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

Arithmetic quantum unique ergodicity for products of hyperbolic 22- and 33-manifolds

Zvi Shem-Tov Einstein Institute of Mathematics, The Hebrew University of Jerusalem, Israel [email protected]  and  Lior Silberman Department of Mathematics, University of British Columbia, Vancouver  BC  V6T 1Z2, Canada [email protected]
Abstract.

We prove the arithemtic quantum unique ergodicity (AQUE) conjecture for sequences of Hecke–Maass forms on quotients Γ\((2))r×((3))s\Gamma\backslash(\mathbb{H}^{(2)})^{r}\times(\mathbb{H}^{(3)})^{s}. An argument by induction on dimension of the orbit allows us to rule out the limit measure concentrating on closed orbits of proper subgroups despite many returns of the Hecke correspondence to neighborhoods of the orbit.

2010 Mathematics Subject Classification:
11F41; 37A45
LS was supported by an NSERC Discovery Grant

1. Introduction

1.1. Statement of result

Fix integers r,s0r,s\geq 0 so that r+s1r+s\geq 1 and consider the symmetric space

S=((2))r×((3))s,S=(\mathbb{H}^{(2)})^{r}\times(\mathbb{H}^{(3)})^{s},

where (n)\mathbb{H}^{(n)} denotes hyperbolic nn-space. Let Γ\Gamma be a lattice in the isometry group G=SL2()r×SL2()sG=\mathrm{SL}_{2}(\mathbb{R})^{r}\times\mathrm{SL}_{2}(\mathbb{C})^{s}. Then the finite-volume manifold Y=Γ\SY=\Gamma\backslash S is equipped with a family of r+sr+s commuting differential operators coming from the Laplace–Beltrami operator in each factor (the sum of which is the Laplace–Beltrami operator of YY). When Γ\Gamma is, in addition, a congruence lattice the manifold YY is also equipped with a ring of discrete averaging operators, the Hecke algebra, which further commute with the differnential operators noted above. Write dy=dvolYdy=\mathop{}\!\mathrm{d}\!\operatorname{vol}_{Y} for the Riemannian volume element normalized to have total volume 1. In this paper we establish the following result:

Theorem 1.

Let ψjL2(Y)\psi_{j}\in L^{2}(Y) be a sequence of normalized joint eigenfunctions of both the ring of invariant differential operators and of the Hecke algebra. Assume that the Laplace-eigenvalues λj\lambda_{j}\to\infty. Then |ψj|2dvolYjwk-*dvolY|\psi_{j}|^{2}\mathop{}\!\mathrm{d}\!\operatorname{vol}_{Y}\xrightarrow[j\to\infty]{\textrm{wk-*}}\mathop{}\!\mathrm{d}\!\operatorname{vol}_{Y}.

The new ingredient in the proof is a method for ruling out certain measures as components of weak-* limits of lifts of this setup to X=Γ\GX=\Gamma\backslash G. To state its realization here let GiG_{i} be one of the factors isomorphic to SL2()\mathrm{SL}_{2}(\mathbb{C}) in the product defining GG. Identifying GiG_{i} with SL2()\mathrm{SL}_{2}(\mathbb{C}) for the moment let Hi=SL2()H_{i}=\mathrm{SL}_{2}(\mathbb{R}), and let Mi={diag(eiθ,eiθ)}M_{i}=\{\operatorname{diag}(e^{i\theta},e^{-i\theta})\} be the group of diagonal matrices with entries of modulus 11. Finally set H=Hi×jiGjH=H_{i}\times\prod_{j\neq i}G_{j} (i.e. multiply HiH_{i} at the iith factor with the full isometry groups GjG_{j} at the other factors). Let g=(gj)jGg=(g_{j})_{j}\in G be such that the entries of the matrix gig_{i} are algebraic numbers.

Theorem 2.

Let μ\mu be a measure on XX which is the weak-* limit of measures

μj=|ϕj|2dvolX\mu_{j}=\left|{\phi_{j}}\right|^{2}\mathop{}\!\mathrm{d}\!\operatorname{vol}_{X}

where the ϕjL2(X)\phi_{j}\in L^{2}(X) are Hecke eigenfunctions. Then

μ(ΓgHMi)=0.\mu(\Gamma gHM_{i})=0\,.

1.2. Context: QUE Conjectures for locally symmetric spaces

More generally, let GG be a semisimple Lie group, KGK\subset G a maximal compact subgroup, Γ<G\Gamma<G a lattice, and consider the locally symmetric space

Y=Γ\G/K,Y=\Gamma\backslash G/K,

with its uniform probability measure dydy. The ring of GG-invariant differential operators on G/KG/K is commutative and commutes with the action of Γ\Gamma, hence acts on functions on YY. Let {ψj}j1L2(Y)\{\psi_{j}\}_{j\geq 1}\subset L^{2}(Y) be an orthonormal sequence of joint eigenfunctions of this ring. We remark that the Laplace–Beltrami operator of YY belongs to the ring and that writing λj\lambda_{j} for the corresponding eigenvalue for ψj\psi_{j} we necessarily have |λj|\left|{\lambda_{j}}\right|\to\infty. The Quantum Unique Ergodicity (QUE) Conjecture for locally symmetric spaces asserts the sequence {ψj}\{\psi_{j}\} becomes equidistributed on YY, in the sense that

Yf(y)|ψj(y)|2𝑑yYf(y)𝑑y,\int_{Y}f(y)|\psi_{j}(y)|^{2}dy\to\int_{Y}f(y)dy,

for all test functions fCc(Y)f\in C_{c}(Y). In other words, the sequence |ψj|2dy|\psi_{j}|^{2}dy of probability measures converges in the weak-* topology to the uniform measure dydy. Next, when Γ\Gamma is a congruence lattice the space YY also affords a commutative family of discrete averaging operators, the Hecke operators, which commute with the invariant differential operators. The Arithmetic QUE Conjecture (AQUE) is the restricted form of the QUE Conjecture for sequences of Hecke eigenfunctions, that is for seuqences where each fjf_{j} is simultaneously an eigenfunction of the ring of invariant differential operators and of the ring of Hecke operators.

Investigation of Arithmetic QUE goes back to the work [11] of Rudnick–Sarnak, who formulated the QUE conjecture for hyperbolic surfaces and, more generally, for compact manifolds of negative sectional curvature. A major breakthrough on this problem was due to Lindenstrauss, who in [8] established AQUE for congruence hyperbolic surfaces, that is congruence quotients Y=Γ\(2)Y=\Gamma\backslash\mathbb{H}^{(2)} of the hyperbolic plane. More precisely, Lindenstrauss proved that weak-* limits as above were proportional to the uniform measure. This fully resolved the Conjecture in the case of uniform lattices, i.e. when the quotient is compact. For non-uniform lattices, however, the possibility remained that the constant of proportionality was strictly less than one ("escape of mass"), an alternative ruled out later by Soundararajan [16]. Silberman–Venkatesh [13, 14] obtains generalizations of some of this work to the case of general semisimple groups GG, formulating the QUE and AQUE conjectures in the context of locally symmetric spaces and obtaining further cases of AQUE.

1.3. Discussion

As stated above, we establish AQUE for congruence quotients of products of hyperbolic 2- and 3-spaces. The case Y=SL2([i])\(3)Y=\mathrm{SL}_{2}(\mathbb{Z}[i])\backslash\mathbb{H}^{(3)} is already new and contains most of the the novel ideas of this paper; reading the paper with this assumption in mind will give the reader most of the insight but avoid the technical legedermain needed to handle number fields with multiple infinite places. For the rest of the introduction we concentrate on this case.

Thus let G=SL2()G=\mathrm{SL}_{2}(\mathbb{C}) acts transitively by isometries on the space S=(3)S=\mathbb{H}^{(3)} with point stablizer the maximal compact subgroup K=SU(2)K=\mathrm{SU}(2). Let Γ=SL2([i])\Gamma=\mathrm{SL}_{2}(\mathbb{Z}[i]), the group of unimodular matrices with Gaussian integer entries, which is indeed a lattice in GG. Then Y=Γ\SY=\Gamma\backslash S is a finite-volume hyperbolic 33-manifold and X=Γ\GX=\Gamma\backslash G is its frame bundle. The action of A={diag(a,a1)a>0}A=\{\operatorname{diag}(a,a^{-1})\mid a>0\} on XX from the right corresponds to the geodesic flow on the frame bundle, with the commuting subgroup M={diag(eiθ,eiθ)}M=\{\operatorname{diag}(e^{i\theta},e^{-i\theta})\} acting by rotating the frame around the tangent vector.

Let H=SL2()GH=\mathrm{SL}_{2}(\mathbb{R})\subset G. Then YH=SL2()\H/SO(2)YY_{H}=\mathrm{SL}_{2}(\mathbb{Z})\backslash H/\mathrm{SO}(2)\subset Y is a finite-volume hyperbolic surface embedded in YY; its unit tangent bundle is the finite-volume HH-orbit XH=SL2()\HXX_{H}=\mathrm{SL}_{2}(\mathbb{Z})\backslash H\subset X. Very relevant to us will be the subset SL2()\HM\mathrm{SL}_{2}(\mathbb{Z})\backslash HM, correspoding to the pullback of the frame bundle of the hyperbolic 33-fold to the surface. Geometrically the unit tangent bundle of YHY_{H} embeds in the frame bundle of YY in multiple ways invariant by the geodesic flow, parametrized by choosing a normal vector to the tangent vector at one point. The set of choices is thus parametrized by the group MM rotating the frame at the chosen point around the tangent vector.

The Casimir element in the universal enveloping algebra of the Lie algebra GG acts on functions on XX and on YY, where it is propotional to the Laplace–Beltrami operator. In addition there is a family of discrete averaing operators acting on functions on XX and commuting with the right GG action, hence also on functions on YY (thought of as KK-invariant functions on XX). These Hecke operators can be constructed as follows: for each gSL2((i))g\in\mathrm{SL}_{2}(\mathbb{Q}(i)) the set [g]=Γ\ΓgΓΓ\SL2((i))[g]=\Gamma\backslash\Gamma g\Gamma\subset\Gamma\backslash\mathrm{SL}_{2}(\mathbb{Q}(i)) is finite, so for a function ff on XX we can set

(1.1) (Tgf)(x)=s[g]f(sx).(T_{g}f)(x)=\sum_{s\in[g]}f(sx)\,.

Since these operators are GG-equivariant they also commute with the Casimir element, and hence with the Laplace–Beltrami operator on YY. They are bounded (the L2L^{2} operator norm of TgT_{g} is at at most the cardinality of [g][g]) and it is not hard to check that the adjoint of TgT_{g} is Tg1T_{g^{-1}}. It is a non-trivial fact that the TgT_{g} commute with each other, and hence are a commuting family of bounded normal operators. For the sequel we will take a different point of view that gives better control of these operators; see the construction in Section 2.5.

The proof of 1 is given in Section 3. Our strategy is the one laid down by Lindenstrauss and followed by later work on AQUE.

  1. (1)

    The microlocal lift of [13] (see also [7] for the case s=0s=0; the original such constructions in much greater generality are due to Shnirelman, Zelditch and Colin de Verdière [15, 18, 4]) shows that any weak-* limit on YY as in 1 is the projection from XX of a limit μ\mu as in 2 which, in addition, is an AMAM-invariant measure.

  2. (2)

    We show that almost every AA-ergodic component of μ\mu has positive entropy under the AA-action. These arguments are standard (see [14] generalizing [1]) but we need to adjust them to handle the fact that MM isn’t finite and that the group is essentially defined over a number field. What is shown is that the measure of an ϵ\epsilon-neighbourhood of a (compact piece) of an orbit xAMxAM in XX decays at least as fast as ϵh\epsilon^{h} for some h>0h>0.

  3. (3)

    The classification of AA-invariant measures in [5] implies that all the ergodic components of μ\mu other than the GG-invariant measure are contained in sets of the form xHMxHM where xHxH is a finite-volume HH-orbit in XX.

  4. (4)

    In a closed HH-orbit ΓgHX\Gamma gH\subset X the entries of gg are algebraic numbers, so by 2 the exceptional possibilies are contained in a countable family of sets each of which has measure zero. It follows that μ\mu is GG-invariant.

  5. (5)

    It was shown in Zaman’s M.Sc. Thesis [17] that the measure μ\mu is a probability measure (a generalization of Soundararajan’s proof [16] for the case of SL2()\SL2()\mathrm{SL}_{2}(\mathbb{Z})\backslash\mathrm{SL}_{2}(\mathbb{R})).

The bulk of the paper is then devoted to realizing step 4. Before discussing how that is achieved, let us remark on its necessity. The group G=SL2()G=\mathrm{SL}_{2}(\mathbb{R}) has no proper semisimple subgroups, so the issue of ruling out components supported on orbits of such subgroups did not arise in the original work of Lindenstrauss. In followup generalizations to higher-rank groups the difficulty was addressed by choosing the lattice Γ\Gamma so that the subgroups HH that could arise did not have finite-volume orbits on XX and GG itself was chosen \mathbb{R}-split so that MM was finite and could be ignored.

We now go further. The group G=SL2()G=\mathrm{SL}_{2}(\mathbb{C}) has H=SL2()H=\mathrm{SL}_{2}(\mathbb{R}) as a relevant subgroup, forcing us to confront the problem of ruling out components supported on finite-volume HH-orbits head-on. It is also not \mathbb{R}-split, forcing us to contend with the infinitude of MM. Indeed, equip a finite-volume HH-orbit xHxH with the HH-invariant probability measure ν\nu. Translating ν\nu by mMm\in M gives further AA-invariant probability measures supported on the subsets xHmxHm, and averaging these measures to gives an AA-invariant probability measure supported on xHMxHM. The construction in step 1 is such that if ν\nu occurs in the ergodic decomposition of μ\mu then so do its MM-translates, so we need to rule out the averaged measure on xHMxHM as occuring in μ\mu (fortunately there are only countable many such subsets). Now previous technology as mentioned in step 2 could show that (ϵ\epsilon-neighborhoods of pieces of) subgroup orbits such as xHxH have small measure111This technology does work in our context too if one uses the new amplifier of Section 5. However, that μ(xH)=0\mu(xH)=0 also automatically follows from the MM-invariance on μ\mu, much in the same way that the \mathbb{R}-invariant of Lebesgue measure shows that it has no atoms. While showing that μ(xHm)=0\mu(xHm)=0 is insufficient (there are uncountably many such subsets), still showing μ(xH)=0\mu(xH)=0 for limits of Hecke eigenfunctions was an important conceptual step in this work. but xHMxHM is not of this form. Instead the subgroup generated by HMHM is all of GG, so the only orbit of a closed subgroup containing xHMxHM is all of XX.

1.4. Sketch of the proof of 2

Let UxHMU\subset xHM be a bounded open set. Then showing that μ(xHM)=0\mu(xHM)=0 amounts to bounding the mass Hecke eigenfunctions can give to ϵ\epsilon-neighbourhoods UϵU_{\epsilon}. In other words, our goal is to prove

(1.2) Uϵ|ϕ(x)|2𝑑x=oϵ(1),\int_{{U_{\epsilon}}}|\phi(x)|^{2}dx=o_{\epsilon}(1),

for any Hecke-eigenfunction ϕ\phi, uniformly in ϕ\phi.

We find a Hecke operator (“amplifier”) τ\tau which acts on ϕ\phi with a large eigenvalue Λ\Lambda yet geometrically “smears” Uϵ{U_{\epsilon}} around the space. The operator τ\tau is a linear combination of operators TgT_{g} so there is a subset supp(τ)SL2([i])\SL2((i))\operatorname{supp}(\tau)\subset\mathrm{SL}_{2}(\mathbb{Z}[i])\backslash\mathrm{SL}_{2}(\mathbb{Q}(i)) so that Λϕ(x)=(τϕ)(x)=ssupp(τ)τ(s).ϕ(sx)\Lambda\phi(x)=(\tau\star\phi)(x)=\sum_{s\in\operatorname{supp}(\tau)}\tau(s).\phi(sx). Choosing τ\tau so that |τ(s)|1\left|{\tau(s)}\right|\leq 1 for each ssupp(τ)s\in\operatorname{supp}(\tau), squaring, applying Cauchy–Schwartz and integrating over Uϵ{U_{\epsilon}} gives

(1.3) μϕ(Uϵ)\displaystyle\mu_{\phi}({U_{\epsilon}}) #supp(τ)Λ2ssupp(τ)μϕ(s.Uϵ)\displaystyle\leq\frac{\#\operatorname{supp}(\tau)}{\Lambda^{2}}\sum_{s\in\operatorname{supp}(\tau)}\mu_{\phi}(s.{U_{\epsilon}})
#supp(τ)Λ2maxssupp(τ)#{ssupp(τ)s.Uϵs.Uϵ}μϕ(ssupp(τ)s.Uϵ)\displaystyle\leq\frac{\#\operatorname{supp}(\tau)}{\Lambda^{2}}\max_{s\in\operatorname{supp}(\tau)}\#\{s^{\prime}\in\operatorname{supp}(\tau)\mid s.{U_{\epsilon}}\cap s^{\prime}.{U_{\epsilon}}\neq\emptyset\}\mu_{\phi}(\bigcup_{s\in\operatorname{supp}(\tau)}s.{U_{\epsilon}})
#supp(τ)Λ2maxssupp(τ)#{ssupp(τ)s.Uϵs.Uϵ}.\displaystyle\leq\frac{\#\operatorname{supp}(\tau)}{\Lambda^{2}}\max_{s\in\operatorname{supp}(\tau)}\#\{s^{\prime}\in\operatorname{supp}(\tau)\mid s.{U_{\epsilon}}\cap s^{\prime}.{U_{\epsilon}}\neq\emptyset\}\,.

