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

Equidistribution of rational subspaces and their shapes

Menny Aka Department of Mathematics, ETH Zürich, Ramistrasse 101, Zürich, Switzerland Andrea Musso D-GESS, ETH Zürich, Ramistrasse 101, Zürich, Switzerland  and  Andreas Wieser Einstein Institute of Mathematics, Givat Ram, Hebrew University of Jerusalem, Jerusalem, Israel
Abstract.

To any kk-dimensional subspace of n\mathbb{Q}^{n} one can naturally associate a point in the Grassmannian Grn,k()\mathrm{Gr}_{n,k}(\mathbb{R}) and two shapes of lattices of rank kk and nkn-k respectively. These lattices originate by intersecting the kk-dimensional subspace and its orthogonal with the lattice n\mathbb{Z}^{n}. Using unipotent dynamics we prove simultaneous equidistribution of all of these objects under congruence conditions when (k,n)(2,4)(k,n)\neq(2,4).

A.W. was supported by ERC grant HomDyn, ID 833423, SNF grant 178958, and the SNF Doc. Mobility grant 195737.

1. Introduction

In this paper, we study the joint distribution of rational subspaces of a fixed discriminant (also called height by some authors) and of two naturally associated lattices: the integer lattice in the subspace and in its orthogonal complement together with some natural refinements.

Let QQ be a positive definite integral quadratic form on n\mathbb{Q}^{n} and let LGrn,k()L\in\mathrm{Gr}_{n,k}(\mathbb{Q}) be a rational kk-dimensional subspace. Here, Grn,k\mathrm{Gr}_{n,k} is the projective variety of kk-dimensional subspaces of the nn-dimensional linear space. The discriminant discQ(L)\mathrm{disc}_{Q}(L) of LL with respect to QQ is the discriminant of the restriction of QQ to the integer lattice L()=LnL(\mathbb{Z})=L\cap\mathbb{Z}^{n}. In a formula,

discQ(L)=det(v1,v1Qv1,vkQvk,v1Qvk,vkQ)\displaystyle\mathrm{disc}_{Q}(L)=\det\begin{pmatrix}\langle v_{1},v_{1}\rangle_{Q}&\cdots&\langle v_{1},v_{k}\rangle_{Q}\\ \vdots&&\vdots\\ \langle v_{k},v_{1}\rangle_{Q}&\cdots&\langle v_{k},v_{k}\rangle_{Q}\end{pmatrix}

where ,Q\langle\cdot,\rangle_{Q} is the bilinear form induced by QQ and v1,,vkv_{1},\ldots,v_{k} is a basis of L()L(\mathbb{Z}). We consider the finite set

Qn,k(D):={LGrn,k():discQ(L)=D}.\displaystyle\mathcal{H}^{n,k}_{Q}(D)\mathrel{\mathop{\mathchar 58\relax}}=\{L\in\mathrm{Gr}_{n,k}(\mathbb{Q})\mathrel{\mathop{\mathchar 58\relax}}\mathrm{disc}_{Q}(L)=D\}.

We attach to any LGrn,k()L\in\mathrm{Gr}_{n,k}(\mathbb{Q}) the restriction of QQ to L()L(\mathbb{Z}) represented in a basis. This is an integral quadratic form in kk-variables which is well-defined up to a change of basis i.e. (in the language of quadratic forms) up to equivalence. In particular, it defines a well-defined point – also called the shape of L()L(\mathbb{Z})

[L()]𝒮k\displaystyle[L(\mathbb{Z})]\in\mathcal{S}_{k}

where 𝒮k\mathcal{S}_{k} is the space of positive definite real quadratic forms on n\mathbb{R}^{n} up to similarity (i.e. up to equivalence and positive multiples). We may identify 𝒮k\mathcal{S}_{k} as

𝒮kOk()\PGLk()/ ​​PGLk()\displaystyle\mathcal{S}_{k}\simeq\mathchoice{\text{\lower 2.15277pt\hbox{$\mathrm{O}_{k}(\mathbb{R})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathrm{PGL}_{k}(\mathbb{R})$\!}\big{/} \lower 2.15277pt\hbox{\!\!$\mathrm{PGL}_{k}(\mathbb{Z})$}}}{\mathrm{O}_{k}(\mathbb{R})\,\backslash\,\mathrm{PGL}_{k}(\mathbb{R})\,/\,\mathrm{PGL}_{k}(\mathbb{Z})}{\mathrm{O}_{k}(\mathbb{R})\,\backslash\,\mathrm{PGL}_{k}(\mathbb{R})\,/\,\mathrm{PGL}_{k}(\mathbb{Z})}{\mathrm{O}_{k}(\mathbb{R})\,\backslash\,\mathrm{PGL}_{k}(\mathbb{R})\,/\,\mathrm{PGL}_{k}(\mathbb{Z})}

which in particular equips 𝒮k\mathcal{S}_{k} with a probability measure m𝒮km_{\mathcal{S}_{k}} arising from the Haar measures of the groups on the right. We will simply call m𝒮km_{\mathcal{S}_{k}} the Haar probability measure on 𝒮k\mathcal{S}_{k}.

Analogously, one may define the point [L()]𝒮nk[L^{\perp}(\mathbb{Z})]\in\mathcal{S}_{n-k} where LL^{\perp} is the orthogonal complement of LL with respect to QQ. Overall, we obtain a triple of points (L,[L()],[L()])(L,[L(\mathbb{Z})],[L^{\perp}(\mathbb{Z})]). The goal of this work is to study the distribution of these points in Grn,k()×𝒮k×𝒮nk\mathrm{Gr}_{n,k}(\mathbb{R})\times\mathcal{S}_{k}\times\mathcal{S}_{n-k} as discQ(L)\mathrm{disc}_{Q}(L) grows. In what follows, Grn,k()\mathrm{Gr}_{n,k}(\mathbb{R}) is given the unique SOQ()\mathrm{SO}_{Q}(\mathbb{R})-invariant probability measure mGrn,k()m_{\mathrm{Gr}_{n,k}(\mathbb{R})}.

Conjecture 1.1.

Let k,nk,n\in\mathbb{N} be integers such that k2k\geq 2 and nk2n-k\geq 2. Then the sets

{(L,[L()],[L()]):LQn,k(D)}\displaystyle\{(L,[L(\mathbb{Z})],[L^{\perp}(\mathbb{Z})])\mathrel{\mathop{\mathchar 58\relax}}L\in\mathcal{H}^{n,k}_{Q}(D)\}

equidistribute111Implicitly, we mean with respect to the product ’Haar’ measure, i.e, the product measure mGrn,k()m𝒮km𝒮nkm_{\mathrm{Gr}_{n,k}(\mathbb{R})}\otimes m_{\mathcal{S}_{k}}\otimes m_{\mathcal{S}_{n-k}}. in Grn,k()×𝒮k×𝒮nk\mathrm{Gr}_{n,k}(\mathbb{R})\times\mathcal{S}_{k}\times\mathcal{S}_{n-k} as DD\to\infty along DD\in\mathbb{N} satisfying Qn,k(D)\mathcal{H}^{n,k}_{Q}(D)\neq\emptyset.

Remark 1.2.

There exists an analogous conjecture for k=1k=1, nk2n-k\geq 2 where one only considers the pairs (L,[L()])(L,[L^{\perp}(\mathbb{Z})]) (and similarly for nk=1n-k=1, k2k\geq 2). This has been studied extensively by the first named author with Einsiedler and Shapira in [AES-dim3, AES-higherdim] where the conjecture is settled for n6n\geq 6 (i.e. nk5n-k\geq 5), for n=4,5n=4,5 under a weak congruence condition and for n=3n=3 under a stronger congruence condition on DD. We remark that, as it is written, [AES-dim3, AES-higherdim] treat only the case where QQ is the sum of squares (that we will sometimes call the standard form), but the arguments carry over without major difficulties. Using effective methods from homogeneous dynamics, Einsiedler, Rühr and Wirth [Shapes-effective] proved an effective version of the conjecture when n=4,5n=4,5 removing in particular all congruence conditions. The case n=3n=3 relies on a deep classification theorem for joinings by Einsiedler and Lindenstrauss [joiningsfinal]; effective versions of that theorem are well out of reach of current methods from homogeneous dynamics. Assuming the Generalized Riemann Hypothesis, Blomer and Brumley [BlomerBrumley] have recently removed the congruence condition in [AES-dim3].

Remark 1.3.

The case k=2k=2 and nk=2n-k=2 of Conjecture 1.1 has been settled in [2in4] by the first and the last named author together with Einsiedler under a (relatively strong) congruence condition when QQ is the sum of four squares. The result in the paper is in fact stronger as it considers two additional shapes that one can naturally associate to LL essentially thanks to the local isomorphism between SO4()\mathrm{SO}_{4}(\mathbb{R}) and SO3()×SO3()\mathrm{SO}_{3}(\mathbb{R})\times\mathrm{SO}_{3}(\mathbb{R}). The arguments carry over without major difficulties to consider norm forms on quaternion algebras (equivalently, the forms QQ for which disc(Q)\mathrm{disc}(Q) is a square in ×\mathbb{Q}^{\times}). In [2in4General], the first and last named author will extend the results of [2in4] to treat arbitrary quadratic forms.

In this article, we prove Conjecture 1.1 in the remaining cases, partially under congruence conditions. For integers D,D,\ell we write D[]D^{[\ell]} for the \ell-power free part of DD i.e. the largest divisor dd of DD with ada^{\ell}\nmid d for any a>1a>1.

Theorem 1.4 (Equidistribution of subspaces and shapes).

Let 2kn2\leq k\leq n be integers with knkk\leq n-k and nk>3n-k>3, and let pp be an odd prime with pdisc(Q)p\nmid\mathrm{disc}(Q). Let DiD_{i}\in\mathbb{N} be a sequence of integers with Di[k]D_{i}^{[k]}\to\infty and Qn,k(Di)\mathcal{H}_{Q}^{n,k}(D_{i})\neq\emptyset for every ii. Then the sets

{(L,[L()],[L()]):LQn,k(Di)}\displaystyle\{(L,[L(\mathbb{Z})],[L^{\perp}(\mathbb{Z})])\mathrel{\mathop{\mathchar 58\relax}}L\in\mathcal{H}_{Q}^{n,k}(D_{i})\}

equidistribute in Grn,k()×𝒮k×𝒮nk\mathrm{Gr}_{n,k}(\mathbb{R})\times\mathcal{S}_{k}\times\mathcal{S}_{n-k} as ii\to\infty assuming the following conditions:

  • pDip\nmid D_{i} if k{3,4}k\in\{3,4\}.

