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

\marginsize

3cm3cm1cm2cm

pp-adic equidistribution of modular geodesics and of Heegner points on Shimura curves

Patricio Pérez-Piña Pontificia Universidad Católica de Chile perezpinha.patricio@gmail.com
Abstract.

We propose a pp-adic version of Duke’s Theorem on the equidistribution of closed geodesics on modular curves. Our approach concerns quadratic fields split at pp as well as a pp-adic covering of the modular curve. We also prove an equidistribution result of Heegner points in the pp-adic space attached to Shimura curves.

1. Introduction

In [Duk88], W. Duke proves the equidistribution, inside the complex points of the modular curve endowed with the hyperbolic measure, of arithmetic objects attached to orders in quadratic extensions of \mathbb{Q}. The nature of these objects depends on the behavior at infinity of the extension. For quadratic imaginary fields these objects are Galois orbits of CM points and in the case of real quadratic fields his equidistribution theorem concerns packets of closed geodesics.

Let pp be a rational prime. Since CM points are defined over number fields, their Galois orbits can be naturally embedded inside the pp-adic space of p\mathbb{C}_{p}-points of the modular curve. In [HMRL20] and [HMRL20], the authors studied the distribution of such Galois orbits obtaining pp-adic analogues to Duke’s theorem in the case of imaginary quadratic fields. In contrast with Duke’s result, different limit measures can appear, depending on the discriminants of the orders involved in the sequence of CM points.

The aim of this article is to state and prove a pp-adic version of Duke’s theorem on the equidistribution of packets of closed geodesics. Also, we prove a pp-adic version of the equidistribution of Galois orbits of Heegner points on Shimura curves, when the prime pp divides the discriminant of the underlying quaternion algebra. The complex version of the latter result is another extension of Duke’s theorem for which we refer to section 6 in [HM06] or to [Zha05].

We recall Duke’s result following section 2 of [ELMV12]. Let \mathcal{H} be the complex upper half plane and denote by Y0Y_{0} the modular curve of level 11. Then we have an identification between the quotient space PSL2()\\mathrm{PSL}_{2}(\mathbb{Z})\backslash\mathcal{H} and Y0()Y_{0}(\mathbb{C}). Moreover, the quotient space PSL2()\PGL2()+\mathrm{PSL}_{2}(\mathbb{Z})\backslash\mathrm{PGL}_{2}(\mathbb{R})^{+} is identified with T1(Y0())T^{1}(Y_{0}(\mathbb{C})), the unit tangent bundle of Y0()Y_{0}(\mathbb{C}). Here and elsewhere, the superscript + indicates that we consider matrices of positive determinant. Suppose that τ\tau is a real quadratic irrational. We can view τ\tau as an element in the boundary of \mathcal{H} and we can consider the geodesic γτ\gamma_{\tau} in \mathcal{H} connecting τ\tau and its Galois conjugate. The projection of γτ\gamma_{\tau} to PSL2()\{\mathrm{PSL}_{2}(\mathbb{Z})}\backslash{\mathcal{H}} is a closed geodesic. We lift this closed geodesic to a compact geodesic orbit γ~τ\widetilde{\gamma}_{\tau} in T1(Y0())T^{1}(Y_{0}(\mathbb{C})) that depends only on the PSL2()\mathrm{PSL}_{2}(\mathbb{Z})-orbit of τ\tau. Also, we denote by 𝒪τ\mathcal{O}_{\tau} the ring of matrices in M2()M_{2}(\mathbb{Z}) having the column (τ   1)t(\tau\,\,\,1)^{t} as an eigenvector. For KK a real quadratic field, we denote by RQ(K)RQ(K) the set of τ\tau as before such that 𝒪τ\mathcal{O}_{\tau} is isomorphic to the maximal order 𝒪K\mathcal{O}_{K}. The set PSL2()\RQ(K)\mathrm{PSL}_{2}(\mathbb{Z})\backslash RQ(K) is finite and its cardinality equals the class number hKh_{K} of KK. We put 𝒢K=τγ~τ\mathcal{G}_{K}=\bigcup_{\tau}\widetilde{\gamma}_{\tau}, with τ\tau running through representatives of PSL2()\RQ(K)\mathrm{PSL}_{2}(\mathbb{Z})\backslash RQ(K). Define by μK\mu_{K} the unique probability measure in T1(Y0())T^{1}(Y_{0}(\mathbb{C})) supported on 𝒢K\mathcal{G}_{K} that is invariant under the geodesic flow. Endow T1(Y0())T^{1}(Y_{0}(\mathbb{C})) with the unique probability measure μL\mu_{L} coming from a Haar measure on PGL2+()\mathrm{PGL}_{2}^{+}(\mathbb{R}) and a counting measure on PSL2()\mathrm{PSL}_{2}(\mathbb{Z}).

Theorem 1.1 (Duke).

As disc(K)disc(K)\to\infty, the sequence μK\mu_{K} converges to μL\mu_{L} in the weak-* topology. In other words, the collection 𝒢K\mathcal{G}_{K} of hKh_{K} closed geodesics becomes equidistributed in T1(Y0())T^{1}(Y_{0}(\mathbb{C})) with respect to the measure μL\mu_{L}.

We proceed to explain our main result on geodesics. As we have seen, in the complex case, geodesics appear when dealing with real quadratic fields or in other words when the infinite place is split in the extension. Following this point of view, we exchange the role of \infty and pp and we focus on quadratic extensions such that pp splits. The role played by PSL2()\mathrm{PSL}_{2}(\mathbb{Z}) and PGL2+()\mathrm{PGL}_{2}^{+}(\mathbb{R}) will be held by Γ+\colonequalsPGL2+([1/p])\Gamma^{+}\colonequals\mathrm{PGL}^{+}_{2}(\mathbb{Z}[1/p]) and PGL2(×p)\mathrm{PGL}_{2}(\mathbb{R}\times\mathbb{Q}_{p}), respectively. Let \mathcal{H}^{\prime} be the set of points τ\tau in \mathcal{H} such that τ\tau generates an imaginary quadratic extension over \mathbb{Q} in which pp splits. The quotient Γ+\\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma^{+}$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{H}^{\prime}$}}}{\Gamma^{+}\,\backslash\,\mathcal{H}^{\prime}}{\Gamma^{+}\,\backslash\,\mathcal{H}^{\prime}}{\Gamma^{+}\,\backslash\,\mathcal{H}^{\prime}} was studied by Ihara who showed that it is related to the special fiber of the modular curve X0(p)X_{0}(p).

Theorem 1.2 (Chapter 5, Theorem 1 in [Iha08]).

Fix 𝔓\mathfrak{P}, a prime above pp in ¯\overline{\mathbb{Z}}, the integral closure of \mathbb{Z} in ¯\overline{\mathbb{Q}}. Identify ¯/𝔓\overline{\mathbb{Z}}/\mathfrak{P} with 𝔽p¯\overline{\mathbb{F}_{p}} and recall that by the Theory of Complex Multiplication, for any τ\tau\in\mathcal{H}^{\prime}, we have j(τ)¯j(\tau)\in\overline{\mathbb{Z}}. The assignment

Γ+τGal(𝔽p¯/𝔽p)(j(τ)mod𝔓)\Gamma^{+}\tau\mapsto\mathrm{Gal}(\overline{\mathbb{F}_{p}}/\mathbb{F}_{p})\cdot(j(\tau)\mod\mathfrak{P})

defines a bijection between Γ+\\Gamma^{+}\backslash\mathcal{H}^{\prime} and Galois orbits in 𝔽p¯SS\overline{\mathbb{F}_{p}}-SS, where SSSS is the set of supersingular jj-invariants.

The pp-adic covering of Y0()Y_{0}(\mathbb{C}) that we consider is the locally compact space

T(p)(Y0())\colonequalsΓ+\(×PGL2(p)).T^{(p)}(Y_{0}(\mathbb{C}))\colonequals\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma^{+}$}\big{\backslash}\raise 2.15277pt\hbox{$(\mathcal{H}\times\mathrm{PGL}_{2}(\mathbb{Q}_{p}))$}}}{\Gamma^{+}\,\backslash\,(\mathcal{H}\times\mathrm{PGL}_{2}(\mathbb{Q}_{p}))}{\Gamma^{+}\,\backslash\,(\mathcal{H}\times\mathrm{PGL}_{2}(\mathbb{Q}_{p}))}{\Gamma^{+}\,\backslash\,(\mathcal{H}\times\mathrm{PGL}_{2}(\mathbb{Q}_{p}))}.

The natural projection T(p)(Y0())Y0()T^{(p)}(Y_{0}(\mathbb{C}))\to Y_{0}(\mathbb{C}) is given by taking a right quotient by PGL2(p)\mathrm{PGL}_{2}(\mathbb{Z}_{p}) (see section 5). Let 𝒯p\mathcal{T}_{p} denote the Bruhat-Tits tree of PGL2(p)\mathrm{PGL}_{2}(\mathbb{Q}_{p}). After identifying the boundary of the tree 𝒯p\mathcal{T}_{p} with 1(p)\mathbb{P}^{1}(\mathbb{Q}_{p}), the splitness condition on pp allows us to see \mathcal{H}^{\prime} as boundary elements of 𝒯p\mathcal{T}_{p}. For τ\tau in \mathcal{H}^{\prime}, we define in Section 5 a cycle Δτ\Delta_{\tau} inside the space T(p)(Y0())T^{(p)}(Y_{0}(\mathbb{C})) than depends only on the Γ+\Gamma^{+}-orbit of τ\tau. The projection of Δτ\Delta_{\tau} on Γ+\\Gamma^{+}\backslash\mathcal{H} is just Γ+τ\Gamma^{+}\tau. The fiber over Γ+τ\Gamma^{+}\tau under the natural projection Γ+\(×𝒯p)Γ+\\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma^{+}$}\big{\backslash}\raise 2.15277pt\hbox{$(\mathcal{H}^{\prime}\times\mathcal{T}_{p})$}}}{\Gamma^{+}\,\backslash\,(\mathcal{H}^{\prime}\times\mathcal{T}_{p})}{\Gamma^{+}\,\backslash\,(\mathcal{H}^{\prime}\times\mathcal{T}_{p})}{\Gamma^{+}\,\backslash\,(\mathcal{H}^{\prime}\times\mathcal{T}_{p})}\to\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma^{+}$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{H}^{\prime}$}}}{\Gamma^{+}\,\backslash\,\mathcal{H}^{\prime}}{\Gamma^{+}\,\backslash\,\mathcal{H}^{\prime}}{\Gamma^{+}\,\backslash\,\mathcal{H}^{\prime}} is Γ+\(Γ+τ×𝒯p)\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma^{+}$}\big{\backslash}\raise 2.15277pt\hbox{$(\Gamma^{+}\tau\times\mathcal{T}_{p})$}}}{\Gamma^{+}\,\backslash\,(\Gamma^{+}\tau\times\mathcal{T}_{p})}{\Gamma^{+}\,\backslash\,(\Gamma^{+}\tau\times\mathcal{T}_{p})}{\Gamma^{+}\,\backslash\,(\Gamma^{+}\tau\times\mathcal{T}_{p})}. This is a graph having the structure of a (p+1)(p+1)-volcano, that can be identified with the pp-isogeny graph attached to any elliptic curve coming from a PSL2()\mathrm{PSL}_{2}(\mathbb{Z})-orbit in Γ+τ\Gamma^{+}\tau. The projection of Δτ\Delta_{\tau} to this graph is its unique closed loop (known as its rim). Observe that there’s a bijection between Γ+\(Γ+τ×𝒯p)\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma^{+}$}\big{\backslash}\raise 2.15277pt\hbox{$(\Gamma^{+}\tau\times\mathcal{T}_{p})$}}}{\Gamma^{+}\,\backslash\,(\Gamma^{+}\tau\times\mathcal{T}_{p})}{\Gamma^{+}\,\backslash\,(\Gamma^{+}\tau\times\mathcal{T}_{p})}{\Gamma^{+}\,\backslash\,(\Gamma^{+}\tau\times\mathcal{T}_{p})} and {τ}×Γτ+\𝒯p\{\tau\}\times\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma^{+}_{\tau}$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{T}_{p}$}}}{\Gamma^{+}_{\tau}\,\backslash\,\mathcal{T}_{p}}{\Gamma^{+}_{\tau}\,\backslash\,\mathcal{T}_{p}}{\Gamma^{+}_{\tau}\,\backslash\,\mathcal{T}_{p}}, where Γτ+\Gamma^{+}_{\tau} denotes the stabilizer of τ\tau in Γ+\Gamma^{+}. The aforementioned loop is formed by taking the unique geodesic in 𝒯p\mathcal{T}_{p} joining τ\tau and its Galois conjugate and projecting it to Γτ+\𝒯p\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma^{+}_{\tau}$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{T}_{p}$}}}{\Gamma^{+}_{\tau}\,\backslash\,\mathcal{T}_{p}}{\Gamma^{+}_{\tau}\,\backslash\,\mathcal{T}_{p}}{\Gamma^{+}_{\tau}\,\backslash\,\mathcal{T}_{p}}. This resembles the complex case where the closed geodesics on Y0()Y_{0}(\mathbb{C}) are projections of geodesics in \mathcal{H} joining two real quadratic numbers that are conjugated. In [BDIS02], such loops in Γτ+\𝒯p\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma^{+}_{\tau}$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{T}_{p}$}}}{\Gamma^{+}_{\tau}\,\backslash\,\mathcal{T}_{p}}{\Gamma^{+}_{\tau}\,\backslash\,\mathcal{T}_{p}}{\Gamma^{+}_{\tau}\,\backslash\,\mathcal{T}_{p}} are called Shintani-cycles, for that reason we refer to Δτ\Delta_{\tau} as the Ihara-Shintani cycle attached to Γ+τ\Gamma^{+}\tau. In this way, we think of the cycles Δτ\Delta_{\tau} as the pp-adic analogues to the compact geodesic orbits on the modular surface and the space T(p)(Y0())T^{(p)}(Y_{0}(\mathbb{C})) as a pp-adic unit tangent bundle of Y0()Y_{0}(\mathbb{C}).

For τ\tau\in\mathcal{H}^{\prime}, we denote by 𝒪τ\mathcal{O}_{\tau} the ring formed by the elements in M2([1/p])M_{2}(\mathbb{Z}[1/p]) having the column (τ   1)t(\tau\,\,\,1)^{t} as an eigenvector. Let KK be a quadratic imaginary field at which pp splits and let 𝒪\mathcal{O} be a [1/p]\mathbb{Z}[1/p]-order inside KK. We denote by IS(𝒪)IS(\mathcal{O}) the collection of Γ+\Gamma^{+}-orbits of those τ\tau\in\mathcal{H}^{\prime} for which (τ)=K\mathbb{Q}(\tau)=K and 𝒪τ\mathcal{O}_{\tau} is isomorphic to the order 𝒪\mathcal{O}. The group Pic(𝒪)Pic(\mathcal{O}) act simply transitive on IS(𝒪)IS(\mathcal{O}) and so this set is finite. Let AA be the diagonal group of PGL2\mathrm{PGL}_{2}. The cycle Δτ\Delta_{\tau} comes equipped with a unique A(p)A(\mathbb{Q}_{p})-right invariant probability measure ντ\nu_{\tau}. Denote by μ𝒪\mu_{\mathcal{O}} the unique probability measure proportional to Γ+τIS(𝒪)ντ\sum_{\Gamma^{+}\tau\in IS(\mathcal{O})}\nu_{\tau}.

Equip T(p)(Y0())T^{(p)}(Y_{0}(\mathbb{C})) with the unique probability measure coming from a Haar measure on PGL2(×p)\mathrm{PGL}_{2}(\mathbb{R}\times\mathbb{Q}_{p}) and the counting measure in Γ\Gamma. Denote this measure by μ\mu. Our pp-adic analogue to Duke’s theorem on the equidistribution of closed geodesics is the following statement.

Theorem A.

As disc(𝒪)disc(\mathcal{O})\to-\infty, the sequence μ𝒪\mu_{\mathcal{O}} converges to μ\mu in the weak-* topology. In other words, for every continous and compactly supported function f:T(p)(Y0())f\colon T^{(p)}(Y_{0}(\mathbb{C}))\to\mathbb{C},

limdisc(𝒪)f𝑑μ𝒪f𝑑μ.\lim_{disc(\mathcal{O})\to-\infty}\int fd\mu_{\mathcal{O}}\to\int fd\mu.

Now we discuss our result on Heegner points on Shimura curves. Fix \mathcal{B} a non-split indefinite quaternion algebra defined over \mathbb{Q} whose discriminant NN^{-} is divisible by pp. Also fix \mathcal{R} an Eichler-order in \mathcal{B} of level N+N^{+}. Let XX be the Shimura curve attached to (,)(\mathcal{B},\mathcal{R}) as in section 6. Let p2\mathbb{Q}_{p^{2}} be the unique quadratic unramified extension of p\mathbb{Q}_{p}. The compact space X(p2)X(\mathbb{Q}_{p^{2}}) admits a uniformization by PGL2(p)\mathrm{PGL}_{2}(\mathbb{Q}_{p}) due to Cerednik and Drinfeld. Pushing a Haar measure of PGL2(p)\mathrm{PGL}_{2}(\mathbb{Q}_{p}) we can endow X(p2)X(\mathbb{Q}_{p^{2}}) with a natural probability measure μX\mu_{X}. Let 𝒪\mathcal{O} be an order in a quadratic imaginary field in which pp is inert and whose discriminant is coprime to pNN+pN^{-}N^{+}. Let Heeg(𝒪)Heeg(\mathcal{O}) denote the collection of Heegner points in X(p2)X(\mathbb{Q}_{p^{2}}) with endomorphism ring isomorphic to 𝒪\mathcal{O} defined in Section 6. We obtain the following equidistribution theorem.

Theorem B.

As disc(𝒪)disc(\mathcal{O})\to-\infty, the collection Heeg(𝒪)Heeg(\mathcal{O}) becomes equidistributed on X(p2)X(\mathbb{Q}_{p^{2}}) with respect to the measure μX\mu_{X}. In other words, for every continuous function f:X(p2)f\colon X(\mathbb{Q}_{p^{2}})\to\mathbb{C},

limdisc(𝒪)1#Heeg(𝒪)PHeeg(𝒪)f(P)X(p2)f𝑑μX.\lim_{disc(\mathcal{O})\to-\infty}\frac{1}{\#Heeg(\mathcal{O})}\sum_{P\in Heeg(\mathcal{O})}f(P)\to\int_{X(\mathbb{Q}_{p^{2}})}fd\mu_{X}.