The best kind of smearing is thus if there were very few non-empty intersections between the translates s.Uϵs.{U_{\epsilon}}, where a bound would follow as long as we could arrage for Λ2\Lambda^{2} to be large compared with the size of the support of τ\tau, but unfortunately that is too got to hope fore. Thankfully with better geometric and spectral arguments than the naive one given above one requires less stringent control over the intersections. For increasingly sophisticated implementations of this strategy see [11, 1, 14, 2, 12], all in settings where UU is a piece of an orbit xLxL for a subgroup LL. Then an intersection between s.Uϵs.{U_{\epsilon}} and Uϵ{U_{\epsilon}} (say) implies that ss is O(ϵ)O(\epsilon)-close to the bounded neighbourhood xUU1x1xUU^{-1}x^{-1} in xLx1xLx^{-1}; choosing the ss in the support of of τ\tau to have sufficiently small denominators relative to ϵ\epsilon then forces all the sSL2((i))s\in\mathrm{SL}_{2}(\mathbb{Q}(i)) causing the intersections to jointly be contained in a single conjugate of LL. If LL is very small (e.g. is a torus) or if the lattice Γ\Gamma is chosen so that the rational points must generate a torus, there will be very few intersections; it is sometimes even possible to choose the support of τ\tau to avoid them entirely.

Here UU1UU^{-1} (let alone UϵUϵ1U_{\epsilon}U_{\epsilon}^{-1}) is an open subset of GG (a reflection of the fact that HMHM generates GG), so the number of intersections is large: sxUU1x1s\in xUU^{-1}x^{-1} holds generically, and no good bound on the number of intersections is possible. Instead for the first time we go beyond estimating the number of intersections and instead bound the measure of the pieces through an induction argument on their dimension. For this lift the picture to GG, replacing xXx\in X with a representative gGg\in G and LL with the submanifold L=gHML=gHM which is, in fact, an irreducible real algebraic subvariety of GG. An intersection sLLsL\cap L with sSL2((i))s\in\mathrm{SL}_{2}(\mathbb{Q}(i)) is then also a subvariety, and hence one of two possibilities must hold: either the intersection has strictly smaller dimension, in which case we call the intersection (and, by abuse of language, the element ss) transverse, or it is not, in which case we call sLsL (and ss) parallel to LL and must actually have sL=LsL=L by the irreducibility of LL. By induction we may assume that the measures of all transverse intersections s.UϵUϵs.{U_{\epsilon}}\cap{U_{\epsilon}} are small (one needs to show that s.UϵUϵs.{U_{\epsilon}}\cap{U_{\epsilon}} is contained in a decreasing neighbourhood of s.UUs.U\cap U). On the other hand the parallel elements stabilize the subvariety LL and this forces them to lie in a proper subgroup SSL2((i))S\subset\mathrm{SL}_{2}(\mathbb{Q}(i)) (here we gained over the naive attempt which considered the much larger subgroup generated by SL2((i))LL1\mathrm{SL}_{2}(\mathbb{Q}(i))\cap LL^{-1}).

For example, the stabilizer of HMHM in under the left action of GG on itself is exactly HH, so parallel intersections can only arise from sSL2()s\in\mathrm{SL}_{2}(\mathbb{Q}). We now introduce a final idea, allowing us to deal with a situation where this group is not a torus. Let p1(4)p\equiv 1\,(4) be a prime number so that p=(a+bi)(abi)p=(a+bi)(a-bi) for integers a,ba,b and set

gp=diag(a+bi,1a+bi).g_{p}=\operatorname{diag}(a+bi,\frac{1}{a+bi})\,.

It turns out that for any Hecke eigenfunction ϕ\phi, the eigenvalue of λp\lambda_{p} of one of the two operators TgpT_{g_{p}}, Tgp2T_{g_{p}^{2}} is at least comparable to the square root of the size of its support. In addition, the supports [gp],[gp2][g_{p}],[g_{p}^{2}] of these operators do not intersect SL2()\mathrm{SL}_{2}(\mathbb{Q}): complex conjguation exchanges the Gaussian prime a+bia+bi with the distinct Gaussian prime abia-bi, so the ratio of matrices from the two double cosets must contain both primes and cannot be real), so these operators themselves do not cause intersections. Making use of these as building blocks and combining the contribution from many primes we then construct an amplifier which has good spectral properties and at the same time avoids intersections caused by SL2()\mathrm{SL}_{2}(\mathbb{Q}).

Each closed orbit xHMxHM is the projection of L=gHML=gHM where the entries of gg are algebraic but need not be rational, so greater care must be taken to define the complex conjugation which limits the intersection and correspondgly the arithmetic progression from which we select the primes needs to be smaller. Also, we need to consider the stabilizers of the subvarieties of LL that will appear in the recursive argument. We show these stabilizers are either contained in a conjguate of HH (so that a corresponding complex conjugation exists) or are tori (which are already known to cause few intersections).

1.5. Organization of the paper

In Section 2 we fix our notation and examine the algebraic algebraic structure of forms of SL2\mathrm{SL}_{2} over number fields, constructing the complex conjugations which control returns of Hecke translates to real submanifolds. We take the adelic point view of the Hecke operators, which is more convenient than the one used for the introduction. In Section 3 we then reduce 1 to 2. Before giving the proof of 2 in Section 6 we devote Section 4 to classifying the stabilizers of the subvarieties that can occur in the recursion and Section 5 to constructing the amplifier.

Acknowledgemnents.

We would like to thank Elon Lindenstrauss and Manfred Einsiedler for very useful suggestions and fruitful discussions. This work forms part of the PhD thesis of the first author at the Hebrew University of Jerusalem.

2. Notations and background

For a comprehensive reference on the theory of algebraic groups over number fields see [10]. The theory of automorphic forms over such groups is developed, for example, in the textbook [3], The particular case of forms of SL2\mathrm{SL}_{2} over number fields is articulated in detail in [6].

2.1. Unit groups of quaternion algebras

Varieties (including algebraic groups) will be named in blackboard font,

Fix a number field FF, and let V=|F|V=\left|{F}\right| be the set of places of FF, VVV_{\infty}\subset V the set of archimedian places, divided into real and complex places as V=VVV_{\infty}=V_{\mathbb{R}}\sqcup V_{\mathbb{C}}. For a place vVv\in V write FvF_{v} for the completion of FF at vv. If the place is finite write 𝒪vFv\mathcal{O}_{v}\subset F_{v} for the maximal compact subring {xFv|x|v1}\{x\in F_{v}\mid\left|{x}\right|_{v}\leq 1\}, and let ϖv𝒪v\varpi_{v}\in\mathcal{O}_{v} be a uniformizer. We normalize the associated absolute value so that qv=|ϖ1|vq_{v}=\left|{\varpi^{-1}}\right|_{v} is the cardinality of the residue field 𝒪v/ϖ𝒪v\mathcal{O}_{v}/\varpi\mathcal{O}_{v}. For a rational prime pp let Fp=v|pFvF_{p}=\prod_{v|p}F_{v} and similarly set F=vFvF_{\infty}=\prod_{v\mid\infty}F_{v}. We write 𝔸f\mathbb{A}_{\mathrm{f}} for the ring of finite adeles v<(Fv:𝒪v)\prod^{\prime}_{v<\infty}(F_{v}:\mathcal{O}_{v}) and 𝔸F\mathbb{A}_{F} for the ring of adeles F×𝔸fF_{\infty}\times\mathbb{A}_{\mathrm{f}}.

If 𝔾\mathbb{G} is an FF-variety and EE is an FF-algebra write 𝔾(E)\mathbb{G}(E) for the EE-points of 𝔾\mathbb{G}, equipped with the analytic topology if EE is a local field extending FF. For a place vv of FF we also write Gv=𝔾(Fv)G_{v}=\mathbb{G}(F_{v}). We further write G=GG=G_{\infty} for the manifold (Lie group) vGv\prod_{v\mid\infty}G_{v}.

Let DD be a quaternion algebra over FF, that is either the matrix algebra M2(F)M_{2}(F) or a division algebra of dimension 44 over FF (for a first reading let F=(i)F=\mathbb{Q}(i), D=M2(F)D=M_{2}(F). Let 𝔻\mathbb{D} be the variety such that 𝔻(E)=DFE\mathbb{D}(E)=D\otimes_{F}E, and let det:𝔻𝔸1\det\colon\mathbb{D}\to\mathbb{A}^{1} be the reduced norm. Let 𝔾=𝔻1\mathbb{G}=\mathbb{D}^{1} be the algebraic group of elements of reduced norm 11:

𝔾(E)={gDFEdet(g)=1}.\mathbb{G}(E)=\{g\in D\otimes_{F}E\mid\det(g)=1\}\,.

For each place vVv\in V we have the algebra Dv=DFFvD_{v}=D\otimes_{F}F_{v}. When vv is complex necessarily DvM2()D_{v}\simeq M_{2}(\mathbb{C}) and thus GvSL2()G_{v}\simeq SL_{2}(\mathbb{C}). When vv is real DvD_{v} is either the split algebra M2()M_{2}(\mathbb{R}) or Hamilton’s quaternions \mathbb{H}, and correspondingly GvG_{v} is one of the groups SL2()\mathrm{SL}_{2}(\mathbb{R}) and 1SU(2)\mathbb{H}^{1}\simeq\mathrm{SU}(2). We suppose there are ss complex places and r+tr+t real places divided into rr of the first form and tt of the second so that

G=GSL2()r×SL2()s×SU(2)tG=G_{\infty}\simeq\mathrm{SL}_{2}(\mathbb{R})^{r}\times\mathrm{SL}_{2}(\mathbb{C})^{s}\times\mathrm{SU}(2)^{t}

.

2.2. Factoring GG_{\infty} over a number field; complex conjugation

We have the factorizations G=vGvG_{\infty}=\prod_{v\mid\infty}G_{v} (and similarly we can also write Gp=vpGvG_{p}=\prod_{v\mid p}G_{v} where pp is a finite place of \mathbb{Q}). Let g𝔾(F)g\in\mathbb{G}(F) and suppose we have some information on in the image of gg in GvG_{v} for an archimedean place vVv\in V_{\mathbb{C}}. A-priori there seems to be no way to translate this to information about gg at a particular finite place over pp. However, it turns out that this is possible if we restrict our attention to a positive density subset of the primes: for such primes we will enumerate the respective places of FF over the two places ,p\infty,p of \mathbb{Q} in such a way that the factors in some sense correspond.

Moreover, for a complex place vVv\in V_{\mathbb{C}}, the usual extension of scalars realizes GvG_{v} as the group of complex points of the \mathbb{C}-group 𝔾×F\mathbb{G}\times_{F}\mathbb{C}. However, when thought of as a Lie group this group has closed subgroups which are not complex – the key example for us being the subgroup SL2()\mathrm{SL}_{2}(\mathbb{R}) of SL2()\mathrm{SL}_{2}(\mathbb{C}). We thus would like to think of the group GvG_{v} as the group of real points of an algebraic group defined over an extension of FF.

In this section we address both concerns simultaneously: given the complex place vv we construct a number field EE equipped with a real place ww and an algebraic group 𝔾~\tilde{\mathbb{G}} defined over EE factoring as an EE-group into a product 𝔾~=i=1r+t+2s𝔾~i\tilde{\mathbb{G}}=\prod_{i=1}^{r+t+2s}\tilde{\mathbb{G}}_{i} such that:

  1. (1)

    The group 𝔾(F)\mathbb{G}(F) embeds in 𝔾~(E)\tilde{\mathbb{G}}(E).

  2. (2)

    We have an isomorphism

    𝔾~(Ew)=i=1r+t+s𝔾~i(Ew)G\tilde{\mathbb{G}}(E_{w})=\prod_{i=1}^{r+t+s}\tilde{\mathbb{G}}_{i}(E_{w})\to G_{\infty}

    where the factors correspond.

  3. (3)

    For a rational prime pp splitting completely in a particular quaratic extension NN of EE we have a place EpE\hookrightarrow\mathbb{Q}_{p} such that For 1is1\leq i\leq s we have 𝔾~i(p)(SL2(p))2\tilde{\mathbb{G}}_{i}(\mathbb{Q}_{p})\simeq(\mathrm{SL}_{2}(\mathbb{Q}_{p}))^{2}, for s<ir+t+ss<i\leq r+t+s we have 𝔾~i(p)SL2(p)\tilde{\mathbb{G}}_{i}(\mathbb{Q}_{p})\simeq\mathrm{SL}_{2}(\mathbb{Q}_{p}) and the resulting factorization of 𝔾~(p)\tilde{\mathbb{G}}(\mathbb{Q}_{p}) is isomorphic to the factorization of GpG_{p} into 2s+r+t2s+r+t copies of SL2(p)\mathrm{SL}_{2}(\mathbb{Q}_{p}) coming from the 2s+r+t2s+r+t places over FF over p\mathbb{Q}_{p}.

  4. (4)

    The latter two identifications are compatible with first embedding, in such a way that if an element γ𝔾(F)\gamma\in\mathbb{G}(F) embeds to a real-valued matrix in 𝔾~1(Ev)\tilde{\mathbb{G}}_{1}(E_{v}) (say) then its image in 𝔾~1(p)\tilde{\mathbb{G}}_{1}(\mathbb{Q}_{p}) lies in the diagonal subgroup of 𝔾~1(p)(SL2(p))2\tilde{\mathbb{G}}_{1}(\mathbb{Q}_{p})\simeq(\mathrm{SL}_{2}(\mathbb{Q}_{p}))^{2}. This will allow us to choose Hecke operators which "avoid" a real subgroup in a particular complex place.

Thus let vv be a complex place of FF and let NN be any finite Galois extension of \mathbb{Q} containing FF (in Section 3 NN will the Galois closure of a splitting field of DD; in Section 5 we make a different choice depending on the subvariety we are trying to avoid). Let ww be a place of NN extending vv, let cwGal(N/)c_{w}\in\operatorname{Gal}(N/\mathbb{Q}) the element acting as complex conjugation in the completion of NN at vv, let E=NcwE=N^{c_{w}} be the fixed field of cwc_{w}, and also write ww for the restriction of ww to EE – a real place of that field. The situation is simpler when EE contains FF (e.g. our running example of F=N=(i)F=N=\mathbb{Q}(i) where E=E=\mathbb{Q}) but we will not assume this is the case.

Writing F=[x]/(f)F=\mathbb{Q}[x]/(f) for an irreducible polynomial f[x]f\in\mathbb{Q}[x] and factoring ff in the extensions NN and EE of \mathbb{Q} we see that the EE-algebra A=FEA=F\otimes_{\mathbb{Q}}E is étale, we have AiEiA\simeq\prod_{i}E_{i} which we can interpret both as an isomorphism of FF-algebras and as an isomorphism of EE-algebras. From the first point view composing each embedding FEiF\hookrightarrow E_{i} with the embedding w:Eiw\colon E_{i}\to\mathbb{C} we obtain an enumeration of the archimedean places of FF, with EiNE_{i}\simeq N for complex places (say those are indexed by 1is1\leq i\leq s) and EiEE_{i}\simeq E for real places (say for s<is+r+ts<i\leq s+r+t). Without loss of generality we assume the inclusion FEiF\hookrightarrow E_{i}\hookrightarrow\mathbb{C} is the place vv fixed above.

Now thinking of AA as a EE-algebra let 𝔾~=𝔾×FA\tilde{\mathbb{G}}=\mathbb{G}\times_{F}A thought of as an algebraic group over EE. Equivalently 𝔾~=(ResF𝔾)×E\tilde{\mathbb{G}}=\left(\operatorname{Res}^{F}_{\mathbb{Q}}\mathbb{G}\right)\times_{\mathbb{Q}}E (Weil restriction of scalars). The factorization of AA then gives a factorization 𝔾~=i𝔾~i\tilde{\mathbb{G}}=\prod_{i}\tilde{\mathbb{G}}_{i} as groups over EE, where 𝔾~i(Ew)𝔾(Fv)\tilde{\mathbb{G}}_{i}(E_{w})\simeq\mathbb{G}(F_{v}) if vv is the iith archimedean place of FF under the identification above. Having done this we also write GiG_{i} for GvG_{v}, especially in situations where we would like vv to vary over finite places.

The embedding FAF\hookrightarrow A gives an inclusion 𝔾(F)𝔾~(E)\mathbb{G}(F)\to\tilde{\mathbb{G}}(E).