  • Dimodp-D_{i}\mod p is a square in 𝔽p×\mathbb{F}_{p}^{\times} if k=2k=2.

Moreover, the analogous statement holds when the roles of kk and nkn-k are reversed.

Remark 1.5.

Maass [Maass56, Maass59] in the 60’s and Schmidt [Schmidt-shapes] in the 90’s have considered problems of this kind. They prove that the set of pairs (L,[L()])(L,[L(\mathbb{Z})]) equidistributes in Grn,k()×𝒮k\mathrm{Gr}_{n,k}(\mathbb{R})\times\mathcal{S}_{k} where LGrn,k()L\in\mathrm{Gr}_{n,k}(\mathbb{Q}) varies over the rational subspaces with discriminant at most DD. In this averaged setup, Horesh and Karasik [HoreshKarasik] recently verified Conjecture 1.1. Indeed, their version is polynomially effective in DD.

Remark 1.6 (Congruence conditions).

As in the previous works referenced in Remarks 1.2 and 1.3, our proof is of dynamical nature and follows from an equidistribution result for certain orbits in an adelic homogeneous space. The congruence conditions at the prime pp assert roughly speaking that one can use non-trivial dynamics at one fixed place for all DD. The acting groups we consider here are (variations of) the p\mathbb{Q}_{p}-points of

𝐇L={gSOQ:g.LL}\displaystyle\mathbf{H}_{L}=\{g\in\mathrm{SO}_{Q}\mathrel{\mathop{\mathchar 58\relax}}g.L\subset L\}^{\circ}

for LGrn,k()L\in\mathrm{Gr}_{n,k}(\mathbb{Q}). In particular, the cases k=2k=2 and k>2k>2 are very different from a dynamical viewpoint:

  • For k>2k>2 the group 𝐇L\mathbf{H}_{L} is semisimple. The knowledge about measures on homogeneous spaces invariant under unipotents is vast – see Ratner’s seminal works [ratner91-measure, ratner-p-adic]. In our situation, we use an SS-arithmetic version of a theorem by Mozes and Shah [mozesshah], proven by Gorodnik and Oh [gorodnikoh], which describes weak-limits of measures with invariance under a semisimple group. Roughly speaking, the theorem implies that any sequence of orbits under a semisimple subgroup is either equidistributed or sits (up to a small shift) inside an orbit of a larger subgroup. The flexibility that this method provides allows us to in fact prove a significantly stronger result; see Theorem 1.11 below.

  • For k=2k=2 and nk3n-k\geq 3, the group 𝐇L\mathbf{H}_{L} is reductive. Thus, one can apply the results mentioned in the previous bullet point only to the commutator subgroup of 𝐇L\mathbf{H}_{L} which is non-maximal and has intermediate subgroups.

One of the novelties of this article is a treatment of this reductive case where we use additional invariance under the center to rule out intermediate subgroups ’on average’ (see §4.3). Here as well as for the second component of the triples in Theorem 1.4 we need equidistribution of certain adelic torus orbits; this is a generalized version of a theorem of Duke [duke88] building on a breakthrough of Iwaniec [iwaniecforduke] – see for example [Dukeforcubic, harcosmichelII, W-linnik]. Furthermore, to prove simultaneous equidistribution of the tuples in Theorem 1.4 we apply a new simple disjointness trick – see the following remark.

Remark 1.7 (Disjointness).

In upcoming work, the first and last named author prove together with Einsiedler, Luethi and Michel [EffDisjoint] an effective version of Conjecture 1.1 when k2k\neq 2. This removes in particular the congruence conditions. The technique consists of a method to ’bootstrap’ effective equidistribution in the individual factors to simultaneous effective equidistribution (in some situations).

In the current article, we use an ineffective analogue of this to prove Theorem 1.4, namely, the very well-known fact that mixing systems are disjoint from trivial systems (see also Lemma 4.2). This simple trick has (to our knowledge) not yet appeared in the literature in a similar context. It is particularly useful when k=2k=2 and nk3n-k\geq 3 in which case we cannot rely solely on methods from unipotent dynamics (see Remark 1.6).

Remark 1.8 (On the power assumption).

The assumption in Theorem 1.4 toward the power free part of the discriminants should only be considered a simplifying assumption. Its purpose is automatically to rule out situations where for most subspaces LQn,k(D)L\in\mathcal{H}^{n,k}_{Q}(D) the quadratic form Q|L()Q|_{L(\mathbb{Z})} (or Q|L()Q|_{L^{\perp}(\mathbb{Z})}) is highly imprimitive (i.e. a multiple of a quadratic form of very small discriminant). We expect that such discriminants do not exist regardless of their factorization. A conjecture in this spirit is phrased in Appendix B. Moreover, Schmidt’s work [Schmidt-count] suggests that |Qn,k(D)|=Dn21+o(1)|\mathcal{H}^{n,k}_{Q}(D)|=D^{\frac{n}{2}-1+o(1)} in which case one could remove the assumption Di[k]D_{i}^{[k]}\to\infty in Theorem 1.4.

1.1. A strengthening

In the following we present a strengthening of Conjecture 1.1 inspired by the notion of grids introduced in [AES-higherdim] and by Bersudsky’s construction of a moduli space [BersudskyModuli] which refines the results of [AES-higherdim].

Consider the set of pairs (L,Λ)(L,\Lambda) where LnL\subset\mathbb{R}^{n} is a kk-dimensional subspace and where Λn\Lambda\subset\mathbb{R}^{n} is a lattice of full rank with the property that LΛL\cap\Lambda is a lattice in LL (LL is Λ\Lambda-rational). We define an equivalence relation on these pairs by setting (L,Λ)(L,Λ)(L,\Lambda)\sim(L^{\prime},\Lambda^{\prime}) whenever the following conditions are satisfied:

  1. (1)

    L=LL=L^{\prime},

  2. (2)

    There exists gGLn()g\in\mathrm{GL}_{n}(\mathbb{R}) with det(g)>0\det(g)>0 such that gg acts on LL and LL^{\perp} as scalar multiplication and gΛ=Λg\Lambda=\Lambda^{\prime}.

We write [L,Λ][L,\Lambda] for the class of (L,Λ)(L,\Lambda); elements of such a class are said to be homothetic along LL or LL-homothetic to (L,Λ)(L,\Lambda). We refer to the set 𝒴\mathcal{Y} of such equivalence classes as the moduli space of basis extensions. Indeed, one can think of a lattice Λ\Lambda such that LΛL\cap\Lambda is a lattice as one choice of complementing the lattice LΛL\cap\Lambda into a basis of n\mathbb{R}^{n}. The equivalence relation is not very transparent in this viewpoint, see Section 6 for further discussion.

The moduli space 𝒴\mathcal{Y} is designed to incorporate subspaces as well as both shapes. Clearly, we have a well-defined map

(1.1) [L,Λ]𝒴LGrn,k()\displaystyle[L,\Lambda]\in\mathcal{Y}\mapsto L\in\mathrm{Gr}_{n,k}(\mathbb{R})

The restriction of QQ to LΛL\cap\Lambda yields a well-defined element of 𝒮k\mathcal{S}_{k}. Similarly, one may check that LL^{\perp} intersects the dual lattice Λ#\Lambda^{\#} in a lattice, the second shape is given by the restriction of QQ to LΛ#L^{\perp}\cap\Lambda^{\#}.

We note that there is a natural identification of 𝒴\mathcal{Y} with a double quotient of a Lie group (cf. Lemma 6.3) so that we may again speak of the ’Haar measure’ on 𝒴\mathcal{Y}.

Conjecture 1.9.

Let k,nk,n\in\mathbb{N} be integers such that k3k\geq 3 and nk3n-k\geq 3. Then the sets

{([L,n]:LQn,k(D)}𝒴\displaystyle\{([L,\mathbb{Z}^{n}]\mathrel{\mathop{\mathchar 58\relax}}L\in\mathcal{H}^{n,k}_{Q}(D)\}\subset\mathcal{Y}

equidistribute with respect to the Haar measure as DD\to\infty along DD\in\mathbb{N} satisfying Qn,k(D)\mathcal{H}^{n,k}_{Q}(D)\neq\emptyset.

Remark 1.10 (From Conjecture 1.9 to Conjecture 1.1).

When QQ is unimodular (i.e. disc(Q)=1\mathrm{disc}(Q)=1), Conjecture 1.9 implies Conjecture 1.1. Otherwise, Conjecture 1.9 implies equidistribution of the triples (L,[L()],[L(n)#])(L,[L(\mathbb{Z})],[L^{\perp}\cap(\mathbb{Z}^{n})^{\#}]) where (n)#(\mathbb{Z}^{n})^{\#} is the dual lattice to n\mathbb{Z}^{n} under the quadratic form QQ:

(n)#={xn:x,yQ for all yn}.\displaystyle(\mathbb{Z}^{n})^{\#}=\{x\in\mathbb{Q}^{n}\mathrel{\mathop{\mathchar 58\relax}}\langle x,y\rangle_{Q}\in\mathbb{Z}\text{ for all }y\in\mathbb{Z}^{n}\}.

This is not significantly different as the lattice L(n)#L^{\perp}\cap(\mathbb{Z}^{n})^{\#} contains LnL^{\perp}\cap\mathbb{Z}^{n} with index at most disc(Q)\mathrm{disc}(Q); it is nevertheless insufficient to deduce Conjecture 1.1. In §6 we introduce tuples [L,ΛL][L,\Lambda_{L}] which satisfy an analogue of Conjecture 1.9; this adapted conjecture implies Conjecture 1.1

We prove the following towards Conjecture 1.9.

Theorem 1.11.