A related result by Disegni [Dis22] concerns the case where the power of pp in the conductor of the Heegner points tends to infinity. In such a situation there is no accumulation measure on X(p)X(\mathbb{C}_{p}) describing the asymptotic distribution of their Galois orbits

The main tool for our purposes is Theorem 4.6 in [ELMV11]. In consequence, our strategy is to describe the previous objects and ambient spaces in an SS-arithmetic and adelic context. In doing so we will appeal to the language of (oriented) optimal embeddings to parametrize our packets as orbits under certain Pic groups which will lead to a description of these as toric orbits inside a homogeneous space. In section 2 we cover the basics about the Bruhat-Tits tree of PGL2\mathrm{PGL}_{2} making special emphasis on how quadratic algebras act on it. In section 3 we recall the language of adelic and SS-arithmetic homogeneous spaces as presented for instance in [ELMV11]. In section 4 we elaborate on the definition of cycles Δψ\Delta_{\psi} in a general setting, where ψ\psi runs over embeddings of quadratic extension to a quaternion algebra. More specifically, in section 4.1 and 4.2 we discuss the structure of Eichler orders and oriented optimal embeddings respectively. In section 4.3 we show how the group Pic(𝒪)Pic(\mathcal{O}) acts simply and transitively on the set of oriented optimal embeddings (with respect to 𝒪\mathcal{O}), showing that the cycles Δψ\Delta_{\psi} defined in section 4.4 are finite and they can be recovered from a single oriented optimal embedding. Section 5 is a specialization of the previous sections and it is where we properly define the cycles Δτ\Delta_{\tau} for τ\tau in Γ+\\Gamma^{+}\backslash\mathcal{H}^{\prime} and discuss the analogies with the geodesics in Duke’s theorem. Section 6 deals with the different specialization related to Heegner points on the pp-adic points of indefinite Shimura curves. Finally, section 7 presents Theorem 4.6 in [ELMV11]. We explain how to apply it to show the equidistribution of the objects defined in sections 5 and 6, proving in particular Theorem A and Theorem B.

Acknowledgments

I would like to thank my advisor Ricardo Menares for trusting me with this project and for his support throughout this work. I also would like to thank Philippe Michel and the Analytic Number Theory group at EPFL. I was very lucky to enjoy their hospitality during the academic period 2022-2023. This work was supported by ANID Doctorado Nacional No 21200911 and by the Bourse d’excellence de la Confédération suisse No. 2022.0414.

2. Dynamics of tori acting on the Bruhat-Tits tree

Our main references for this section are [Ser03], [Voi21] and [Cas14]. Let \ell be a rational prime. Let 𝒯\mathcal{T}_{\ell} be the Bruhat-Tits tree of PGL2()\mathrm{PGL}_{2}(\mathbb{Q}_{\ell}). Its set of vertices 𝒱(𝒯)\mathcal{V}(\mathcal{T}_{\ell}) corresponds to homothety classes of \mathbb{Z}_{\ell}-lattices in 2\mathbb{Q}_{\ell}^{2}. The class of a lattice Λ\Lambda is denoted by [Λ][\Lambda] and for nn\in\mathbb{Z} we set vn\colonequals[n×]v_{n}\colonequals[\ell^{n}\mathbb{Z}_{\ell}\times\mathbb{Z}_{\ell}]. Two vertices are connected by an edge if they admit representatives Λ\Lambda and Λ\Lambda^{\prime} such that ΛΛΛ\ell\Lambda\subseteq\Lambda^{\prime}\subseteq\Lambda and Λ/Λ\Lambda/\Lambda^{\prime} is cyclic of order \ell. With these definitions, 𝒯\mathcal{T}_{\ell} is a (+1)(\ell+1)-regular tree.

If ee is an edge connecting two vertices, the choice of one of them as its source (denoted s(e)s(e)) and the other one as its target (denoted t(e)t(e)) make ee an oriented edge. We denote by (𝒯)\overrightarrow{\mathcal{E}}(\mathcal{T}_{\ell}) the set of oriented edges. Let ee_{\infty} be the oriented edge such that s(e)=v0s(e_{\infty})=v_{0} and t(e)=v1t(e_{\infty})=v_{-1}. The group PGL2()\mathrm{PGL}_{2}(\mathbb{Q}_{\ell}) acts transitively on 𝒱(𝒯)\mathcal{V}(\mathcal{T}_{\ell}) and (𝒯)\overrightarrow{\mathcal{E}}(\mathcal{T}_{\ell}). The subgroups PGL2()\mathrm{PGL}_{2}(\mathbb{Z}_{\ell}) and

Γ0()={(abcd)PGL2():c}\Gamma_{0}(\ell\mathbb{Z}_{\ell})=\left\{\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\mathrm{PGL}_{2}(\mathbb{Z}_{\ell}):\ell\mid c\right\}

are the stabilizer of v0v_{0} and ee_{\infty} respectively. Consequently, we have identifications

𝒱(𝒯)=PGL2()/PGL2() and (𝒯)=PGL2()/Γ0().\mathcal{V}(\mathcal{T}_{\ell})=\mathrm{PGL}_{2}(\mathbb{Q}_{\ell})/\mathrm{PGL}_{2}(\mathbb{Z}_{\ell})\mbox{ and }\overrightarrow{\mathcal{E}}(\mathcal{T}_{\ell})=\mathrm{PGL}_{2}(\mathbb{Q}_{\ell})/\Gamma_{0}(\ell\mathbb{Z}_{\ell}).

A path of length n0n\geq 0 in 𝒯\mathcal{T}_{\ell} is sequence of nn adjacent vertices without backtracking. Paths of length nn starting from v0v_{0} are identified with 1(/n)\mathbb{P}^{1}(\mathbb{Z}/\ell^{n}\mathbb{Z}): To give such a path is equivalent to give a vertex vv at distance nn from v0v_{0}. The group PGL2()\mathrm{PGL}_{2}(\mathbb{Z}_{\ell}) acts transitively on such vertices and one such vertex is vnv_{-n}. Its stabilizer in PGL2()\mathrm{PGL}_{2}(\mathbb{Z}_{\ell}) is

Γ0(n)={(abcd)PGL2():nc}.\Gamma_{0}(\ell^{n}\mathbb{Z}_{\ell})=\left\{\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\mathrm{PGL}_{2}(\mathbb{Z}_{\ell}):\ell^{n}\mid c\right\}.

Therefore we have PGL2()\mathrm{PGL}_{2}(\mathbb{Z}_{\ell})-equivariant identifications between

(2.1) PGL2()/Γ0(n)1(/n),\mathrm{PGL}_{2}(\mathbb{Z}_{\ell})/\Gamma_{0}(\ell^{n}\mathbb{Z}_{\ell})\cong\mathbb{P}^{1}(\mathbb{Z}/\ell^{n}\mathbb{Z}),

and paths of length nn starting from v0v_{0} such that the path ending at vnv_{-n} corresponds to \infty. The action of PGL2()\mathrm{PGL}_{2}(\mathbb{Z}_{\ell}) on the right hand side of (2.1) is by fractional linear transformations.

A branch in 𝒯\mathcal{T}_{\ell} is a sequence of adjacent vertices without backtracking. Two branches are equivalent if they differ by a finite initial sequence. An equivalence class of branches is called an end of 𝒯\mathcal{T}_{\ell} and we denote the set of ends by 𝒯\partial\mathcal{T}_{\ell}. We can see 𝒯\partial\mathcal{T}_{\ell} as the boundary of the tree and identify it with 1()\mathbb{P}^{1}(\mathbb{Q}_{\ell}): The group PGL2()PGL_{2}(\mathbb{Q}_{\ell}) acts transitively on the set of ends and the stabilizer of the end coming from the branch {vn}n0\{v_{-n}\}_{n\geq 0} is the subgroup B()PGL2()B(\mathbb{Q}_{\ell})\subseteq\mathrm{PGL}_{2}(\mathbb{Q}_{\ell}) of upper triangular elements. We have PGL2()\mathrm{PGL}_{2}(\mathbb{Q}_{\ell})-equivariant identifications

(2.2) 𝒯PGL2()/B()PGL2()/B()1()=1(),\partial\mathcal{T}_{\ell}\cong\mathrm{PGL}_{2}(\mathbb{Q}_{\ell})/B(\mathbb{Q}_{\ell})\cong\mathrm{PGL}_{2}(\mathbb{Z}_{\ell})/B(\mathbb{Z}_{\ell})\cong\mathbb{P}^{1}(\mathbb{Z}_{\ell})=\mathbb{P}^{1}(\mathbb{Q}_{\ell}),

such that that the branch coming from {vn}n0\{v_{-n}\}_{n\geq 0} corresponds to \infty. As before, under this identification the action of PGL2()\mathrm{PGL}_{2}(\mathbb{Q}_{\ell}) on 1()\mathbb{P}^{1}(\mathbb{Q}_{\ell}) corresponds to the action by fractional linear transformations.

A geodesic is the path without backtracking in the tree connecting two different points in the boundary. The geodesic joining the points 0 and \infty is given by the vertices {vn}n\{v_{n}\}_{n\in\mathbb{Z}} and we denote it by 𝒢\mathcal{G}. In general 𝒢xy\mathcal{G}_{x}^{y} denotes the geodesic joining xx and yy.

Let vv and vv^{\prime} be two vertices in 𝒯\mathcal{T}_{\ell}. Being 𝒯\mathcal{T}_{\ell} a tree, there exists a unique finite path connecting vv and vv^{\prime}. The distance d(v,v)d(v,v^{\prime}) between vv and vv^{\prime} is the length of this path. For instance, two vertices are at distance 11 if they are connected by an edge. Let ee an edge and η\eta a path (no necessarily of finite length) in the tree. The distance d(v,η)d(v,\eta) between vv and η\eta is minwηd(v,w)\min_{w\in\eta}d(v,w) and the distance d(e,η)d(e,\eta) between ee and η\eta is maxwed(w,η)\max_{w\in e}{d(w,\eta)}.

We denote by APGL2()A\subseteq PGL_{2}(\mathbb{Q}_{\ell}) the subgroup of diagonal elements.

Proposition 2.1.

The following hold

  1. (1)

    The group AA fixes the points 0 and \infty in 𝒯\partial\mathcal{T}_{\ell}. For every n0n\geq 0, it acts transitively on the set of vertices vv with d(v,𝒢)=nd(v,\mathcal{G})=n.

  2. (2)

    The group PGL2()\mathrm{PGL}_{2}(\mathbb{Z}_{\ell}) has v0v_{0} as a unique fixed vertex and for every n0n\geq 0 it acts transitively on the sets of vertices vv with d(v,v0)=nd(v,v_{0})=n.

  3. (3)

    The group Γ0(),(010)\left\langle\Gamma_{0}(\ell\mathbb{Z}_{\ell}),\begin{pmatrix}0&1\\ \ell&0\end{pmatrix}\right\rangle has a unique fixed edge e={v0,v1}e_{\infty}=\{v_{0},v_{-1}\}. For every n0n\geq 0, it acts transitively on the set of edges ee with d(e,e)=nd(e,e_{\infty})=n.

Proof.

See chapter I, sections 3-5 in [Cas14]. ∎

Let 𝒦M2()\mathcal{K}\subseteq M_{2}(\mathbb{Q}_{\ell}) be a quadratic \mathbb{Q}_{\ell}-algebra. We denote by 𝒪𝒦\mathcal{O}_{\mathcal{K}} the maximal compact and open subring of 𝒦\mathcal{K}. The order of conductor n\ell^{n} is +n𝒪𝒦\mathbb{Z}_{\ell}+\ell^{n}\mathcal{O}_{\mathcal{K}}. We set 𝔪𝒦\colonequals𝒪𝒦\mathfrak{m}_{\mathcal{K}}\colonequals\ell\mathcal{O}_{\mathcal{K}}, 𝒪𝒦(0)=𝒪𝒦×\mathcal{O}_{\mathcal{K}}^{(0)}=\mathcal{O}_{\mathcal{K}}^{\times} and for n1n\geq 1 we define 𝒪𝒦(n)=1+𝔪𝒦n\mathcal{O}_{\mathcal{K}}^{(n)}=1+\mathfrak{m}_{\mathcal{K}}^{n}, the nn-th principal subgroup of units.

Corollary 2.2.

Let 𝒦M2()\mathcal{K}\subseteq M_{2}(\mathbb{Q}_{\ell}) be a quadratic \mathbb{Q}_{\ell}-algebra. Consider the action of 𝒦¯×\colonequals𝒦×/\overline{\mathcal{K}}^{\times}\colonequals\mathcal{K}^{\times}/\mathbb{Q}_{\ell}^{*} on 𝒯\mathcal{T}_{\ell} through the embedding 𝒦×/PGL2()\mathcal{K}^{\times}/\mathbb{Q}_{\ell}^{*}\subseteq PGL_{2}(\mathbb{Q}_{\ell}).

  1. (1)

    If 𝒦/\mathcal{K}/\mathbb{Q}_{\ell} splits, then 𝒦¯×\overline{\mathcal{K}}^{\times} fixes two points x,y𝒯x,y\in\partial\mathcal{T}_{\ell}. For n0n\geq 0, 𝒦¯×\overline{\mathcal{K}}^{\times} acts transitively on the sets of vertices ww with d(w,𝒢xy)=nd(w,\mathcal{G}_{x}^{y})=n. Moreover, the stabilizer of ww is the image in 𝒦¯×\overline{\mathcal{K}}^{\times} of the order of conductor n\ell^{n} in 𝒦\mathcal{K} where d(v,𝒢xy)=nd(v,\mathcal{G}_{x}^{y})=n.

  2. (2)

    If 𝒦/\mathcal{K}/\mathbb{Q}_{\ell} is an unramified field extension, then 𝒦¯×\overline{\mathcal{K}}^{\times} has a unique fixed vertex vv and for n0n\geq 0, it acts transitively on the sets of vertices vv^{\prime} with d(v,v)=nd(v^{\prime},v)=n. Moreover, if d(v,v)=nd(v^{\prime},v)=n, the stabilizer of vv^{\prime} is the image in 𝒦¯×\overline{\mathcal{K}}^{\times} of 𝒪𝒦(n)\mathcal{O}_{\mathcal{K}}^{(n)}.

  3. (3)

    If 𝒦/\mathcal{K}/\mathbb{Q}_{\ell} is a ramified field extension, then 𝒦¯×\overline{\mathcal{K}}^{\times} has a unique fixed edge ee and for n0n\geq 0, it acts transitively on the set of edges ee^{\prime} with d(e,e)=nd(e,e^{\prime})=n. Moreover, if d(e,e)=nd(e^{\prime},e)=n, the stabilizer of ee^{\prime} is the image of the subgroup 𝒪𝒦(2n)\mathcal{O}_{\mathcal{K}}^{(2n)}.

Proof.

The first item follows directly from Proposition 2.1 since there exists gPGL2()g\in PGL_{2}(\mathbb{Q}_{\ell}) such that 𝒦¯×=gAg1\overline{\mathcal{K}}^{\times}=gAg^{-1} and in particular 𝒢xy=g𝒢\mathcal{G}_{x}^{y}=g\mathcal{G}. Also working up to conjugation, for (2)(2) and (3)(3) it is enough to prove it for the algebra 𝒦\mathcal{K} (isomorphic to (d)\mathbb{Q}_{\ell}(\sqrt{d})) given by elements of the form (xydyx)\begin{pmatrix}x&y\\ dy&x\end{pmatrix} where dd\in\mathbb{Z} is a non-zero square mod \ell in case (2) and for case (3) d\ell\mid\mid d. Now we need to observe that in case (2) (resp. case (3)) 𝒦¯×\overline{\mathcal{K}}^{\times} is contained in PGL2()\mathrm{PGL}_{2}(\mathbb{Z}_{\ell}) (resp. Γ0(),(010)\left\langle\Gamma_{0}(\ell\mathbb{Z}_{\ell}),\begin{pmatrix}0&1\\ \ell&0\end{pmatrix}\right\rangle). The statement about the fixed elements follows immediately from Proposition 2.1.

In case (2), transitivity follows since 𝒦¯×\overline{\mathcal{K}}^{\times} acts transitively on 1(/n)\mathbb{P}^{1}(\mathbb{Z}/\ell^{n}\mathbb{Z}) for every nn. For (3) one can assume d\ell\mid\mid d and note that (01d0)\begin{pmatrix}0&1\\ d&0\end{pmatrix} preserves ee_{\infty} by interchanching v0v_{0} and v1v_{-1}. Therefore, it is enough to show that 𝒦¯×Γ0()\overline{\mathcal{K}}^{\times}\cap\Gamma_{0}(\ell\mathbb{Z}_{\ell}) acts transitively on the vertices in the connected component of 𝒯e\mathcal{T}_{\ell}-e_{\infty} connected to v1v_{-1} and at a given distance to v1v_{-1}. Now for (xydyx)\begin{pmatrix}x&y\\ dy&x\end{pmatrix} in the given intersection, x×x\in\mathbb{Z}_{\ell}^{\times} and its action on (1   0)t(1\,\,\,0)^{t} is given by (1dyx1)t(1\,\,\,\,dyx^{-1})^{t}. Hence this orbit in 1(/n+1)\mathbb{P}^{1}(\mathbb{Z}/\ell^{n+1}\mathbb{Z}), which corresponds to the orbit of v(n+1)v_{-(n+1)} in 𝒯e\mathcal{T}_{\ell}-e_{\infty}, has n\ell^{n} distinct elements which is exactly the number of vertices at distance nn from v1v_{-1} in this component.

Once transitivity is proven, the fact about stabilizers follows since these algebras are commutative and by a counting argument: #(𝒪𝒦×/𝒪𝒦(n))/(/(n))=n1(+1)\#\mathchoice{\text{\raise 2.15277pt\hbox{$(\mathcal{O}_{\mathcal{K}}^{\times}/\mathcal{O}_{\mathcal{K}}^{(n)})$}\big{/}\lower 2.15277pt\hbox{$(\mathbb{Z}_{\ell}^{*}/\mathbb{Z}_{\ell}^{(n)})$}}}{(\mathcal{O}_{\mathcal{K}}^{\times}/\mathcal{O}_{\mathcal{K}}^{(n)})\,/\,(\mathbb{Z}_{\ell}^{*}/\mathbb{Z}_{\ell}^{(n)})}{(\mathcal{O}_{\mathcal{K}}^{\times}/\mathcal{O}_{\mathcal{K}}^{(n)})\,/\,(\mathbb{Z}_{\ell}^{*}/\mathbb{Z}_{\ell}^{(n)})}{(\mathcal{O}_{\mathcal{K}}^{\times}/\mathcal{O}_{\mathcal{K}}^{(n)})\,/\,(\mathbb{Z}_{\ell}^{*}/\mathbb{Z}_{\ell}^{(n)})}=\ell^{n-1}(\ell+1) in case (2):

𝒪𝒦×/𝒪𝒦(n)𝔽2×𝒪𝒦(1)/𝒪𝒦(n)\mathcal{O}_{\mathcal{K}}^{\times}/\mathcal{O}_{\mathcal{K}}^{(n)}\cong\mathbb{F}_{\ell^{2}}^{*}\times\mathcal{O}_{\mathcal{K}}^{(1)}/\mathcal{O}_{\mathcal{K}}^{(n)}

which has cardinality (21)2(n1)(\ell^{2}-1)\ell^{2(n-1)}. Then the quotient #(𝒪𝒦×/×)/(𝒪𝒦(n)/(n))\#\mathchoice{\text{\raise 2.15277pt\hbox{$(\mathcal{O}_{\mathcal{K}}^{\times}/\mathbb{Z}_{\ell}^{\times})$}\big{/}\lower 2.15277pt\hbox{$(\mathcal{O}_{\mathcal{K}}^{(n)}/\mathbb{Z}_{\ell}^{(n)})$}}}{(\mathcal{O}_{\mathcal{K}}^{\times}/\mathbb{Z}_{\ell}^{\times})\,/\,(\mathcal{O}_{\mathcal{K}}^{(n)}/\mathbb{Z}_{\ell}^{(n)})}{(\mathcal{O}_{\mathcal{K}}^{\times}/\mathbb{Z}_{\ell}^{\times})\,/\,(\mathcal{O}_{\mathcal{K}}^{(n)}/\mathbb{Z}_{\ell}^{(n)})}{(\mathcal{O}_{\mathcal{K}}^{\times}/\mathbb{Z}_{\ell}^{\times})\,/\,(\mathcal{O}_{\mathcal{K}}^{(n)}/\mathbb{Z}_{\ell}^{(n)})} has cardinality (21)2(n1)/(1)n1=(+1)n1,(\ell^{2}-1)\ell^{2(n-1)}/(\ell-1)\ell^{n-1}=(\ell+1)\ell^{n-1}, the number of points at distance nn from v0v_{0}.