Finally let pp be a rational prime which splits completely in NN (hence also in FNF\subset N). Choosing any place wp:Npw_{p}\colon N\to\mathbb{Q}_{p} lying over pp all embeddings of FF into p\mathbb{Q}_{p} factor through wpw_{p} (as above, with different embeddings of FF into NN). Restricting wpw_{p} to EE we then have

EiEp{pEi=Ep×pEi=NE_{i}\otimes_{E}\mathbb{Q}_{p}\simeq\begin{cases}\mathbb{Q}_{p}&E_{i}=E\\ \mathbb{Q}_{p}\times\mathbb{Q}_{p}&E_{i}=N\end{cases}

since when Ei=NE_{i}=N there are two places of NN lying over the place wpw_{p} of EE. Furthermore, Gal(N/E)={cw,id}\operatorname{Gal}(N/E)=\{c_{w},\operatorname{id}\} acts transitively on these places so that cwc_{w} swaps them.

Finally let γ𝔾(F)\gamma\in\mathbb{G}(F) and suppose that its image in 𝔾~1(Ew)\tilde{\mathbb{G}}_{1}(E_{w}) lies in the subgroup g1𝔾~1(Ew)cwg11g_{1}\tilde{\mathbb{G}}_{1}(E_{w})^{c_{w}}g_{1}^{-1} for some g1𝔾~1(E)g_{1}\in\tilde{\mathbb{G}}_{1}(E) (we think of 𝔾~1(Ew)cw\tilde{\mathbb{G}}_{1}(E_{w})^{c_{w}} as the "standard" copy of SL2()\mathrm{SL}_{2}(\mathbb{R}) in 𝔾~1(Ew)SL2()\tilde{\mathbb{G}}_{1}(E_{w})\simeq\mathrm{SL}_{2}(\mathbb{C})). Since the image of g11γg1in𝔾~1(E)g_{1}^{-1}\gamma g_{1}in\tilde{\mathbb{G}}_{1}(E) is fixed by cwc_{w}, this persists when we embed our group in p\mathbb{Q}_{p} using the place wpw_{p}, and it follows that the image of γ\gamma in 𝔾~1(Ewp)\tilde{\mathbb{G}}_{1}(E_{w_{p}}) lies in g1𝔾~1(Ewp)cwg11g_{1}\tilde{\mathbb{G}}_{1}(E_{w_{p}})^{c_{w}}g_{1}^{-1}, in other words in a particular conjugate of the diagonal subgroup 𝔾~1(Ewp)cw\tilde{\mathbb{G}}_{1}(E_{w_{p}})^{c_{w}} (so-called because when we identified 𝔾~1(Ewp)cw(SL2(p))2\tilde{\mathbb{G}}_{1}(E_{w_{p}})^{c_{w}}\simeq\left(\mathrm{SL}_{2}(\mathbb{Q}_{p})\right)^{2} the Galois automorphism cwc_{w} acts by exchanging the factors).

Before going any further, it may be helpful to have two concrete examples of the setup so far.

Example 3.

Let F=(ξ)F=\mathbb{Q}(\xi) where ξ3=2\xi^{3}=2 and let D=M2(F)D=M_{2}(F) be the matrix algebra over F. The field FF has one real place vv^{\prime} (coming from the unique real cube root of 22) and one complex place vv (coming from the two non-real cube roots), so the corresponding Lie group is SL2()×SL2()\mathrm{SL}_{2}(\mathbb{C})\times\mathrm{SL}_{2}(\mathbb{R}) with the symmetric space 3×2\mathbb{H}^{3}\times\mathbb{H}^{2}. Let us now examine the alternative realization and its use.

For this let ω\omega be a cube root of unity so that N=F(ω)=F(3)N=F(\omega)=F(\sqrt{-3}) is the normal closure of FF in which the Galois conjugates of ξ\xi are ξ,ξω,ξω2\xi,\xi\omega,\xi\omega^{2}.

We choose the complex embedding vv so that222We use the mumber theory convention e(z)=exp(πiz)e(z)=\exp(\pi iz) v(ξ)=23e(1/3)v(\xi)=\sqrt[3]{2}e(1/3) and extend this to NN by w(σ)=e(1/3)w(\sigma)=e(1/3). The Galois group Gal(N/)\operatorname{Gal}(N/\mathbb{Q}) is the full permutation group on the cube roots of 22 in NN; among its elements let cwc_{w} be the one fixing ξσ2\xi\sigma^{2} and exchanging ξ,ξσ\xi,\xi\sigma. Note that w(ξσ2=23w(\xi\sigma^{2}=\sqrt[3]{2}\in\mathbb{C} so cwc_{w} is complex conjguation at ww, and one can also check that cw(σ)=σ2c_{w}(\sigma)=\sigma^{2}. We thus obtain the subfield E=Ncw=(ξσ2)NE=N^{c_{w}}=\mathbb{Q}(\xi\sigma^{2})\subset N, a field which is abstractly isomorphic to FF but disjoint from it as a subfield of NN.

Next, we have FE=E[x]/(x32)E[x]/(x2+ξσ2x+ξ2σ)(xξσ2)E1E2F\otimes_{\mathbb{Q}}E=E[x]/(x^{3}-2)\simeq E[x]/(x^{2}+\xi\sigma^{2}x+\xi^{2}\sigma)(x-\xi\sigma^{2})\simeq E_{1}\oplus E_{2} where E1N=E(ω)E_{1}\simeq N=E(\omega) and E2EE_{2}\simeq E (and also E2FE_{2}\simeq F!). Extending scalars in DD we obtain the algebra

D~\displaystyle\tilde{D} =DF(E1E2)\displaystyle=D\otimes_{F}(E_{1}\oplus E_{2})
(DFE1)(DFE2)\displaystyle\simeq(D\otimes_{F}E_{1})\oplus(D\otimes_{F}E_{2}) (DFN)(DFE)\displaystyle\simeq(D\otimes_{F}N)\oplus(D\otimes_{F}E)

It follows that D~M2(N)M2(E)\tilde{D}\simeq M_{2}(N)\oplus M_{2}(E). Restricting to elements of determinant one produces from the algebras DD, D~\tilde{D} the FF-group 𝔾(K)={gM2(K)det(g)=1}\mathbb{G}(K)=\{g\in M_{2}(K)\mid\det(g)=1\} and the EE-group 𝔾~(K)={(g1,g2)M2(K[ω])×M2(K)det(g1)=det(g2)=1}\tilde{\mathbb{G}}(K)=\{(g_{1},g_{2})\in M_{2}(K[\omega])\times M_{2}(K)\mid\det(g_{1})=\det(g_{2})=1\} for any field extension KK of FF, EE respectively. It may seem odd to distinguish between EE and FF (which are after all isomorphic fields in this example), so let us put this distinction use. We have G=SL2()×SL2()G_{\infty}=\mathrm{SL}_{2}(\mathbb{C})\times\mathrm{SL}_{2}(\mathbb{R}), but also 𝔾~(Ew)SL2(Nw)×SL2(Ew)SL2()×SL2()\tilde{\mathbb{G}}(E_{w})\simeq\mathrm{SL}_{2}(N_{w})\times\mathrm{SL}_{2}(E_{w})\simeq\mathrm{SL}_{2}(\mathbb{C})\times\mathrm{SL}_{2}(\mathbb{R}). Furthermore the isomorphism of the two groups G𝔾~(Ew)G_{\infty}\simeq\tilde{\mathbb{G}}(E_{w}) preserves the diagonal inclusion of SL2(F)\mathrm{SL}_{2}(F) in both of them – exactly because the complex place ww of NN restricts to the complex place of FF but the real place of EE. Furthermore complex conjugation in the first factor is now algebraic: it is the obvious automorphism of the algebra K[ω]=K[x]/(x2+x+1)K[\omega]=K[x]/(x^{2}+x+1).

Thus let pp be a rational prime which splits completely in NN. Let v1,v2,v3v_{1},v_{2},v_{3} be the three places of FF lying over pp, giving us the group Gp=SL2(Fv1)×SL2(Fv2)×SL2(Fv3)G_{p}=\mathrm{SL}_{2}(F_{v_{1}})\times\mathrm{SL}_{2}(F_{v_{2}})\times\mathrm{SL}_{2}(F_{v_{3}}) together with the diagonal embedding of 𝔾(F)=SL2(F)\mathbb{G}(F)=\mathrm{SL}_{2}(F) in G×GpG_{\infty}\times G_{p}. Later we will obtain elements γ𝔾(F)\gamma\in\mathbb{G}(F) whose image in the factor GvSL2()G_{v}\simeq\mathrm{SL}_{2}(\mathbb{C}) lie in SL2()\mathrm{SL}_{2}(\mathbb{R}), that is are fixed by the complex conjugation, and we will want to choose elements of GpG_{p} which "avoid" those γ\gamma in some sense. For this let w1w_{1} be a place of EE lying over pp. Observe that since pp splits in NN, p\mathbb{Q}_{p} contains the cube roots of unity, Ew1[ω]ppE_{w_{1}}[\omega]\simeq\mathbb{Q}_{p}\oplus\mathbb{Q}_{p} and this isomorphism gives us the other two places of FF in

𝔾~(Ew1)SL2(pp)×SL2(p)Gp.\tilde{\mathbb{G}}(E_{w_{1}})\simeq\mathrm{SL}_{2}(\mathbb{Q}_{p}\oplus\mathbb{Q}_{p})\times\mathrm{SL}_{2}(\mathbb{Q}_{p})\simeq G_{p}\,.

However, all this is compatible with the action of cwc_{w} in the first factor, mapping ω\omega to ω2=1ω\omega^{2}=-1-\omega, which amounts to exchanging the two copies of p\mathbb{Q}_{p}. Accordingly let γ𝔾(F)\gamma\in\mathbb{G}(F) be such that v(γ)v(\gamma) is real. Then the image of γ\gamma in 𝔾~(Ew)\tilde{\mathbb{G}}(E_{w}) is fixed by cwc_{w}, so the same must hold for the image in 𝔾~(E)\tilde{\mathbb{G}}(E) and hence the image in 𝔾~(Ew1)\tilde{\mathbb{G}}(E_{w_{1}}). In short, we have chosen the factors in the isomorphism Gp(SL2(p))3G_{p}\simeq(\mathrm{SL}_{2}(\mathbb{Q}_{p}))^{3} so that those γ\gamma which are real in the first complex place are diagonal (have the same image) in the first two pp-adic places. We could then "avoid" (in the precise sense we need) such γ\gamma by choosing Hecke operators (see below) living at exactly one of the two places exchanged by cwc_{w}.

Example 4.

Continuing with the same field FF, let D=(ξ,ξF)D=\left(\dfrac{-\xi,-\xi}{F}\right) be the quaternion algebra with FF-basis 1,i,j,k1,i,j,k such that i2=j2=ξi^{2}=j^{2}=-\xi and ij=ji=kij=-ji=k. The reduced norm of this algebra is then the quadratic form

(2.1) det(a+bi+cj+dk)=a2+ξb2+ξc2+ξ2d2\det(a+bi+cj+dk)=a^{2}+\xi b^{2}+\xi c^{2}+\xi^{2}d^{2}

which is definite in the real embedding vv^{\prime} for which v(ξ)=23v^{\prime}(\xi)=\sqrt[3]{2}. Then DvD_{v^{\prime}} is a noncommutative real division algebra, in other words Hamilton’s quaternions – whereas we still have DvM2()D_{v}\simeq M_{2}(\mathbb{C}) since the only complex division algebra is \mathbb{C}.

Viewing the quaternions as a 2-dimensional vector spaces and letting the invertible quaternions act on themselves by multiplication shows that the group of norm-11 quaternions isomorphic to SU(2)\mathrm{SU}(2), and hence for the group 𝔾\mathbb{G} of norm-11 elements in DD we have

GSL2()×SU(2)G_{\infty}\simeq\mathrm{SL}_{2}(\mathbb{C})\times\mathrm{SU}(2)

With the corresponding symmetric space 3\mathbb{H}^{3}. The rest of the discussion in the previous example continues unchanged: for any prime pp we can still realize Gp𝔾~(Ew1)G_{p}\simeq\tilde{\mathbb{G}}(E_{w_{1}}) so that complex conjugation in the first factor of 𝔾~\tilde{\mathbb{G}} corresponds to both complex conjugation in GvG_{v} and to swapping the first two factors in Gp(SL2(p))3G_{p}\simeq\left(\mathrm{SL}_{2}(\mathbb{Q}_{p})\right)^{3}.

This elucidates the extra difficulty of proving our results in the case of lattices such as SL2(𝒪F)\mathrm{SL}_{2}(\mathcal{O}_{F}) which do act cocompactly on 3\mathbb{H}^{3} but in a more complicated fashion than a Bianchi group such as SL2([i])\mathrm{SL}_{2}(\mathbb{Z}[i]). Consider an element of SL2(F)\mathrm{SL}_{2}(F) which is real in the complex embedding. When F=[i]F=\mathbb{Z}[i] (say) complex conjugation is a Galois automorphism of FF and it is clear how it acts on SL2(Fv1)×SL2(Fv2)\mathrm{SL}_{2}(F_{v_{1}})\times\mathrm{SL}_{2}(F_{v_{2}}) for the two places of FF lying over a rational prime pp splitting in FF; elements of SL2((i))\mathrm{SL}_{2}(\mathbb{Q}(i)) which are real in the complex embedding lie in SL2()\mathrm{SL}_{2}(\mathbb{Q}) and clearly embed diagonally. On the other hand in the present example FF has no Galois automorphism and so showing that elements of D1D^{1} that are real at the complex place have identical images at two of the three pp-adic places most naturally involves going beyond FF.

2.3. The real group

We return to the isomorphism G(SL2())s(SL2())r(SU(2))tG_{\infty}\simeq\left(\mathrm{SL}_{2}(\mathbb{C})\right)^{s}\left(\mathrm{SL}_{2}(\mathbb{R})\right)^{r}\left(\mathrm{SU}(2)\right)^{t} and fix some subgroups of this group.

At each infinite vv where DvD_{v} splits, let AvGvA_{v}\subset G_{v} be the group corresponding to the group of diagonal matrices with positive real entries under the isomorphism above. Also let KvK_{v} be a compatible maximal compact subgroup (corresponding to the subgroup SU(2)\mathrm{SU}(2) at a complex place, SO(2)\mathrm{SO}(2) at a real place). From these let Mv=ZKv(Av)M_{v}=Z_{K_{v}}(A_{v}) be the centralizer of AvA_{v} in KvK_{v}, so that MvM_{v} is the group {±I}SL2()\{\pm I\}\subset\mathrm{SL}_{2}(\mathbb{R}) at a real place or the group U(1)SL2()U(1)\subset\mathrm{SL}_{2}(\mathbb{C}) of diagonal matrices with inverse entries both of modulus 11. Observe that in either case the group AvMvA_{v}M_{v} consists of the FvF_{v}-points of a maximal FvF_{v}-split algebraic torus of 𝔾×FFv\mathbb{G}\times_{F}F_{v}.

At the real places vv where DvD_{v} remains a division algebra Kv=GvSU(2)K_{v}=G_{v}\simeq\mathrm{SU}(2) is a maximal subgroup. With this choice K=K=vKvK=K_{\infty}=\prod_{v\mid\infty}K_{v} is a maximal compact subgroup of GG and G/K((3))s((2))rG/K\simeq\left(\mathbb{H}^{(3)}\right)^{s}\left(\mathbb{H}^{(2)}\right)^{r}.

We fix (arbitrarily) once and for all a left-invariant Riemannian metric on GG (equivalently, a positive definite quadratic form on LieG\operatorname{Lie}G). This induces a left-invariant metric on X=Γ\GX=\Gamma\backslash G where the quotient map is non-expansive. For a subset UGU\subset G we write Uϵ{U_{\epsilon}} for its ϵ\epsilon-neighborhood with respect to this metric. This metric is used in Sections 3 and 6, but in both cases we can first restrict our attention to compact subsets of GG. The choice of metric thus affects some overall constants but not the bottom line. For example in Section 3 we establish bounds of the form μ(Uϵ)Cϵh\mu({U_{\epsilon}})\leq C\epsilon^{h}; with a different metric Uϵ{U_{\epsilon}} would be contained in UcϵU_{c\epsilon} with respect to the new one and the bound would be identical except for the value of the constant CC.

2.4. pp-adic and adelic groups

Let RDR\subset D be an order, that is a subring which an 𝒪F\mathcal{O}_{F}-lattice in DD. Then for every finite place vv of FF, Rv=R𝒪F𝒪vR_{v}=R\otimes_{\mathcal{O}_{F}}\mathcal{O}_{v} is an order in DvD_{v}, that is a compact open 𝒪v\mathcal{O}_{v}-subalgebra of DvD_{v}. Its group of units Rv1R_{v}^{1} is then a compact subgroup of GvG_{v} which is a maximal compact subgroup KvK_{v} at almost all places.

For all but finitely many places, DvM2(Fv)D_{v}\simeq M_{2}(F_{v}) and then GvSL2(Fv)G_{v}\simeq\mathrm{SL}_{2}(F_{v}) and KvSL2(𝒪v)K_{v}\simeq\mathrm{SL}_{2}(\mathcal{O}_{v}) (we don’t choose KvK_{v} at the finitely many places where DvD_{v} is a division algebra or where RvR_{v} is not a maximal order).