Let k,nk,n be integers with 3knk3\leq k\leq n-k and let pp be an odd prime with pdisc(Q)p\nmid\mathrm{disc}(Q). Let DiD_{i}\in\mathbb{N} be a sequence of integers with Di[k]D_{i}^{[k]}\to\infty and Qn,k(Di)\mathcal{H}_{Q}^{n,k}(D_{i})\neq\emptyset for every ii. Then the sets

{([L,n]:LQn,k(Di)}\displaystyle\{([L,\mathbb{Z}^{n}]\mathrel{\mathop{\mathchar 58\relax}}L\in\mathcal{H}_{Q}^{n,k}(D_{i})\}

equidistribute in 𝒴\mathcal{Y} as ii\to\infty assuming in addition that pDip\nmid D_{i} if k{3,4}k\in\{3,4\}.

Remark 1.12.

As mentioned in Remark 1.6, the assumption k3k\geq 3 and nk3n-k\geq 3 asserts that the acting group underlying the problem is semisimple. There are instances where one could overcome this obstacle: Khayutin [Khayutin-kugasato] proves equidistribution of grids when (k,n)=(1,3)(k,n)=(1,3) as conjectured in [AES-higherdim] using techniques from geometric invariant theory.

1.2. Further refinements and questions

For an integral quadratic form qq in kk variables a primitive representation of qq by QQ is a \mathbb{Z}-linear map ι:kn\iota\mathrel{\mathop{\mathchar 58\relax}}\mathbb{Z}^{k}\to\mathbb{Z}^{n} such that Q(ι(v))=q(v)Q(\iota(v))=q(v) for all vkv\in\mathbb{Z}^{k} and such that ι(k)n=ι(k)\mathbb{Q}\iota(\mathbb{Z}^{k})\cap\mathbb{Q}^{n}=\iota(\mathbb{Z}^{k}). One can identify primitive representations of qq with subspaces LGrn,k()L\in\mathrm{Gr}_{n,k}(\mathbb{Q}) such that Q|L()Q|_{L(\mathbb{Z})} is equivalent to qq. Given this definition, one could ask about the distribution of the pairs

(1.2) {(L,[L()]):LGrn,k() and Q|L() is equivalent to q}\displaystyle\{(L,[L^{\perp}(\mathbb{Z})])\mathrel{\mathop{\mathchar 58\relax}}L\in\mathrm{Gr}_{n,k}(\mathbb{Q})\text{ and }Q|_{L(\mathbb{Z})}\text{ is equivalent to }q\}

inside Grn,k()×𝒮nk\mathrm{Gr}_{n,k}(\mathbb{R})\times\mathcal{S}_{n-k} when disc(q)\mathrm{disc}(q)\to\infty. The condition disc(q)\mathrm{disc}(q)\to\infty here is not sufficient; for instance, when qq represents 11 and QQ represents 11 only on, say, ±vn\pm v\in\mathbb{Z}^{n} then any primitive representation of qq by QQ must contain ±v\pm v. However, the subspaces in Grn,k()\mathrm{Gr}_{n,k}(\mathbb{R}) containing ±v\pm v form a Zariski closed subset. Assuming that the minimal value represented by qq goes to infinity, the above question is very strongly related to results of Ellenberg and Venkatesh [localglobalEV] as are indeed our techniques in this article. In principle, these techniques should apply to show that under congruence conditions as in Theorems 1.4 and 1.11 the pairs in (1.2) are equidistributed when qiq_{i} is a sequence of quadratic forms primitively representable by QQ whose minimal values tend to infinity.

As alluded to in Remark 1.12 it would be interesting to know if Khayutin’s technique applies to show the analogue of Theorem 1.11 when, say, (k,n)=(2,5),(2,4)(k,n)=(2,5),(2,4). The two cases are from a dynamical perspective quite different as noted in Remark 1.6.

Furthermore, we note that this paper has various clear directions of possible generalization. Most notably, this paper can be extended to indefinite forms. Let QQ be an indefinite integral quadratic form on n\mathbb{Q}^{n} of signature (r,s)(r,s). Here, we observe that SOQ()\mathrm{SO}_{Q}(\mathbb{R}) does not act transitively on Grn,k()\mathrm{Gr}_{n,k}(\mathbb{R}). Indeed, the degenerate subspaces form a Zariski closed subset (the equation being disc(Q|L)=0\mathrm{disc}(Q|_{L})=0). The complement is a disjoint union of finitely many open sets on which SOQ()\mathrm{SO}_{Q}(\mathbb{R}) acts transitively; for each tuple (r,s)(r^{\prime},s^{\prime}) with r+s=kr^{\prime}+s^{\prime}=k and rr,ssr^{\prime}\leq r,\ s^{\prime}\leq s such an open set is given by the subspaces LL for which Q|LQ|_{L} has signature (r,s)(r^{\prime},s^{\prime}). The analogue of the above conjectures and theorems can then be formulated by replacing Grn,k()\mathrm{Gr}_{n,k}(\mathbb{R}) with one of these open sets. The proofs generalize without major difficulties to this case; we refrain from doing so here for simplicity of the exposition. Other directions of generalization include the number field case which is not addressed in any of the works prior to this article and is hence interesting in other dimensions as well.

Acknowledgments The authors would like to thank Michael Bersudsky, Manfred Einsiedler and Manuel Luethi for useful discussions. We are also thankful towards the anonymous referee who made various valuable suggestions towards improving the exposition.

1.3. Organization of the paper

This article consists of two parts. In Part 11 – the ’dynamical’ part – we establish the necessary results concerning equidistribution of certain adelic orbits. It is structured as follows:

  • In §2, we prove various results concerning stabilizer subgroups of subspaces.

  • In §3, we prove the homogeneous analogue of Theorem 1.11. The key ingredient of our proof is an SS-arithmetic extension of a theorem of Mozes and Shah [mozesshah] proven by Gorodonik and Oh [gorodnikoh]. The arguments used in this section only work when the dimension and codimension (i.e. kk and nkn-k) are at least 3.

  • \bullet

    In §4, we prove the homogeneous analogue of Theorem 1.4 for two dimensional subspaces (i.e. for k=2k=2). Contrary to the case of dimension and codimension at least 3, the groups whose dynamics we use are not semisimple (see Remark 1.6). In particular, the theorem of Gorodonik and Oh [gorodnikoh] is not sufficient and more subtle arguments, relying on Duke’s Theorem [duke88] and the trick mentioned in Remark 1.7, are required.

In Part 22, we deduce Theorems 1.4 and 1.11 from the homogeneous dynamics results proven in §3 (k>2k>2) and §4 (k=2k=2) of the first part. More precisely, it is structured as follows:

  1. \bullet

    In §5, we prove that the discriminant of the orthogonal complement of a subspace is equal to the discriminant of the subspace up to an essentially negligible factor.

  2. \bullet

    In §6, we study the moduli space of base extensions and show that it surjects onto Grn,k()×𝒮k×𝒮nk\mathrm{Gr}_{n,k}(\mathbb{R})\times\mathcal{S}_{k}\times\mathcal{S}_{n-k}. From this, we prove that a slight strengthening of Theorem 1.11 implies Theorem 1.4. In these considerations, it is useful to consider subspaces together with an orientation.

  3. \bullet

    In §7, we finally establish Theorems 1.4 and 1.11. The technique here is by now standard – we interpret the sets in Theorem 1.11 as projections of the adelic orbits in Part 11 (or a slight adaptation thereof).

In the appendix, we establish various complementary facts.

  1. \bullet

    In Appendix A, we discuss non-emptiness conditions for the set Qn,k(D)\mathcal{H}^{n,k}_{Q}(D) when the quadratic form QQ is the sum of squares. In particular, we prove that Qn,k(D)\mathcal{H}^{n,k}_{Q}(D)\neq\emptyset for all n5n\geq 5. The techniques here are completely elementary and we do not provide any counting results.

  2. \bullet

    In Appendix B, we prove various facts complementing the discussion in §5. For instance, we prove that if LGrn,k()L\in\mathrm{Gr}_{n,k}(\mathbb{Q}) is a subspace where k<nkk<n-k then the quadratic form on the orthogonal complement Q|L()Q|_{L^{\perp}(\mathbb{Z})} is primitive up to negligible factors.

1.4. Notation

Let VV_{\mathbb{Q}} be the set of places of \mathbb{Q} and denote by v\mathbb{Q}_{v} for any vVv\in V_{\mathbb{Q}} the completion at vv. Given a subset SVS\subset V_{\mathbb{Q}} we define the ring S\mathbb{Q}_{S} to be the restricted direct product of p\mathbb{Q}_{p} for pSp\in S with respect to the subgroups p\mathbb{Z}_{p} for pS{}p\in S\setminus\{\infty\}. Moreover, we set S:=[1p:pS{}}\mathbb{Z}^{S}\mathrel{\mathop{\mathchar 58\relax}}=\mathbb{Z}[\frac{1}{p}\mathrel{\mathop{\mathchar 58\relax}}p\in S\setminus\{\infty\}\}. When S=VS=V_{\mathbb{Q}} we denote S\mathbb{Q}_{S} by 𝔸\mathbb{A} and call it the ring of adeles. When instead S=V{}S=V_{\mathbb{Q}}\setminus\{\infty\} we denote S\mathbb{Q}_{S} by 𝔸f\mathbb{A}_{f} and call it the ring of finite adeles. Finally, we let ^=pV{}p\hat{\mathbb{Z}}=\prod_{p\in V_{\mathbb{Q}}\setminus\{\infty\}}\mathbb{Z}_{p}.