In case (3),

(𝒪𝒦×/𝒪𝒦(2n))𝔽×𝒪𝒦(1)/𝒪𝒦(2n),(\mathcal{O}_{\mathcal{K}}^{\times}/\mathcal{O}_{\mathcal{K}}^{(2n)})\cong\mathbb{F}_{\ell}^{*}\times\mathcal{O}_{\mathcal{K}}^{(1)}/\mathcal{O}_{\mathcal{K}}^{(2n)},

which has cardinality (1)2n1(\ell-1)\ell^{2n-1}. Then

(𝒪𝒦(1)/(1))/(𝒪𝒦(2n)/(n))=(𝒪𝒦(1)/𝒪𝒦(2n))/((1)/(n)),(\mathcal{O}_{\mathcal{K}}^{(1)}/\mathbb{Z}_{\ell}^{(1)})/(\mathcal{O}_{\mathcal{K}}^{(2n)}/\mathbb{Z}_{\ell}^{(n)})=(\mathcal{O}_{\mathcal{K}}^{(1)}/\mathcal{O}_{\mathcal{K}}^{(2n)})/(\mathbb{Z}_{\ell}^{(1)}/\mathbb{Z}_{\ell}^{(n)}),

has cardinality 2n1/n1=n\ell^{2n-1}/\ell^{n-1}=\ell^{n}, the number of vertices in the connected component of v1v_{-1} and at distance nn to it. ∎

3. Adelic and SS-arithmetic homogeneous spaces

Fix pp be a rational prime. Let 𝔸\mathbb{A} be the ring of adeles over \mathbb{Q}. Let BB be a quaternion algebra over \mathbb{Q} that splits over pp. Denote by 𝑮\boldsymbol{G} the algebraic group PB×\colonequalsB×/𝔾mPB^{\times}\colonequals B^{\times}/\mathbb{G}_{m} of projectivized units of BB. We use 𝑮~\widetilde{\boldsymbol{G}} for the algebraic group B1B^{1} of units of reduced norm 11 in BB.

Let AA be a ring. If vv a place of \mathbb{Q}, we denote by AvA_{v} the algebra AvA\otimes\mathbb{Z}_{v} if vv is finite or AA\otimes\mathbb{R} if v=v=\infty. Let SS be a finite set of places of \mathbb{Q} containing the place pp. We use the notations 𝔸(S)=wSw\mathbb{A}^{(S)}=\prod^{\prime}_{w\not\in S}\mathbb{Q}_{w}, 𝔸f=𝔸()\mathbb{A}_{f}=\mathbb{A}^{(\infty)}, 𝔸#=𝔸(p,)\mathbb{A}_{\#}=\mathbb{A}^{(p,\infty)} and S=vSv\mathbb{Q}_{S}=\prod_{v\in S}\mathbb{Q}_{v}. Anagolously , we use ^\widehat{\mathbb{Z}}, ^#\widehat{\mathbb{Z}}_{\#} to denote the closure of \mathbb{Z} in 𝔸f\mathbb{A}_{f} and in 𝔸(p)𝔸#\mathbb{A}^{(p)}\subset\mathbb{A}_{\#} respectively. In the same direction, A^\widehat{A} and A^#\widehat{A}_{\#} will denote A^A\otimes\widehat{\mathbb{Z}} and A^#A\otimes\widehat{\mathbb{Z}}_{\#} respectively. Finally, if tt is an adelic element in 𝔸\mathbb{A}, tft_{f}, t#t_{\#}, tSt_{S} and tvt_{v} will denote the respective projection of tt in 𝔸f\mathbb{A}_{f}, 𝔸#\mathbb{A}_{\#}, S\mathbb{Q}_{S} and v\mathbb{Q}_{v}. Similarly for 𝔸\mathbb{A}-points over 𝑮\boldsymbol{G} and 𝑮~\widetilde{\boldsymbol{G}}.

The group 𝑮~\widetilde{\boldsymbol{G}} satisfies the following strong approximation theorem.

Theorem 3.1.

The subgroup 𝐆~()𝐆~(p)\widetilde{\boldsymbol{G}}(\mathbb{Q})\widetilde{\boldsymbol{G}}(\mathbb{Q}_{p}) is dense in 𝐆~(𝔸)\widetilde{\boldsymbol{G}}(\mathbb{A}).

Proof.

See [Vig80], Théoreme 4.3, p.81. ∎

In particular, if W𝑮~(𝔸(S))W\subseteq\widetilde{\boldsymbol{G}}(\mathbb{A}^{(S)}) is an open subgroup, then 𝑮~(𝔸)=𝑮~()𝑮~(S)W\widetilde{\boldsymbol{G}}(\mathbb{A})=\widetilde{\boldsymbol{G}}(\mathbb{Q})\widetilde{\boldsymbol{G}}(\mathbb{Q}_{S})W.

Corollary 3.2.

Let HB×(𝔸(S))H\subseteq B^{\times}(\mathbb{A}^{(S)}) be an open subgroup such that nr(B×(𝔸))=nr(B×()B×(S)H)\mathrm{nr}(B^{\times}(\mathbb{A}))=\mathrm{nr}(B^{\times}(\mathbb{Q})B^{\times}(\mathbb{Q}_{S})H). Then we also have the decomposition B×(𝔸)=B×()B×(S)HB^{\times}(\mathbb{A})=B^{\times}(\mathbb{Q})B^{\times}(\mathbb{Q}_{S})H. In particular 𝐆(𝔸)=𝐆()𝐆(S)H¯\boldsymbol{G}(\mathbb{A})=\boldsymbol{G}(\mathbb{Q})\boldsymbol{G}(\mathbb{Q}_{S})\overline{H} with H¯\overline{H} the image of HH in 𝐆(𝔸(S))\boldsymbol{G}(\mathbb{A}^{(S)}).

Proof.

By the decomposition nr(B×(𝔸))=nr(B×()B×(S)H)\mathrm{nr}(B^{\times}(\mathbb{A}))=\mathrm{nr}(B^{\times}(\mathbb{Q})B^{\times}(\mathbb{Q}_{S})H), given t𝑮(𝔸)t\in\boldsymbol{G}(\mathbb{A}), there exist xB×()x_{\mathbb{Q}}\in B^{\times}(\mathbb{Q}), xSB×(S)x_{S}\in B^{\times}(\mathbb{Q}_{S}) and xHHx_{H}\in H such that x1txH1xS1𝑮~(𝔸)x_{\mathbb{Q}}^{-1}tx_{H}^{-1}x_{S}^{-1}\in\widetilde{\boldsymbol{G}}(\mathbb{A}). We obtain

B×(𝔸)=B×()𝑮~(𝔸)B×(S)HB^{\times}(\mathbb{A})=B^{\times}(\mathbb{Q})\widetilde{\boldsymbol{G}}(\mathbb{A})B^{\times}(\mathbb{Q}_{S})H

which allows us to conclude since

𝑮~(𝔸)=𝑮~()𝑮~(S)(H𝑮~(𝔸(S)))B×()B×(S)H.\widetilde{\boldsymbol{G}}(\mathbb{A})=\widetilde{\boldsymbol{G}}(\mathbb{Q})\widetilde{\boldsymbol{G}}(\mathbb{Q}_{S})(H\cap\widetilde{\boldsymbol{G}}(\mathbb{A}^{(S)}))\subseteq B^{\times}(\mathbb{Q})B^{\times}(\mathbb{Q}_{S})H.

Since 𝑮\boldsymbol{G} has no non-trivial \mathbb{Q}-characters, the subgroup 𝑮()\boldsymbol{G}(\mathbb{Q}) is a lattice in 𝑮(𝔸)\boldsymbol{G}(\mathbb{A}) by Theorem 5.5 in [PR94]. We will use the notation [𝑮][\boldsymbol{G}] for the locally compact space of finite measure 𝑮()\𝑮(𝔸)\mathchoice{\text{\lower 2.15277pt\hbox{$\boldsymbol{G}(\mathbb{Q})$}\big{\backslash}\raise 2.15277pt\hbox{$\boldsymbol{G}(\mathbb{A})$}}}{\boldsymbol{G}(\mathbb{Q})\,\backslash\,\boldsymbol{G}(\mathbb{A})}{\boldsymbol{G}(\mathbb{Q})\,\backslash\,\boldsymbol{G}(\mathbb{A})}{\boldsymbol{G}(\mathbb{Q})\,\backslash\,\boldsymbol{G}(\mathbb{A})}. If 𝑲\boldsymbol{K} is a compact subgroup of 𝑮(𝔸)\boldsymbol{G}(\mathbb{A}), we use [𝑮]𝑲[\boldsymbol{G}]_{\boldsymbol{K}} to denote the locally compact space [𝑮]/𝑲\mathchoice{\text{\raise 2.15277pt\hbox{$[\boldsymbol{G}]$}\big{/}\lower 2.15277pt\hbox{$\boldsymbol{K}$}}}{[\boldsymbol{G}]\,/\,\boldsymbol{K}}{[\boldsymbol{G}]\,/\,\boldsymbol{K}}{[\boldsymbol{G}]\,/\,\boldsymbol{K}}. The symbol []𝑲[\cdot]_{\boldsymbol{K}} denotes the natural projection 𝑮(𝔸)[𝑮]𝑲\boldsymbol{G}(\mathbb{A})\to[\boldsymbol{G}]_{\boldsymbol{K}} so that the class of g𝑮(𝔸)g\in\boldsymbol{G}(\mathbb{A}) in [𝑮]𝑲[\boldsymbol{G}]_{\boldsymbol{K}} is [g]𝑲[g]_{\boldsymbol{K}}.

We denote by G the group 𝑮(S)\boldsymbol{G}(\mathbb{Q}_{S}). Assume that 𝑲\boldsymbol{K} is the projection in 𝑮(𝔸)\boldsymbol{G}(\mathbb{A}) from an open subgroup HB×(𝔸(S))H\subseteq B^{\times}(\mathbb{A}^{(S)}) satisfying the hypothesis of Corollary 3.2 and set Γ=𝑮()𝑲G\Gamma=\boldsymbol{G}(\mathbb{Q})\cap\boldsymbol{K}\subseteq G. By the conclusion of the cited corollary, given g𝑮(𝔸)g\in\boldsymbol{G}(\mathbb{A}) there exists x𝑮()x\in\boldsymbol{G}(\mathbb{Q}) such that xg#𝑲xg_{\#}\in\boldsymbol{K}. Like this we obtain an identification

(3.1) [𝑮]𝑲Γ\G,[\boldsymbol{G}]_{\boldsymbol{K}}\cong\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma$}\big{\backslash}\raise 2.15277pt\hbox{$G$}}}{\Gamma\,\backslash\,G}{\Gamma\,\backslash\,G}{\Gamma\,\backslash\,G},

sending [g]𝑲[g]_{\boldsymbol{K}} to the Γ\Gamma-orbit of xgSxg_{S} with xx as before.

4. The cycles

4.1. Eichler orders

Let pp and BB be as in the previous section and consider NN^{-} as the product of the primes where BB ramifies and N+N^{+} as a positive integer coprime to NpN^{-}p. An Eichler [1/p]\mathbb{Z}[1/p]-order RR in BB is the intersection of two uniquely determined, not necessarily distinct, maximal [1/p]\mathbb{Z}[1/p]-orders in BB. The level of RR is the index of RR inside any of the two maximal orders that define it. By a local global-principle (Theorem 9.1.1 in [Voi21]) RR is determined by its local components RqR_{q} for qpq\neq p so we focus on these local orders. Being also the intersection of two uniquely determined, not necessarily distinct, maximal q\mathbb{Z}_{q}-orders in BqB_{q} they are Eichler orders in BqB_{q} and its level is defined analogously.

If qNq\nmid N^{-}, BB splits at qq and we can use an identification BqM2(q)B_{q}\cong M_{2}(\mathbb{Q}_{q}) and the tree 𝒯q\mathcal{T}_{q} to understand these type of orders. Indeed, maximal orders in M2(q)M_{2}(\mathbb{Q}_{q}) are in bijection with 𝒱(𝒯q)\mathcal{V}(\mathcal{T}_{q}) by sending the class of a lattice Λ\Lambda to the order Endq(Λ)\colonequals{xM2(q)xΛΛ}\mathrm{End}_{\mathbb{Z}_{q}}(\Lambda)\colonequals\{x\in M_{2}(\mathbb{Q}_{q})\mid x\Lambda\subseteq\Lambda\}. Consequently, the Eichler order RqR_{q} can be identified with a pair of vertices in 𝒯q\mathcal{T}_{q}. If RqR_{q} has level qnq^{n}, then the two corresponding vertices are at distance nn. In this fashion, we can identify Eichler orders of level qnq^{n} in BqB_{q} with paths in 𝒯q\mathcal{T}_{q} of length nn. We denote by ηRq\eta_{R_{q}} the path that corresponds to RqR_{q}. Once given an orientation to this path (select a source s(ηRq)s(\eta_{R_{q}}) and target t(ηRq)t(\eta_{R_{q}})), the stabilizer in PGL2(q)PBq×\mathrm{PGL}_{2}(\mathbb{Q}_{q})\cong PB_{q}^{\times} of this oriented path is R¯q×\overline{R}^{\times}_{q} (see [Voi21] chapter 23). Since an Eichler order of level 11 (resp. qq) correspond a vertex (resp. an edge), we denote in those special cases the associated path as vRqv_{R_{q}} and eRqe_{R_{q}}, respectively.

Remark 1.

All of this shows that the set of maximal [1/p]\mathbb{Z}[1/p]-orders in BqB_{q} (qNpq\nmid N^{-}p) forms the set of vertices of a tree which we also denote by 𝒯q\mathcal{T}_{q} by an abuse of notation. The action of the group PBq×=Bq×/q×PB_{q}^{\times}=B_{q}^{\times}/\mathbb{Q}_{q}^{\times} on 𝒯q\mathcal{T}_{q} corresponds to conjugation.

4.2. Oriented optimal embeddings

Fix an Eichler [1/p]\mathbb{Z}[1/p]-order RR in BB. Let KK be a quadratic field and 𝒪\mathcal{O} a [1/p]\mathbb{Z}[1/p]-order in KK. There exists a unique integer dd with p2dp^{2}\nmid d such that 𝒪=𝒪d[1/p]\mathcal{O}=\mathcal{O}_{d}[1/p], where 𝒪d\colonequals[d+d2]\mathcal{O}_{d}\colonequals\mathbb{Z}\left[\frac{d+\sqrt{d}}{2}\right], the \mathbb{Z}-order of discriminant dd. We assume that (d,N)=1(d,N^{-})=1 and if \ell is a prime with 2N+\ell^{2}\mid N^{+}, then (d,)=1(d,\ell)=1. An algebra embedding ψ:KB\psi\colon K\to B is said to be optimal (with respect to 𝒪\mathcal{O} and RR) if ψ(K)R=ψ(𝒪)\psi(K)\cap R=\psi(\mathcal{O}). Necessarily KK is inert at primes diving NN^{-}, split at primes dividing N+N^{+} and imaginary in the case that BB is definite.

For \ell a prime, note that ψ\psi induces a morphism (which we denote by the same letter) ψ:KB\psi\colon K_{\ell}\to B_{\ell} by tensoring with \mathbb{Q}_{\ell}. For N\ell\nmid N^{-} the map ψ\psi induces an action of K×K_{\ell}^{\times} on 𝒯\mathcal{T}_{\ell} through composition with the natural map B×PB×B_{\ell}^{\times}\to PB_{\ell}^{\times}. If K¯×\overline{K}_{\ell}^{\times} denotes K×/×K^{\times}_{\ell}/\mathbb{Q}_{\ell}^{\times}, there is also an induced action of K¯×\overline{K}_{\ell}^{\times}. If N+\ell\mid N^{+}, since ψ(K)\psi(K_{\ell}) is a split algebra over \mathbb{Q}_{\ell}, by Corollary 2.2 (1), the group K×K_{\ell}^{\times} preserves a geodesic in 𝒯\mathcal{T}_{\ell} that will be denoted by 𝒢ψ\mathcal{G}_{\psi}. In addition, the optimality condition ensures that the stabilizer of ηR\eta_{R_{\ell}} in K×K_{\ell}^{\times} is 𝒪××\mathcal{O}_{\ell}^{\times}\mathbb{Q}_{\ell}^{\times}.

An orientation of RR at N\ell\mid N^{-} is a ring morphism ν:R𝔽2\nu_{\ell}\colon R_{\ell}\to\mathbb{F}_{\ell^{2}}. There exist two of them and we fix one. An orientation of RR at N+\ell\mid N^{+} is the choice of a maximal order SS_{\ell} in BB_{\ell} containing RR_{\ell}. By the first paragraph, this is equivalent to give an orientation to ηR\eta_{R_{\ell}}. There are also two possibilities and we fix one. We orient ηR\eta_{R_{\ell}} by declaring that s(ηR)=vSs(\eta_{R_{\ell}})=v_{S_{\ell}}.

For N\ell\mid N^{-}, an orientation of 𝒪\mathcal{O} at \ell is a ring morphism μ:𝒪𝔽2\mu_{\ell}\colon\mathcal{O}\to\mathbb{F}_{\ell^{2}}. There are two of them an we fix one. Let ψ:KB\psi\colon K\to B an optimal embedding as before. We say that ψ\psi is oriented at N\ell\mid N^{-} if it respects the orientations of 𝒪\mathcal{O} and RR at \ell, i.e. if the diagram

𝒪{\mathcal{O}_{\ell}}R{R_{\ell}}𝔽2{\mathbb{F}_{\ell^{2}}}ψ\scriptstyle{\psi}μ\scriptstyle{\mu_{\ell}}ν\scriptstyle{\nu_{\ell}}

is commutative.