Let 𝔾(𝔸f)=v<(Gv:Kv)\mathbb{G}(\mathbb{A}_{\mathrm{f}})=\prod^{\prime}_{v<\infty}(G_{v}:K_{v}) be the restricted direct product of the GvG_{v} and let 𝔾(𝔸)=G×𝔾(𝔸f)\mathbb{G}(\mathbb{A})=G_{\infty}\times\mathbb{G}(\mathbb{A}_{\mathrm{f}}). Then the diagonal embedding 𝔾(F)𝔾(𝔸)\mathbb{G}(F)\hookrightarrow\mathbb{G}(\mathbb{A}) realizes 𝔾(F)\mathbb{G}(F) as a lattice there. Fix an open compact subgroup Kf𝔾(𝔸f)K_{\mathrm{f}}\subset\mathbb{G}(\mathbb{A}_{\mathrm{f}}). Then there exists a finite set SS of finite places (including all places where KvK_{v} was left undefined above) such that Kf=H×S∌v<KvK_{\mathrm{f}}=H\times\prod_{S\not\ni v<\infty}K_{v} for some open compact subgroup H<vSGvH<\prod_{v\in S}G_{v}. We will generally ignore all places in SS.

Let Γ=𝔾(F)Kf\Gamma=\mathbb{G}(F)\cap K_{\mathrm{f}}. Its image in GG_{\infty} is a lattice, and we have the identification333We use here that 𝔾\mathbb{G} is a form of the simply connected algebraic group SL2\mathrm{SL}_{2}; in general the adelic quotient would correspond to a disjoint union of quotients Γ\G\Gamma\backslash G

X=Γ\G𝔾(F)\𝔾(𝔸)/KfX=\Gamma\backslash G\simeq\mathbb{G}(F)\backslash\mathbb{G}(\mathbb{A})/K_{\mathrm{f}}

given by mapping the coset Γg\Gamma g_{\infty} with the double coset [g,1]=𝔾(F)(g,1)Kf[g_{\infty},1]=\mathbb{G}(F)(g_{\infty},1)K_{\mathrm{f}}.

Conversely every congruence subgroup of GG (with the respect to the FF-rational structure 𝔾\mathbb{G}) contains a lattice Γ\Gamma as above. Since equidistribution modulo Γ\Gamma implies equidistribution modulo any overgroup (assuming the eigenfunctions are properly invariant) it suffices to consider the case above.

We further set Y=X/K=Γ\SY=X/K=\Gamma\backslash S where S=G/KS=G/K is the symmetric space of GG. Equipping XX with the GG-invariant probability measure we study functions ("automorphic forms") in L2(X)L^{2}(X), and identify L2(Y)L^{2}(Y) with the subset of KK-invariant functions. For each noncompact factor GiG_{i} of GG let ωi\omega_{i} be the Casimir element in the universal enveloping algebra U(𝔤i)U(\mathfrak{g}_{i}\otimes_{\mathbb{R}}\mathbb{C}). Then ωi\omega_{i} acts on the right on smooth functions on GG and XX, equivariantly with respect to the right GG-actions. It thus descends to a differential operator on functions on SS where it coincides with the Laplace–Beltrami operator on the irreducible symmetric space Gi/KiG_{i}/K_{i} (with the Laplace–Beltrami operator of SS being i=1r+sωi\sum_{i=1}^{r+s}\omega_{i}). A Maass form on YY is a functions ψL2(Y)\psi\in L^{2}(Y) which is a joint eigenfunctions of the ωi\omega_{i}.

2.5. Hecke operators

For every finite place vv, the convolution algebra of locally constant compactly supported functions on the totally disconnected group GvG_{v} acts on the right by convolution on the space of smooth functions on 𝔾(F)GA\mathbb{G}(F)\\ GA. At a place vv where KvK_{v} is defined and contained in KfK_{\mathrm{f}}, the subalgebra of bi-KvK_{v}-invariant functions v=(Gv:Kv)=Cc(Kv\Gv/Kv)\mathcal{H}_{v}=\mathcal{H}(G_{v}:K_{v})=C_{\mathrm{c}}^{\infty}(K_{v}\backslash G_{v}/K_{v}) preserves the subspace of right-KfK_{\mathrm{f}} invariant functions, and hence acts on the space of functions on 𝔾(F)\𝔾(𝔸)/Kf=Γ\G\mathbb{G}(F)\backslash\mathbb{G}(\mathbb{A})/K_{\mathrm{f}}=\Gamma\backslash G. We note that the actions of the different v\mathcal{H}_{v} commute with each other and with the right GG-action on our space, and call the algebra of operators on functions generated by all of them the Hecke algebra.

Let 𝒫\mathcal{P} the set of rational primes pp which split completely in NN, and such that every place vv of FF above pp has KvK_{v} as a factor of KfK_{\mathrm{f}}, a set of positive natural density in the primes footnoteThose are the primes whose Artin symbol in Gal(f)=Gal(N/)\operatorname{Gal}(f)=\operatorname{Gal}(N/\mathbb{Q}) is trivial and the claim follows immediately from the effective Chebotarev Density Theorem.. For such pp let p=vpv\mathcal{H}_{p}=\otimes_{v\mid p}\mathcal{H}_{v} be the "Hecke algebra at pp". We will only consider Hecke operators in the restricted Hecke algebra =p𝒫p\mathcal{H}=\otimes^{\prime}_{p\in\mathcal{P}}\mathcal{H}_{p} generated by the p\mathcal{H}_{p}.

Since the right actions of GvG_{v} and GiG_{i} on 𝔾(F)\𝔾(𝔸F)\mathbb{G}(F)\backslash\mathbb{G}(\mathbb{A}_{F}) commute, the Hecke operators commute with the differential operators of the previous section. A Hecke–Maass form is a Maas form ψL2(Y)\psi\in L^{2}(Y) which also a joint eigenfunction of the Hecke algebra. Since the group actions commute it also follows that if ϕL2(X)\phi\in L^{2}(X) is any other element of the irreducible representation generated by ϕ\phi then ψ\psi is also an eigenfunction of the Hecke algebra with the same eigenvalues as ψ\psi.

3. Homogeneity

In this section we invoke the expected machinery to deduce 1 from 2. The main new ingredient (which was already known to experts) is the statement of a so-called “diophantine lemma” (10 below) for groups defined over a number field.

As described in the introduction let {ψj}j1L2(Y)\{\psi_{j}\}_{j\geq 1}\subset L^{2}(Y) be a normalized sequence of Hecke–Maass forms with Laplace eigenvalues tending to infinity. Let μ¯j\bar{\mu}_{j} be the corresponding probability measures on YY (recall that those are the measures with density |ψj(x)|2\left|{\psi_{j}(x)}\right|^{2} with respect to the Riemannian measure); our ultimate goal is to show that the μ¯j\bar{\mu}_{j} converge to the normalized Riemannian volume on YY, or equivalently that this is the only subsequential limit. Accordingly (passing a subsequence) we assume the μ¯j\bar{\mu}_{j} converse weak-* to a measure μ¯\bar{\mu} on YY (which is a probability measure even when YY is non-compact as shown in [17]).

Again passing to a subsequence there is an infinite place vv such GvG_{v} is non-compact and such that the Laplace–Beltrami eigenvalues of ψj\psi_{j} with respect to the Laplace operator at vv tend to infinity; without loss of generality we may assume it is the first place in our enumeration. Then by the microlocal lift of [13] there are Hecke eigenfunctions ϕjL2(X)\phi_{j}\in L^{2}(X) such that any subsequential weak-* limit μ\mu of the associated measures μj\mu_{j} has the following properties:

  1. (1)

    μ\mu projects to μ¯\bar{\mu} under the map XYX\to Y (in particular, μ\mu is a probability measure).

  2. (2)

    μ\mu is A1A_{1}-invariant.

Passing to a subsequence yet again we may assume that the μj\mu_{j} themselves converge, and would like to show that the limit μ\mu is the GG-invariant probability measure on XX. At this point the differential operators exit the stage: in the sequel we only use the fact that ϕj\phi_{j} are normalized Hecke eigenfunctions on XX and that μ\mu has the two properties above.

The rest of the section is divided as follows: in Section 3.1 we show that every ergodic component of μ\mu has positive entropy with respect to the action of A1A_{1}. In Section 3.2 we then invoke a measure rigidity theorem of Einsiedler–Lindenstrauss to and interpret its results, classifying the possible ergodic components of μ\mu.

3.1. Positive Entropy

Let aA1a\in A_{1} be non-trivial. Under the isomorphism G1SL2(F1)G_{1}\simeq\mathrm{SL}_{2}(F_{1}), aa is a diagonal matrix with distinct real positive entries, so T1=ZG1(a)T_{1}=Z_{G_{1}}(a) is the group A1M1A_{1}M_{1} of all diagonal matrices in G1G_{1}. Setting G,2=i2GiG_{\infty,\geq 2}=\prod_{i\geq 2}G_{i}, the centralizer of aa in GG is then T=T1i2GiT=T_{1}\cdot\prod_{i\geq 2}G_{i}.

For a compact neighborhood of the identity UTU\subset T recall our notation Uϵ{U_{\epsilon}} for an ϵ\epsilon-neighborhood of UU in GG. We will establish the following result:

Proposition 5.

There is a constant h>0h>0 such that for any compact subset ΩG\Omega\subset G (expected to be large) and any UU, we have for any ϵ\epsilon small enough (depending on Ω\Omega, CC) and any Hecke eigenfunction ϕjL2(X)\phi_{j}\in L^{2}(X) that for all gΩg\in\Omega,

μj(ΓgUϵ)Ω,Cϵh.\mu_{j}\left(\Gamma g{U_{\epsilon}}\right)\ll_{\Omega,C}\epsilon^{h}\,.

The key point is that the implied constant is independent of ϕ\phi, so that μ\mu satisfies the same inequality.

This result is essentially contained in [14] (on some level already in [1]) except that [14] assumes that GG is \mathbb{R}-split, which is the not the case for SL2()\mathrm{SL}_{2}(\mathbb{C}). In fact all that is needed there is that the centralizer TvT_{v} at some infinite place GvG_{v} is a torus, and even weaker hypotheses suffices (which the author plan to addressed in future work).

3.1.1. Diophantine Lemma

We begin by reviewing a notion of "denominator" for elements of FF and 𝔾(F)\mathbb{G}(F). The construction is the natural generalization to number fields of the notion used in [14] for groups over the rationals. For further discussion see 11.

Definition 6.

The denominator of xvFvx_{v}\in F_{v} is the natural number denomv(xv)=max{|xv|v,1}\operatorname{denom}_{v}(x_{v})=\max\{\left|{x_{v}}\right|_{v},1\}. Equivalently if xv=ϖvkyx_{v}=\varpi_{v}^{k}y with y𝒪v×y\in\mathcal{O}_{v}^{\times} then

denomv(xv)={qvkk01k0.\operatorname{denom}_{v}(x_{v})=\begin{cases}q_{v}^{-k}&k\leq 0\\ 1&k\geq 0\end{cases}\,.

We now extend this definition. First, for x𝔸fx\in\mathbb{A}_{\mathrm{f}} or x𝔸Fx\in\mathbb{A}_{F} we set

denomx=v<denomv(xv),\operatorname{denom}{x}=\prod_{v<\infty}\operatorname{denom}_{v}(x_{v})\,,

where all but finitely many of the factors are 11 since xv𝒪vx_{v}\in\mathcal{O}_{v} for all but finitely many vv. Second for xFx\in F let denom(x)\operatorname{denom}(x) be the denominator of its image in 𝔸F\mathbb{A}_{F}.

We further extend the definition to matrix algebras over the above rings. Specifically for xvMN(Fv)x_{v}\in M_{N}(F_{v}) let denomv(xv)\operatorname{denom}_{v}(x_{v}) be the largest of the denominators of the matrix entries (equivalently this is the denominator of the fractional ideal they generate), and again extend this to MN(𝔸f)M_{N}(\mathbb{A}_{\mathrm{f}}), MN(𝔸F)M_{N}(\mathbb{A}_{F}) and MN(F)M_{N}(F) by multiplying over the places and restriction, respectively.

Finally, the product formula v|x|v=1\prod_{v}\left|{x}\right|_{v}=1 for xF×x\in F^{\times} implies v|x|vdenomx1\prod_{v\mid\infty}\left|{x}\right|_{v}\operatorname{denom}{x}\geq 1 for all nonzero xFx\in F and hence also for xGLn(F)x\in GL_{n}(F).

Remark 7.

We could have defined the denominator of xMN(𝔸f)x\in M_{N}(\mathbb{A}_{\mathrm{f}}) by taking the largest of the denominators of its entries (call that denom\operatorname{denom}^{\prime} for the nonce); the two notions are equivalent in that denom(x)denom(x)denom(x)n2\operatorname{denom}^{\prime}(x)\leq\operatorname{denom}(x)\leq\operatorname{denom}^{\prime}(x)^{n^{2}} for all xx. Our choice agrees with defining for xMN(F)x\in M_{N}(F) the denominator as the denominator of the fractional ideal generated by the matrix entries. In the sequel the precise choice of denominator affects the exponents in 10 and thus the precise entropy hh we obtain in 13 but does not change the positivity of hh, which suffices for our purposes (and ultimately by determining the limit exactly we prove the measure has maximal entropy anyway).

The following is an immediate calculation and we omit the proof.

Lemma 8.

Let xv,yvMN(Fv)x_{v},y_{v}\in M_{N}(F_{v}). Then

denomv(xv+yv),denomv(xvyv)denomv(xv)denomv(yv),\operatorname{denom}_{v}(x_{v}+y_{v}),\operatorname{denom}_{v}(x_{v}y_{v})\leq\operatorname{denom}_{v}(x_{v})\operatorname{denom}_{v}(y_{v})\,,

and in particular if yvGLN(𝒪v)y_{v}\in\mathrm{GL}_{N}(\mathcal{O}_{v}) then denomv(xvyv)=denomv(xv)\operatorname{denom}_{v}(x_{v}y_{v})=\operatorname{denom}_{v}(x_{v}). Furthermore there is a constant CC depending only on nn such that

denomv(xv1)denomv(x)C\operatorname{denom}_{v}(x_{v}^{-1})\leq\operatorname{denom}_{v}(x)^{C}

for xSLN(Fv)x\in\mathrm{SL}_{N}(F_{v}).

Corollary 9.

Multiplying place-by-place the same inequalities hold for the denominators in MN(𝔸f)M_{N}(\mathbb{A}_{\mathrm{f}}) and MN(F)M_{N}(F).

Finally if 𝔾\mathbb{G} is a linear algebraic FF-group fixing an FF-embedding ρ:𝔾SLN\rho\colon\mathbb{G}\to\mathrm{SL}_{N} allows us to define the denominators of elements of 𝔾(Fv)\mathbb{G}(F_{v}), 𝔾(𝔸)\mathbb{G}(\mathbb{A}), 𝔾(F)\mathbb{G}(F). The same reasoning as above would show that changing the embedding gives an equivalent definition in the sense above (i.e. up to multiplying by constants and raising to powers). In our case letting DD act on itself by multiplication gives an embedding GSL4(F)G\to\mathrm{SL}_{4}(F) and using the order RR to define the integral structure we further have ρ(Kv)SL4(𝒪v)\rho(K_{v})\subset\mathrm{SL}_{4}(\mathcal{O}_{v}) whenever KvK_{v} is defined. It follows that our local denominator is a bi-KvK_{v}-invariant function on GvG_{v}, so every basic Hecke operator (the characteristic function of a double coset KvgvKvK_{v}g_{v}K_{v}) has a well-defined denominator. Furthermore (identifying GvSL2(Fv)G_{v}\simeq\mathrm{SL}_{2}(F_{v})) this double coset has a representative av=(ϖvm00ϖvm)a_{v}=\begin{pmatrix}\varpi_{v}^{m}&0\\ 0&\varpi_{v}^{-m}\end{pmatrix} for some m0m\geq 0 in which case we call 2m2m the radius of the Hecke operator; its denominator is then qvmq_{v}^{m}.

Given constants c1,c2c_{1},c_{2} the set of potential Hecke operators is the set of gf𝔾(𝔸F)g_{\mathrm{f}}\in\mathbb{G}(\mathbb{A}_{F}) which are supported away from SS and of denominators at most c1ϵc2c_{1}\epsilon^{-c2}. We will be using Hecke operators of uniformly bounded radius so the main effect here is to bound the set of places vv under conisderation. Say that a potential Hecke operator causes an intersection at gΩg_{\infty}\in\Omega if there is γ𝔾(F)\gamma\in\mathbb{G}(F) such that

γgUϵKfgUϵgfKf,\gamma g_{\infty}{U_{\epsilon}}K_{\mathrm{f}}\cap g_{\infty}{U_{\epsilon}}g_{\mathrm{f}}K_{\mathrm{f}}\neq\emptyset\,,

in which case we say γ\gamma is involved in the intersection.

Proposition 10.

One can choose c1,c2,ϵ0>0c_{1},c_{2},\epsilon_{0}>0 such that such that if ϵ<ϵ0\epsilon<\epsilon_{0} then for all gΩg_{\infty}\in\Omega there is an algebraic FF-torus 𝕊𝔾\mathbb{S}\subset\mathbb{G} such that all γ𝔾(F)\gamma\in\mathbb{G}(F) involved in intersections at gg_{\infty} lie in 𝕊(F)\mathbb{S}(F).

Proof.