Let 𝐆<SLN\mathbf{G}<\mathrm{SL}_{N} be a connected algebraic group defined over \mathbb{Q}. We identify 𝐆(S)=𝐆(S)SLN(S)\mathbf{G}(\mathbb{Z}^{S})=\mathbf{G}(\mathbb{Q}_{S})\cap\mathrm{SL}_{N}(\mathbb{Z}^{S}) with its diagonally embedded copy in 𝐆(S)\mathbf{G}(\mathbb{Q}_{S}). If 𝐆\mathbf{G} has no non-trivial \mathbb{Q}-characters (for instance when the radical of 𝐆\mathbf{G} is unipotent), the Borel-Harish-Chandra Theorem (see [platonov, Thm. 5.5]) yields that 𝐆(S)\mathbf{G}(\mathbb{Z}^{S}) is a lattice in 𝐆(S)\mathbf{G}(\mathbb{Q}_{S}) whenever S\infty\in S. In particular, the quotient 𝐆(S)/𝐆(S)\mathbf{G}(\mathbb{Q}_{S})/\mathbf{G}(\mathbb{Z}^{S}) is a finite volume homogeneous space. For g𝐆(S)g\in\mathbf{G}(\mathbb{Q}_{S}) and vSv\in S, gvg_{v} denotes the vv-adic component of gg.

Whenever 𝐆\mathbf{G} is semisimple, we denote by 𝐆(S)+\mathbf{G}(\mathbb{Q}_{S})^{+} the image of the simply connected cover in 𝐆(S)\mathbf{G}(\mathbb{Q}_{S}) (somewhat informally, this can be thought of as the part of 𝐆(S)\mathbf{G}(\mathbb{Q}_{S}) which is generated by unipotents).

1.4.1. Quadratic forms

In this whole article, (V,Q)(V,Q) is a fixed non-degenerate quadratic space over \mathbb{Q} of dimension nn. The induced bilinear form is denoted by ,Q\langle\cdot,\cdot\rangle_{Q}. We assume throughout that (V,Q)(V,Q) is positive definite. We also identify VV with n\mathbb{Q}^{n} and suppose that ,Q\langle\cdot,\cdot\rangle_{Q} takes integral values on n×n\mathbb{Z}^{n}\times\mathbb{Z}^{n} in which case we say that QQ is integral. Equivalently, the matrix representation MQM_{Q} in the standard basis of n\mathbb{Z}^{n} has integral entries.

We denote by OQ\mathrm{O}_{Q} resp. SOQ\mathrm{SO}_{Q} the orthogonal resp. special orthogonal group for QQ. Recall that SOQ\mathrm{SO}_{Q} is abelian if dim(V)=2\dim(V)=2 and semisimple otherwise. We denote by SpinQ\mathrm{Spin}_{Q} the spin group for QQ which is the simply connected cover of SOQ\mathrm{SO}_{Q} if dim(V)>2\dim(V)>2. Explicitly, the spin group may be constructed from the Clifford algebra of QQ. We remark that this article contains certain technicalities that will use the Clifford algebra – we refer to [knus] for a thorough discussion. The spin group comes with an isogeny of \mathbb{Q}-groups ρQ:SpinQSOQ\rho_{Q}\mathrel{\mathop{\mathchar 58\relax}}\mathrm{Spin}_{Q}\to\mathrm{SO}_{Q} which satisfies that for any field KK of characteristic zero we have an exact sequence (cf. [knus, p. 64])

SpinQ(K)SOQ(K)K×/(K×)2.\mathrm{Spin}_{Q}(K)\rightarrow\mathrm{SO}_{Q}(K)\rightarrow K^{\times}/(K^{\times})^{2}.

where the second homomorphism is given by the spinor norm. The isogeny ρQ\rho_{Q} induces an integral structure on SpinQ\mathrm{Spin}_{Q}. For instance, SpinQ()\mathrm{Spin}_{Q}(\mathbb{Z}) consists of elements gSpinQ()g\in\mathrm{Spin}_{Q}(\mathbb{Q}) for which ρQ(g)SOQ()\rho_{Q}(g)\in\mathrm{SO}_{Q}(\mathbb{Z}). To simplify notation, we will write g.vg.v for the action of SpinQ\mathrm{Spin}_{Q} on a vector in nn-dimensional linear space. Here, the action is naturally induced by the isogeny ρQ\rho_{Q} (and the standard representation of SOQ\mathrm{SO}_{Q}).

Furthermore, we let Grn,k\mathrm{Gr}_{n,k} denote the Grassmannian of kk-dimensional subspaces of VV. Note that this is a homogeneous variety for SOQ\mathrm{SO}_{Q} and (through the isogeny ρQ\rho_{Q}) also for SpinQ\mathrm{Spin}_{Q}. If we assume that QQ is positive definite (as we always do), the action of SOQ()\mathrm{SO}_{Q}(\mathbb{R}) on Grn,k()\mathrm{Gr}_{n,k}(\mathbb{R}) is transitive. Furthermore, in this case the spinor norm on SOQ()\mathrm{SO}_{Q}(\mathbb{R}) takes only positive values so that SpinQ()\mathrm{Spin}_{Q}(\mathbb{R}) surjects onto SOQ()\mathrm{SO}_{Q}(\mathbb{R}) and in particular also acts transitively.

We denote the standard positive definite form (i.e. the sum of nn squares) by Q0Q_{0} and write SOn\mathrm{SO}_{n} for its special orthogonal group. As Q0Q_{0} and QQ have the same signature, there exists ηQGLn()\eta_{Q}\in\mathrm{GL}_{n}(\mathbb{R}) with det(ηQ)>0\det(\eta_{Q})>0 such that ηQtηQ=MQ\eta_{Q}^{t}\eta_{Q}=M_{Q} or equivalently

(1.3) Q0(ηQx)=Q(x)\displaystyle Q_{0}(\eta_{Q}x)=Q(x)

holds for all xnx\in\mathbb{R}^{n} (similarly for the induced bilinear forms). In particular, ηQ\eta_{Q} maps pairs of vectors in VV which are orthogonal with respect to QQ onto pairs of vectors which are orthogonal with respect to Q0Q_{0}. Also, ηQ1SOn()ηQ=SOQ()\eta_{Q}^{-1}\mathrm{SO}_{n}(\mathbb{R})\eta_{Q}=\mathrm{SO}_{Q}(\mathbb{R}).

1.4.2. Quadratic forms on sublattices and discriminants

For any finitely generated \mathbb{Z}-lattice Γ<n\Gamma<\mathbb{Q}^{n} (of arbitrary rank) the restriction of QQ to Γ\Gamma induces a quadratic form. We denote by qΓq_{\Gamma} the representation of this form in a choice of basis of Γ\Gamma. Hence, qΓq_{\Gamma} is well-defined up to equivalence (and not proper equivalence) of quadratic forms (i.e. up to change of basis).

If Γ<n\Gamma<\mathbb{Z}^{n}, qΓq_{\Gamma} is an integral quadratic form and we denote by gcd(qΓ)\gcd(q_{\Gamma}) the greatest common divisor of its coefficients (which is independent of the choice of basis). Note that gcd(qΓ)\gcd(q_{\Gamma}) is sometimes also referred to as the content of qΓq_{\Gamma}. We write q~Γ=1gcd(qΓ)qΓ\tilde{q}_{\Gamma}=\frac{1}{\gcd(q_{\Gamma})}q_{\Gamma} for the primitive multiple of qΓq_{\Gamma}. If LnL\subset\mathbb{Q}^{n} is a subspace, we sometimes write qLq_{L} instead of qL()q_{L(\mathbb{Z})} for simplicity.

The discriminant discQ(Γ)\mathrm{disc}_{Q}(\Gamma) of a finitely generated \mathbb{Z}-lattice Γ<n\Gamma<\mathbb{Q}^{n} is the discriminant of qΓq_{\Gamma}. As at the beginning of the introduction, we write discQ(L)\mathrm{disc}_{Q}(L) instead of discQ(L())\mathrm{disc}_{Q}(L(\mathbb{Z})) for any subspace LnL\subset\mathbb{Q}^{n}. Given a prime pp we also define

(1.4) discp,Q(L)=disc(Q|L(p))p/(p×)2\displaystyle\mathrm{disc}_{p,Q}(L)=\mathrm{disc}(Q|_{L(\mathbb{Z}_{p})})\in\mathbb{Z}_{p}/(\mathbb{Z}_{p}^{\times})^{2}

where L(p)=L(p)pnL(\mathbb{Z}_{p})=L(\mathbb{Q}_{p})\cap\mathbb{Z}_{p}^{n}. We have the following useful identity

(1.5) discQ(L)=ppνp(discp,Q(L))\displaystyle\mathrm{disc}_{Q}(L)=\prod_{p}p^{\nu_{p}(\mathrm{disc}_{p,Q}(L))}

where the product is taken over all primes pp and νp\nu_{p} denotes the standard pp-adic valuation. Note that only primes dividing the discriminant contribute non-trivially.

1.4.3. Choice of a reference subspace

We fix an integer knk\leq n for which we always assume that one of the following holds:

  • k3k\geq 3 and nk3n-k\geq 3,

  • k=2k=2 and nk3n-k\geq 3, or

  • k3k\geq 3 and nk=2n-k=2.

Let L0VL_{0}\subset V be given by

(1.6) L0=k×{(0,,0)}V.\displaystyle L_{0}=\mathbb{Q}^{k}\times\{(0,\ldots,0)\}\subset V.

We adapt the choice of ηQ\eta_{Q} to this reference subspace L0L_{0} and suppose that the first kk column vectors in ηQ1\eta_{Q}^{-1} are an orthonormal basis of L0L_{0}. This choice asserts that ηQ\eta_{Q} maps L0()L_{0}(\mathbb{R}) to L0()L_{0}(\mathbb{R}) and hence L0()L_{0}^{\perp}(\mathbb{R}) to {(0,,0)}×nk\{(0,\ldots,0)\}\times\mathbb{R}^{n-k},

1.4.4. Ambient groups

The following subgroups of SLn\mathrm{SL}_{n} will be useful throughout this work:

𝐏n,k\displaystyle\mathbf{P}_{n,k} ={(AB0D)SLn:det(A)=det(D)=1}\displaystyle=\Big{\{}\begin{pmatrix}A&B\\ 0&D\end{pmatrix}\in\mathrm{SL}_{n}\mathrel{\mathop{\mathchar 58\relax}}\det(A)=\det(D)=1\Big{\}}
𝐃n,k\displaystyle\mathbf{D}_{n,k} ={(A00D)SLn:det(A)=det(D)=1}.\displaystyle=\Big{\{}\begin{pmatrix}A&0\\ 0&D\end{pmatrix}\in\mathrm{SL}_{n}\mathrel{\mathop{\mathchar 58\relax}}\det(A)=\det(D)=1\Big{\}}.

where AA is a k×kk\times k-matrix, DD is an (nk)×(nk)(n-k)\times(n-k)-matrix and BB is a k×(nk)k\times(n-k)-matrix. We denote by π1\pi_{1} resp. π2\pi_{2} the projection of 𝐏n,k\mathbf{P}_{n,k} onto the upper-left resp. bottom-right block. We also define the group

𝐆=SpinQ×𝐏n,k.\displaystyle\mathbf{G}=\mathrm{Spin}_{Q}\times\mathbf{P}_{n,k}.

