent for special cycles. Preprint, available at arXiv:2310.01677, 2023.
  • [Sha24] Syed Waqar Ali Shah. On constructing zeta elements for Shimura varieties. Preprint, available at arXiv:2409.03517, 2024.
  • [SV17] Yiannis Sakellaridis and Akshay Venkatesh. Periods and harmonic analysis on spherical varieties. Astérisque, (396):viii+360, 2017.
  • [SV24] Marco Sangiovanni Vincentelli. Crafting Euler Systems: Beyond the Motivic Mold. PhD thesis, 2024.
  • [SW22] Yiannis Sakellaridis and Jonathan Wang. Intersection complexes and unramified LL-factors. J. Amer. Math. Soc., 35:799–910, 2022.