Suppose we have b1,b2Uϵb_{1},b_{2}\in{U_{\epsilon}} and kfKfk_{\mathrm{f}}\in K_{\mathrm{f}} such that γgb1=gb2gfkf\gamma g_{\infty}b_{1}=g_{\infty}b_{2}g_{\mathrm{f}}k_{\mathrm{f}}, or equivalently

γ=gb2b11g1gfkf.\gamma=g_{\infty}b_{2}b_{1}^{-1}g_{\infty}^{-1}g_{\mathrm{f}}k_{\mathrm{f}}.

By hypothesis denom(γ)=denom(gfkf)=denom(gf)c1ϵc2\operatorname{denom}(\gamma)=\operatorname{denom}(g_{\mathrm{f}}k_{\mathrm{f}})=\operatorname{denom}(g_{\mathrm{f}})\leq c_{1}\epsilon^{-c_{2}}. In addition since Ω\Omega is compact we have that gb2b11g1g_{\infty}b_{2}b_{1}^{-1}g_{\infty}^{-1} is OΩ(ϵ)O_{\Omega}(\epsilon)-close to an element of CC1CC^{-1}.

The matrix commutator [γ1,γ2]=γ1γ2γ2γ1[\gamma_{1},\gamma_{2}]=\gamma_{1}\gamma_{2}-\gamma_{2}\gamma_{1} (interpreted via the fixed embedding in SL4SL_{4}) is a polynomial function 𝔾2SL4\mathbb{G}^{2}\to\mathrm{SL}_{4} with coefficients in FF. Thus we have a constant cc so that if γ1,γ2𝔾(F)\gamma 1,\gamma_{2}\in\mathbb{G}(F) then denom([γ1,γ2])cdenom(γ1)cdenom(γ2)c\operatorname{denom}([\gamma_{1},\gamma_{2}])\leq c\operatorname{denom}(\gamma_{1})^{c}\operatorname{denom}(\gamma_{2})^{c}. Now suppose that γ1,γ2\gamma_{1},\gamma_{2} are both involved in intersections at gg_{\infty}. Then then γ1\gamma_{1} and γ2\gamma_{2} are CΩϵC_{\Omega}\epsilon-close to elements of gCC1g1g_{\infty}CC^{-1}g_{\infty}^{-1}. Since the commutator is a smooth function on G×GG\times G the element [γ1,γ2]G[\gamma_{1},\gamma_{2}]\in G is OΩ,C(epsilon)O_{\Omega,C}(epsilon)-close to the commutator of two elements drawn from gCC1g1g_{\infty}CC^{-1}g_{\infty}^{-1}. Since T1T_{1} is commutative we conclude that |[γ1,γ2]|1=OΩ,C(ϵ)\left|{[\gamma_{1},\gamma_{2}]}\right|_{1}=O_{\Omega,C}(\epsilon), and that for i2i\geq 2 we have |[γ1,γ2]|i=OΩ,C(1)\left|{[\gamma_{1},\gamma_{2}]}\right|_{i}=O_{\Omega,C}(1) if we also assume ϵ<1\epsilon<1.

Suppose γ1,γ2\gamma_{1},\gamma_{2} do not commute. By the product formula we then have

denom([γ1,γ2])i|[γ1,γ2]|i1,\operatorname{denom}([\gamma_{1},\gamma_{2}])\cdot\prod_{i}\left|{[\gamma_{1},\gamma_{2}]}\right|_{i}\geq 1\,,

that is

cc1ϵcc2OΩ,C(ϵ)OΩ,C(1)r+s+t11.cc_{1}\epsilon^{-cc_{2}}O_{\Omega,C}(\epsilon)\cdot O_{\Omega,C}(1)^{r+s+t-1}\geq 1\,.

Now if c2<1cc_{2}<\frac{1}{c} then for any c1>0c_{1}>0 there is ϵ0<0\epsilon_{0}<0 so that for ϵ<ϵ0\epsilon<\epsilon_{0} this is impossible. We conclude that, with these choices, the γ\gamma that are involved in intersections commute, so the FF-subgroup of 𝔾\mathbb{G} generated by those γ\gamma is commutative. If this subgroup is finite its elements are semisimple (and contained in a torus), and otherwise the only commutative connected subgroups of SL2\mathrm{SL}_{2} (even over its algebraic closure) are either tori or unipotent. Note that tori are self-centralizing whereas the component group of the centralizer of a unipotenet subgroup is represented by the center of SL2\mathrm{SL}_{2}.

This concludes the argument when DD is a division algebra, since in that case 𝔾(F)\mathbb{G}(F) consists entirely of semisimple elements and the maximal commutative FF-subgroups are all tori. When D=M2(F)D=M_{2}(F) we need to rule out the possibility that some γ\gamma causing an intersection is unipotent.

For this choose CC and ϵ0\epsilon_{0} small enough so that every gB(C,ϵ0)B(C,ϵ0)1g\in B(C,\epsilon_{0})B(C,\epsilon_{0})^{-1} is close enough to the identity to have |Tr(g)2|i<1\left|{\operatorname{Tr}(g)-2}\right|_{i}<1 at each infinite place. Now suppose that γ𝔾(F)\gamma\in\mathbb{G}(F) is involved in an intersection at gg_{\infty}, and that γ\gamma is unipotent. Then |Tr(γ)2|1>1\left|{\operatorname{Tr}(\gamma)-2}\right|_{1}>1 so Tr(γ)=2\operatorname{Tr}(\gamma)=2 (the alternative was that Tr(γ)=2\operatorname{Tr}(\gamma)=-2).

Writing γ=gb2b11g1\gamma=g_{\infty}b_{2}b_{1}^{-1}g_{\infty}^{-1} we get that |Tr(b2b11)2|1=2\left|{Tr(b_{2}b_{1}^{-1})-2}\right|_{1}=2. Now at the factor G1G_{1} each bib_{i} is ϵ\epsilon-close to an element of T1T_{1}, so this element has trace 2+O(ϵ)2+O(\epsilon) (as measured by ||1\left|{\cdot}\right|_{1}); since T1T_{1} is a fixed torus this element is O(ϵ)O(\epsilon)-close to the identity. Finally since Ω\Omega is compact conjugation by gg_{\infty} does not change this fact: the image of γ\gamma itself in G1G_{1} is O(ϵ)O(\epsilon)-close to the identity. Now the same argument as before shows that γ=Id\gamma=\operatorname{Id} if c2<1c_{2}<1 and ϵ\epsilon is small enough (the entries of γId\gamma-\operatorname{Id} are elements of FF whose denominators are not too large and whose archimedean absolute values are bounded and ||1\left|{\cdot}\right|_{1} absolute value is small). It follows that γ=Id\gamma=\operatorname{Id} and we are done. ∎

Remark 11.

The concrete argument using the commutator already appears in [1, Lem. 3.3] and two more general versions appears in [14, Sec. 4] for group over \mathbb{Q}. The new observation here is that one only needs information about the γ\gamma at a single real place (assuming boundedness at the other real places). In particular we cannot just restrict scalars to \mathbb{Q} and apply the earlier result.

Remark 12.

For any FF-group 𝔾\mathbb{G} the following is true: let Ti<GiT_{i}<G_{i} be an FiF_{i}-subgroup. Then there is an exponent c2c_{2} so that for all compact CTi×jiGiC\subset T_{i}\times\prod_{j\neq i}G_{i} and all compact ΩG\Omega\subset G there are c1,ϵ0>0c_{1},\epsilon_{0}>0 such that for all gΩg_{\infty}\in\Omega if ϵ<ϵ0\epsilon<\epsilon_{0} then all γ𝔾(F)\gamma\in\mathbb{G}(F) involved in intersections of potential Hecke translates of gUϵg_{\infty}{U_{\epsilon}} are contained in 𝕊(L)\mathbb{S}(L) where LFiL\subset F_{i} is a finite extension of FF and 𝕊𝔾\mathbb{S}\subset\mathbb{G} is an LL-subgroup which is conjugate to TiT_{i} over FiF_{i}.

3.1.2. Bounds on the mass of tubes

The arguments of [14, §5] (in fact already of [1] for the group at hand) now establish the following, using the diophantine lemma we proved above instead of the analogous lemmas in those papers, noting that our notion of denominator is equivalent to the notion we’d obtain over \mathbb{Q} for the group ResF𝔾\operatorname{Res}^{F}_{\mathbb{Q}}\mathbb{G}.

Proposition 13.

Fix compact subsets ΩG\Omega\subset G, CT1×G,2C\subset T_{1}\times G_{\infty,\geq 2}. Then for any normalized Hecke eigenfunction ϕL2(X)\phi\in L^{2}(X) (pulled back to a function on GG) and any gΩg_{\infty}\in\Omega we have

gUϵ|ϕ(x)|2dvol(x)=O(ϵh)\int_{g_{\infty}{U_{\epsilon}}}\left|{\phi(x)}\right|^{2}\mathop{}\!\mathrm{d}\!\operatorname{vol}(x)=O(\epsilon^{h})

for a universal constant h>0h>0 where the implied constant depends on Ω,C\Omega,C but not on ϕ\phi.

Corollary 14.

A1A_{1} acts on every ergodic component of μ\mu with positive entropy.

3.2. Measure rigidity

In this section we interpret μ\mu as a measure on 𝔾(F)\𝔾(𝔸)\mathbb{G}(F)\backslash\mathbb{G}(\mathbb{A}). Let vv be any finite place at which GvG_{v} is non-compact. Then [8, Thm. 8.1] shows that μ\mu is GvG_{v}-recurrent: if B𝔾(F)\𝔾(𝔸)B\subset\mathbb{G}(F)\backslash\mathbb{G}(\mathbb{A}) is any set of positive measure then for almost every xBx\in B the set of returns {gvGvxgvB}\{g_{v}\in G_{v}\mid xg_{v}\in B\} is unbounded.

Lemma 15.

For μ\mu-almost every x𝔾(F)\𝔾(𝔸)x\in\mathbb{G}(F)\backslash\mathbb{G}(\mathbb{A}) the group {hM1×G,2×Gvxz=z}\{h\in M_{1}\times G_{\infty,\geq 2}\times G_{v}\mid xz=z\} is finite.

Proof.

Let x=𝔾(F)gx=\mathbb{G}(F)g with g=(g,gf)𝔾(𝔸)g=(g_{\infty},g_{\mathrm{f}})\in\mathbb{G}(\mathbb{A}) where (by the density of 𝔾(F)\mathbb{G}(F) in 𝔾(𝔸f)\mathbb{G}(\mathbb{A}_{\mathrm{f}})) we may assume gfKfg_{\textrm{f}}\in K_{\mathrm{f}}, et h𝔾(𝔸)h\in\mathbb{G}(\mathbb{A}) have h1M1h_{1}\in M_{1}, and suppose that 𝔾(F).gh=𝔾(F).g\mathbb{G}(F).g\cdot h=\mathbb{G}(F).g. Then there exists γ𝔾(F)\gamma\in\mathbb{G}(F) such that gh=γgg\cdot h=\gamma g or equivalently

γ=ghg1.\gamma=ghg^{-1}\,.

Thus conjugation by gg (and the inclusion 𝔾(F)G1\mathbb{G}(F)\hookrightarrow G_{1}) embeds the group in the statement in 𝔾(F)g1(A1M1)g11\mathbb{G}(F)\cap g_{1}(A_{1}M_{1})g_{1}^{-1}. As in Section 3.1.1 above this intersection consists of the F1F_{1}-points of the diagonal F1F_{1}-torus of G1G_{1}, which is one-dimensional, so if the group in the statement was infinite it would be Zariski-dense in the conjugate torus; equivalently the Zariski closure of group would be an FF-torus 𝕋𝔾\mathbb{T}\subset\mathbb{G} with g11𝕋g1g_{1}^{-1}\mathbb{T}g_{1} diagonal.

For each torus 𝕋\mathbb{T} the set of g1G1g_{1}\in G_{1} that diagonalize it lies in two A1M1A_{1}M_{1}-cosets (due to the effect of the Weyl group). Since there are countably many such tori the set of xx for which the group is infinite is contained in the countable union of images in 𝔾(F)\𝔾(𝔸)\mathbb{G}(F)\backslash\mathbb{G}(\mathbb{A}) of the sets of the form

(g1AkMk)×G,2×Kf,(g_{1}A_{k}M_{k})\times G_{\infty,\geq 2}\times K_{\mathrm{f}}\,,

and we need to show this is a null set. Since our measure is right-KfK_{\mathrm{f}}-invariant we may instead consider the image of the set in X=Γ\GX=\Gamma\backslash G_{\infty}, and it remains to recall that the positive entropy argument showed that the images of the sets (g1AkMk)×G,2(g_{1}A_{k}M_{k})\times G_{\infty,\geq 2}, have μ\mu-measure zero in a strong quantitative form (the argument gave a uniform bound for the mass of ϵ\epsilon-neighborhoods of compact parts of such images). ∎

At this point we have verified the hypotheses of the following measure rigidity result.

Definition 16.

A measure ν\nu on XX is homogeneous if it is the unique HH-invariant probability measure on a closed HH-orbit for a closed subgroup H<GH<G.

Theorem 17 (Einsiedler–Lindenstrauss [5, Thm. 1.5]).

Let μ\mu be an A1A_{1}-invariant probability measure on XX such that

  1. (1)

    μ\mu has positive entropy on a.e. ergodic component with respect to the action of A1A_{1}.

  2. (2)

    μ\mu is GvG_{v}-recurrent.

  3. (3)

    For μ\mu-a.e. zΓp\G×Gpz\in\Gamma_{p}\backslash G\times G_{p} the group

    {hM1×G,2×Gvzh=z}\{h\in M_{1}\times G_{\infty,\geq 2}\times G_{v}\mid zh=z\}

    is finite.

Then μ\mu is a convex combination of homogeneous measures. Furthermore, for each such component ν\nu the associated group HH contains a semisimple algebraic subgroup of G1G_{1} of real rank 11 which further contains A1A_{1} and conversely HH is a finite-index subgroup of an algebraic subgroup of GG (here we think of G1G_{1} and GG as real algebraic groups).

By strong approximation the lattice Γ\Gamma is irreducible in GG, so that any G1G_{1}-invariant measure is in fact GG-invariant. Thus every component ν\nu for which HH contains G1G_{1} is the GG-invariant measure. In particular this happens when G1SL2()G_{1}\simeq\mathrm{SL}_{2}(\mathbb{R}) (the corresponding place of FF is real) since SL2()\mathrm{SL}_{2}(\mathbb{R}) has no proper semisimple subgroups, at which point 1 follows directly. For the remainder of the paper we will thus assume that G1SL2()G_{1}\simeq\mathrm{SL}_{2}(\mathbb{C}) (the place is complex) and that HH contains a conjugate of SL2()\mathrm{SL}_{2}(\mathbb{R}) there (these being the only proper semisimple subgroups).

We now fix a particular representative for this conjugacy class. Let ~1𝔾~1\tilde{\mathbb{H}}_{1}\subset\tilde{\mathbb{G}}_{1} be the fixed points of the automorphism cwc_{w}. Indeed the isomorphism G1SL2()G_{1}\simeq\mathrm{SL}_{2}(\mathbb{C}) has cwc_{w} act via complex conjugation so H1=~1(Ew)H_{1}=\tilde{\mathbb{H}}_{1}(E_{w}) is is exactly SL2()SL_{2}(\mathbb{R}), the subgroup of matrices fixed by complex conjugation. Recall that our choice of A1A_{1}, the positive diagonal subgroup makes it a subgroup of this H1H_{1}, in fact a real Cartan subgroup there.

Proposition 18.

Let ν\nu be a component of μ\mu as in the Theorem. Then the support of ν\nu is contained in the image in XX of a submanifold of GG of the form

L=g1H1M1G,2L=g_{1}H_{1}M_{1}G_{\infty,\geq 2}

where g1𝔾~1(¯Ew)g_{1}\in\tilde{\mathbb{G}}_{1}(\bar{\mathbb{Q}}\cap E_{w}).

Proof.

Let ν\nu be a component of μ\mu as in the theorem, supported on an HH-orbit xHXxH\subset X where HH contains a conjugate of the fixed group H1H_{1}. Let H~G=𝔾~(Ew)\tilde{H}\subset G=\tilde{\mathbb{G}}(E_{w}) be the semisimple EwE_{w}\simeq\mathbb{R}-algebraic subgroup of which HH is a finite-index subgroup. Note that H~G1=HG1\tilde{H}\cap G_{1}=H\cap G_{1} because SL2()\mathrm{SL}_{2}(\mathbb{R}) is a maximal proper closed subgroup of SL2()\mathrm{SL}_{2}(\mathbb{C}), and for the same reason H1H_{1} is the projection to G1G_{1} of any closed subgroup of GG that contains H1H_{1}.

Writing x=Γgx=\Gamma g for some gGg\in G the translate xHg1xHg^{-1} is the H=gHg1H^{\prime}=gHg^{-1}-orbit of the identity and supports an HH^{\prime}-invariant measure; in other words ΓH\Gamma\cap H^{\prime} is a lattice in HH^{\prime}. By the Borel Density Theorem this lattice is Zariski-dense in H~\tilde{H}^{\prime} making H~\tilde{H} defined over EE in view of Γ𝔾(F)𝔾~(E)\Gamma\subset\mathbb{G}(F)\subset\tilde{\mathbb{G}}(E).