By 𝐆¯\bar{\mathbf{G}} we denote the Levi subgroup of 𝐆\mathbf{G} with B=0B=0 i.e.

𝐆¯\displaystyle\bar{\mathbf{G}} =SpinQ×𝐃n,kSpinQ×SLk×SLnk.\displaystyle=\mathrm{Spin}_{Q}\times\mathbf{D}_{n,k}\simeq\mathrm{Spin}_{Q}\times\mathrm{SL}_{k}\times\mathrm{SL}_{n-k}.
Remark 1.13.

Concerning the aforementioned groups we will need two well known facts. Firstly, 𝐃n,k\mathbf{D}_{n,k} is a maximal subgroup of 𝐏n,k\mathbf{P}_{n,k} (meaning that there is no connected \mathbb{Q}-group 𝐌\mathbf{M} with 𝐃n,k𝐌𝐏n,k\mathbf{D}_{n,k}\subsetneq\mathbf{M}\subsetneq\mathbf{P}_{n,k}) – see for example [emvforSld, Prop. 3.2]. Secondly, for any quadratic form qq in dd variables SOq\mathrm{SO}_{q} is maximal in SLd\mathrm{SL}_{d} – see for example [LiebeckSeitzMax] for a modern discussion of maximal subgroups of the classical groups.

1.4.5. Landau notation

In classical Landau notation, we write fgf\asymp g for two positive functions if there exist constants c,C>0c,C>0 with cfgCfcf\leq g\leq Cf. If the constants depend on another quantity aa, we sometimes write fagf\asymp_{a}g to emphasize the dependence.

Part I

Homogeneous results


\@afterheading

For an overview of the contents of this part, we refer the reader to §1.3.

2. Stabilizer groups

Recall that throughout the article QQ is a positive definite integral quadratic form on V=nV=\mathbb{Q}^{n}. In particular, any subspace of n\mathbb{Q}^{n} is non-degenerate with respect to QQ.

2.1. Stabilizers of subspaces

For any subspace L¯nL\subset\overline{\mathbb{Q}}^{n} we define the following groups:

  1. \bullet

    𝐇L<SpinQ\mathbf{H}_{L}<\mathrm{Spin}_{Q} is the identity component of the stabilizer group of LL in SpinQ\mathrm{Spin}_{Q} for the action of SpinQ\mathrm{Spin}_{Q} on Grn,k\mathrm{Gr}_{n,k}.

  2. \bullet

    𝐇L<SOQ\mathbf{H}^{\prime}_{L}<\mathrm{SO}_{Q} is the identity component of the stabilizer group of LL in SOQ\mathrm{SO}_{Q} for the action of SOQ\mathrm{SO}_{Q} on Grn,k\mathrm{Gr}_{n,k}.

Note that we have an isogeny 𝐇L𝐇L\mathbf{H}_{L}\to\mathbf{H}^{\prime}_{L}. Furthermore, the restriction to LL resp. LL^{\perp} yields an isomorphism of \mathbb{Q}-groups

(2.1) 𝐇LSOQ|L×SOQ|L.\displaystyle\mathbf{H}^{\prime}_{L}\to\mathrm{SO}_{Q|_{L}}\times\mathrm{SO}_{Q|_{L^{\perp}}}.

To see this, one needs to check that the image consists indeed of special orthogonal transformations. This follows from the fact that the determinant of the restrictions is a morphism with finite image and hence its kernel must be everything by connectedness. In particular, we have the following cases:

  • If k3k\geq 3 and nk3n-k\geq 3, 𝐇L\mathbf{H}^{\prime}_{L} (and hence also 𝐇L\mathbf{H}_{L}) is semisimple.

  • If k=2k=2 and nk3n-k\geq 3 (or k3k\geq 3 and nk=2n-k=2), 𝐇L\mathbf{H}^{\prime}_{L} is reductive.

  • If k=2k=2 and nk=2n-k=2 (which is not a case this paper covers), 𝐇L\mathbf{H}^{\prime}_{L} is abelian.

Remark 2.1 (Special Clifford groups and (2.1)).

While it might seem appealing to suspect that 𝐇L\mathbf{H}_{L} is simply-connected, this is actually false. The following vague and lengthy explanation is not needed in the sequel. Denote by 𝐌\mathbf{M} the special Clifford group of QQ and similarly by 𝐌1\mathbf{M}_{1} resp. 𝐌2\mathbf{M}_{2} the special Clifford groups of Q|LQ|_{L} resp. Q|LQ|_{L^{\perp}} (for the duration of this remark) – cf. [knus]. These are reductive groups whose center is a one-dimensional \mathbb{Q}-isotropic torus. We identify 𝐌1,𝐌2\mathbf{M}_{1},\mathbf{M}_{2} as subgroups of 𝐌\mathbf{M} and write 𝐂\mathbf{C} for the center of 𝐌\mathbf{M} which is in fact equal to 𝐌1𝐌2\mathbf{M}_{1}\cap\mathbf{M}_{2}. The natural map ϕ:𝐌1×𝐌2𝐌\phi\mathrel{\mathop{\mathchar 58\relax}}\mathbf{M}_{1}\times\mathbf{M}_{2}\to\mathbf{M} has kernel {(x,y)𝐂×𝐂:xy=1}\{(x,y)\in\mathbf{C}\times\mathbf{C}\mathrel{\mathop{\mathchar 58\relax}}xy=1\} so that

𝐌1×𝐌2/{(x,y)𝐂×𝐂:xy=1}{g𝐌:g preserves L}.\displaystyle\mathchoice{\text{\raise 2.15277pt\hbox{$\mathbf{M}_{1}\times\mathbf{M}_{2}$}\big{/}\lower 2.15277pt\hbox{$\{(x,y)\in\mathbf{C}\times\mathbf{C}\mathrel{\mathop{\mathchar 58\relax}}xy=1\}$}}}{\mathbf{M}_{1}\times\mathbf{M}_{2}\,/\,\{(x,y)\in\mathbf{C}\times\mathbf{C}\mathrel{\mathop{\mathchar 58\relax}}xy=1\}}{\mathbf{M}_{1}\times\mathbf{M}_{2}\,/\,\{(x,y)\in\mathbf{C}\times\mathbf{C}\mathrel{\mathop{\mathchar 58\relax}}xy=1\}}{\mathbf{M}_{1}\times\mathbf{M}_{2}\,/\,\{(x,y)\in\mathbf{C}\times\mathbf{C}\mathrel{\mathop{\mathchar 58\relax}}xy=1\}}\simeq\{g\in\mathbf{M}\mathrel{\mathop{\mathchar 58\relax}}g\text{ preserves }L\}^{\circ}.

Furthermore, we have the spinor norm which is a character χ:𝐌𝐆m\chi\mathrel{\mathop{\mathchar 58\relax}}\mathbf{M}\to\mathbf{G}_{m} whose kernel is the spin group. Similarly, we have spinor norms χ1,χ2\chi_{1},\chi_{2} for 𝐌1\mathbf{M}_{1} resp. 𝐌2\mathbf{M}_{2} which are simply the restrictions of χ\chi. The above yields that

𝐇L{(g1,g2)𝐌1×𝐌2:χ(g1)χ(g2)=1}/ker(ϕ)\displaystyle\mathbf{H}_{L}\simeq\mathchoice{\text{\raise 2.15277pt\hbox{$\{(g_{1},g_{2})\in\mathbf{M}_{1}\times\mathbf{M}_{2}\mathrel{\mathop{\mathchar 58\relax}}\chi(g_{1})\chi(g_{2})=1\}$}\big{/}\lower 2.15277pt\hbox{$\ker(\phi)$}}}{\{(g_{1},g_{2})\in\mathbf{M}_{1}\times\mathbf{M}_{2}\mathrel{\mathop{\mathchar 58\relax}}\chi(g_{1})\chi(g_{2})=1\}\,/\,\ker(\phi)}{\{(g_{1},g_{2})\in\mathbf{M}_{1}\times\mathbf{M}_{2}\mathrel{\mathop{\mathchar 58\relax}}\chi(g_{1})\chi(g_{2})=1\}\,/\,\ker(\phi)}{\{(g_{1},g_{2})\in\mathbf{M}_{1}\times\mathbf{M}_{2}\mathrel{\mathop{\mathchar 58\relax}}\chi(g_{1})\chi(g_{2})=1\}\,/\,\ker(\phi)}

which is isogenous (but not isomorphic) to SpinQ|L×SpinQ|L\mathrm{Spin}_{Q|_{L}}\times\mathrm{Spin}_{Q|_{L^{\perp}}}.

The first result we prove states that the group 𝐇L\mathbf{H}_{L} totally determines the subspace LL (up to orthogonal complements). More precisely,

Proposition 2.2.

Let L1,L2VL_{1},L_{2}\leq V be non-degenerate222Recall that a non-trivial subspace WVW\subset V is non-degenerate if disc(Q|W)0\mathrm{disc}(Q|_{W})\neq 0 or equivalently if there is no non-zero vector wWw\in W so that w,w=0\langle w,w^{\prime}\rangle=0 for all wWw^{\prime}\in W. This notion is stable under extension of scalars. subspaces. If 𝐇L1=𝐇L2\mathbf{H}_{L_{1}}=\mathbf{H}_{L_{2}}, then L1=L2L_{1}=L_{2} or L1=L2L_{1}=L_{2}^{\perp}.

The proposition follows directly from the following simple lemma:

Lemma 2.3.