Take N+\ell\mid N^{+} and assume that d\ell\nmid d. An orientation of 𝒪\mathcal{O} at \ell is a group isomorphism μ:K×/𝒪××\mu_{\ell}\colon K_{\ell}^{\times}/\mathcal{O}_{\ell}^{\times}\mathbb{Q}_{\ell}^{\times}\to\mathbb{Z}. There are two of them and we fix one. Since d\ell\nmid d, 𝒪\mathcal{O}_{\ell} is maximal and by Corollary 2.2 (1), the path ηR\eta_{R_{\ell}} is contained in 𝒢ψ\mathcal{G}_{\psi}. The transformation Tψ\colonequalsψ(μ1(1))T_{\psi}\colonequals\psi(\mu_{\ell}^{-1}(1)) acts transitively on 𝒢ψ\mathcal{G}_{\psi} as a translation. We say that ψ\psi is oriented if Tψn(vS)T_{\psi}^{n}(v_{S_{\ell}}) and vSv_{S_{\ell}} are the two vertices connected by ηR\eta_{R_{\ell}}.

Keep the assumption that N+\ell\mid N^{+} but assume now that d\ell\mid d. Since KK splits at \ell, the prime \ell does not divides the square-free part of dd so 2nd\ell^{2n}\mid\mid d for some n1n\geq 1. Remember that by our initial asumptions this implies that N+\ell\mid\mid N^{+} so ηR=eR\eta_{R_{\ell}}=e_{R_{\ell}} is an edge with stabilizer in K×K_{\ell}^{\times} given by 𝒪××=𝒪K(n)×\mathcal{O}_{\ell}^{\times}\mathbb{Q}_{\ell}^{\times}=\mathcal{O}_{K_{\ell}}^{(n)}\mathbb{Q}_{\ell}^{\times}. Since n1n\geq 1, the edge eRe_{R_{\ell}} is not contained in 𝒢ψ\mathcal{G}_{\psi}. Its furthest vertex from 𝒢ψ\mathcal{G}_{\psi} is located at distance nn and by consequence the other extreme is located at distance n1n-1. We will say that ψ\psi is oriented if the furthest vertex is vSv_{S_{\ell}}.

Remark 2.

In simple words, for N+\ell\mid N^{+} bud d\ell\nmid d, ψ\psi is oriented if ηR\eta_{R_{\ell}} is pointing in the same direction as the flow defined by the transformation TψT_{\psi} in 𝒢ψ\mathcal{G}_{\psi}. On the other hand, if d\ell\mid d, being oriented is equivalent to say that eRe_{R_{\ell}} is pointing towards 𝒢ψ\mathcal{G}_{\psi}.

Lemma 4.1.

The group R¯×\overline{R}^{\times} respects orientations when acting by conjugation on optimal embeddings.

Proof.

Let ψ\psi be an oriented optimal embedding. For xR¯×x\in\overline{R}^{\times}, clearly xψx1x\psi x^{-1} still satisfies the optimality condition so we focus on the orientations. For N\ell\mid N^{-}, we show that xPB×x\in PB^{\times}_{\ell} respects the orientations if and only if xR¯×x\in\overline{R}_{\ell}^{\times}. There exits some jBj\in B_{\ell} such that B=ψ(K)+jψ(K)B_{\ell}=\psi(K_{\ell})+j\psi(K_{\ell}) with j2=pj^{2}=p and αj=jα\alpha j=j\alpha^{\prime} for every αψ(K)\alpha\in\psi(K_{\ell}), where we denote by α\alpha^{\prime} the conjugate of α\alpha in ψ(K)\psi_{\ell}(K_{\ell}). Since (d,N)=1(d,N^{-})=1, 𝒪\mathcal{O}_{\ell} is maximal in KK_{\ell} and therefore R=ψ(𝒪)+jψ(𝒪)R_{\ell}=\psi(\mathcal{O}_{\ell})+j\psi(\mathcal{O}_{\ell}) with maximal ideal 𝔪=ψ(p𝒪)+jψ(𝒪)\mathfrak{m}=\psi(p\mathcal{O}_{\ell})+j\psi(\mathcal{O}_{\ell}) (see [Voi21] Theorem 13.3.11). Write xx in the form x=α+jβx=\alpha+j\beta with α\alpha and β\beta in ψ(K)\psi(K_{\ell}). Then x¯=α+jβ\overline{x}=\alpha^{\prime}+j\beta and nr(x)=N(α)pN(β)\mathrm{nr}(x)=N(\alpha)-pN(\beta) with N()=Nψ(K)/()N(\cdot)=N_{\psi(K_{\ell})/\mathbb{Q}_{\ell}}(\cdot). This allows us to compute for aψ(𝒪)a\in\psi(\mathcal{O}_{\ell}) the equality

xax1=1N(α)pN(β)(N(α)apN(β)a+jαβ(aa)).xax^{-1}=\frac{1}{N(\alpha)-pN(\beta)}(N(\alpha)a-pN(\beta)a^{\prime}+j\alpha^{\prime}\beta(a-a^{\prime})).

From this we see that xax1xax^{-1} is congruent to aa mod 𝔪\mathfrak{m}, if αψ(𝒪)×\alpha\in\psi(\mathcal{O}_{\ell})^{\times} and to aa^{\prime} if not. The fact that αψ(𝒪)×\alpha\in\psi(\mathcal{O}_{\ell})^{\times} occurs if and only if xx belongs to R¯×\overline{R}_{\ell}^{\times} allows us to conclude in this case.

For N+\ell\mid N^{+}, we show that xR¯×x\in\overline{R}_{\ell}^{\times} respects orientation. Note that xSx1=Sx{S_{\ell}}x^{-1}={S_{\ell}} since R¯×S¯×\overline{R}_{\ell}^{\times}\subseteq\overline{S}_{\ell}^{\times} so that xvS=vSxv_{S_{\ell}}=v_{S_{\ell}} and xηR=ηRx\eta_{R_{\ell}}=\eta_{R_{\ell}}. Also note that 𝒢xψx1=x𝒢ψ\mathcal{G}_{x\psi x^{-1}}=x\mathcal{G}_{\psi}. Since PB×PB_{\ell}^{\times} preserves distances, these equalities are enough to show that xψx1x\psi x^{-1} is oriented noting that when d\ell\nmid d, Txψx1=xTψx1T_{x\psi x^{-1}}=xT_{\psi}x^{-1}. ∎

4.3. The action of Pic(𝒪)Pic(\mathcal{O}).

Keep the notations from the previous sections. If qNq\mid N^{-}, there exists a unique maximal order in BqB_{q} and therefore RqR_{q} is maximal. Now theorem 13.3.11 in [Voi21] implies that nr(Rq×)=q×\mathrm{nr}(R_{q}^{\times})=\mathbb{Z}_{q}^{\times} since it contains a copy of the ring of integers of the quadratic unramified extension of q\mathbb{Q}_{q}. Let qNpq\nmid N^{-}p, then BB splits at qq and qnq^{n} is the level of RqR_{q} if and only if qnN+q^{n}\mid\mid N^{+}. From Proposition 23.4.3 in [Voi21], after choosing an isomorphism of BqM2(q)B_{q}\cong M_{2}(\mathbb{Q}_{q}), RqR_{q} is conjugate to the order (qqqnqq)\begin{pmatrix}\mathbb{Z}_{q}&\mathbb{Z}_{q}\\ q^{n}\mathbb{Z}_{q}&\mathbb{Z}_{q}\end{pmatrix} so we also have nr(Rq×)=q×\mathrm{nr}(R_{q}^{\times})=\mathbb{Z}_{q}^{\times} in this case.

From now on, SS will denote the finite set of places {,p}\{\infty,p\} if BB is indefinite and the singleton {p}\{p\} otherwise. Let 𝑲\boldsymbol{K} to be the image in 𝑮(𝔸)\boldsymbol{G}(\mathbb{A}) of R^#×\widehat{R}_{\#}^{\times} if BB is indefinite or the image of R^#×B×\widehat{R}_{\#}^{\times}B_{\infty}^{\times} otherwise. Since nr(Rq×)=q×\mathrm{nr}(R_{q}^{\times})={\mathbb{Z}}_{q}^{\times} for qpq\neq p, the open subgroup 𝑲\boldsymbol{K} satisfies the hypothesis in Corollary 3.2. Indeed, we have the equalities nr(B×(𝔸))=𝔸×\mathrm{nr}(B^{\times}(\mathbb{A}))=\mathbb{A}^{\times} or nr(B×(𝔸))=𝔸f××>0\mathrm{nr}(B^{\times}(\mathbb{A}))=\mathbb{A}_{f}^{\times}\times\mathbb{R}_{>0} in the indefinite or definite case respectively (Lemma 13.4.9 in [Voi21]). Finally, the Hasse-Schilling theorem (Theorem 14.7.4 in [Voi21]) and the decomposition 𝔸×=×^××\mathbb{A}^{\times}=\mathbb{Q}^{\times}\widehat{\mathbb{Z}}^{\times}\mathbb{R}^{\times} implies the claim.

Therefore, the identification (3.1) holds with Γ=𝑮()𝑲=R¯×\Gamma=\boldsymbol{G}(\mathbb{Q})\cap\boldsymbol{K}=\overline{R}^{\times} and we have

(4.1) [𝑮]𝑲Γ\𝑮(S).[\boldsymbol{G}]_{\boldsymbol{K}}\cong\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma$}\big{\backslash}\raise 2.15277pt\hbox{$\boldsymbol{G}(\mathbb{Q}_{S})$}}}{\Gamma\,\backslash\,\boldsymbol{G}(\mathbb{Q}_{S})}{\Gamma\,\backslash\,\boldsymbol{G}(\mathbb{Q}_{S})}{\Gamma\,\backslash\,\boldsymbol{G}(\mathbb{Q}_{S})}.

We recall that the identification is given by sending [g]𝑲[g]_{\boldsymbol{K}} to the Γ\Gamma-class of xgSxg_{S}, where x𝑮()x\in\boldsymbol{G}(\mathbb{Q}) is such that xg#𝑲xg_{\#}\in\boldsymbol{K}.

Let 𝑻K\boldsymbol{T}_{K} be the algebraic torus resK/𝔾m/𝔾m\mathrm{res}_{K/\mathbb{Q}}\mathbb{G}_{m}/\mathbb{G}_{m}. By the local-global principle for lattices ([Voi21] Theorem 9.1.1) and using that pp is invertible in 𝒪\mathcal{O}, we can identify Pic(𝒪)Pic(\mathcal{O}) as

𝑻K()\𝑻K(𝔸f)/ ​​𝒪^×=𝑻K()\𝑻K(𝔸#)/ ​​𝒪^#×=𝑻K()\𝑻K(𝔸)/ ​​𝒪^#×𝑻K(S).\mathchoice{\text{\lower 2.15277pt\hbox{$\boldsymbol{T}_{K}(\mathbb{Q})$}\big{\backslash}\raise 2.15277pt\hbox{$\boldsymbol{T}_{K}(\mathbb{A}_{f})$\!}\big{/} \lower 2.15277pt\hbox{\!\!$\widehat{\mathcal{O}}^{\times}$}}}{\boldsymbol{T}_{K}(\mathbb{Q})\,\backslash\,\boldsymbol{T}_{K}(\mathbb{A}_{f})\,/\,\widehat{\mathcal{O}}^{\times}}{\boldsymbol{T}_{K}(\mathbb{Q})\,\backslash\,\boldsymbol{T}_{K}(\mathbb{A}_{f})\,/\,\widehat{\mathcal{O}}^{\times}}{\boldsymbol{T}_{K}(\mathbb{Q})\,\backslash\,\boldsymbol{T}_{K}(\mathbb{A}_{f})\,/\,\widehat{\mathcal{O}}^{\times}}=\mathchoice{\text{\lower 2.15277pt\hbox{$\boldsymbol{T}_{K}(\mathbb{Q})$}\big{\backslash}\raise 2.15277pt\hbox{$\boldsymbol{T}_{K}(\mathbb{A}_{\#})$\!}\big{/} \lower 2.15277pt\hbox{\!\!$\widehat{\mathcal{O}}^{\times}_{\#}$}}}{\boldsymbol{T}_{K}(\mathbb{Q})\,\backslash\,\boldsymbol{T}_{K}(\mathbb{A}_{\#})\,/\,\widehat{\mathcal{O}}^{\times}_{\#}}{\boldsymbol{T}_{K}(\mathbb{Q})\,\backslash\,\boldsymbol{T}_{K}(\mathbb{A}_{\#})\,/\,\widehat{\mathcal{O}}^{\times}_{\#}}{\boldsymbol{T}_{K}(\mathbb{Q})\,\backslash\,\boldsymbol{T}_{K}(\mathbb{A}_{\#})\,/\,\widehat{\mathcal{O}}^{\times}_{\#}}=\mathchoice{\text{\lower 2.15277pt\hbox{$\boldsymbol{T}_{K}(\mathbb{Q})$}\big{\backslash}\raise 2.15277pt\hbox{$\boldsymbol{T}_{K}(\mathbb{A})$\!}\big{/} \lower 2.15277pt\hbox{\!\!$\widehat{\mathcal{O}}^{\times}_{\#}\boldsymbol{T}_{K}(\mathbb{Q}_{S})$}}}{\boldsymbol{T}_{K}(\mathbb{Q})\,\backslash\,\boldsymbol{T}_{K}(\mathbb{A})\,/\,\widehat{\mathcal{O}}^{\times}_{\#}\boldsymbol{T}_{K}(\mathbb{Q}_{S})}{\boldsymbol{T}_{K}(\mathbb{Q})\,\backslash\,\boldsymbol{T}_{K}(\mathbb{A})\,/\,\widehat{\mathcal{O}}^{\times}_{\#}\boldsymbol{T}_{K}(\mathbb{Q}_{S})}{\boldsymbol{T}_{K}(\mathbb{Q})\,\backslash\,\boldsymbol{T}_{K}(\mathbb{A})\,/\,\widehat{\mathcal{O}}^{\times}_{\#}\boldsymbol{T}_{K}(\mathbb{Q}_{S})}.

We denote by opt(𝒪,R)\mathrm{opt}(\mathcal{O},R) the set of all oriented optimal embeddings KBK\to B and by [opt(𝒪,R)][\mathrm{opt}(\mathcal{O},R)] we mean the set of R¯×\overline{R}^{\times}-conjugacy classes in opt(𝒪,R)\mathrm{opt}(\mathcal{O},R). The class of an embedding ψ\psi is denoted by [ψ][\psi]. Now we define an action of Pic(𝒪)Pic(\mathcal{O}) on [opt(𝒪,R)][\mathrm{opt}(\mathcal{O},R)]. Let 𝔞Pic(𝒪)\mathfrak{a}\in Pic(\mathcal{O}) and assume that it is related to the class of t𝑻K(𝔸)t\in\boldsymbol{T}_{K}(\mathbb{A}). The embedding ψ\psi induces an adelic map K𝔸B𝔸K\otimes\mathbb{A}\to B\otimes\mathbb{A} which in turns induces an injection

𝑻K()\𝑻K(𝔸)/ ​​𝒪^#×[𝑮]𝑲=[G].\mathchoice{\text{\lower 2.15277pt\hbox{$\boldsymbol{T}_{K}(\mathbb{Q})$}\big{\backslash}\raise 2.15277pt\hbox{$\boldsymbol{T}_{K}(\mathbb{A})$\!}\big{/} \lower 2.15277pt\hbox{\!\!$\widehat{\mathcal{O}}^{\times}_{\#}$}}}{\boldsymbol{T}_{K}(\mathbb{Q})\,\backslash\,\boldsymbol{T}_{K}(\mathbb{A})\,/\,\widehat{\mathcal{O}}^{\times}_{\#}}{\boldsymbol{T}_{K}(\mathbb{Q})\,\backslash\,\boldsymbol{T}_{K}(\mathbb{A})\,/\,\widehat{\mathcal{O}}^{\times}_{\#}}{\boldsymbol{T}_{K}(\mathbb{Q})\,\backslash\,\boldsymbol{T}_{K}(\mathbb{A})\,/\,\widehat{\mathcal{O}}^{\times}_{\#}}\to[\boldsymbol{G}]_{\boldsymbol{K}}=[G].

Remembering the identification (4.1), the class of t𝑻K(𝔸)t\in\boldsymbol{T}_{K}(\mathbb{A}) is sent to the Γ\Gamma-class of xψ(tS)x\psi(t_{S}), where x𝑮()x\in\boldsymbol{G}(\mathbb{Q}) is such that xψ(t#)𝑲x\psi(t_{\#})\in\boldsymbol{K}. We define the action of 𝔞\mathfrak{a} on [ψ][\psi] as 𝔞[ψ]\colonequals[xψx1]\mathfrak{a}\star[\psi]\colonequals[x\psi x^{-1}].

Lemma 4.2.

The action of 𝔞\mathfrak{a} on the class of ψ\psi is well defined and gives rise to an action of Pic(𝒪)Pic(\mathcal{O}) on [opt(𝒪,R)][\mathrm{opt}(\mathcal{O},R)]

Proof.

It is well defined because the class of ψ(t#)\psi(t_{\#}) depends only on the class of 𝔞\mathfrak{a} and on the other hand conjugating ψ\psi by an element of γΓ\gamma\in\Gamma changes xx by xγ1x\gamma^{-1} since γ𝑲\gamma\in\boldsymbol{K}. After conjugating γψγ1\gamma\psi\gamma^{-1} by xγ1γx\gamma^{-1}\gamma we obtain again xψx1x\psi x^{-1}.

Since x𝑮()x\in\boldsymbol{G}(\mathbb{Q}), xψx1x\psi x^{-1} is indeed a new embedding that send KB()K\to B(\mathbb{Q}). For the optimality it is enough to show that for qpq\neq p one has xψ(Kq)x1Rq=xψ(𝒪q)x1x\psi(K_{q})x^{-1}\cap R_{q}=x\psi(\mathcal{O}_{q})x^{-1}. We know ψ(Kq)Rq=ψ(𝒪q)\psi(K_{q})\cap R_{q}=\psi(\mathcal{O}_{q}) but since im(ψ)\mathrm{im}(\psi) is commutative, after conjugating by xψ(tq)x\psi(t_{q}) we have xψ(Kq)x1(xψ(tq))Rq(xψ(tq))1=xψ(𝒪q)x1x\psi(K_{q})x^{-1}\cap(x\psi(t_{q}))R_{q}(x\psi(t_{q}))^{-1}=x\psi(\mathcal{O}_{q})x^{-1} and now we use (xψ(tq))Rq(xψ(tq))1=Rq(x\psi(t_{q}))R_{q}(x\psi(t_{q}))^{-1}=R_{q} since xψ(tq)R¯q×x\psi(t_{q})\in\overline{R}_{q}^{\times}.