The intersection G1H~=g1(G1H)g11G_{1}\cap\tilde{H}^{\prime}=g_{1}(G_{1}\cap H)g_{1}^{-1} is also defined over EE, where we write g1𝔾~1(Ew)g_{1}\in\tilde{\mathbb{G}}_{1}(E_{w}) for the first coordinate of gg. The group g1(G1H)g11g_{1}(G_{1}\cap H)g_{1}^{-1} and the fixed group H1H_{1} are then conjugate subgroups of G1G_{1} defined over EE, so by the Lemma below they are conjugate by an element of 𝔾~1(Ew¯)\tilde{\mathbb{G}}_{1}(E_{w}\cap\bar{\mathbb{Q}}); let g1g_{1}^{\prime} be such an element. We then have

G1H=(g11g1)H1(g11g1)1.G_{1}\cap H=(g_{1}^{-1}g_{1}^{\prime})H_{1}(g_{1}^{-1}g_{1}^{\prime})^{-1}\,.

Conjugating A1<HA_{1}<H (an avatar of the A1A_{1}-invariance of μ\mu), gives a Cartan subgroup (g11g1)1A1(g11g1)H1(g_{1}^{-1}g_{1}^{\prime})^{-1}A_{1}(g_{1}^{-1}g_{1}^{\prime})\subset H_{1}, and since all Cartan subgroups of a semisimple Lie group are conjugate we obtain h1H1h_{1}\in H_{1} such that

h1A1h11=(g11g1)1A1(g11g1)1).h_{1}A_{1}h_{1}^{-1}=(g_{1}^{-1}g_{1}^{\prime})^{-1}A_{1}(g_{1}^{-1}g_{1}^{\prime})^{-1})\,.

so that h11(g11g1)1NG1(A1)=NH1(A1)M1h_{1}^{-1}(g_{1}^{-1}g_{1}^{\prime})^{-1}\in N_{G_{1}}(A_{1})=N_{H_{1}}(A_{1})M_{1} and thus that (g11g1)1H1M1(g_{1}^{-1}g_{1}^{\prime})^{-1}\in H_{1}M_{1}.

Finally in terms of all those elements,

ΓgH\displaystyle\Gamma gH Γg1(g11g)(G1H)G,2\displaystyle\subset\Gamma g_{1}(g_{1}^{-1}g)(G_{1}\cap H)G_{\infty,\geq 2}
=Γg1(g11g1)H1(g11g1)1G,2\displaystyle=\Gamma g_{1}(g_{1}^{-1}g_{1}^{\prime})H_{1}(g_{1}^{-1}g_{1}^{\prime})^{-1}G_{\infty,\geq 2}
Γg1H1H1MG,2\displaystyle\subset\Gamma g_{1}^{\prime}H_{1}H_{1}MG_{\infty,\geq 2}
=Γg1H1MG,2.\displaystyle=\Gamma g_{1}^{\prime}H_{1}MG_{\infty,\geq 2}\,.

Now there are countably many submanifolds of the form above, so to prove our main theorem and show that μ\mu is the GG-invariant measure it suffices to rule out each submanifold separately.

Lemma 19.

Let EE\subset\mathbb{R} be a number field with a fixed real place, and let 𝔾\mathbb{G} be an algebraic group defined EE. Suppose the two EE-subgroups 1,2\mathbb{H}_{1},\mathbb{H}_{2} of 𝔾\mathbb{G} are conjugate in 𝔾()\mathbb{G}(\mathbb{R}). Then they are already conjugate over a finite extension of EE contained in \mathbb{R}.

Proof.

Let 𝔾\mathbb{P}\subset\mathbb{G} be the transporter of 1\mathbb{H}_{1} to 2\mathbb{H}_{2} (the elements conjugating one subgroup to the other), an EE-subvariety by [9, Prop. 1.79]. The statement "\mathbb{P} is not empty" is then a first-order statement in the language of fields. Since the theory of real closed fields (augmented by the diagram of EE) eliminates quantifiers, the truth of the statement over \mathbb{R} (the hypothesis above) implies its truth over the algebraic closure of EE in \mathbb{R}, the real closed field ¯\bar{\mathbb{Q}}\cap\mathbb{R}. Finally any point of \mathbb{P} over this field already lies in a finite extension of EE. ∎

4. Submanifolds with small stabilizers

In the previous section we showed that components of the limit measure μ\mu other than the uniform measure are supported in images in X=Γ\GX=\Gamma\backslash G_{\infty} of submanifolds of GG_{\infty} of the form L=g1H1M1G,2L=g_{1}H_{1}M_{1}G_{\infty,\geq 2}. In this section we will use elementary arguments to show that such manifolds LL and their submanifolds cannot be left-invariant by subgroup of G1G_{1} which are "too large" in a technical sense. In the next section we will then construct Hecke operators which avoid such subgroups, allowing us to prove 2 in the last section.

Since the calculations will take place in the group G1G_{1} rather than the entire group GG_{\infty}, for this section only we write 𝔾\mathbb{G} for this group (recall that it is defined over EE) and GG for the group 𝔾1(Ew)SL2()\mathbb{G}_{1}(E_{w})\simeq\mathrm{SL}_{2}(\mathbb{C}) that we usually denote G1G_{1}. Similarly we will drop the subscript 11 for algebraic and Lie subgroups of GG, especially HSL2()H\simeq\mathrm{SL}_{2}(\mathbb{R}) and MU(1)M\simeq U(1). We write \mathbb{R} for EwE_{w} and =Nw\mathbb{C}=N_{w}.

Definition 20.

Let S<GS<G be an (abstract) subgroup. We say that SS is small (in GG) if there exist a finite extension KK\subset\mathbb{C} of EE and a finite-index subgroup S<SS^{\prime}<S such that SS^{\prime} is contained in the KK-points 𝔹(K)\mathbb{B}(K) of a KK-subgroup 𝔹\mathbb{B} of 𝔾\mathbb{G} of one of the following forms:

  1. (1)

    Multiplicative type: 𝔹\mathbb{B} is diagonable over \mathbb{C}.

  2. (2)

    SL2\mathrm{SL}_{2}-type: 𝔹\mathbb{B} is 𝔾(K)\mathbb{G}(K)-conjugate to a subgroup of ×EK\mathbb{H}\times_{E}K.

Remark 21.

Observe that 𝔾()SL2()×SL2()\mathbb{G}(\mathbb{C})\simeq\mathrm{SL}_{2}(\mathbb{C})\times\mathrm{SL}_{2}(\mathbb{C}) in such a way that the image of ()\mathbb{H}(\mathbb{C}) is the subgroup {(g,g)gSL2()}\{(g,g)\mid g\in\mathrm{SL}_{2}(\mathbb{C})\} (with cwc_{w} acting on the complexified group by exchanging the two factors). This allows us to check that the subgroup SU(2)={gG=SL2()cw(g)=g1t}\mathrm{SU}(2)=\{g\in G=\mathrm{SL}_{2}(\mathbb{C})\mid c_{w}(g)={}^{t}g^{-1}\} is also of SL2\mathrm{SL}_{2}-type: its complexification is the subgroup {(g,g1t)gSL2(C)}\{(g,{}^{t}g^{-1})\mid g\in\mathrm{SL}_{2}(C)\} which is conjugate to \mathbb{H} by the element (I2,(0110))\left(I_{2},\begin{pmatrix}0&-1\\ 1&0\end{pmatrix}\right).

We will be interested in the stabilizers of algebraic varieties. On the one hand they will arise for us as abstract subgroups stabilizing sets of points. On the other hand, we will also be analyzing them as algebraic groups. To connect the two points of view we rely on the following:

Theorem 22 ([9, Cor. 1.81]).

Let 𝔾\mathbb{G} be an algebraic group acting on an algebraic variety 𝕏\mathbb{X}. Let <𝕏\mathbb{P}<\mathbb{X} be a closed subvariety. Then the stabilizer 𝕊={g𝔾g𝕏=𝕏}\mathbb{S}=\{g\in\mathbb{G}\mid g\mathbb{X}=\mathbb{X}\} is a closed subgroup of 𝔾\mathbb{G}. Furthermore if KK is an extension of the field of definition such that 𝕏(K)\mathbb{X}(K) is Zariski-dense in 𝕏\mathbb{X} then 𝕊(K)={g𝔾(K)g𝕏(K)=𝕏(K)}\mathbb{S}(K)=\{g\in\mathbb{G}(K)\mid g\mathbb{X}(K)=\mathbb{X}(K)\}.

Definition 23.

We say that a submanifold LGL\subset G has small stabilizers if its Zariski closure 𝕃𝔾\mathbb{L}\subset\mathbb{G} is defined over a finite extension F/EF^{\prime}/E contained in \mathbb{R} and we have

  • For every FF^{\prime}-subvariety 𝕃\mathbb{P}\subset\mathbb{L} such that ()\mathbb{P}(\mathbb{R}) is Zariski-dense in \mathbb{P}, the stabilizer

    SP={s𝔾(F)s()=()},S_{P}=\{s\in\mathbb{G}(F^{\prime})\mid s\mathbb{P}(\mathbb{R})=\mathbb{P}(\mathbb{R})\},

    is a small subgroup of GG.

We begin our analysis of stabilizers of subvarieties of the submanifolds gHMgHM with the observation that HMHM is, in fact, an EE-subvariety, as the reader can verify by writing explicit algebraic equations in the real and complex parts of the matrix entries. Each translate gHMgHM is then also a real subvariety which, in the case of interest where gg has algebraic entries, is defined over a number field.

As a preliminary step we rule out one possible kind of stabilizer:

Lemma 24.

Let PL=gHMP\subset L=gHM be a real subvariety, and let 𝕊\mathbb{S} be its stabilizer and S=𝕊()S=\mathbb{S}(\mathbb{R})^{\circ} (identity component in the analytic topology). Then SS is one of:

  1. (1)

    the real points of an algebraic torus; or

  2. (2)

    compact (as a Lie group); or

  3. (3)

    contained in a conjugate of H=SL2()H=\mathrm{SL}_{2}(\mathbb{R}).

Proof.

We have dimSdimPdimL=4\dim S\leq\dim P\leq\dim L=4, and since LL itself is connected but not the coset of a subgroup, the dimension of SS is actually at most 33.

Other than possibilities 1-3 above the only other closed connected subgroups of SL2()\mathrm{SL}_{2}(\mathbb{C}) of dimension at most 33 are solvable with unipotent radical of dimension 22, so we suppose SS^{\circ} is of that form. Conjugating SS (and translating LL and PP appropriately) we may assume that SS contains all the matrices of the form u(z)=(1z01)u(z)=\begin{pmatrix}1&z\\ 0&1\end{pmatrix} for zz\in\mathbb{C}, which we can now rule out by by showing that no orbit of this group is contained in any submanifold LL of as above, let alone in a submanifold PP.

Thus fix l=ghml=ghm for some gGg\in G, hHh\in H, mMm\in M. Setting y=ghy=gh that u(z)ghmgHMu(z)ghm\in gHM is equivalent to

y1u(z)yHM,y^{-1}u(z)y\in HM\,,

and for y=(abcd)y=\begin{pmatrix}a&b\\ c&d\end{pmatrix} this reads

g1u(z)g=I2+(cdd2c2cd)z.g^{-1}u(z)g=I_{2}+\begin{pmatrix}cd&d^{2}\\ -c^{2}&-cd\end{pmatrix}z\,.

Now the only complex subspace of the tangent space at the identity of HMHM is the Lie alebra of AMAM, that is the diagonal matrices, whereas the image of the derivative of the map zy1u(z)yz\mapsto y^{-1}u(z)y is never of this form since c,dc,d can’t both vanish. ∎

We can now prove the expected assertion:

Proposition 25.

Let FF^{\prime}\subset\mathbb{R} be a finite extension of EE, and let g𝔾(F)g\in\mathbb{G}(F^{\prime}). Then the submanifold L=gHML=gHM of GG has small stabilizers.

Proof.

We observed earlier LL is Zariski-closed and defined over FF^{\prime}; let 𝕃\mathbb{L} be the underlying subvariety, \mathbb{P} an FF^{\prime}-subvariety with dense real points, and 𝕊\mathbb{S} its stabilizer. We need to show that 𝕊(F)\mathbb{S}(F^{\prime}) is small,

If we are in the first possibility in 24 then 𝕊\mathbb{S} is diagonable, hence of multiplicative type. In the second possibility SS is contained in a maximal compact subgroup, i.e. a conjugate of SU(2)\mathrm{SU}(2), so either SS is conjugate to SU(2)\mathrm{SU}(2) (over a number field by 19) and hence of SL2\mathrm{SL}_{2}-type as observed in 21, or SS is contained in a torus and hence is of multiplicative type.

The only remaining possibility is that SS is conjugate to a subgroup of SL2()\mathrm{SL}_{2}(\mathbb{R}). This case is more complicated because it is not a-priori clear that SS is conjugate to a subgroup defined over a number field, and we divide further into cases according to the unipotent radical of 𝕊\mathbb{S}.

  1. (1)

    If the unipotent radical of 𝕊\mathbb{S} is non-trivial it is at most 11-dimensional by 24. Conjugating by an element of 𝔾(F)\mathbb{G}(F^{\prime}) we may assume it is the group of upper-triangular unipotent matrices, the normalizer of which is the group of upper-triangular matrices and in particular up to FF^{\prime}-conjugacy 𝕊\mathbb{S} is contained in ×EF\mathbb{H}\times_{E}F^{\prime}.

  2. (2)

    Otherwise 𝕊\mathbb{S} is semisimple and (since its dimension is at most 3) it is either a torus or a form of SL2\mathrm{SL}_{2}, in which case 𝕊()\mathbb{S}(\mathbb{R}) is isomorphic to one of the groups SL2()\mathrm{SL}_{2}(\mathbb{R}), SU(2)\mathrm{SU}(2). But all subgroups of SL2()\mathrm{SL}_{2}(\mathbb{C}) of each form are conjugate, and since the representative subgroups are defined over EE the conjugacy is over FF^{\prime} as before.

5. Constructing an amplifier

Given a Hecke eigenfunction ϕL2(X)\phi\in L^{2}(X) and a small subgroup SS of G1G_{1} in section we construct a Hecke operator which acts on ϕ\phi with large eigenvalue (making it an “amplifier” for ϕ\phi) while at the same time avoiding the orbits of SS in products of trees 𝔾1(p)/𝔾1(p)\mathbb{G}_{1}(\mathbb{Q}_{p})/\mathbb{G}_{1}(\mathbb{Z}_{p}).

We begin with the observation that having constructed the factorization 𝔾~=i𝔾~i\tilde{\mathbb{G}}=\prod_{i}\tilde{\mathbb{G}}_{i} over EE, we can now extend the field NN to contain the field KK evincing the smallness of SS as in 20 in such a way that EE contains the field of definition FF^{\prime} of SS. Thus for the rest of the paper the fields EE and NN and the factorizations of GG will depend on SS (and ultimately on the subvariety of GG we are trying to avoid) rather than just on 𝔾\mathbb{G} as in the first run through Section 2. Since we are ruling out the subvarities one-by-one this is not a problem.

Recall that from that section that the group 𝔾1\mathbb{G}_{1} is defined over EE so that 𝔾1(E)=SL2(E1)\mathbb{G}_{1}(E)=\mathrm{SL}_{2}(E_{1}) where since we assume G1=SL2()G_{1}=\mathrm{SL}_{2}(\mathbb{C}) we have E1=NE_{1}=N, and in that case for each prime pp splitting completely in NN we can fix a place wpw_{p} of NN lying over pp so that E1ENwpp×pE_{1}\otimes_{E}N_{w_{p}}\simeq\mathbb{Q}_{p}\times\mathbb{Q}_{p} or equivalently so that Gp,1=𝔾1(Ewp)SL2(p)×SL2(p)G_{p,1}=\mathbb{G}_{1}(E_{w_{p}})\simeq\mathrm{SL}_{2}(\mathbb{Q}_{p})\times\mathrm{SL}_{2}(\mathbb{Q}_{p}) (with cwc_{w} exchanging the two factors). Then

Gp=vpGv=i𝔾i(Ewp)=Gp,1×Gp,2,G_{p}=\prod_{v\mid p}G_{v}=\prod_{i}\mathbb{G}_{i}(E_{w_{p}})=G_{p,1}\times G_{p,\geq 2}\,,

and omitting finitely many primes we may also assume that with this identification the maximal compact subgroups Kp,iK_{p,i} isomorphic to SL2(p)\mathrm{SL}_{2}(\mathbb{Z}_{p}) or SL2(p)2\mathrm{SL}_{2}(\mathbb{Z}_{p})^{2} of the factors are factors of the adelic open compact subgroup KfK_{\mathrm{f}}. Let KpK_{p} be their product.

Definition 26.

A Hecke operator τp\tau\in\mathcal{H}_{p} one-sided if it is supported in the image of the first factor of Gp,1G_{p,1}.