Let LVL\subset V be a non-degenerate subspace and let WVW\subset V be a non-trivial non-degenerate subspace invariant under 𝐇L\mathbf{H}^{\prime}_{L}. Then W{L,L,V}W\in\{L,L^{\perp},V\}.

Proof.

We first observe the following: over ¯\bar{\mathbb{Q}}, 𝐇L\mathbf{H}^{\prime}_{L} acts transitively on the set of anisotropic lines in LL and in LL^{\perp}. Indeed, by Witt’s theorem [Cassels, p. 20] the special orthogonal group in dimension at least 22 acts transitively on vectors of the same quadratic value. In any two lines one can find vectors of the same quadratic value by taking roots.

Let wWw\in W be anisotropic and write w=w1+w2w=w_{1}+w_{2} for w1Lw_{1}\in L and w2Lw_{2}\in L^{\perp}. As ww is anisotropic, one of w1w_{1} or w2w_{2} must also be anisotropic; we suppose that w1w_{1} is anisotropic without loss of generality. Let h𝐇L(¯)h\in\mathbf{H}^{\prime}_{L}(\bar{\mathbb{Q}}) be such that hw1w1hw_{1}\neq w_{1} and hw2=w2hw_{2}=w_{2}. Then

u:=hww=hw1w1LW.\displaystyle u\mathrel{\mathop{\mathchar 58\relax}}=hw-w=hw_{1}-w_{1}\in L\cap W.

We claim that we can choose hh so that uu is anisotropic. Indeed, as w1w_{1} is anisotropic its orthogonal complement in LL is non-degenerate (as LL is non-degenerate). We can thus choose hh to map w1w_{1} to a vector orthogonal to it by the above variant of Witt’s theorem. Then

Q(u)=Q(hw1)+Q(w1)=2Q(w1)0.\displaystyle Q(u)=Q(hw_{1})+Q(w_{1})=2Q(w_{1})\neq 0.

Now note that LWL\cap W is 𝐇L\mathbf{H}^{\prime}_{L}-invariant. By a further application of the above variant of Witt’s theorem and the fact that LL is spanned by anisotropic vectors (LL is non-degenerate), we obtain that LW=LL\cap W=L or equivalently LWL\subset W. Thus, we may write W=LWW=L\oplus W^{\prime} where WW^{\prime} is an orthogonal complement to LL in WW and in particular contained in LL^{\perp}. The subspace WW^{\prime} must be non-degenerate as WW and LL are and hence is trivial or contains anisotropic vectors. If WW^{\prime} is trivial, W=LW=L and we are done. Otherwise, we apply the above variant of Witt’s theorem and obtain that W=LW^{\prime}=L^{\perp} and W=VW=V. ∎

An analogous statement holds for the relationship between quadratic forms and their special stabilizer groups.

Proposition 2.4.

Let Q1,Q2Q_{1},Q_{2} be rational quadratic forms on VV. If SOQ1=SOQ2\mathrm{SO}_{Q_{1}}=\mathrm{SO}_{Q_{2}}, then Q1=rQ2Q_{1}=rQ_{2} for some rr\in\mathbb{Q}.

For a proof see [AES-higherdim, Lemma 3.3].

2.1.1. Maximality

We now aim to prove that for any non-degenerate subspace LL the connected \mathbb{Q}-groups 𝐇L\mathbf{H}^{\prime}_{L} and 𝐇L\mathbf{H}_{L} are maximal subgroups. Here, maximal is meant among connected and proper subgroups (as it was in Remark 1.13).

Proposition 2.5.

For any non-degenerate subspace LVL\subset V the groups 𝐇L\mathbf{H}^{\prime}_{L} and 𝐇L\mathbf{H}_{L} are maximal.

The result above is well-known and due to Dynkin, who classified the maximal subgroups of the classical groups in [DynkinMaximal] (see also the work of Liebeck and Seitz such as [LiebeckSeitzMax]). We will give an elementary proof.

Proof.

Note that it suffices to prove the statement for 𝐇L\mathbf{H}^{\prime}_{L}. As LL is non-degenerate, we may choose an orthogonal basis of VV consisting of an orthogonal basis of LL and an orthogonal basis of LL^{\perp}. Let

MQ=(M100M4)withM1,M4diagonal matricesM_{Q}=\begin{pmatrix}M_{1}&0\\ 0&M_{4}\end{pmatrix}\ \text{with}\ M_{1},M_{4}\ \text{diagonal matrices}

be the matrix representation of QQ in this basis. Computing the Lie algebras of SOQ\mathrm{SO}_{Q} and 𝐇L\mathbf{H}^{\prime}_{L} we obtain:

𝔤:=Lie(SOQ)={AMat(n):ATMQ+MQA=0}\mathfrak{g}\mathrel{\mathop{\mathchar 58\relax}}=\mathrm{Lie}(\mathrm{SO}_{Q})=\{A\in\mathrm{Mat}(n)\mathrel{\mathop{\mathchar 58\relax}}A^{T}M_{Q}+M_{Q}A=0\}

and

𝔥:=Lie(𝐇L)={AMat(n):A=(A100A4)andAiTMi+MiAi=0,i=1,4}.\mathfrak{h}\mathrel{\mathop{\mathchar 58\relax}}=\mathrm{Lie}(\mathbf{H}^{\prime}_{L})=\Big{\{}A\in\mathrm{Mat}(n)\mathrel{\mathop{\mathchar 58\relax}}A=\begin{pmatrix}A_{1}&0\\ 0&A_{4}\end{pmatrix}\ \text{and}\ A_{i}^{T}M_{i}+M_{i}A_{i}=0,i=1,4\Big{\}}.

We may split 𝔤\mathfrak{g} in a direct sum 𝔥𝔯\mathfrak{h}\oplus\mathfrak{r} where 𝔯\mathfrak{r} is an invariant subspace under the adjoint action of 𝐇L\mathbf{H}^{\prime}_{L} on 𝔤\mathfrak{g}. Explicitly, we may set

𝔯={(0A2A30):A2TM1+M4A3=0}.\mathfrak{r}=\Big{\{}\begin{pmatrix}0&A_{2}\\ A_{3}&0\\ \end{pmatrix}\mathrel{\mathop{\mathchar 58\relax}}A_{2}^{T}M_{1}+M_{4}A_{3}=0\Big{\}}.

We claim that the representation of 𝐇L\mathbf{H}^{\prime}_{L} on 𝔯\mathfrak{r} is irreducible. Note that we may as well show that the representation of SOQ|L×SOQ|L\mathrm{SO}_{Q|_{L}}\times\mathrm{SO}_{Q|_{L^{\perp}}} on Mat(k,nk)\mathrm{Mat}(k,n-k) given by

((σ1,σ2),A)σ1Aσ21((\sigma_{1},\sigma_{2}),A)\mapsto\sigma_{1}A\sigma_{2}^{-1}

is irreducible. Over ¯\bar{\mathbb{Q}} we may apply Lemma 2.6 below from which this follows.

Now let 𝐌\mathbf{M} be a connected group containing 𝐇L\mathbf{H}^{\prime}_{L} and let 𝔪\mathfrak{m} be its Lie algebra. Note that 𝔪𝔯\mathfrak{m}\cap\mathfrak{r} is an invariant subspace under the adjoint action of 𝐇L\mathbf{H}^{\prime}_{L} on 𝔯\mathfrak{r}. Since this representation is irreducible, 𝔪𝔯={0}\mathfrak{m}\cap\mathfrak{r}=\{0\} or 𝔪𝔯=𝔯\mathfrak{m}\cap\mathfrak{r}=\mathfrak{r}. In the former case, we have that 𝔪=𝔥\mathfrak{m}=\mathfrak{h} and in the latter 𝔪=𝔤\mathfrak{m}=\mathfrak{g}. It follows that 𝐇L\mathbf{H}^{\prime}_{L} is maximal and the proof is complete. ∎

Lemma 2.6.

For any k,m3k,m\geq 3 the action of SOk×SOm\mathrm{SO}_{k}\times\mathrm{SO}_{m} on Mat(k,m)\mathrm{Mat}(k,m) by right- resp. left-multiplication is irreducible.

Proof.

We write a very elementary proof for the sake of completeness. First, assume that k,m3k,m\geq 3. Note that the standard representation of SOk\mathrm{SO}_{k} (resp. SOm\mathrm{SO}_{m}) is irreducible as333Note that whenever k=2k=2 any isotropic vector is a fixed vector. k3k\geq 3 (resp. m3m\geq 3). It follows that the representation of SOk×SOm\mathrm{SO}_{k}\times\mathrm{SO}_{m} on the tensor product of the respective standard representations is also irreducible (see, for instance, [TensorProduct_Irreducible, Theorem 3.10.2]); the latter is isomorphic to the representation in the lemma. ∎

2.2. The isotropy condition

We establish here congruence conditions which imply isotropy of the stabilizer groups 𝐇L\mathbf{H}_{L}. Recall that a p\mathbb{Q}_{p}-group 𝐆\mathbf{G} is strongly isotropic if for every connected non-trivial normal subgroup 𝐍<𝐆\mathbf{N}<\mathbf{G} defined over p\mathbb{Q}_{p}, the group 𝐍(p)\mathbf{N}(\mathbb{Q}_{p}) is not compact. We say that a \mathbb{Q}-group 𝐆\mathbf{G} is strongly isotropic at a prime pp if 𝐆\mathbf{G} is strongly isotropic as a p\mathbb{Q}_{p}-group.

Proposition 2.7.

Let (V,Q)(V^{\prime},Q^{\prime}) be any non-degenerate quadratic space over p\mathbb{Q}_{p}. Then QQ^{\prime} is isotropic if and only if SpinQ\mathrm{Spin}_{Q^{\prime}} is strongly isotropic.

Proof.