To see that xψx1x\psi x^{-1} is oriented we have to check locally at NN+\ell\mid N^{-}N^{+}. If N\ell\mid N^{-}, ψ(K¯)×R¯×\psi(\overline{K}_{\ell})^{\times}\subseteq\overline{R}_{\ell}^{\times} so xR¯×x\in\overline{R}_{\ell}^{\times} and therefore xψx1x\psi x^{-1} is oriented. If N+\ell\mid N^{+}, observe that ψ(tq)𝒢ψ=𝒢ψ\psi(t_{q})\mathcal{G}_{\psi}=\mathcal{G}_{\psi} by defintion of 𝒢ψ\mathcal{G}_{\psi}. Also, we have xψ(t)ηR=ηRx\psi(t_{\ell})\eta_{R_{\ell}}=\eta_{R_{\ell}} since xψ(t)R¯×x\psi(t_{\ell})\in\overline{R}_{\ell}^{\times}. Then, since xψ(t)x\psi(t_{\ell}) preserves distances, ηR\eta_{R_{\ell}} is at the same distance to x𝒢ψ=𝒢xψx1x\mathcal{G}_{\psi}=\mathcal{G}_{x\psi x^{-1}}. Since xψ(t)x\psi(t_{\ell}) preserves the orientation of ηR\eta_{R_{\ell}}, it follows that xψx1x\psi x^{-1} is oriented noting in the case d\ell\nmid d that ψ(t)Tψψ(t)1=Tψ\psi(t_{\ell})T_{\psi}\psi(t_{\ell})^{-1}=T_{\psi}.

Suppose that 𝔞\mathfrak{a} and 𝔟\mathfrak{b} are related to the adelic elements tt and ss. Consider x,y𝑮()x,y\in\boldsymbol{G}(\mathbb{Q}) such that xψ(s#),yxψ(t#)x1𝑲x\psi(s_{\#}),yx\psi(t_{\#})x^{-1}\in\boldsymbol{K}. The equality yxψ(t#s#)=yxψ(t#)x1xψ(s#)𝑲yx\psi(t_{\#}s_{\#})=yx\psi(t_{\#})x^{-1}x\psi(s_{\#})\in\boldsymbol{K} shows that 𝔞(𝔟[ψ])=𝔞𝔟[ψ]\mathfrak{a}\star(\mathfrak{b}\star[\psi])=\mathfrak{a}\mathfrak{b}\star[\psi].

Proposition 4.3.

The group Pic(𝒪)Pic(\mathcal{O}) acts simply transitively on [opt(𝒪,R)][\mathrm{opt}(\mathcal{O},R)].

Proof.

To show that the action is simple take γΓ\gamma\in\Gamma such that γ1xψ(γ1x)1=ψ\gamma^{-1}x\psi(\gamma^{-1}x)^{-1}=\psi. In particular γ1x\gamma^{-1}x belongs to the normalizer of ψ(K)×\psi(K)^{\times} which is ψ(K)×sψ(K)×\psi(K)^{\times}\cup s\psi(K)^{\times} with any sB×()s\in B^{\times}(\mathbb{Q}) such that sψs1=ψσs\psi s^{-1}=\psi\circ\sigma, where σ\sigma is the non-trivial \mathbb{Q}-automorphism of KK. Since γ1x\gamma^{-1}x normalized each value of ψ\psi we must have γ1xψ(K)×\gamma^{-1}x\in\psi(K)^{\times}. See Corollaire 2.3, p.6 in [Vig80].This implies x=γψ(t)x=\gamma\psi(t) with tK×t\in K^{\times} and therefore ψ(tt#)γ𝑲=𝑲\psi(tt_{\#})\in\gamma\boldsymbol{K}=\boldsymbol{K}. This shows that 𝔞\mathfrak{a} is trivial in Pic(𝒪)Pic(\mathcal{O}) and therefore the action is simple.

Now we show the transitivity. Let ψ\psi and ψ\psi^{\prime} be two optimal embeddings for 𝒪\mathcal{O}. As a consequence of Skolem-Noether there exists x𝑮()x\in\boldsymbol{G}(\mathbb{Q}) such that ψ=xψx1\psi^{\prime}=x\psi x^{-1}. To prove transitivity we need to show that there exists an adelic element tt such that xψ(t#)𝑲x\psi(t_{\#})\in\boldsymbol{K}. Since both ψ\psi and ψ\psi^{\prime} are optimal for 𝒪\mathcal{O}, we have for qpq\neq p

(4.2) ψ(𝒪q)=ψ(Kq)Rq=ψ(Kq)x1Rqx.\psi(\mathcal{O}_{q})=\psi(K_{q})\cap R_{q}=\psi(K_{q})\cap x^{-1}R_{q}x.

If qNq\mid N^{-}, we have ψ(Kq)×R¯q×\psi(K_{q})^{\times}\subseteq\overline{R}_{q}^{\times}. Then the condition xψ(tq)R¯q×x\psi(t_{q})\in\overline{R}_{q}^{\times} is equivalent to xR¯q×x\in\overline{R}_{q}^{\times}. Since ψ=xψx1\psi^{\prime}=x\psi x^{-1} and ψ\psi and ψ\psi^{\prime} are oriented, we have indeed that xR¯q×x\in\overline{R}_{q}^{\times}.

If qNq\nmid N^{-} remember that ψ\psi induces an action of Kq×K_{q}^{\times} on 𝒯q\mathcal{T}_{q} and in this case ψ(Kq)\psi(K_{q}) is a split algebra over q\mathbb{Q}_{q} so that there exists a unique geodesic 𝒢ψ\mathcal{G}_{\psi} preserved by this action. The relation (4.2) shows that ηRq\eta_{R_{q}} and x1ηRqx^{-1}\eta_{R_{q}} have the same stabilizer in Kq×K_{q}^{\times} under this action. If in addition qN+q\nmid N^{+}, ηRq=vRq\eta_{R_{q}}=v_{R_{q}} and x1ηRq=x1vRqx^{-1}\eta_{R_{q}}=x^{-1}v_{R_{q}} are actually vertices and Corollary 2.2 shows they are at the same distance to 𝒢ψ\mathcal{G}_{\psi} (since same stabilizer implies same distance) and then they are in the same orbit. Therefore there exists some tqKq×t_{q}\in K_{q}^{\times} satisfying xψ(tq)vRq=vRqx\psi(t_{q})v_{R_{q}}=v_{R_{q}}. This implies xψ(tq)R¯q×x\psi(t_{q})\in\overline{R}_{q}^{\times}.

Now assume that qnN+q^{n}\mid\mid N^{+} and qdq\nmid d. The relation (4.2) and the fact that ψ\psi and ψ\psi^{\prime} are oriented implies that ηRq\eta_{R_{q}} and x1ηRqx^{-1}\eta{R_{q}} are contained in 𝒢ψ\mathcal{G}_{\psi} and they point in the same direction as the flow defined by TψT_{\psi}. Therefore there exists some nn\in\mathbb{Z} such that xTψnηRq=ηRqxT_{\psi}^{n}\eta_{R_{q}}=\eta_{R_{q}} as an oriented path. Since Tψn=ψ(μq1(n))T_{\psi}^{n}=\psi(\mu_{q}^{-1}(n)) we have that xψ(tq)R¯q×x\psi(t_{q})\in\overline{R}_{q}^{\times} for any tqK×t_{q}\in K_{\ell}^{\times} lying in μq1(n)\mu_{q}^{-1}(n) when taken mod 𝒪q×q×\mathcal{O}_{q}^{\times}\mathbb{Q}_{q}^{\times}.

Finally if qN+q\mid\mid N^{+} and qdq\mid d, again the condition of being oriented embeddings and relation (4.2) implies that the edges ηRq=eRq\eta_{R_{q}}=e_{R_{q}} and x1ηRq=x1eRqx^{-1}\eta_{R_{q}}=x^{-1}e_{R_{q}} are at the same distance to 𝒢ψ\mathcal{G}_{\psi} and pointing in the same direction relative to 𝒢ψ\mathcal{G}_{\psi}. Corollary 2.2 (1) implies that they are in the same orbit under the action of Kq×K_{q}^{\times} and so there exists tqKq×t_{q}\in K_{q}^{\times} such that xψ(tq)eRq=eRqx\psi(t_{q})e_{R_{q}}=e_{R_{q}} as oriented edges, implying that xψ(tq)R¯q×x\psi(t_{q})\in\overline{R}_{q}^{\times}. ∎

4.4. The cycles Δψ\Delta_{\psi}

Keep the previous notations and fix HH a subgroup of GG that is the image of the units of an algebra in BSB_{S} isomorphic to KSK_{S}. Let ψopt(𝒪,R)\psi\in opt(\mathcal{O},R), we denote by Δψ\Delta_{\psi} the image in Γ\G\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma$}\big{\backslash}\raise 2.15277pt\hbox{$G$}}}{\Gamma\,\backslash\,G}{\Gamma\,\backslash\,G}{\Gamma\,\backslash\,G} of a set of the form ψ(KS)×g\psi(K_{S})^{\times}g with gGg\in G such that ψ(KS)×=gHg1\psi(K_{S})^{\times}=gHg^{-1} (such gg exists by Skolem-Noether). Since HH has index 2#S2^{\#S} in its normalizer in GG, there are only 2#S2^{\#S} such sets and they depend only on the class of ψ\psi in [opt(𝒪,R)][opt(\mathcal{O},R)]. In the following sections we will make use of the action of GG on certain spaces to select a single Δψ\Delta_{\psi} among the 2#S2^{\#S} possibilities.

Let h=h(𝒪)h^{\prime}=h^{\prime}(\mathcal{O}) be the cardinality of Pic(𝒪)Pic(\mathcal{O}) and consider t1,,tht_{1},...,t_{h^{\prime}} ideles in K^#×𝔸K×\widehat{K}^{\times}_{\#}\subseteq\mathbb{A}_{K}^{\times} such that K^#×=i=1hKti1𝒪^#×\widehat{K}^{\times}_{\#}=\bigsqcup_{i=1}^{h^{\prime}}K^{*}t_{i}^{-1}\widehat{\mathcal{O}}^{\times}_{\#}. Note that this is equivalent to ask for a full set of representatives for Pic(𝒪)Pic(\mathcal{O}) under its identification with K×\K^#×/𝒪^#×\mathchoice{\text{\lower 2.15277pt\hbox{$K^{\times}$}\big{\backslash}\raise 2.15277pt\hbox{$\widehat{K}^{\times}_{\#}$\!}\big{/} \lower 2.15277pt\hbox{\!\!$\widehat{\mathcal{O}}_{\#}^{\times}$}}}{K^{\times}\,\backslash\,\widehat{K}^{\times}_{\#}\,/\,\widehat{\mathcal{O}}_{\#}^{\times}}{K^{\times}\,\backslash\,\widehat{K}^{\times}_{\#}\,/\,\widehat{\mathcal{O}}_{\#}^{\times}}{K^{\times}\,\backslash\,\widehat{K}^{\times}_{\#}\,/\,\widehat{\mathcal{O}}_{\#}^{\times}}.

Proposition 4.4.

Fix ψ0opt(𝒪,R)\psi_{0}\in opt(\mathcal{O},R). Then, the projection of 𝐓ψ0g\boldsymbol{T}_{\psi_{0}}g in [𝐆]𝐊=Γ\G[\boldsymbol{G}]_{\boldsymbol{K}}=\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma$}\big{\backslash}\raise 2.15277pt\hbox{$G$}}}{\Gamma\,\backslash\,G}{\Gamma\,\backslash\,G}{\Gamma\,\backslash\,G} is of the form ψ[opt(𝒪,R)]Δψ\bigsqcup_{\psi\in[opt(\mathcal{O},R)]}\Delta_{\psi}.

Proof.

The image of [𝑻K][\boldsymbol{T}_{K}] under ψ0\psi_{0} in [𝑮]𝑲[\boldsymbol{G}]_{\boldsymbol{K}} factors through

K×\𝔸K×/ ​​𝒪^#×S=i=1hK(ti1𝒪^#××(KS)×).\mathchoice{\text{\lower 2.15277pt\hbox{$K^{\times}$}\big{\backslash}\raise 2.15277pt\hbox{$\mathbb{A}_{K}^{\times}$\!}\big{/} \lower 2.15277pt\hbox{\!\!$\widehat{\mathcal{O}}^{\times}_{\#}\mathbb{Q}_{S}^{*}$}}}{K^{\times}\,\backslash\,\mathbb{A}_{K}^{\times}\,/\,\widehat{\mathcal{O}}^{\times}_{\#}\mathbb{Q}_{S}^{*}}{K^{\times}\,\backslash\,\mathbb{A}_{K}^{\times}\,/\,\widehat{\mathcal{O}}^{\times}_{\#}\mathbb{Q}_{S}^{*}}{K^{\times}\,\backslash\,\mathbb{A}_{K}^{\times}\,/\,\widehat{\mathcal{O}}^{\times}_{\#}\mathbb{Q}_{S}^{*}}=\bigsqcup_{i=1}^{h^{\prime}}K^{*}(t_{i}^{-1}\widehat{\mathcal{O}}^{\times}_{\#}\times(K_{S})^{\times}).

The image of ti1𝒪^#××(KS)×t_{i}^{-1}\widehat{\mathcal{O}}^{\times}_{\#}\times(K_{S})^{\times} under ψ0\psi_{0} in [𝑮]𝑲[\boldsymbol{G}]_{\boldsymbol{K}} is ψ0(ti1)𝑲×(ψ0(KS))×\psi_{0}(t_{i}^{-1})\boldsymbol{K}\times(\psi_{0}(K_{S}))^{\times}. Under the identification in (4.1) this is

Γxiψ0(KS)×g=Γxiψ0(KS)×xi1xig,\Gamma x_{i}\psi_{0}(K_{S})^{\times}g=\Gamma x_{i}\psi_{0}(K_{S})^{\times}x_{i}^{-1}x_{i}g,

where xi𝑮()x_{i}\in\boldsymbol{G}(\mathbb{Q}) is such that xiψ0(ti1)𝑲x_{i}\psi_{0}(t_{i}^{-1})\in\boldsymbol{K}. This finishes the proof since Proposition 4.3 shows that the collection xiψ0xi1x_{i}\psi_{0}x_{i}^{-1} runs over a complete set of representatives for [opt(𝒪,R)][opt(\mathcal{O},R)]. ∎

5. Ihara-Shintani Cycles

In this section we specialize to the case B=M2()B=M_{2}(\mathbb{Q}), S={p,}S=\{p,\infty\}, R=M2([1/p])R=M_{2}(\mathbb{Z}[1/p]) and KK\subseteq\mathbb{C} is imaginary and splits at pp. In this case 𝑲=PGL2(^#)\boldsymbol{K}=\mathrm{PGL}_{2}(\widehat{\mathbb{Z}}_{\#}) and Γ=PGL2([1/p])\Gamma=\mathrm{PGL}_{2}(\mathbb{Z}[1/p]) so we have natural identifications

[𝑮]𝑲PSO2()Γ\()×PGL2(p)Γ+\×PGL2(p),[\boldsymbol{G}]_{\boldsymbol{K}\cdot{\mathrm{PSO}_{2}(\mathbb{R})}}\cong\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma$}\big{\backslash}\raise 2.15277pt\hbox{$(\mathbb{C}-\mathbb{R})\times\mathrm{PGL}_{2}(\mathbb{Q}_{p})$}}}{\Gamma\,\backslash\,(\mathbb{C}-\mathbb{R})\times\mathrm{PGL}_{2}(\mathbb{Q}_{p})}{\Gamma\,\backslash\,(\mathbb{C}-\mathbb{R})\times\mathrm{PGL}_{2}(\mathbb{Q}_{p})}{\Gamma\,\backslash\,(\mathbb{C}-\mathbb{R})\times\mathrm{PGL}_{2}(\mathbb{Q}_{p})}\cong\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma^{+}$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{H}\times\mathrm{PGL}_{2}(\mathbb{Q}_{p})$}}}{\Gamma^{+}\,\backslash\,\mathcal{H}\times\mathrm{PGL}_{2}(\mathbb{Q}_{p})}{\Gamma^{+}\,\backslash\,\mathcal{H}\times\mathrm{PGL}_{2}(\mathbb{Q}_{p})}{\Gamma^{+}\,\backslash\,\mathcal{H}\times\mathrm{PGL}_{2}(\mathbb{Q}_{p})},

where Γ+=ΓPGL2+()\Gamma^{+}=\Gamma\cap\mathrm{PGL}_{2}^{+}(\mathbb{R}).

Let 𝔓\mathfrak{P} be a prime above pp in ¯\overline{\mathbb{Z}}, the integral closure of \mathbb{Z} in ¯\overline{\mathbb{Q}}. This defines an embedding ¯p\overline{\mathbb{Q}}\hookrightarrow\mathbb{C}_{p} and in particular KpK\hookrightarrow\mathbb{Q}_{p}.

As in section 4.2, let 𝒪\mathcal{O} denotes a [1/p]\mathbb{Z}[1/p] order in KK. For any ψopt(𝒪,R)\psi\in opt(\mathcal{O},R), the torus ψ(K)×\psi(K_{\infty})^{\times} acts on \mathbb{C}-\mathbb{R} by fractional linear transformations, having two fixed points conjugate to each other. We denote by τ\tau its fixed point in \mathcal{H}\subseteq\mathbb{C}-\mathbb{R}. On the other hand, the group ψ(Kp)×\psi(K_{p})^{\times} has two fixed points in 1(p)\mathbb{P}^{1}(\mathbb{Q}_{p}) (when identified with 𝒯p\partial\mathcal{T}_{p}, these are the two end points of the geodesic preserved by Corollary 2.2) which are also τ\tau and its conjugate, since ψ(K)×\psi(K)^{\times} fixes them.

Let AA denote the diagonal group of PGL2\mathrm{PGL}_{2} and consider H=PSO2()×A(p)H=\mathrm{PSO}_{2}(\mathbb{R})\times A(\mathbb{Q}_{p}). Let gGg\in G be such that ψ(KS)×=gHg1\psi(K_{S})^{\times}=gHg^{-1} (it exists by Skolem-Noether). It is always the case that both gig_{\infty}\cdot i and gp0g_{p}\cdot 0 assume the value τ\tau or τ\tau^{\prime}. If we multiply gg by some element in NG(H)N_{G}(H), the normalizer of HH in GG, we cover all 4=[NG(H):H]4=[N_{G}(H):H] possibilities. Then we normalize the choice of Δψ\Delta_{\psi} in Γ\G\Gamma\backslash G by requiring that gi=τg_{\infty}\cdot i=\tau and gp0=τg_{p}\cdot 0=\tau.