Identifying Gp,1G_{p,1} with SL2(p)2\mathrm{SL}_{2}(\mathbb{Q}_{p})^{2} and KpK_{p} with SL2(p)2\mathrm{SL}_{2}(\mathbb{Z}_{p})^{2} the basic Hecke operators τpj\tau_{p^{j}} will be the one-sided Hecke operator correspoding to the characteristic function of the double coset

SL2(p)(pj00pj)SL2(p),\mathrm{SL}_{2}(\mathbb{Z}_{p})\begin{pmatrix}p^{j}&0\\ 0&p^{-j}\end{pmatrix}\mathrm{SL}_{2}(\mathbb{Z}_{p})\,,

that is summing over the sphere of radius 2j2j in the Bruhat-Tits tree of the first factor of Gp,1G_{p,1}.

Lemma 27.

Let S𝔾1(E)S\subset\mathbb{G}_{1}(E) be a small subgroup. Then there is a constant CC (depending on SS) such that for all primes p>Cp>C (splitting completely in NN) the SS-orbit of the identity coset in Gp/KpG_{p}/K_{p}:

  1. (1)

    Does not meet the support of any one-sided τpj\tau_{p^{j}}, if SS is of SL2\mathrm{SL}_{2}-type.

  2. (2)

    Meets the support of any one-sided τpj\tau_{p^{j}} at most CC times, if is of multiplcative type.

Proof.

Suppose first that SS is of multiplicative type and let g𝔾(K)g\in\mathbb{G}(K) be such that g(K)g1Sg\mathbb{H}(K)g^{-1}\cap S is of finite index CC in SS with coset representatives {s1,,sC}\{s_{1},\ldots,s_{C}\}. Since the prime pp splits completely in the field KNK\subset N, the element hh embeds in 𝔾1(Ewp)\mathbb{G}_{1}(E_{w_{p}}); omitting finitely many primes we may also assume that gg and the sis_{i} all lies in Kp,1K_{p,1}.

Now any element of the ss-orbit has the form sig(h,h)g1s_{i}g(h,h)g^{-1} for some hSL2(p)h\in\mathrm{SL}_{2}(\mathbb{Q}_{p}). Assuming si,gK1,ps_{i},g\in K_{1,p} if this meets the support of a one-sided Hecke operator the second coordinate of this element must lie in the compact subgroup SL2(p)\mathrm{SL}_{2}(\mathbb{Z}_{p}) which by the KAKKAK decomposition in SL2(p)\mathrm{SL}_{2}(\mathbb{Q}_{p}) forces hSL2(p)h\in\mathrm{SL}_{2}(\mathbb{Z}_{p}) the element is in the identity coset.

Suppose now instead that g𝕋(K)g1Sg\mathbb{T}(K)g^{-1}\cap S is of finite index CC in SS where 𝕋\mathbb{T} is the diagonal torus of 𝔾1\mathbb{G}_{1}. Again assume that gKpg\in K_{p}; since our torus is KK-split and pp splits completely in KK the orbit of g𝕋(p×p)g1g\mathbb{T}(\mathbb{Q}_{p}\times\mathbb{Q}_{p})g^{-1} is the product of two apartments (=geodesics) passing through the origin of the trees SL2(p)/SL2(p)\mathrm{SL}_{2}(\mathbb{Q}_{p})/\mathrm{SL}_{2}(\mathbb{Z}_{p}). In particular in the first coordinate the orbit meets the translates by si1s_{i}^{-1} of each sphere (the support of τpj\tau_{p^{j}}) at most twice. ∎

Next we construct the standard amplifier in SL2(p)\mathrm{SL}_{2}(\mathbb{Q}_{p}). The fundamental observation (often attributed to Iwaniec) is that it is impossible for the eigenvalues of τp\tau_{p}, τp2\tau_{p^{2}} to be simultaneously small.

Lemma 28.

Given ϕ\phi we can choose τ\tau to be either τp\tau_{p} or τp2\tau_{p^{2}} so that the corresponding eigenvalue λ\lambda satisfies |λ|#supp(τ)1/2\left|{\lambda}\right|\gg\#\operatorname{supp}(\tau)^{1/2}.

Proof.

Let λpj\lambda_{p^{j}} be the eigenvalue of τpj\tau_{p^{j}} acting on ϕ\phi. A direct calculation in the tree gives the convolution identity

(5.1) τp2\displaystyle\tau_{p}^{2} =p(p+1)+(p1)τp+τp2,\displaystyle=p(p+1)+(p-1)\tau_{p}+\tau_{p^{2}}\,,
(5.2) τp22\displaystyle\tau_{p^{2}}^{2} =p3(p+1)+p2(p1)τp+p(p1)τp2+(p1)τp3+τp4.\displaystyle=p^{3}(p+1)+p^{2}(p-1)\tau_{p}+p(p-1)\tau_{p^{2}}+(p-1)\tau_{p^{3}}+\tau_{p^{4}}\,.

The first identity implies (λp)2=p(p+1)+(p1)λp+λp2(\lambda_{p})^{2}=p(p+1)+(p-1)\lambda_{p}+\lambda_{p^{2}}, so at least one of |λp|p\left|{\lambda_{p}}\right|\gg p or |λp2|p2\left|{\lambda_{p^{2}}}\right|\gg p^{2} must hold with the implied constants absolute. ∎

Combining the spectral calculation and the control of intersections we have

Corollary 29 (Local construction).

Let S<𝔾(F)S<\mathbb{G}(F) be a subgroup, and assume the image of SS under the projection to 𝔾1(E)\mathbb{G}_{1}(E) is small. Then there exist an absolute constant LL, a constant CC (depending on SS), and a set of positive density PS𝒫P_{S}\subset\mathcal{P} such that for every prime pPSp\in P_{S} there exists a finite set JppJ_{p}\subset\mathcal{H}_{p} of basic Hecke operators such that:

  1. (1)

    For each hpJph_{p}\in J_{p},

    p#supp(hp)pp^{\ell}\ll\#\operatorname{supp}(h_{p})\ll p^{\ell}

    for some L\ell\ll L where the constants are absolute (we can take L=4L=4).

  2. (2)

    For xpGp/Kpx_{p}\in G_{p}/K_{p} other than the origin we have |(hphp)(xp)|p1\left|{(h_{p}\star h_{p}^{*})(x_{p})}\right|\ll p^{\ell-1}.

  3. (3)

    For each hpJph_{p}\in J_{p}, the number of intersections of the SS-orbit in Gp/KpG_{p}/K_{p} and the support of any of hph_{p},hph_{p}^{*},hphph_{p}\star h_{p}^{*} is bounded above by CC.

  4. (4)

    For each Hecke eigenfunction ϕL2(X)\phi\in L^{2}(X), at least one hpJph_{p}\in J_{p} acts on ϕ\phi with eigenvalue λp\lambda_{p} satisfying

    |λp|(#supphp)1/2.\left|{\lambda_{p}}\right|\gg\left(\#\operatorname{supp}h_{p}\right)^{1/2}\,.
Proof.

27 and 28 together show that the claim holds for all primes which are large enough (depending on SS) and split completely in a number field NN which depends on SS (and 𝔾\mathbb{G}). The Chebotarev Density Theorem shows that the set of split primes has positive density. ∎

Our global amplifier will amplify a specified Hecke-eigenfunction ϕ\phi. However, we need some control of its action on its orthogonal complement as well. Accordingly for a self-adjoint Hecke operator τ\tau\in\mathcal{H} we denote by c=c(τ)c=c(\tau) the smallest non-negative constant such that the spherical transform of τ\tau on the unitary dual is bounded below by c-c. In particular the spectrum of τ\tau acting on L2(X)L^{2}(X) is contained in [c,)[-c,\infty) and therefore (a fact which can be taken as a not-quite-equivalent definition) we have for all RL2(X)R\in L^{2}(X) that

(5.3) τ.R,RcR22.\left<\tau.R,R\right>\geq-c\left\|{R}\right\|_{2}^{2}.
Proposition 30 (Global construction).

Continuing with the hypotheses of 29 let also ϵ>0\epsilon>0. Then there is Q=Q(S,ϵ)Q=Q(S,\epsilon) such that for every Hecke-eigenfunction ϕL2(X)\phi\in L^{2}(X), there exists τSpan(pQJp)\tau\in\operatorname{Span}_{\mathbb{C}}\left(\cup_{p\leq Q}J_{p}\right) acting on ϕ\phi with eigenvalue Λ\Lambda satisfying

  1. (1)

    supp(τ)\operatorname{supp}(\tau) meets the SS-orbit of the identity in pGp/Kp\prod_{p}G_{p}/K_{p} in at most ϵΛτ\frac{\epsilon\Lambda}{\left\|{\tau}\right\|_{\infty}} points.

  2. (2)

    c(τ)ϵΛc(\tau)\leq\epsilon\Lambda.

Proof.

For Q1Q\in\mathbb{Z}_{\geq 1} and L\ell\leq L let 𝒫\mathcal{P}_{\ell} be the set of primes p[Q,2Q]PSp\in[Q,2Q]\cap P_{S} for which there is an operator hph_{p} with eigenvalue λp\lambda_{p} as in the conclusion of 29 with support of size comparable to pp^{\ell}, and fix \ell so that 𝒫=𝒫\mathcal{P}=\mathcal{P}_{\ell} consists of at least 1/L1/L of the primes on in [Q,2Q]PS[Q,2Q]\cap P_{S}. We then have #𝒫Q/logQ\#\mathcal{P}\gg Q/\log Q where the constant depends on SS through the density of PSP_{S}.

For each p𝒫p\in\mathcal{P} let ζp\zeta_{p} be a complex number of magnitude 11 such that ζpλp\zeta_{p}\lambda_{p} is a positive real number, and finally set

τ1=(p𝒫ζphp)(p𝒫ζphp)\tau_{1}=(\sum_{p\in\mathcal{P}}\zeta_{p}h_{p})*(\sum_{p\in\mathcal{P}}\zeta_{p}h_{p})^{*}

and

τ=τϕ=τ1τ1(1)δ,\tau=\tau_{\phi}=\tau_{1}-\tau_{1}(1)\delta,

where δ\delta is the identity element of the (full) Hecke algebra, and τ1(1)\tau_{1}(1) is the value of the function τ\tau at the identity coset. In other words τ\tau is obtained from τ1\tau_{1} by restricting away from a single point.

It is clear that τ\tau is self-adjoint. To compute its \ell^{\infty} norm (as a function on Kf\𝔾(𝔸f)/KfK_{\mathrm{f}}\backslash\mathbb{G}(\mathbb{A}_{\mathrm{f}})/Kf) we start with the fact that, as functions on that space, we have the pointwise identity

τ1pPhphp+p<q(hphq+hqhp).\tau_{1}\leq\sum_{p\in P}h_{p}\star h_{p}+\sum_{p<q\in\P}\left(h_{p}\star h_{q}^{*}+h_{q}\star h_{p}^{*}\right).

Since the summands on the right have disjoint supports, it suffices to estimate each separately. First, if pqp\neq q then the support of the convlution is the product of the supports since GpG_{p} and GqG_{q} are disjoint commuting subgroups, so those terms are bounded by 22. At a single prime pp the local construction gives the bound p1p^{\ell-1} for the values of hphph_{p}\star h_{p} away from the origin, and we conclude that τQ1\left\|{\tau}\right\|_{\infty}\ll Q^{\ell-1}.

Next, the eigenvalue of τ1\tau_{1} is clearly

(p𝒫|λp|)2(pPs#supp(hp))2(#𝒫Q/2)2Q2+log2Q,(\sum_{p\in\mathcal{P}}\left|{\lambda_{p}}\right|)^{2}\gg\left(\sum_{p\in Ps}\sqrt{\#\operatorname{supp}(h_{p})}\right)^{2}\gg\left(\#\mathcal{P}Q^{\ell/2}\right)^{2}\gg\frac{Q^{2+\ell}}{\log^{2}Q},

and the argument of the previous paragraph has shown that

τ1(1)=p𝒫#supp(hp)Q1+/2logQ.\tau_{1}(1)=\sum_{p\in\mathcal{P}}\#\operatorname{supp}(h_{p})\ll\frac{Q^{1+\ell/2}}{\log Q}\,.

Subtracting the two gives

ΛQ2+log2Q.\Lambda\gg\frac{Q^{2+\ell}}{\log^{2}Q}\,.

For primes p,qp,q let Tp,qT_{p,q} be the set of points in supphphq\operatorname{supp}h_{p}\star h_{q}^{*} that meet the SS-orbit of the identity. When p=qp=q these sets are of uniformly bounded size by Item 3, whereas when pqp\neq q this is the product of the corresponding subsets of Gp/KpG_{p}/K_{p} and Gq/KqG_{q}/K_{q} so again uniformly bounded. We conclude that the product of τ\left\|{\tau_{\infty}}\right\| with the total number of intersections satisfies

(5.4) τp,q𝒫#Tp,qΛτ#𝒫2ΛQ1.\frac{\left\|{\tau_{\infty}}\right\|\sum_{p,q\in\mathcal{P}}\#T_{p,q}}{\Lambda}\ll\left\|{\tau_{\infty}}\right\|\frac{\#\mathcal{P}^{2}}{\Lambda}\ll Q^{-1}\,.

Similarly since τ1\tau_{1} is self-adjoint, c(τ)=τ1(1)c(\tau)=\tau_{1}(1) and hence

(5.5) c(τ)Λ=τ1(1)λQ1/2logQ.\frac{c(\tau)}{\Lambda}=\frac{\tau_{1}(1)}{\lambda}\ll Q^{-1-\ell/2}\log Q\,.

Finally if we take QQ large enough we can can ensure both right-hand-sides of Eqs. 5.4 and 5.5 are less than ϵ\epsilon. ∎

6. Non-concentration on homogenous submanifolds

We have shown that to rule out non-Haar components it suffices to rule out components supported in images in in X=Γ\GX=\Gamma\backslash G of submanifolds gHMGgHM\subset G, which we finally do in this section. As in Section 3 instead of bounding μ(ΓgHM)\mu(\Gamma gHM) we will bound μ(ΓUϵ)\mu(\Gamma U_{\epsilon}) where UgHMU\subset gHM is a bounded neighborhood. Unlike the positive entropy arguments we now need to treat UU as a subset of gHMgHM rather than try for additional uniformity by fixing a single UHMU\subset HM and translating by gg later.

Since will calculate in GG and 𝔾(𝔸)\mathbb{G}(\mathbb{A}) but ultimately make statements about subsets of XX we introduce the notation πG:GX=Γ\G\pi_{G}\colon G\to X=\Gamma\backslash G and π𝔸:𝔾(𝔸)X=𝔾(F)\𝔾(𝔸)/Kf\pi_{\mathbb{A}}\colon\mathbb{G}(\mathbb{A})\to X=\mathbb{G}(F)\backslash\mathbb{G}(\mathbb{A})/K_{\mathrm{f}} for the quotient maps.

Definition 31.

An algebraic piece of GG will be a triple (U,𝕃,F)(U,\mathbb{L},F^{\prime}) where F/EF^{\prime}/E is a finite extension contained in EwE_{w}, 𝕃𝔾~1\mathbb{L}\subset\tilde{\mathbb{G}}_{1} is an irreducible subvariety defined over FF^{\prime} and irreducible over EwE_{w}, and U(𝕃×𝔾~2)(Ew)U\subset(\mathbb{L}\times\tilde{\mathbb{G}}_{\geq 2})(E_{w}) is Zariski-dense, and also bounded and relatively open in the analytic topology.

This parametrization is redundant (UU determines 𝕃\mathbb{L}) but it is easier to directly keep track of 𝕃\mathbb{L} and its field of definition. We also write 𝕃~=𝕃×𝔾~2\tilde{\mathbb{L}}=\mathbb{L}\times\tilde{\mathbb{G}}_{\geq 2}

Consider now a translate of an algebraic piece UU by some element gf𝔾(𝔸f)g_{\mathrm{f}}\in\mathbb{G}(\mathbb{A}_{\mathrm{f}}). Let γ𝔾(F)\gamma\in\mathbb{G}(F) be such that γgfKf\gamma g_{\mathrm{f}}\in K_{\mathrm{f}}. Then

π𝔸(Ugf)=π𝔸(γUgf)=πG(γU)\pi_{\mathbb{A}}(Ug_{\mathrm{f}})=\pi_{\mathbb{A}}(\gamma Ug_{\mathrm{f}})=\pi_{G}(\gamma U)

For a subset UGU\subset G write U¯\bar{U} for its closure in the analytic topology and U¯1\bar{U}^{1} for its projection to G1G_{1}.

Definition 32.

We say the translate UgfUg_{\mathrm{f}} is transverse to UU if for each γ𝔾(F)\gamma\in\mathbb{G}(F) with γgfKf\gamma g_{\mathrm{f}}\in K_{\mathrm{f}} the FF^{\prime}-Zariski closure of the intersection γU¯1U¯1G1\gamma\bar{U}^{1}\cap\bar{U}^{1}\subset G_{1} is of smaller dimension than 𝕃\mathbb{L}. Abusing notation we say that gfg_{\mathrm{f}} itself is transverse to UU, and otherwise say that it is parallel to UU.

Observe that the definition only depends on the coset of gfg_{\mathrm{f}} in 𝔾(𝔸f)/Kf\mathbb{G}(\mathbb{A}_{\mathrm{f}})/K_{\mathrm{f}}.

Lemma 33.