If QQ^{\prime} is isotropic, VV^{\prime} contains an hyperbolic plane HH (see [Cassels, Chapter 2. Lemma 2.1]). Then SpinQ\mathrm{Spin}_{Q^{\prime}} contains SpinQ|H\mathrm{Spin}_{Q^{\prime}|_{H}} which is a split torus. Hence, SpinQ\mathrm{Spin}_{Q^{\prime}} is isotropic. Conversely, if QQ^{\prime} is anisotropic then SpinQ(p)\mathrm{Spin}_{Q^{\prime}}(\mathbb{Q}_{p}) is compact as the hypersurface Q(x)=1Q^{\prime}(x)=1 is compact. This proves that QQ^{\prime} is isotropic if and only if SpinQ\mathrm{Spin}_{Q^{\prime}} is isotropic. This is sufficient to prove the proposition if dim(V)=2\dim(V^{\prime})=2 (as the torus SpinQ\mathrm{Spin}_{Q^{\prime}} is one-dimensional) and if dim(V)>2\dim(V^{\prime})>2 is not equal to 44 as SpinQ\mathrm{Spin}_{Q^{\prime}} is absolutely almost simple in these cases.

Suppose that dim(V)=4\dim(V^{\prime})=4. We freely use facts about Clifford algebras and spin groups from [knus] (mostly Chapter 99 therein). Recall that SpinQ\mathrm{Spin}_{Q^{\prime}} is equal to the norm one elements of the even Clifford algebra 𝒞0\mathcal{C}^{0} of QQ^{\prime}. If the center 𝒵\mathcal{Z} of 𝒞0\mathcal{C}^{0} is a field over p\mathbb{Q}_{p}, 𝒞0\mathcal{C}^{0} is a quaternion algebra over 𝒵\mathcal{Z} and SpinQ\mathrm{Spin}_{Q^{\prime}} is simple. In this case, the proof works as in the case of dim(V)4\dim(V^{\prime})\neq 4.

So suppose that the center is split which is equivalent to disc(Q)\mathrm{disc}(Q^{\prime}) being a square in p\mathbb{Q}_{p}. Thus, there is a quaternion algebra \mathcal{B} over p\mathbb{Q}_{p} such that (V,Q)(V^{\prime},Q^{\prime}) is similar to (,Nr)(\mathcal{B},\mathrm{Nr}) where Nr\mathrm{Nr} is the norm on \mathcal{B}. Then SpinQSL1()×SL1()\mathrm{Spin}_{Q^{\prime}}\simeq\mathrm{SL}_{1}(\mathcal{B})\times\mathrm{SL}_{1}(\mathcal{B}) which is a product of two p\mathbb{Q}_{p}-simple groups. Note that \mathcal{B} or SL1()\mathrm{SL}_{1}(\mathcal{B}) are isotropic if and only if QQ^{\prime} is isotropic. This concludes the proof of the proposition. ∎

Via (2.1) we obtain the following.

Corollary 2.8.

Let LGrn,k()L\in\mathrm{Gr}_{n,k}(\mathbb{Q}) and pp be an odd prime. Then, 𝐇L\mathbf{H}_{L} is strongly isotropic at pp if and only if the quadratic spaces (L,Q|L)(L,Q|_{L}) and (L,Q|L)(L^{\perp},Q|_{L^{\perp}}) are isotropic over p\mathbb{Q}_{p}.

Using standard arguments (as in [AES-higherdim, Lemma 3.7] for example) we may deduce the following explicit characterization of isotropy.

Proposition 2.9.

Let LGrn,k()L\in\mathrm{Gr}_{n,k}(\mathbb{Q}) be a rational subspace and let pp be an odd prime. Then, 𝐇L\mathbf{H}_{L} is strongly isotropic at pp if any of the following conditions hold:

  • k5k\geq 5 and nk5n-k\geq 5.

  • 3k<53\leq k<5, nk5n-k\geq 5, and pdiscQ(L)p\nmid\mathrm{disc}_{Q}(L).

  • k5k\geq 5, 3nk<53\leq n-k<5, and pdiscQ(L)p\nmid\mathrm{disc}_{Q}(L^{\perp}).

  • 3k<53\leq k<5, 3nk<53\leq n-k<5, pdiscQ(L)p\nmid\mathrm{disc}_{Q}(L), and pdiscQ(L)p\nmid\mathrm{disc}_{Q}(L^{\perp}).

  • k=2k=2, nk5n-k\geq 5, and discQ(L)(𝔽p×)2-\mathrm{disc}_{Q}(L)\in(\mathbb{F}_{p}^{\times})^{2} (i.e. discQ(L)-\mathrm{disc}_{Q}(L) is a non-zero square modulo pp).

  • k=2k=2, 3nk<53\leq n-k<5, pdiscQ(L)p\nmid\mathrm{disc}_{Q}(L^{\perp}), and discQ(L)(𝔽p×)2-\mathrm{disc}_{Q}(L)\in(\mathbb{F}_{p}^{\times})^{2}.

  • k5k\geq 5, nk=2n-k=2, and discQ(L)(𝔽p×)2-\mathrm{disc}_{Q}(L^{\perp})\in(\mathbb{F}_{p}^{\times})^{2}.

  • 3k<53\leq k<5, nk=2n-k=2, pdiscQ(L)p\nmid\mathrm{disc}_{Q}(L), and discQ(L)(𝔽p×)2-\mathrm{disc}_{Q}(L^{\perp})\in(\mathbb{F}_{p}^{\times})^{2}.

While the list is lengthy, let us note that half of it consists in interchanging the roles of kk and nkn-k as well as LL and LL^{\perp}. Also, whenever pdisc(Q)p\nmid\mathrm{disc}(Q) the conditions pdiscQ(L)p\nmid\mathrm{disc}_{Q}(L) and pdiscQ(L)p\nmid\mathrm{disc}_{Q}(L^{\perp}) are equivalent (see Proposition 5.4 and its corollary). When k=4k=4 or nk=4n-k=4 the above criteria are sufficient but not necessary. For example, the form x12+x22+x32+px42x_{1}^{2}+x_{2}^{2}+x_{3}^{2}+px_{4}^{2} is isotropic though its discriminant is divisible by pp.

2.3. Diagonal embeddings of stabilizer groups

In this section, we define a diagonally embedded copy 𝚫𝐇L<SpinQ×𝐏n,k\mathbf{\Delta H}_{L}<\mathrm{Spin}_{Q}\times\mathbf{P}_{n,k} of the stabilizer group of any subspace LGrn,k()L\in\mathrm{Gr}_{n,k}(\mathbb{Q}).

With the arithmetic application in Part 22 in mind, we must allow for any rational subspace a choice of a full rank \mathbb{Z}-lattice ΛLn\Lambda_{L}\subset\mathbb{Q}^{n} with

nΛL(n)#:={vn:v,w for all wn}.\displaystyle\mathbb{Z}^{n}\subset\Lambda_{L}\subset(\mathbb{Z}^{n})^{\#}\mathrel{\mathop{\mathchar 58\relax}}=\{v\in\mathbb{Q}^{n}\mathrel{\mathop{\mathchar 58\relax}}\langle v,w\rangle\in\mathbb{Z}\text{ for all }w\in\mathbb{Z}^{n}\}.

If QQ is unimodular (i.e. disc(Q)=1\mathrm{disc}(Q)=1) then ΛL=n=(n)#\Lambda_{L}=\mathbb{Z}^{n}=(\mathbb{Z}^{n})^{\#}. We emphasize that for the arguments in the current Part 11, this choice of intermediate lattice ΛL\Lambda_{L} is inconsequential and the reader may safely assume ΛL=n\Lambda_{L}=\mathbb{Z}^{n} at first.

Let gLGLn()g_{L}\in\mathrm{GL}_{n}(\mathbb{Q}) be such that gLn=ΛLg_{L}\mathbb{Z}^{n}=\Lambda_{L}, its first kk columns are a basis of LΛLL\cap\Lambda_{L} and such that det(gL)>0\det(g_{L})>0. In words, the columns of gLg_{L} complement a basis of LΛLL\cap\Lambda_{L} into an oriented basis of ΛL\Lambda_{L}. We then have a well-defined morphism with finite kernel

(2.2) ΨL:𝐇L𝐏n,k,hgL1ρQ(h)gL.\displaystyle\Psi_{L}\mathrel{\mathop{\mathchar 58\relax}}\mathbf{H}_{L}\to\mathbf{P}_{n,k},\ h\mapsto g_{L}^{-1}\rho_{Q}(h)g_{L}.

Note that the morphism depends on the choice of ΛL\Lambda_{L}, but we omit this dependency here to simplify notation. It also depends on the choice of basis; a change of basis conjugates ΨL\Psi_{L} by an element of 𝐏n,k()\mathbf{P}_{n,k}(\mathbb{Z}).

One can restrict the action of an element of 𝐇L\mathbf{H}_{L} to LL and represent the so-obtained special orthogonal transformation in the basis contained in gLg_{L}. This yields an epimorphism (as in (2.1))

ψ1,L:𝐇LSOqLΛL.\displaystyle\psi_{1,L}\mathrel{\mathop{\mathchar 58\relax}}\mathbf{H}_{L}\to\mathrm{SO}_{q_{{}_{L\cap\Lambda_{L}}}}.

Explicitly, the epimorphism is given by

ψ1,L:h𝐇Lπ1(gL1ρQ(h)gL)=π1ΨL(h)SOqLΛL.\displaystyle\psi_{1,L}\mathrel{\mathop{\mathchar 58\relax}}h\in\mathbf{H}_{L}\mapsto\pi_{1}(g_{L}^{-1}\rho_{Q}(h)g_{L})=\pi_{1}\circ\Psi_{L}(h)\in\mathrm{SO}_{q_{{}_{L\cap\Lambda_{L}}}}.