For τ\tau in \mathcal{H}^{\prime}, we take KτK_{\tau} to be the \mathbb{Q}-algebra of matrices gM2()g\in M_{2}(\mathbb{Q}) such that gτ=τg\cdot\tau=\tau, together with the zero matrix. The map sending g=(ABCD)g=\begin{pmatrix}A&B\\ C&D\end{pmatrix} to its eigenvalue Cτ+DC\tau+D (with eigenvector (τ   1)t(\tau\,\,\,1)^{t}) is an isomorphism onto (τ)\mathbb{Q}(\tau). Denote its inverse as ψτ\psi_{\tau} and let IS(𝒪)IS(\mathcal{O}) be as in the Introduction. If Γ+τ\Gamma^{+}\tau belongs to IS(𝒪)IS(\mathcal{O}), the map ψτ\psi_{\tau} is an element in opt(𝒪,R)opt(\mathcal{O},R) whose fixed point in \mathcal{H} is τ\tau. We could have also taken the embedding coming from the eigenvalue Cτ+DC\tau^{\prime}+D but this one is Γ\Gamma-conjugate to ψτ\psi_{\tau} since there exists gPGL2()Γg\in\mathrm{PGL}_{2}(\mathbb{Z})\subset\Gamma taking τ\tau to τ\tau^{\prime}.

For τ\tau in {\mathcal{H}^{\prime}}, we define the Ihara-Shintani cycle Δτ\Delta_{\tau} attacthed to Γ+τ\Gamma^{+}\tau to be the projection of Δψτ\Delta_{\psi_{\tau}} to [𝑮]𝑲PSO2()[\boldsymbol{G}]_{\boldsymbol{K}\cdot PSO_{2}(\mathbb{R})}. It corresponds to the Γ+\Gamma^{+} orbit of {τ}×Kτ,p×g\{\tau\}\times K_{\tau,p}^{\times}g in ×PGL2(p)\mathcal{H}\times\mathrm{PGL}_{2}(\mathbb{Q}_{p}). If Γ+τIS(𝒪)\Gamma^{+}\tau\in IS(\mathcal{O}), the map Kp×ΔτK_{p}^{\times}\to\Delta_{\tau} sending xx to the Γ+\Gamma^{+}-class of (τ,ψτ(x)g)(\tau,\psi_{\tau}(x)g) induces a uniformization 𝒪×\Kp×/p×Δτ\mathchoice{\text{\lower 2.15277pt\hbox{$\mathcal{O}^{\times}$}\big{\backslash}\raise 2.15277pt\hbox{$K_{p}^{\times}$\!}\big{/} \lower 2.15277pt\hbox{\!\!$\mathbb{Q}_{p}^{\times}$}}}{\mathcal{O}^{\times}\,\backslash\,K_{p}^{\times}\,/\,\mathbb{Q}_{p}^{\times}}{\mathcal{O}^{\times}\,\backslash\,K_{p}^{\times}\,/\,\mathbb{Q}_{p}^{\times}}{\mathcal{O}^{\times}\,\backslash\,K_{p}^{\times}\,/\,\mathbb{Q}_{p}^{\times}}\cong\Delta_{\tau}. The measure ντ\nu_{\tau} from the introduction is the push-forward of a finite measure in 𝒪×\Kp×/p×\mathchoice{\text{\lower 2.15277pt\hbox{$\mathcal{O}^{\times}$}\big{\backslash}\raise 2.15277pt\hbox{$K_{p}^{\times}$\!}\big{/} \lower 2.15277pt\hbox{\!\!$\mathbb{Q}_{p}^{\times}$}}}{\mathcal{O}^{\times}\,\backslash\,K_{p}^{\times}\,/\,\mathbb{Q}_{p}^{\times}}{\mathcal{O}^{\times}\,\backslash\,K_{p}^{\times}\,/\,\mathbb{Q}_{p}^{\times}}{\mathcal{O}^{\times}\,\backslash\,K_{p}^{\times}\,/\,\mathbb{Q}_{p}^{\times}} coming from a Haar measure in Kp×K_{p}^{\times}.

Let dd be as in section 4.2. Suppose that Γ+τIS(𝒪)\Gamma^{+}\tau\in IS(\mathcal{O}). Since τ\tau\in\mathcal{H}^{\prime} and KKτK\cong K_{\tau} we have that pdp\nmid d and pp splits into two primes inside 𝒪d\mathcal{O}_{d}. Both of these primes are proper so that they define elements in Cl(𝒪d)Cl(\mathcal{O}_{d}). Let kk be the order of 𝔭=𝔓𝒪d\mathfrak{p}=\mathfrak{P}\cap\mathcal{O}_{d} in this group so we can write 𝔭k=u𝒪d\mathfrak{p}^{k}=u\mathcal{O}_{d} for some u𝒪du\in\mathcal{O}_{d}.

Lemma 5.1.

We have 𝒪×=𝒪d×,u,p\mathcal{O}^{\times}=\langle\mathcal{O}_{d}^{\times},u,p\rangle.

Proof.

Note that x𝒪×x\in\mathcal{O}^{\times} if and only if pnx𝒪dp^{n}x\in\mathcal{O}_{d} and pmx𝒪dp^{m}\in x\mathcal{O}_{d} for some n,m0n,m\geq 0. Therefore, after multiplying by a power of pp, we can assume that x𝒪dx\in\mathcal{O}_{d} and x𝒪dx\mathcal{O}_{d} is an ideal dividing pm𝒪dp^{m}\mathcal{O}_{d}, so that it is of the shape 𝔭i𝔭j\mathfrak{p}^{i}\mathfrak{p}^{\prime j} with i,j0i,j\geq 0 by uniqueness of the prime decomposition (principal ideals are proper). Here 𝔭\mathfrak{p}^{\prime} denotes the conjugate of 𝔭\mathfrak{p} under the non-trivial automorphism of KK. If i,jki,j\geq k we can divide xx by a power of uu or uu^{\prime} in order to have 0i,j<k0\leq i,j<k. If i=ji=j we see that xx differs from an element in 𝒪d×\mathcal{O}_{d}^{\times} by a power of pp. If this is not the case, after taking classes in Cl(𝒪d)Cl(\mathcal{O}_{d}) we obtain a contradiction with the minimality of kk. All of this proves that 𝒪×=𝒪d×,u,p\mathcal{O}^{\times}=\langle\mathcal{O}_{d}^{\times},u,p\rangle since uu=pkuu^{\prime}=p^{k}. ∎

Denote the stabilizer of τ\tau in Γ+\Gamma^{+} as Γτ+\Gamma_{\tau}^{+}. Then Γτ+=𝒪τ×\Gamma_{\tau}^{+}=\mathcal{O}_{\tau}^{\times}. Idenfity K𝔭=pK_{\mathfrak{p}}=\mathbb{Q}_{p} and Kpp×pK_{p}\cong\mathbb{Q}_{p}\times\mathbb{Q}_{p} by sending xKx\in K to (x,x)(x,x^{\prime}). When 𝒪d×={±1}\mathcal{O}_{d}^{\times}=\{\pm 1\}, Lemma 5.1 allows us to see that Γτ+\Gamma_{\tau}^{+} is isomorphic to (u/u)(u/u^{\prime})^{\mathbb{Z}} under ψτ\psi_{\tau}. Observe that ordp(u/u)=ord𝔭(u/u)=ord𝔭(u)=k>0\mathrm{ord}_{p}(u/u^{\prime})=\mathrm{ord}_{\mathfrak{p}}(u/u^{\prime})=\mathrm{ord}_{\mathfrak{p}}(u)=k>0.

Proposition 5.2.

For τ\tau\in\mathcal{H}^{\prime}, the image of {τ}×𝒯p\{\tau\}\times\mathcal{T}_{p} in Γ+\(×𝒯p)\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma^{+}$}\big{\backslash}\raise 2.15277pt\hbox{$(\mathcal{H}\times\mathcal{T}_{p})$}}}{\Gamma^{+}\,\backslash\,(\mathcal{H}\times\mathcal{T}_{p})}{\Gamma^{+}\,\backslash\,(\mathcal{H}\times\mathcal{T}_{p})}{\Gamma^{+}\,\backslash\,(\mathcal{H}\times\mathcal{T}_{p})} has the structure of a (p+1)(p+1)-volcano. It can be identified with the pp-isogeny graph of the elliptic curve with ordinary reduction associated with PSL2()τ\mathrm{PSL}_{2}(\mathbb{Z})\tau.

Proof.

Assume Γ+τIS(𝒪)\Gamma^{+}\tau\in IS(\mathcal{O}). The image is Γ+\Γ+({τ}×𝒯p){τ}×𝒪τ×\𝒯p\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma^{+}$}\big{\backslash}\raise 2.15277pt\hbox{$\Gamma^{+}(\{\tau\}\times\mathcal{T}_{p})$}}}{\Gamma^{+}\,\backslash\,\Gamma^{+}(\{\tau\}\times\mathcal{T}_{p})}{\Gamma^{+}\,\backslash\,\Gamma^{+}(\{\tau\}\times\mathcal{T}_{p})}{\Gamma^{+}\,\backslash\,\Gamma^{+}(\{\tau\}\times\mathcal{T}_{p})}\approx\{\tau\}\times\mathchoice{\text{\lower 2.15277pt\hbox{$\mathcal{O}_{\tau}^{\times}$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{T}_{p}$}}}{\mathcal{O}_{\tau}^{\times}\,\backslash\,\mathcal{T}_{p}}{\mathcal{O}_{\tau}^{\times}\,\backslash\,\mathcal{T}_{p}}{\mathcal{O}_{\tau}^{\times}\,\backslash\,\mathcal{T}_{p}} and point (1) in Corollary 2.2 together with the previous paragraph implies that 𝒪τ×\𝒯p\mathchoice{\text{\lower 2.15277pt\hbox{$\mathcal{O}_{\tau}^{\times}$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{T}_{p}$}}}{\mathcal{O}_{\tau}^{\times}\,\backslash\,\mathcal{T}_{p}}{\mathcal{O}_{\tau}^{\times}\,\backslash\,\mathcal{T}_{p}}{\mathcal{O}_{\tau}^{\times}\,\backslash\,\mathcal{T}_{p}} is a (p+1)(p+1)-volcano with a cycle of lenght kk as a rim. This rim is the projection of the geodesic preserved by Kp×K_{p}^{\times} and the levels are given by distance to the geodesic.

For the second claim we work first with the set of vertices. We have an identification between Γ+\(×PGL2(p))\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma^{+}$}\big{\backslash}\raise 2.15277pt\hbox{$(\mathcal{H}\times\mathrm{PGL}_{2}(\mathbb{Q}_{p}))$}}}{\Gamma^{+}\,\backslash\,(\mathcal{H}\times\mathrm{PGL}_{2}(\mathbb{Q}_{p}))}{\Gamma^{+}\,\backslash\,(\mathcal{H}\times\mathrm{PGL}_{2}(\mathbb{Q}_{p}))}{\Gamma^{+}\,\backslash\,(\mathcal{H}\times\mathrm{PGL}_{2}(\mathbb{Q}_{p}))} and PSL2()\(×PGL2(p))\mathchoice{\text{\lower 2.15277pt\hbox{$\mathrm{PSL}_{2}(\mathbb{Z})$}\big{\backslash}\raise 2.15277pt\hbox{$(\mathcal{H}\times\mathrm{PGL}_{2}(\mathbb{Z}_{p}))$}}}{\mathrm{PSL}_{2}(\mathbb{Z})\,\backslash\,(\mathcal{H}\times\mathrm{PGL}_{2}(\mathbb{Z}_{p}))}{\mathrm{PSL}_{2}(\mathbb{Z})\,\backslash\,(\mathcal{H}\times\mathrm{PGL}_{2}(\mathbb{Z}_{p}))}{\mathrm{PSL}_{2}(\mathbb{Z})\,\backslash\,(\mathcal{H}\times\mathrm{PGL}_{2}(\mathbb{Z}_{p}))} by sending Γ+(τ,g)\Gamma^{+}(\tau,g) to PSL2()(γτ,γg)\mathrm{PSL}_{2}(\mathbb{Z})(\gamma\tau,\gamma g), where γΓ+\gamma\in\Gamma^{+} satisfies γgPGL2(p)\gamma g\in\mathrm{PGL}_{2}(\mathbb{Z}_{p}). Therefore we can also identify

(5.1) Γ+\(×𝒱(𝒯p))Γ+\(×PGL2(p))/ ​​PGL2(p)PSL2()\.\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma^{+}$}\big{\backslash}\raise 2.15277pt\hbox{$(\mathcal{H}\times\mathcal{V}(\mathcal{T}_{p}))$}}}{\Gamma^{+}\,\backslash\,(\mathcal{H}\times\mathcal{V}(\mathcal{T}_{p}))}{\Gamma^{+}\,\backslash\,(\mathcal{H}\times\mathcal{V}(\mathcal{T}_{p}))}{\Gamma^{+}\,\backslash\,(\mathcal{H}\times\mathcal{V}(\mathcal{T}_{p}))}\cong\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma^{+}$}\big{\backslash}\raise 2.15277pt\hbox{$(\mathcal{H}\times\mathrm{PGL}_{2}(\mathbb{Q}_{p}))$\!}\big{/} \lower 2.15277pt\hbox{\!\!$\mathrm{PGL}_{2}(\mathbb{Z}_{p})$}}}{\Gamma^{+}\,\backslash\,(\mathcal{H}\times\mathrm{PGL}_{2}(\mathbb{Q}_{p}))\,/\,\mathrm{PGL}_{2}(\mathbb{Z}_{p})}{\Gamma^{+}\,\backslash\,(\mathcal{H}\times\mathrm{PGL}_{2}(\mathbb{Q}_{p}))\,/\,\mathrm{PGL}_{2}(\mathbb{Z}_{p})}{\Gamma^{+}\,\backslash\,(\mathcal{H}\times\mathrm{PGL}_{2}(\mathbb{Q}_{p}))\,/\,\mathrm{PGL}_{2}(\mathbb{Z}_{p})}\cong\mathchoice{\text{\lower 2.15277pt\hbox{$\mathrm{PSL}_{2}(\mathbb{Z})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{H}$}}}{\mathrm{PSL}_{2}(\mathbb{Z})\,\backslash\,\mathcal{H}}{\mathrm{PSL}_{2}(\mathbb{Z})\,\backslash\,\mathcal{H}}{\mathrm{PSL}_{2}(\mathbb{Z})\,\backslash\,\mathcal{H}}.

Consider Γ+(τ,v)Γ+\(×𝒱(𝒯p))\Gamma^{+}(\tau,v)\in\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma^{+}$}\big{\backslash}\raise 2.15277pt\hbox{$(\mathcal{H}^{\prime}\times\mathcal{V}(\mathcal{T}_{p}))$}}}{\Gamma^{+}\,\backslash\,(\mathcal{H}^{\prime}\times\mathcal{V}(\mathcal{T}_{p}))}{\Gamma^{+}\,\backslash\,(\mathcal{H}^{\prime}\times\mathcal{V}(\mathcal{T}_{p}))}{\Gamma^{+}\,\backslash\,(\mathcal{H}^{\prime}\times\mathcal{V}(\mathcal{T}_{p}))}. There exist gPGL2(p)g\in\mathrm{PGL}_{2}(\mathbb{Q}_{p}) such that v=gv0v=gv_{0}. Under the previous identifications this orbit corresponds to the elliptic curve associated with PSL2()γτ\mathrm{PSL}_{2}(\mathbb{Z})\gamma\tau, where γgPGL2(p)\gamma g\in\mathrm{PGL}_{2}(\mathbb{Z}_{p}). Certainly this elliptic curve is pmp^{m}-isogenous to the elliptic curve attached to PSL2()τ\mathrm{PSL}_{2}(\mathbb{Z})\tau, for some m0m\geq 0, and it has CM by 𝒪dp2n\mathcal{O}_{dp^{2n}} for some n0n\geq 0. This is the case if ψγτ(K)M2()=ψγτ(𝒪dp2n)\psi_{\gamma\tau}(K)\cap M_{2}(\mathbb{Z})=\psi_{\gamma\tau}(\mathcal{O}_{dp^{2n}}). But ψγτ=γψτγ1\psi_{\gamma\tau}=\gamma\psi_{\tau}\gamma^{-1} so this condition boils down to ψτ(K)γ1M2()γ=ψτ(𝒪dp2n)\psi_{\tau}(K)\cap\gamma^{-1}M_{2}(\mathbb{Z})\gamma=\psi_{\tau}(\mathcal{O}_{dp^{2n}}). We can check this condition locally. Since γPGL2(q)\gamma\in\mathrm{PGL}_{2}(\mathbb{Z}_{q}) for qpq\neq p, it is only neccesary to check that ψτ(Kp)γ1M2(p)γ=ψτ(𝒪dp2np)\psi_{\tau}(K_{p})\cap\gamma^{-1}M_{2}(\mathbb{Z}_{p})\gamma=\psi_{\tau}(\mathcal{O}_{dp^{2n}}\otimes\mathbb{Z}_{p}). Since the projectivized units of γ1M2(p)γ\gamma^{-1}M_{2}(\mathbb{Z}_{p})\gamma form the stabilizer of γ1v0=gv0\gamma^{-1}v_{0}=gv_{0} and 𝒪dp2np\mathcal{O}_{dp^{2n}}\otimes\mathbb{Z}_{p} is the order of conductor pnp^{n} in 𝒪dp\mathcal{O}_{d}\otimes\mathbb{Z}_{p}, by Corollary 2.2 this is equivalent to v=gv0v=gv_{0} being at distance nn from the geodesic fixed by Kp×K_{p}^{\times}.

For the edges, we remark that

Γ+\(×(𝒯p))Γ+\(×PGL2(p))/ ​​Γ0(pp)Γ0(p)\.\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma^{+}$}\big{\backslash}\raise 2.15277pt\hbox{$(\mathcal{H}\times\overrightarrow{\mathcal{E}}(\mathcal{T}_{p}))$}}}{\Gamma^{+}\,\backslash\,(\mathcal{H}\times\overrightarrow{\mathcal{E}}(\mathcal{T}_{p}))}{\Gamma^{+}\,\backslash\,(\mathcal{H}\times\overrightarrow{\mathcal{E}}(\mathcal{T}_{p}))}{\Gamma^{+}\,\backslash\,(\mathcal{H}\times\overrightarrow{\mathcal{E}}(\mathcal{T}_{p}))}\cong\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma^{+}$}\big{\backslash}\raise 2.15277pt\hbox{$(\mathcal{H}\times\mathrm{PGL}_{2}(\mathbb{Q}_{p}))$\!}\big{/} \lower 2.15277pt\hbox{\!\!$\Gamma_{0}(p\mathbb{Z}_{p})$}}}{\Gamma^{+}\,\backslash\,(\mathcal{H}\times\mathrm{PGL}_{2}(\mathbb{Q}_{p}))\,/\,\Gamma_{0}(p\mathbb{Z}_{p})}{\Gamma^{+}\,\backslash\,(\mathcal{H}\times\mathrm{PGL}_{2}(\mathbb{Q}_{p}))\,/\,\Gamma_{0}(p\mathbb{Z}_{p})}{\Gamma^{+}\,\backslash\,(\mathcal{H}\times\mathrm{PGL}_{2}(\mathbb{Q}_{p}))\,/\,\Gamma_{0}(p\mathbb{Z}_{p})}\cong\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma_{0}(p)$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{H}$}}}{\Gamma_{0}(p)\,\backslash\,\mathcal{H}}{\Gamma_{0}(p)\,\backslash\,\mathcal{H}}{\Gamma_{0}(p)\,\backslash\,\mathcal{H}}.