Let γ𝔾(F)\gamma\in\mathbb{G}(F) and gf𝔾(𝔸f)g_{\mathrm{f}}\in\mathbb{G}(\mathbb{A}_{\mathrm{f}}) be such that γkfKf\gamma k_{\mathrm{f}}\in K_{\mathrm{f}}. Suppose further that the FF^{\prime}-Zariski closure of γU¯1U¯1\gamma\bar{U}^{1}\cap\bar{U}^{1} is of dimension dim𝕃\dim\mathbb{L} Then γ𝕃=𝕃\gamma\mathbb{L}=\mathbb{L}.

For a function ϕL2(X)\phi\in L^{2}(X) and a measurable subset NGN\subset G (resp. N𝔾(𝔸F)N\subset\mathbb{G}(\mathbb{A}_{F})) we write ϕN\phi_{N} for the restriction of ϕ\phi to πG(N)\pi_{G}(N) (resp. to π𝔸(N)\pi_{\mathbb{A}}(N)).

Proposition 34.

Let (U,𝕃,F)(U,\mathbb{L},F^{\prime}) be an algebraic piece of GG and let τ\tau be a self-adjoint Hecke operator. Write PP for the set of parallel elements in the support of τ\tau. Then there exists a finite collection 𝒱\mathcal{V} of algebraic pieces (V,𝕃,F′′)(V,\mathbb{L}^{\prime},F^{\prime\prime}) of GG such that the 𝕃\mathbb{L}^{\prime} are irreducible proper FF^{\prime}-subvarieties of 𝕃\mathbb{L} and such that for every δ>0\delta>0 there exists ϵ>0\epsilon>0 such that for any eigenfunction ϕL2(X)\phi\in L^{2}(X) of τ\tau with eigenvalue Λ>0\Lambda>0, we have

ϕUϵ21Λ(#Pτ+c(τ)+τϕUϵ2V𝒱ϕVδ2).\left\|{\phi_{U_{\epsilon}}}\right\|^{2}\leq\frac{1}{\Lambda}\left(\#P\left\|{\tau}\right\|_{\infty}+c(\tau)+\frac{\left\|{\tau}\right\|_{\infty}}{\left\|{\phi_{U_{\epsilon}}}\right\|^{2}}\sum_{V\in\mathcal{V}}\left\|{\phi_{V_{\delta}}}\right\|^{2}\right)\,.

Here c(τ)c(\tau) is the constant defined in (5.3), and pp is the number of elements in supp(τ)\operatorname{supp}(\tau) which are parallel to UU.

Proof.

We estimate the expression

(6.1) τ.ϕUϵ,ϕUϵ\left\langle\tau.\phi_{U_{\epsilon}},\phi_{U_{\epsilon}}\right\rangle

in two different ways: a geometric upper bound and a spectral lower bound.

On the geometric side let S=PT𝔾(𝔸f)S=P\sqcup T\subset\mathbb{G}(\mathbb{A}_{\mathrm{f}}) be a set of representatives for the support of τ\tau, partitioned into the subsets of transverse and parallel elements. By the triangle inequality

τ.ϕUϵ,ϕUϵτgfSX|ϕUϵ(ggf)ϕUϵ(g)¯|dg.\left\langle\tau.\phi_{U_{\epsilon}},\phi_{U_{\epsilon}}\right\rangle\leq\left\|{\tau}\right\|_{\infty}\sum_{g_{\mathrm{f}}\in S}\int_{X}\left|{\phi_{U_{\epsilon}}(gg_{\mathrm{f}})\overline{\phi_{U_{\epsilon}}(g)}}\right|dg.

When gfPg_{\mathrm{f}}\in P we naively apply Cauchy–Schwartz and the unitarity of the right 𝔾(𝔸f)\mathbb{G}(\mathbb{A}_{\mathrm{f}})-action to get

(6.2) X|ϕUϵ(ggf)ϕUϵ(g)¯|𝑑gϕUϵ2\int_{X}\left|{\phi_{U_{\epsilon}}(gg_{\mathrm{f}})\overline{\phi_{U_{\epsilon}}(g)}}\right|dg\leq\left\|{\phi_{U_{\epsilon}}}\right\|^{2}

Now suppose gfTg_{\mathrm{f}}\in T. The images of Uϵ{U_{\epsilon}} and Uϵgf{U_{\epsilon}}g_{\mathrm{f}} in XX intersect if and only if there is γ𝔾(F)\gamma\in\mathbb{G}(F) such that γgfKf\gamma g_{\mathrm{f}}\in K_{\mathrm{f}} and such that γUϵUϵ\gamma{U_{\epsilon}}\cap{U_{\epsilon}}\neq\emptyset. These γ\gamma belong to the intersection 𝔾(F)(UϵUϵ1)×(Kfgf1)\mathbb{G}(F)\cap({U_{\epsilon}}{U_{\epsilon}}^{-1})\times(K_{\mathrm{f}}g_{\mathrm{f}}^{-1}) of a discrete subset and a compact of 𝔾(𝔸)\mathbb{G}(\mathbb{A}) so there are fintiely many of them.

For each such γ\gamma and any δ>0\delta>0 the boundedness (=precompactness) of UU ensures the existance of ϵ>0\epsilon>0 (depending also on γ\gamma) such that

UϵγUϵ(U¯γU¯)δ.{U_{\epsilon}}\cap\gamma{U_{\epsilon}}\subset(\bar{U}\cap\gamma\bar{U})_{\delta}\,.

Since we assumed that gfg_{\mathrm{f}} is transverse the FF^{\prime}-Zariski closure 𝕄\mathbb{M} of U¯1γU¯1\bar{U}^{1}\cap\gamma\bar{U}^{1} in 𝔾~1\tilde{\mathbb{G}}_{1} has dim𝕄<dim𝕃\dim\mathbb{M}<\dim\mathbb{L}. The irreducible components of 𝕄×FEw\mathbb{M}\times_{F^{\prime}}E_{w} are defined over some finite extension F′′F^{\prime\prime} of FF^{\prime}, and we can then cover U¯γU¯\bar{U}\cap\gamma\bar{U} with finitely many algebraic pieces (V,𝕃,F′′)(V,\mathbb{L}^{\prime},F^{\prime\prime}) where 𝕃\mathbb{L}^{\prime} is an irreducible component of 𝕄×FF′′\mathbb{M}\times_{F^{\prime}}F^{\prime\prime}. The union of the δ\delta-neighborhoods of these pieces then covers the intersection UϵγUϵ{U_{\epsilon}}\cap\gamma{U_{\epsilon}}.

Now for one piece VV we have

Vδ|ϕ(ggf)ϕ(g)¯|𝑑g\displaystyle\int_{V_{\delta}}\left|{\phi(gg_{\mathrm{f}})\overline{\phi(g)}}\right|dg 12(Vδ|ϕ(γg)|2𝑑g+Vδ|ϕ(g)|2𝑑g)\displaystyle\leq\frac{1}{2}\left(\int_{V_{\delta}}\left|{\phi(\gamma g)}\right|^{2}dg+\int_{V_{\delta}}\left|{\phi(g)}\right|^{2}dg\right)
=12(γ1Vδ|ϕ(g)|2𝑑g+Vδ|ϕ(g)|2𝑑g).\displaystyle=\frac{1}{2}\left(\int_{\gamma^{-1}{V_{\delta}}}\left|{\phi(g)}\right|^{2}dg+\int_{{V_{\delta}}}\left|{\phi(g)}\right|^{2}dg\right).

Accordingly, let 𝒱\mathcal{V} denote the set of pieces VV and γ1V\gamma^{-1}V arising as components of 𝕄\mathbb{M} as gfg_{\mathrm{f}} ranges over TT and γ\gamma ranges over the finite set causing intersections (note that γ1Vδ=(γ1V)δ\gamma^{-1}{V_{\delta}}=(\gamma^{-1}V)_{\delta}).

Since this set of pieces is finite we can choose ϵ\epsilon small enough for all of them. Combining the bounds for parallel and transverse intersections then gives the geometric-side estimate

(6.3) τ.ϕUϵ,ϕUϵ#PτϕUϵ2+τV𝒱ϕVδ2.\left\langle\tau.\phi_{U_{\epsilon}},\phi_{U_{\epsilon}}\right\rangle\leq\#P\left\|{\tau}\right\|_{\infty}\left\|{\phi_{U_{\epsilon}}}\right\|^{2}+\left\|{\tau}\right\|_{\infty}\sum_{V\in\mathcal{V}}\left\|{\phi_{V_{\delta}}}\right\|^{2}\,.

We next obtain a spectral lower bound. Returning to Eq. 6.1, to the extent ϕUϵ\phi_{U_{\epsilon}} "approximates" ϕ\phi we expect τϕUϵ\tau\phi_{U_{\epsilon}} to be approximately ΛϕUϵ\Lambda\phi_{U_{\epsilon}}. This is not literally true; instead we write ϕUϵ=aϕ+ϕ\phi_{U_{\epsilon}}=a\phi+\phi^{\perp} for some ϕ\phi^{\perp} orthogonal to ϕ\phi. Here a=ϕUϵ,ϕ=ϕUϵ2a=\left\langle\phi_{U_{\epsilon}},\phi\right\rangle=\left\|{\phi_{U_{\epsilon}}}\right\|^{2} since ϕ\phi is normalized. Since also ϕϕUϵ\left\|{\phi^{\perp}}\right\|\leq\left\|{\phi_{U_{\epsilon}}}\right\| we have

τ.ϕUϵ,ϕUϵ\displaystyle\left\langle\tau.\phi_{U_{\epsilon}},\phi_{U_{\epsilon}}\right\rangle =a2τ.ϕ,ϕ+2τ.ϕ,ϕ+τ.ϕ,ϕ\displaystyle=a^{2}\left\langle\tau.\phi,\phi\right\rangle+2\Re\left\langle\tau.\phi,\phi^{\perp}\right\rangle+\left\langle\tau.\phi^{\perp},\phi^{\perp}\right\rangle
(6.4) =a2Λ+2Λϕ,ϕ+τ.ϕ,ϕ\displaystyle=a^{2}\Lambda+2\Lambda\Re\left\langle\phi,\phi^{\perp}\right\rangle+\left\langle\tau.\phi^{\perp},\phi^{\perp}\right\rangle
ϕUϵ4Λc(τ)ϕUϵ2.\displaystyle\geq\left\|{\phi_{U_{\epsilon}}}\right\|^{4}\Lambda-c(\tau)\left\|{\phi_{U_{\epsilon}}}\right\|^{2}\,.

Putting Eqs. 6.3 and 6 together and dividing by ΛϕUϵ2\Lambda\left\|{\phi_{U_{\epsilon}}}\right\|^{2} finally established the Proposition. ∎

We can now prove out main result.

Theorem 35.

Let μ\mu be a weak-* limit of normalized Hecke-eigenfunctions on XX and let (U,𝕃,F)(U,\mathbb{L},F^{\prime}) be an algebraic piece of GG so that 𝕃\mathbb{L} has small stabilizers. Then μ(πG(U))=0\mu(\pi_{G}(U))=0.

Proof.

We show by induction on dim𝕃\dim\mathbb{L} that for every η>0\eta>0 there is ϵ>0\epsilon>0 such that ϕUϵ)2η\left\|{\phi_{U_{\epsilon}}}\right\|)^{2}\leq\eta for all Hecke eigenfunctions ϕ\phi. It would follow that μ(πG(U))η\mu(\pi_{G}(U))\leq\eta for all limits.

When UU and 𝕃\mathbb{L} are empty ("dimension 1-1") there is nothing to prove, so we assume dim𝕃0\dim\mathbb{L}\geq 0. Let S<𝔾(F)S<\mathbb{G}(F) be the stabilizer of 𝕃\mathbb{L} which is small by hypothesis. With α>\alpha> to be chosen later let τ\tau be the Hecke operator constructed in 30 so that τϕ=Λϕ\tau\phi=\Lambda\phi and #Pτ,c(τ)αΛ\#P\left\|{\tau}\right\|_{\infty},c(\tau)\leq\alpha\Lambda where PP is the set of elements in the support of τ\tau which lie on the SS-orbit. By 33 every element in the support of τ\tau parallel to UU lies in the SS-orbit so 34 produces a finite set 𝒱\mathcal{V} of algebraic pieces contained in UU (hence themselves having small stabilizers) and such that for any δ>0\delta>0 there is ϵ>0\epsilon>0 such that

ϕUϵ21Λ(#Pτ+c(τ)+τϕUϵ2V𝒱ϕVδ2).\left\|{\phi_{U_{\epsilon}}}\right\|^{2}\leq\frac{1}{\Lambda}\left(\#P\left\|{\tau}\right\|_{\infty}+c(\tau)+\frac{\left\|{\tau}\right\|_{\infty}}{\left\|{\phi_{U_{\epsilon}}}\right\|^{2}}\sum_{V\in\mathcal{V}}\left\|{\phi_{V_{\delta}}}\right\|^{2}\right)\,.

By induction we can choose δ\delta small enough so that for each V𝒱V\in\mathcal{V} we have ϕVδ21#𝒱α\left\|{\phi_{V_{\delta}}}\right\|^{2}\leq\frac{1}{\#\mathcal{V}}\alpha. With this choice (and the corresponding ϵ\epsilon) we have

ϕUϵ2α(2+1ϕUϵ2),\left\|{\phi_{U_{\epsilon}}}\right\|^{2}\leq\alpha\left(2+\frac{1}{\left\|{\phi_{U_{\epsilon}}}\right\|^{2}}\right)\,,

and hence ϕUδ223α\left\|{\phi_{U_{\delta^{\prime}}}}\right\|_{2}^{2}\leq 3\sqrt{\alpha} (if we always choose ϵ<1\epsilon<1). The Theorem follows upon choosing α\alpha small enough. ∎

Proof of 2.

Without loss of generality ii=1. Let gG=iGig\in G=\prod_{i}G_{i} have its 11st coordinate in 𝔾~1(¯)\tilde{\mathbb{G}}_{1}(\bar{\mathbb{Q}}). Then 25 shows that L=g(H1M1)G,2L=g(H_{1}M_{1})G_{\infty,\geq 2} has small stabilizers. By 35 for every bounded open ULU\subset L we have μ(πG(U))=0\mu(\pi_{G}(U))=0 and covering LL by countably many such UU we conclude that μ(πG(L))=0\mu(\pi_{G}(L))=0. ∎

References

  • [1] Jean Bourgain and Elon Lindenstrauss. Entropy of quantum limits. Comm. Math. Phys., 233(1):153–171, 2003.
  • [2] Shimon Brooks and Elon Lindenstrauss. Joint quasimodes, positive entropy, and quantum unique ergodicity. Invent. Math., 198(1):219–259, 2014.
  • [3] Daniel Bump. Automorphic forms and representations, volume 55 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 1997.
  • [4] Yves Colin de Verdière. Ergodicité et fonctions propres du laplacien. Comm. Math. Phys., 102(3):497–502, 1985.
  • [5] Manfred Einsiedler and Elon Lindenstrauss. On measures invariant under diagonalizable actions: the rank-one case and the general low-entropy method. J. Mod. Dyn., 2(1):83–128, 2008.
  • [6] Stephen S. Gelbart. Automorphic forms on adèle groups. Princeton University Press, Princeton, N.J.; University of Tokyo Press, Tokyo, 1975. Annals of Mathematics Studies, No. 83.
  • [7] Elon Lindenstrauss. On quantum unique ergodicity for Γ\×\Gamma\backslash\mathbb{H}\times\mathbb{H}. Internat. Math. Res. Notices, 2001(17):913–933, 2001.
  • [8] Elon Lindenstrauss. Invariant measures and arithmetic quantum unique ergodicity. Ann. of Math. (2), 163(1):165–219, 2006.
  • [9] J. S. Milne. Algebraic groups, volume 170 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2017. The theory of group schemes of finite type over a field.
  • [10] 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.
  • [11] Zeév Rudnick and Peter Sarnak. The behaviour of eigenstates of arithmetic hyperbolic manifolds. Comm. Math. Phys., 161(1):195–213, 1994.
  • [12] Zvi Shem-Tov. Positive entropy using hecke operators at a single place. Int. Math. Res. Not. IMRN, 09 2020.
  • [13] Lior Silberman and Akshay Venkatesh. Quantum unique ergodicity for locally symmetric spaces. Geom. Funct. Anal., 17(3):960–998, 2007. arXiv:math.RT/407413.
  • [14] Lior Silberman and Akshay Venkatesh. Entropy bounds and quantum unique ergodicity for Hecke eigenfunctions on division algebras. In Probabilistic methods in geometry, topology and spectral theory, volume 739 of Contemp. Math., pages 171–197. Amer. Math. Soc., [Providence], RI, [2019] ©2019.
  • [15] Alexander I. Šnirelman. Ergodic properties of eigenfunctions. Uspekhi Mat. Nauk, 29(6(180)):181–182, 1974.
  • [16] Kannan Soundararajan. Quantum unique ergodicity for SL2()\{\rm SL}_{2}(\mathbb{Z})\backslash\mathbb{H}. Ann. of Math. (2), 172(2):1529–1538, 2010.
  • [17] Asif Zaman. Escape of mass on hilbert modular varieties. Master’s thesis, The University of British Columbia, 2012. M.Sc. thesis.
  • [18] Steven Zelditch. Pseudodifferential analysis on hyperbolic surfaces. J. Funct. Anal., 68(1):72–105, 1986.