This shows that two points in Γ+\(Γ+τ×𝒯p)\Gamma^{+}\backslash(\Gamma^{+}\tau\times\mathcal{T}_{p}) are connected by an edge if and only if the respective elliptic curves are pp-isogenous. This ends the proof of the second claim. ∎

By definition of Δτ\Delta_{\tau}, the vertex gv0gv_{0} belongs to the geodesic preserved by Kp×K_{p}^{\times} so we have that the second coordinate of Δτ\Delta_{\tau} projects to the rim of 𝒪τ×\𝒯p\mathchoice{\text{\lower 2.15277pt\hbox{$\mathcal{O}_{\tau}^{\times}$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{T}_{p}$}}}{\mathcal{O}_{\tau}^{\times}\,\backslash\,\mathcal{T}_{p}}{\mathcal{O}_{\tau}^{\times}\,\backslash\,\mathcal{T}_{p}}{\mathcal{O}_{\tau}^{\times}\,\backslash\,\mathcal{T}_{p}}. This closed cycle in the quotient of the tree is what the authors in [BDIS02] refer to as a pp-adic Shintani cycle. This justifies the name of Δτ\Delta_{\tau} and since the rim is a closed cycle we can think of it as a discrete closed geodesic mimicking Duke’s geodesics in the archimedean case. Note the following consequence of Proposition 5.2 and its proof.

Corollary 5.3.

The image of Δτ\Delta_{\tau} in PSL2()\\mathchoice{\text{\lower 2.15277pt\hbox{$\mathrm{PSL}_{2}(\mathbb{Z})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{H}$}}}{\mathrm{PSL}_{2}(\mathbb{Z})\,\backslash\,\mathcal{H}}{\mathrm{PSL}_{2}(\mathbb{Z})\,\backslash\,\mathcal{H}}{\mathrm{PSL}_{2}(\mathbb{Z})\,\backslash\,\mathcal{H}} under (5.1) corresponds to the kk elliptic curves with CM by 𝒪d\mathcal{O}_{d} and which are pp^{\infty}-isogenous to the elliptic curve associated with PSL2()τ\mathrm{PSL}_{2}(\mathbb{Z})\tau.

The set of elliptic curves described in the above Corollary form a closed orbit under the usual action of 𝔭\mathfrak{p}^{\mathbb{Z}} on the set of elliptic curves with CM by 𝒪d\mathcal{O}_{d}. In this regard, just as Duke’s geodesics form a closed orbit under the geodesic flow, the cycles Δτ\Delta_{\tau} are a closed orbit under a “flow ”defined by pp.

6. Heegner points on Shimura curves

In this section we focus on the cases where BB is a definite quaternion algebra ramified at NN^{-} and KK is imaginary and inert at pp. Remember that RR is a [1/p]\mathbb{Z}[1/p]-Eichler order of level N+N^{+} and Γ=R¯×\Gamma=\overline{R}^{\times}. Denote by Γ+\Gamma^{+} the image in Γ\Gamma of R1R_{1}, the elements in RR with reduced norm equal to 11.

Let \mathcal{B} be the indefinite quaternion algebra over \mathbb{Q} ramified at the primes dividing NpN^{-}p. Let \mathcal{R} be an Eichler order of level N+N^{+} in \mathcal{B}. After fixing an isomorphism ι:M2()\iota_{\infty}\colon\mathcal{B}_{\infty}\cong M_{2}(\mathbb{R}), the group Γ=ι(×)\Gamma_{\infty}=\iota_{\infty}(\mathcal{R}^{\times}) acts on \colonequals\mathcal{H}_{\infty}\colonequals\mathbb{C}\smallsetminus\mathbb{R}. The space Γ\\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma_{\infty}$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{H}$}}}{\Gamma_{\infty}\,\backslash\,\mathcal{H}}{\Gamma_{\infty}\,\backslash\,\mathcal{H}}{\Gamma_{\infty}\,\backslash\,\mathcal{H}}_{\infty} is a compact Riemann surface that corresponds to the complex points of an algebraic variety XX defined over \mathbb{Q} i.e. X()=Γ\X(\mathbb{C})=\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma_{\infty}$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{H}$}}}{\Gamma_{\infty}\,\backslash\,\mathcal{H}}{\Gamma_{\infty}\,\backslash\,\mathcal{H}}{\Gamma_{\infty}\,\backslash\,\mathcal{H}}_{\infty}. We refer to XX as the Shimura curve attached to the data (,)(\mathcal{B},\mathcal{R}).

Let p2\mathbb{Q}_{p^{2}} be the unique unramified quadratic extension of p\mathbb{Q}_{p}. The compact space X(p2)X(\mathbb{Q}_{p^{2}}) admits the following pp-adic uniformization. Let p\colonequalsp2p\mathcal{H}_{p}\colonequals\mathbb{Q}_{p^{2}}-\mathbb{Q}_{p}, the Drinfeld upper-half plane. After fixing an isomorphism BpM2(p)B_{p}\cong M_{2}(\mathbb{Q}_{p}), Γ\Gamma acts on p\mathcal{H}_{p} by fractional linear transformations.

Theorem 6.1 (Cerednik-Drinfeld).

The quotient space Γ+\p\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma^{+}$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{H}_{p}$}}}{\Gamma^{+}\,\backslash\,\mathcal{H}_{p}}{\Gamma^{+}\,\backslash\,\mathcal{H}_{p}}{\Gamma^{+}\,\backslash\,\mathcal{H}_{p}} corresponds to the space of p2\mathbb{Q}_{p^{2}}-points of the Shimura curve XX.

Proof.

See Theorem 4.7 in [BD98] or also [Cer76] or [Dd76]. ∎

Fix some up×(p×)2u\in\mathbb{Z}_{p}^{\times}\smallsetminus(\mathbb{Z}_{p}^{\times})^{2} and τ0\tau_{0} a square root of uu so that p2=p(τ0)\mathbb{Q}_{p^{2}}=\mathbb{Q}_{p}(\tau_{0}). The stabilizer 𝕆p\mathbb{O}_{p} of τ0\tau_{0} in PBp×PB^{\times}_{p} is compact (isomorphic to p2×/p×\mathbb{Q}_{p^{2}}^{\times}/\mathbb{Q}_{p}^{\times}). Since PGL2(p)\mathrm{PGL}_{2}(\mathbb{Q}_{p}) acts transitively on p\mathcal{H}_{p}, we have the identification of locally compacts spaces

[𝑮]𝑲𝕆pΓ\PBp×/ ​​𝕆pΓ\p,[\boldsymbol{G}]_{\boldsymbol{K}\cdot\mathbb{O}_{p}}\cong\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma$}\big{\backslash}\raise 2.15277pt\hbox{$PB_{p}^{\times}$\!}\big{/} \lower 2.15277pt\hbox{\!\!$\mathbb{O}_{p}$}}}{\Gamma\,\backslash\,PB_{p}^{\times}\,/\,\mathbb{O}_{p}}{\Gamma\,\backslash\,PB_{p}^{\times}\,/\,\mathbb{O}_{p}}{\Gamma\,\backslash\,PB_{p}^{\times}\,/\,\mathbb{O}_{p}}\cong\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{H}_{p}$}}}{\Gamma\,\backslash\,\mathcal{H}_{p}}{\Gamma\,\backslash\,\mathcal{H}_{p}}{\Gamma\,\backslash\,\mathcal{H}_{p}},

as in (4.1).

Let ω\omega be any element in ΓΓ+\Gamma-\Gamma^{+}. Since [Γ:Γ+]=2[\Gamma:\Gamma^{+}]=2, it follows that ω\omega defines an involution on X(p2)X(\mathbb{Q}_{p^{2}}) and therefore Γ\p\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{H}_{p}$}}}{\Gamma\,\backslash\,\mathcal{H}_{p}}{\Gamma\,\backslash\,\mathcal{H}_{p}}{\Gamma\,\backslash\,\mathcal{H}_{p}} corresponds to pairs of points in X(p2)X(\mathbb{Q}_{p^{2}}) connected by this involution.

According to the moduli interpretation of XX (see section 4 in [BD98]), the points in X(p2)X(\mathbb{Q}_{p^{2}}) corresponds to abelian surfaces over p2\mathbb{Q}_{p^{2}} with quaternionic multiplication by \mathcal{R} and a N+N^{+}-level structure. Given AA, one such abelian surface, we denote by End(A)\mathrm{End}(A) the algebra of endormophism of AA (over ¯\overline{\mathbb{Q}}) which commute with the quaternionic multiplication and respect the N+N^{+}-level structure. A Heegner point in X(p2)X(\mathbb{Q}_{p^{2}}) is a point whose associated abelian surface AA has End(A)\mathrm{End}(A) isomorphic to an order in a quadratic imaginary field.

As in section 4.2, let 𝒪\mathcal{O} denotes a [1/p]\mathbb{Z}[1/p] order in KK. Let ψopt(𝒪,R)\psi\in opt(\mathcal{O},R). Then ψ(Kp)×\psi(K_{p})^{\times} acts in p\mathcal{H}_{p} with two fixed points τ\tau and τ\tau^{\prime}. We assume that τ\tau is the fixed point satisfying that for every xKp×x\in K_{p}^{\times}, ψ(x)\psi(x) acts on the column vector (τ    1)t(\tau\,\,\,\,1)^{t} as multiplication by xx. This time we take H=𝕆pH=\mathbb{O}_{p} and Δψ\Delta_{\psi} equals the Γ\Gamma-orbit of ψ(Kp)×g\psi(K_{p})^{\times}g with gPBp×g\in PB_{p}^{\times} such that ψ(Kp)×=g𝕆pg1\psi(K_{p})^{\times}=g\mathbb{O}_{p}g^{-1}. We normalize the choice of gg by asking that gτ0=τg\tau_{0}=\tau. The projection of Δψ\Delta_{\psi} in Γ\p\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{H}_{p}$}}}{\Gamma\,\backslash\,\mathcal{H}_{p}}{\Gamma\,\backslash\,\mathcal{H}_{p}}{\Gamma\,\backslash\,\mathcal{H}_{p}} is Γτ\Gamma\tau. It corresponds to the pair of points in X(p2)X(\mathbb{Q}_{p^{2}}) associated with Γ+τ\Gamma^{+}\tau and Γ+ωτ\Gamma^{+}\omega\tau. Let dd be as in section 4.2.

Theorem 6.2.

Under the identification given by Theorem 6.1, the class of Γ+τ\Gamma^{+}\tau corresponds to a Heegner point in X(p2)X(\mathbb{Q}_{p^{2}}) whose associated order is isomorphic to 𝒪d\mathcal{O}_{d}. As ψ\psi varies over [opt(𝒪,R)][opt(\mathcal{O},R)], they are all different.

Proof.

Since pp is inert in KK, Pic(𝒪)Pic(𝒪d)Pic(\mathcal{O})\cong Pic(\mathcal{O}_{d}). Now see Theorem 5.3 in [BD98]. ∎

The previous result justifies that we name by Heeg(𝒪)\mathrm{Heeg}(\mathcal{O}), the collection of Γτ\Gamma\tau in Γ\p\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{H}_{p}$}}}{\Gamma\,\backslash\,\mathcal{H}_{p}}{\Gamma\,\backslash\,\mathcal{H}_{p}}{\Gamma\,\backslash\,\mathcal{H}_{p}} as ψ\psi runs over [opt(𝒪,R)][opt(\mathcal{O},R)].

7. Adelic methods

In this last section we prove Theorem A and Theorem B. They will be a direct consequence of Proposition 7.4.

The group 𝑮(𝔸)\boldsymbol{G}(\mathbb{A}) is unimodular (any left Haar measure is also a right Haar measure) so there is a unique 𝑮(𝔸)\boldsymbol{G}(\mathbb{A})-invariant probability measure dmdm in [𝑮][\boldsymbol{G}] such that for any fCc(𝑮(𝔸))f\in C_{c}(\boldsymbol{G}(\mathbb{A})),

(7.1) 𝑮(𝔸)f𝑑g=[𝑮]𝑮()f(hg)𝑑h𝑑m(g),\int_{\boldsymbol{G}(\mathbb{A})}fdg=\int_{[\boldsymbol{G}]}\int_{\boldsymbol{G}(\mathbb{Q})}f(hg)dhdm(g),

where dhdh is the counting measure in 𝑮()\boldsymbol{G}(\mathbb{Q}) and dgdg a Haar measure on 𝑮(𝔸)\boldsymbol{G}(\mathbb{A}).

Definition 1.

A homogeneous toral subset in [𝐆][\boldsymbol{G}] is a subset of the form Y=𝐓()\𝐓(𝔸)gY=\boldsymbol{T}(\mathbb{Q})\backslash\boldsymbol{T}(\mathbb{A})g, with g𝐆(𝔸)g\in\boldsymbol{G}(\mathbb{A}) and 𝐓𝐆\boldsymbol{T}\subseteq\boldsymbol{G} a maximal torus anisotropic over \mathbb{Q}.

The pushforward of the Haar probability measure on 𝑻()\𝑻(𝔸)\boldsymbol{T}(\mathbb{Q})\backslash\boldsymbol{T}(\mathbb{A}) defines a probability measure μY\mu_{Y} on the homogeneous toral set Y=𝑻()\𝑻(𝔸)gY=\boldsymbol{T}(\mathbb{Q})\backslash\boldsymbol{T}(\mathbb{A})g.

Let KK be a quadratic field and let ψ:KB()\psi\colon K\to B(\mathbb{Q}) be an algebra embedding. Let 𝑻K\boldsymbol{T}_{K} be the algebraic torus resK/𝔾m/𝔾m\mathrm{res}_{K/\mathbb{Q}}\mathbb{G}_{m}/\mathbb{G}_{m}. Then ψ\psi induces a morphism (which we denote by the same letter) ψ:𝑻K𝑮\psi\colon\boldsymbol{T}_{K}\to\boldsymbol{G}. Denote by 𝑻ψ\boldsymbol{T}_{\psi} the image of 𝑻K\boldsymbol{T}_{K} under ψ\psi in 𝑮\boldsymbol{G}. Then every maximal anisotropic torus 𝑻𝑮\boldsymbol{T}\subseteq\boldsymbol{G} defined over \mathbb{Q} is of the form 𝑻ψ\boldsymbol{T}_{\psi} for some embedding ψ\psi. Now, we attach to 𝑻ψ\boldsymbol{T}_{\psi} an order in KK. Fix a \mathbb{Z}-order RR in B()B(\mathbb{Q}). Then define the local orders Λ=ψ(K)gRg1\Lambda_{\ell}=\psi(K_{\ell})\cap g_{\ell}R_{\ell}g_{\ell}^{-1}. The order Λ\Lambda_{\ell} is maximal for almost every \ell. Indeed, let TT be the finite set of places such that gR¯×g_{\ell}\not\in\overline{R}_{\ell}^{\times}, where R¯×\overline{R}_{\ell}^{\times} denotes the image of R×R_{\ell}^{\times} in PB×=B×/×PB_{\ell}^{\times}=B_{\ell}^{\times}/\mathbb{Q}_{\ell}^{\times}. Then if T\ell\not\in T, the intersection ψ(K)gRg1\psi(K_{\ell})\cap g_{\ell}R_{\ell}g_{\ell}^{-1} is just ψ(K)R\psi(K_{\ell})\cap R_{\ell}. But this is the localization of the global order ψ(K)R\psi(K)\cap R and so it is maximal outside a finite set of places. This shows that Λ=KΛ\Lambda=K\cap\prod\Lambda_{\ell} is a global order in KK.

Definition 2 ([Kha19] section 2.4.4 or equivalently [ELMV11]).

The discriminant of a homogeneous toral set of the form Y=𝐓ψgY=\boldsymbol{T}_{\psi}g, with ψ:KB()\psi\colon K\to B(\mathbb{Q}) an embeeding, is the absolute value of the discriminant of the order Λ\Lambda attached to it as in the previous paragraph.

Theorem 7.1.

Let {Yi}\{Y_{i}\} be a sequence of homogeneous toral sets whose discriminants approach \infty as ii\to\infty. Then, any weak* accumulation point of the sequence of measures μYi\mu_{Y_{i}} is a homogeneous probability measure on [𝐆][\boldsymbol{G}], invariant under 𝐆(𝔸)+\boldsymbol{G}(\mathbb{A})^{+}, the image of 𝐆~(𝔸)\widetilde{\boldsymbol{G}}(\mathbb{A}) in 𝐆(𝔸)\boldsymbol{G}(\mathbb{A}).

Proof.

Theorem 4.6 in [ELMV11]

The defect of the limit measure in Theorem 7.1 being 𝑮(𝔸)\boldsymbol{G}(\mathbb{A})-invariant is solved in the following fashion. Assume that {μYi}\{\mu_{Y_{i}}\} weak* converges to μ\mu. We want to prove that for all ff in Cu([𝑮])C_{u}([\boldsymbol{G}]), the space of bounded and uniformly continuous functions, we have

(7.2) [𝑮]f(g)𝑑μ(g)=[𝑮]f(g)𝑑m(g).\int_{[\boldsymbol{G}]}f(g)d\mu(g)=\int_{[\boldsymbol{G}]}f(g)dm(g).

By the discussion made [ALMW22], sections 9 and 10, we are reduced to prove (7.2) in the case of functions invariant under 𝑮(𝔸)+\boldsymbol{G}(\mathbb{A})^{+}. Let 𝑮char\boldsymbol{G}_{char} denote the quotient [𝑮]/𝑮(𝔸)+\mathchoice{\text{\raise 2.15277pt\hbox{$[\boldsymbol{G}]$}\big{/}\lower 2.15277pt\hbox{$\boldsymbol{G}(\mathbb{A})^{+}$}}}{[\boldsymbol{G}]\,/\,\boldsymbol{G}(\mathbb{A})^{+}}{[\boldsymbol{G}]\,/\,\boldsymbol{G}(\mathbb{A})^{+}}{[\boldsymbol{G}]\,/\,\boldsymbol{G}(\mathbb{A})^{+}}. The reduced norm induces a homeomorphism

𝑮char×\𝔸×/ ​​(𝔸×)2,\boldsymbol{G}_{char}\cong\mathchoice{\text{\lower 2.15277pt\hbox{$\mathbb{Q}^{\times}$}\big{\backslash}\raise 2.15277pt\hbox{$\mathbb{A}^{\times}$\!}\big{/} \lower 2.15277pt\hbox{\!\!$(\mathbb{A}^{\times})^{2}$}}}{\mathbb{Q}^{\times}\,\backslash\,\mathbb{A}^{\times}\,/\,(\mathbb{A}^{\times})^{2}}{\mathbb{Q}^{\times}\,\backslash\,\mathbb{A}^{\times}\,/\,(\mathbb{A}^{\times})^{2}}{\mathbb{Q}^{\times}\,\backslash\,\mathbb{A}^{\times}\,/\,(\mathbb{A}^{\times})^{2}},

which gives to 𝑮char\boldsymbol{G}_{char} the structure of a compact abelian group. Since the group of characters is dense in the space of continuous functions, by Weyl’s criterion we are reduced to the case when ff is a character.

If χ\chi is a character of ×\𝔸×/(𝔸×)2\mathchoice{\text{\lower 2.15277pt\hbox{$\mathbb{Q}^{\times}$}\big{\backslash}\raise 2.15277pt\hbox{$\mathbb{A}^{\times}$\!}\big{/} \lower 2.15277pt\hbox{\!\!$(\mathbb{A}^{\times})^{2}$}}}{\mathbb{Q}^{\times}\,\backslash\,\mathbb{A}^{\times}\,/\,(\mathbb{A}^{\times})^{2}}{\mathbb{Q}^{\times}\,\backslash\,\mathbb{A}^{\times}\,/\,(\mathbb{A}^{\times})^{2}}{\mathbb{Q}^{\times}\,\backslash\,\mathbb{A}^{\times}\,/\,(\mathbb{A}^{\times})^{2}}, by continuity there exists some compact open group M𝔸f×M\subseteq\mathbb{A}_{f}^{\times} such that χM=1\chi\mid_{M}=1 and therefore we have a character of the group ×\𝔸×/>0M(/D)×\mathchoice{\text{\lower 2.15277pt\hbox{$\mathbb{Q}^{\times}$}\big{\backslash}\raise 2.15277pt\hbox{$\mathbb{A}^{\times}$\!}\big{/} \lower 2.15277pt\hbox{\!\!$\mathbb{R}_{>0}M$}}}{\mathbb{Q}^{\times}\,\backslash\,\mathbb{A}^{\times}\,/\,\mathbb{R}_{>0}M}{\mathbb{Q}^{\times}\,\backslash\,\mathbb{A}^{\times}\,/\,\mathbb{R}_{>0}M}{\mathbb{Q}^{\times}\,\backslash\,\mathbb{A}^{\times}\,/\,\mathbb{R}_{>0}M}\cong(\mathbb{Z}/D\mathbb{Z})^{\times} for some DD\in\mathbb{Z} and trivial on squares. Therefore, if we assume MM maximal, we are in the presence of a primitive real Dirichlet character that must be attached to some quadratic field KχK_{\chi}.

We have reduced the test function to be of the shape χnrd\chi\circ\mathrm{nrd} where χ\chi is a quadratic Hecke character.

Proposition 7.2.

Let χ\chi be a non-trivial character of 𝐆char\boldsymbol{G}_{char} and Y=𝐓ψgY=\boldsymbol{T}_{\psi}g an homogeneous toral set with ψ:KB\psi\colon K\to B. Then,

Yχ(nrd(t))𝑑μY(t)=χ(g)𝑻ψχ(NK/(t))𝑑μ𝑻ψ(t)={χ(g), if K=Kχ0, if KKχ\int_{Y}\chi(\mathrm{nrd}(t))d\mu_{Y}(t)=\chi(g)\int_{\boldsymbol{T}_{\psi}}\chi(N_{K/\mathbb{Q}}(t))d\mu_{\boldsymbol{T}_{\psi}}(t)=\begin{cases}\chi(g)&\mbox{, if }K=K_{\chi}\\ 0&\mbox{, if }K\neq K_{\chi}\end{cases}
Proof.

Assume that χ\chi is related to Kχ=(D)K_{\chi}=\mathbb{Q}(\sqrt{D}) in the sense of the previous paragraph. If LL denotes (ζ|D|)\mathbb{Q}(\zeta_{|D|}), then KLK\subseteq L is a Galois extension and class field theory provides the diagram

𝔸K×{\mathbb{A}_{K}^{\times}}Gal(L/K){\mathrm{Gal}(L/K)}𝔸×{\mathbb{A}_{\mathbb{Q}}^{\times}}Gal(L/){\mathrm{Gal}(L/\mathbb{Q})}(/D)×{(\mathbb{Z}/D\mathbb{Z})^{\times}}Gal(K/){\mathrm{Gal}(K/\mathbb{Q})}{±1}.{\{\pm 1\}.}NK/\scriptstyle{N_{K/\mathbb{Q}}}rec\scriptstyle{rec}\scriptstyle{\sim}χ\scriptstyle{\chi}\scriptstyle{\sim}

If K=KχK=K_{\chi} and t𝑻𝑲t\in\boldsymbol{T}_{\boldsymbol{K}} the equality χnrd(ψ(t))=χ(NK/(t))=χ(NKχ/(t))=1\chi\circ\mathrm{nrd}(\psi(t))=\chi(N_{K/\mathbb{Q}}(t))=\chi(N_{K_{\chi}/\mathbb{Q}}(t))=1 follows from the previous diagram. When KKχK\neq K_{\chi} there exists a prime \ell split in KK and inert in KχK_{\chi}. Let ss be the idele corresponding to the class of a prime 𝔩\mathfrak{l} above \ell in KK. Then χ(nrd(ψ(s)))=χ(NK/(𝔩))=χ(p)=1\chi(\mathrm{nrd}(\psi(s)))=\chi(N_{K/\mathbb{Q}}(\mathfrak{l}))=\chi(p)=-1 and so the substitution ttst\mapsto ts shows that 𝑻ψχ(nrd(t))𝑑μ𝑻ψ(t)=0\int_{\boldsymbol{T}_{\psi}}\chi(\mathrm{nrd}(t))d\mu_{\boldsymbol{T}_{\psi}}(t)=0. ∎

Let MM be a compact subgroup of 𝑮(𝔸f)\boldsymbol{G}(\mathbb{A}_{f}) with nrd(Mp)(p×)2=p×\mathrm{nrd}(M_{p})(\mathbb{Z}_{p}^{\times})^{2}=\mathbb{Z}_{p}^{\times} for almost every pp. Let 𝒦(M)\mathcal{K}(M) be the finite set of quadratic fields attached to the characters of the finite group ×\𝔸×/(𝔸×)2nrd(M)\mathchoice{\text{\lower 2.15277pt\hbox{$\mathbb{Q}^{\times}$}\big{\backslash}\raise 2.15277pt\hbox{$\mathbb{A}^{\times}$\!}\big{/} \lower 2.15277pt\hbox{\!\!$(\mathbb{A}^{\times})^{2}\mathrm{nrd}(M)$}}}{\mathbb{Q}^{\times}\,\backslash\,\mathbb{A}^{\times}\,/\,(\mathbb{A}^{\times})^{2}\mathrm{nrd}(M)}{\mathbb{Q}^{\times}\,\backslash\,\mathbb{A}^{\times}\,/\,(\mathbb{A}^{\times})^{2}\mathrm{nrd}(M)}{\mathbb{Q}^{\times}\,\backslash\,\mathbb{A}^{\times}\,/\,(\mathbb{A}^{\times})^{2}\mathrm{nrd}(M)}. Equip [𝑮]M[\boldsymbol{G}]_{M} with dmMdm_{M} the pushforward of dmdm under the natural projection.

Proposition 7.3.

Let {Yi=𝐓ψigi}\{Y_{i}=\boldsymbol{T}_{\psi_{i}}g_{i}\} be a sequence of homogeneous toral sets with ψi:KiB\psi_{i}\colon K_{i}\to B such that Ki𝒦(M)K_{i}\notin\mathcal{K}(M). Assume also that its sequence of discriminants approaches \infty as ii\to\infty. Then the projection of YiY_{i} to [𝐆]M[\boldsymbol{G}]_{M} becomes equidistributed with respect to dmMdm_{M}. In other words, the sequence {μYi,M}\{\mu_{{Y_{i}},M}\} converges weak* to dmMdm_{M}.

Proof.

In terms of functionals over C([𝑮]M)C([\boldsymbol{G}]_{M}), the pushforward of μYi\mu_{Y_{i}} corresponds to the restriction of μYi\mu_{Y_{i}} over the space of MM-right invariant functions over [𝑮][\boldsymbol{G}]. Therefore, by theorem 7.1 and its following discussion, to prove equidistribution we need to check

[𝑮]f(g)𝑑μ(g)=[𝑮]f(g)𝑑m(g),\int_{[\boldsymbol{G}]}f(g)d\mu(g)=\int_{[\boldsymbol{G}]}f(g)dm(g),

with μ\mu a weak* accumulation measure of μYi\mu_{Y_{i}} and ff of the form χnrd\chi\circ\mathrm{nrd} with an MM-right invariant character χ\chi. This last condition implies that Kχ𝒦(M)K_{\chi}\in\mathcal{K}(M) and now proposition 7.2 allows us to conclude since after taking a subsequence we can assume that μYi\mu_{Y_{i}} weak* converges to μ\mu and then [G]f(g)𝑑μ(g)=limYif(g)𝑑μYi(g)\int_{[G]}f(g)d\mu(g)=\lim\int_{Y_{i}}f(g)d\mu_{Y_{i}}(g). ∎

The map HΔψH\to\Delta_{\psi} sending hh to Γgh\Gamma gh induces a uniformization of Δψ\Delta_{\psi} by H/(Hg1Γg)H/(H\cap g^{-1}\Gamma g). Using a Haar measure in HH, we equip H/(Hg1Γg)H/(H\cap g^{-1}\Gamma g) with a finite measure and push it to a finite HH-right invariant measure νψ\nu_{\psi} in Δψ\Delta_{\psi}. We equip Ω𝒪=[ψ]opt(𝒪,R)Δψ\Omega_{\mathcal{O}}=\bigsqcup_{[\psi]\in opt(\mathcal{O},R)}\Delta_{\psi} with the unique probability measure μ𝒪\mu_{\mathcal{O}} proportional to opt(𝒪,R)νψ\sum_{opt(\mathcal{O},R)}\nu_{\psi}. There exits a unique probability measure on the quotient Γ\G\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma$}\big{\backslash}\raise 2.15277pt\hbox{$G$}}}{\Gamma\,\backslash\,G}{\Gamma\,\backslash\,G}{\Gamma\,\backslash\,G} induced by the counting measure on Γ\Gamma and a Haar measure on GG satisfying the property analogue to (7.1). It is the pushforward of the measure dmdm by means of the projection [𝑮]Γ\G[\boldsymbol{G}]\to\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma$}\big{\backslash}\raise 2.15277pt\hbox{$G$}}}{\Gamma\,\backslash\,G}{\Gamma\,\backslash\,G}{\Gamma\,\backslash\,G} described in (4.1).

Proposition 7.4.

The collection Ω𝒪\Omega_{\mathcal{O}} becomes equidistributed on Γ\G\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma$}\big{\backslash}\raise 2.15277pt\hbox{$G$}}}{\Gamma\,\backslash\,G}{\Gamma\,\backslash\,G}{\Gamma\,\backslash\,G} as dd\to-\infty. In particular, the collection Ih(𝒪)Ih(\mathcal{O}) becomes equidistributed in Γ+\×PGL2(p)\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma^{+}$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{H}\times\mathrm{PGL}_{2}(\mathbb{Q}_{p})$}}}{\Gamma^{+}\,\backslash\,\mathcal{H}\times\mathrm{PGL}_{2}(\mathbb{Q}_{p})}{\Gamma^{+}\,\backslash\,\mathcal{H}\times\mathrm{PGL}_{2}(\mathbb{Q}_{p})}{\Gamma^{+}\,\backslash\,\mathcal{H}\times\mathrm{PGL}_{2}(\mathbb{Q}_{p})} and Heeg(𝒪)Heeg(\mathcal{O}) becomes equidistributed in Γ\p\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{H}_{p}$}}}{\Gamma\,\backslash\,\mathcal{H}_{p}}{\Gamma\,\backslash\,\mathcal{H}_{p}}{\Gamma\,\backslash\,\mathcal{H}_{p}}.

Proof.

Remember the identification [𝑮]𝑲=Γ\G[\boldsymbol{G}]_{\boldsymbol{K}}=\mathchoice{\text{\lower 2.15277pt\hbox{$\Gamma$}\big{\backslash}\raise 2.15277pt\hbox{$G$}}}{\Gamma\,\backslash\,G}{\Gamma\,\backslash\,G}{\Gamma\,\backslash\,G}. Fix ψ0opt(𝒪,R)\psi_{0}\in opt(\mathcal{O},R) and let Y𝒪Y_{\mathcal{O}} the homogeneous toral set 𝑻ψ0g\boldsymbol{T}_{\psi_{0}}g with gg as before. Since det(R^#×)=^#×\det(\widehat{R}^{\times}_{\#})=\widehat{\mathbb{Z}}_{\#}^{\times}, the set 𝒦(𝑲f)\mathcal{K}(\boldsymbol{K}_{f}) consist only of quadratic fields ramified at pp, i.e (±p)\mathbb{Q}(\sqrt{\pm p}). But our sequence of tori consider KK split or inert at pp, so we will always be in the case KiKχK_{i}\neq K_{\chi} and therefore by Theorem 7.1 together with Proposition 7.3 there’s equidistribution of the image of Y𝒪Y_{\mathcal{O}} in [𝑮]𝑲[\boldsymbol{G}]_{\boldsymbol{K}} if we show that the discriminant goes to \infty. Note that Proposition 5.1 shows that Y𝒪Y_{\mathcal{O}} projects onto Ω𝒪\Omega_{\mathcal{O}}.

Take RR an \mathbb{Z}-order in B()B(\mathbb{Q}) agreeing with RR outside pp. We know ψ(K)R=ψ(𝒪d[1/p])\psi(K)\cap R=\psi(\mathcal{O}_{d}[1/p]). Localizing at qpq\neq p we have gq=1g_{q}=1 which implies Λq=𝒪dq\Lambda_{q}=\mathcal{O}_{d}\otimes\mathbb{Z}_{q}. Then the discriminant of Y𝒪Y_{\mathcal{O}} is the discriminant of 𝒪dpm\mathcal{O}_{dp^{m}} for some m0m\geq 0. Certainly this goes to infinity and we must have equidistribution. ∎

Remark 3.

In the indefinite case, since gpg_{p} is chosen so that ψ(Kp)×=gpAgp1\psi(K_{p})^{\times}=g_{p}Ag_{p}^{-1} we have that gpv0g_{p}v_{0} belongs to the geodesic fixed by ψ(Kp)×\psi(K_{p})^{\times}. This is equivalent to gv0gv_{0} having stabilizer by the maximal order of ψ(Kp)×\psi(K_{p})^{\times} which in turn is equivalent to ψ(Kp)×gpM2(p)gp1\psi(K_{p})^{\times}\cap g_{p}M_{2}(\mathbb{Z}_{p})g_{p}^{-1} being maximal. This shows that the discriminant of YY is exactly the discriminant of 𝒪d\mathcal{O}_{d}.

References

  • [ALMW22] Menny Aka, Manuel Luethi, Philippe Michel, and Andreas Wieser. Simultaneous supersingular reductions of CM elliptic curves. Journal für die reine und angewandte Mathematik (Crelles Journal), 2022(786):1–43, feb 2022.
  • [BD98] Massimo Bertolini and Henri Darmon. Heegner points, pp-adic LL-functions, and the Cerednik-Drinfeld uniformization. Invent. Math., 131(3):453–491, 1998.
  • [BDIS02] Massimo Bertolini, Henri Darmon, Adrian Iovita, and Michael Spiess. Teitelbaum’s exceptional zero conjecture in the anticyclotomic setting. Amer. J. Math., 124(2):411–449, 2002.
  • [Cas14] B. Casselman. The bruhat-tits tree of sl(2). Available at https://ncatlab.org/nlab/files/CasselmanOnBruhatTitsTree2014.pdf, March 2014.
  • [Cer76] I. V. Cerednik. Uniformization of algebraic curves by discrete arithmetic subgroups of PGL2(kw){PGL}{2}(k{w}) with compact quotient spaces. Mat. Sb. (N.S.), 100(142)(1):59–88, 165, 1976.
  • [Dd76] V. G. Drinfel d. Coverings of pp-adic symmetric domains. Funkcional. Anal. i Prilozen., 10(2):29–40, 1976.
  • [Dis22] Daniel Disegni. pp-adic equidistribution of CM points. Comment. Math. Helv., 97(4):635–668, 2022.
  • [Duk88] W. Duke. Hyperbolic distribution problems and half-integral weight Maass forms. Invent. Math., 92(1):73–90, 1988.
  • [ELMV11] Manfred Einsiedler, Elon Lindenstrauss, Philippe Michel, and Akshay Venkatesh. Distribution of periodic torus orbits and Duke’s theorem for cubic fields. Ann. of Math. (2), 173(2):815–885, 2011.
  • [ELMV12] Manfred Einsiedler, Elon Lindenstrauss, Philippe Michel, and Akshay Venkatesh. The distribution of closed geodesics on the modular surface, and Duke’s theorem. Enseign. Math. (2), 58(3-4):249–313, 2012.
  • [HM06] Gergely Harcos and Philippe Michel. The subconvexity problem for Rankin-Selberg LL-functions and equidistribution of Heegner points. II. Invent. Math., 163(3):581–655, 2006.
  • [HMRL20] Sebastián Herrero, Ricardo Menares, and Juan Rivera-Letelier. pp-adic distribution of CM points and Hecke orbits I: Convergence towards the Gauss point. Algebra Number Theory, 14(5):1239–1290, 2020.
  • [Iha08] Yasutaka Ihara. On congruence monodromy problems, volume 18 of MSJ Memoirs. Mathematical Society of Japan, Tokyo, 2008. Reproduction of the Lecture Notes: Volume 1 (1968) [MR0289518], Volume 2 (1969) [MR0289519] (University of Tokyo), with author’s notes (2008).
  • [Kha19] Ilya Khayutin. Joint equidistribution of CM points. Ann. of Math. (2), 189(1):145–276, 2019.
  • [PR94] Vladimir Platonov and Andrei Rapinchuk. Algebraic groups and number theory, volume 139 of Pure and Applied Mathematics. Academic Press, Inc., Boston, MA, 1994. Translated from the 1991 Russian original by Rachel Rowen.
  • [Ser03] Jean-Pierre Serre. Trees. Springer Monographs in Mathematics. Springer-Verlag, Berlin, 2003. Translated from the French original by John Stillwell, Corrected 2nd printing of the 1980 English translation.
  • [Vig80] Marie-France Vignéras. Arithmétique des algèbres de quaternions, volume 800 of Lecture Notes in Mathematics. Springer, Berlin, 1980.
  • [Voi21] John Voight. Quaternion algebras, volume 288 of Graduate Texts in Mathematics. Springer, Cham, [2021] ©2021.
  • [Zha05] Shou-Wu Zhang. Equidistribution of CM-points on quaternion Shimura varieties. Int. Math. Res. Not., (59):3657–3689, 2005.