Similarly to the above, one can obtain an epimorphism 𝐇LSOQ|L\mathbf{H}_{L}\to\mathrm{SO}_{Q|_{L^{\perp}}}. To explicit this, we would like to specify how to obtain a basis of LΛL#L^{\perp}\cap\Lambda_{L}^{\#} from gLg_{L}. For this, observe first that the basis dual to the columns of gLg_{L} is given by the columns of MQ1(gL1)tM_{Q}^{-1}(g_{L}^{-1})^{t}. Note that the last nkn-k columns of MQ1(gL1)tM_{Q}^{-1}(g_{L}^{-1})^{t} are orthogonal to LL so they form a basis of ΛL#L\Lambda_{L}^{\#}\cap L^{\perp}. We hence obtain an epimorphism

ψ2,L:h𝐇Lπ2(gLtMQρQ(h)MQ1(gL1)t)SOqLΛL#.\displaystyle\psi_{2,L}\mathrel{\mathop{\mathchar 58\relax}}h\in\mathbf{H}_{L}\mapsto\pi_{2}(g_{L}^{t}M_{Q}\rho_{Q}(h)M_{Q}^{-1}(g_{L}^{-1})^{t})\in\mathrm{SO}_{q_{L^{\perp}\cap\Lambda_{L}^{\#}}}.

Note that

gLtMQρQ(h)MQ1(gL1)t=gLtρQ(h1)t(gL1)t=(gL1ρQ(h1)gL)t\displaystyle g_{L}^{t}M_{Q}\rho_{Q}(h)M_{Q}^{-1}(g_{L}^{-1})^{t}=g_{L}^{t}\rho_{Q}(h^{-1})^{t}(g_{L}^{-1})^{t}=(g_{L}^{-1}\rho_{Q}(h^{-1})g_{L})^{t}

which shows that

ψ2,L(h)=π2((gL1ρQ(h1)gL)t)=π2(gL1ρQ(h1)gL)t=π2(ΨL(h1))t.\displaystyle\psi_{2,L}(h)=\pi_{2}((g_{L}^{-1}\rho_{Q}(h^{-1})g_{L})^{t})=\pi_{2}(g_{L}^{-1}\rho_{Q}(h^{-1})g_{L})^{t}=\pi_{2}(\Psi_{L}(h^{-1}))^{t}.

We define the group

(2.3) 𝚫𝐇L={(h,ΨL(h)):h𝐇L}SpinQ×𝐏n,k=𝐆.\displaystyle\mathbf{\Delta H}_{L}=\{(h,\Psi_{L}(h))\mathrel{\mathop{\mathchar 58\relax}}h\in\mathbf{H}_{L}\}\subset\mathrm{Spin}_{Q}\times\mathbf{P}_{n,k}=\mathbf{G}.

By the definitions above, the morphism

𝐆𝐆¯,(g1,g2)(g1,π1(g2),π2(g21)t)\displaystyle\mathbf{G}\to\bar{\mathbf{G}},\ (g_{1},g_{2})\mapsto(g_{1},\pi_{1}(g_{2}),\pi_{2}(g_{2}^{-1})^{t})

induces a morphism

𝚫𝐇L{(h,ψ1,L(h),ψ2,L(h)):h𝐇L}=:𝚫𝐇¯L𝐆¯\displaystyle\mathbf{\Delta H}_{L}\to\{(h,\psi_{1,L}(h),\psi_{2,L}(h))\mathrel{\mathop{\mathchar 58\relax}}h\in\mathbf{H}_{L}\}=\mathrel{\mathop{\mathchar 58\relax}}\mathbf{\Delta\bar{H}}_{L}\subset\bar{\mathbf{G}}

which is in fact an isogeny.

3. The dynamical version of the theorem in codimension at least 3

As mentioned in the introduction, our aim is to translate the main theorems into a statement concerning weak limits of orbit measures on an adequate adelic homogeneous space. In this and the next section we shall establish these equidistribution theorems for orbit measures. This section treats the case k,nk3k,n-k\geq 3.

In the following we call a sequence of subspaces LiGrn,k()L_{i}\in\mathrm{Gr}_{n,k}(\mathbb{Q}) admissible if

  1. (1)

    discQ(Li)\mathrm{disc}_{Q}(L_{i})\to\infty as ii\to\infty,

  2. (2)

    disc(q~Li)\mathrm{disc}(\tilde{q}_{L_{i}})\to\infty as ii\to\infty,

  3. (3)

    disc(q~Li)\mathrm{disc}(\tilde{q}_{L_{i}^{\perp}})\to\infty as ii\to\infty, and

  4. (4)

    there exists a prime pp such that 𝐇Li(p)\mathbf{H}_{L_{i}}(\mathbb{Q}_{p}) is strongly isotropic for all ii.

This section establishes the following theorem. Conjecturally, an analogous version should hold when k=2k=2 or nk=2n-k=2 (see Remark 1.12).

Theorem 3.1.

Let LiGrn,k()L_{i}\in\mathrm{Gr}_{n,k}(\mathbb{Q}) be an admissible sequence of rational subspaces (with a choice of lattice ΛLi\Lambda_{L_{i}} as in §2.3), let gi𝐆()g_{i}\in\mathbf{G}(\mathbb{R}) and let μi\mu_{i} be the Haar probability measure on the closed orbit

gi𝚫𝐇Li(𝔸)𝐆()𝐆(𝔸)/𝐆().\displaystyle g_{i}\mathbf{\Delta H}_{L_{i}}(\mathbb{A})\mathbf{G}(\mathbb{Q})\subset\mathchoice{\text{\raise 2.15277pt\hbox{$\mathbf{G}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathbf{G}(\mathbb{Q})$}}}{\mathbf{G}(\mathbb{A})\,/\,\mathbf{G}(\mathbb{Q})}{\mathbf{G}(\mathbb{A})\,/\,\mathbf{G}(\mathbb{Q})}{\mathbf{G}(\mathbb{A})\,/\,\mathbf{G}(\mathbb{Q})}.

Then μi\mu_{i} converges to the Haar probability measure on 𝐆(𝔸)/𝐆()\mathchoice{\text{\raise 2.15277pt\hbox{$\mathbf{G}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathbf{G}(\mathbb{Q})$}}}{\mathbf{G}(\mathbb{A})\,/\,\mathbf{G}(\mathbb{Q})}{\mathbf{G}(\mathbb{A})\,/\,\mathbf{G}(\mathbb{Q})}{\mathbf{G}(\mathbb{A})\,/\,\mathbf{G}(\mathbb{Q})} as ii\to\infty.

The rest of the section is devoted to proving Theorem 3.1. We remark that the notion of admissible sequences here is an ad hoc notion which appeared in other instances (see e.g. [2in4]) to assert a similar goal. The assumptions (1)–(3) in the definition of admissibility are in fact necessary for the above theorem to hold while (4) can conjecturally be removed.

3.1. A general result on equidistribution of packets

The crucial input to our results is an SS-arithmetic extension of a theorem of Mozes and Shah [mozesshah] by Gorodnik and Oh [gorodnikoh]. We state a version of it here for the reader’s convenience.

Let 𝖦\mathsf{G} be a simply-connected connected semisimple algebraic group defined over \mathbb{Q} and Y𝔸=𝖦(𝔸)/𝖦()Y_{\mathbb{A}}=\mathsf{G}(\mathbb{A})/\mathsf{G}(\mathbb{Q}). Let WW be a compact open subgroup of 𝖦(𝔸f)\mathsf{G}(\mathbb{A}_{f}). We denote by Cc(Y𝔸,W)C_{c}(Y_{\mathbb{A}},W) the set of all continuous compactly supported functions on Y𝔸Y_{\mathbb{A}} which are WW invariant. Consider a sequence (𝖧i)i(\mathsf{H}_{i})_{i\in\mathbb{N}} of connected semisimple subgroups of 𝖦\mathsf{G} and let μi\mu_{i} denote the Haar probability measure on the orbit 𝖧i(𝔸)+𝖦()Y𝔸\mathsf{H}_{i}(\mathbb{A})^{+}\mathsf{G}(\mathbb{Q})\subset Y_{\mathbb{A}} where 𝖧i(𝔸)+\mathsf{H}_{i}(\mathbb{A})^{+} is the image of the adelic points of the simply connected cover of 𝖧i\mathsf{H}_{i} in 𝖧i(𝔸)\mathsf{H}_{i}(\mathbb{A}). For given gi𝖦(𝔸)g_{i}\in\mathsf{G}(\mathbb{A}) we are interested in the weak* limits of the sequence of measures giμig_{i}\mu_{i}.

Theorem 3.2 (Gorodnik-Oh [gorodnikoh, Theorem 1.7]).

Assume that there exists a prime pp such that 𝖧i\mathsf{H}_{i} is strongly isotropic at pp for all ii\in\mathbb{N}. Then, for any weak limit of the sequence (giμi)(g_{i}\mu_{i}) with μ(Y𝔸)=1\mu(Y_{\mathbb{A}})=1, there exists a connected \mathbb{Q}-group 𝖬<𝖦\mathsf{M}<\mathsf{G} such that the following hold:

  1. (1)

    For all ii large enough, there exist δi𝖦()\delta_{i}\in\mathsf{G}(\mathbb{Q}) such that:

    δi1𝖧iδi𝖬.\delta_{i}^{-1}\mathsf{H}_{i}\delta_{i}\subset\mathsf{M}.
  2. (2)

    For any compact open subgroup WW of 𝖦(𝔸f)\mathsf{G}(\mathbb{A}_{f}) there exists a finite index normal subgroup M0=M0(W)M_{0}=M_{0}(W) of 𝖬(𝔸)\mathsf{M}(\mathbb{A}) and g𝖦(𝔸)g\in\mathsf{G}(\mathbb{A}) such that μ\mu agrees with the Haar probability measure on gM0𝖦()gM_{0}\mathsf{G}(\mathbb{Q}) when restricted to Cc(Y𝔸,W)C_{c}(Y_{\mathbb{A}},W). Moreover, there exists hi𝖧i(𝔸)+h_{i}\in\mathsf{H}_{i}(\mathbb{A})^{+} such that gihiδigg_{i}h_{i}\delta_{i}\rightarrow g as ii\rightarrow\infty.

  3. (3)

    If the centralizers of 𝖧i\mathsf{H}_{i} are \mathbb{Q}-anisotropic for all ii\in\mathbb{N}, then 𝖬\mathsf{M} is semisimple. Moreover, for any compact open subgroup WW, M0=M0(W)M_{0}=M_{0}(W) in 2 contains 𝖬(𝔸)+𝖬()\mathsf{M}(\mathbb{A})^{+}\mathsf{M}(\mathbb{Q}).

We remark that the theorem as stated in [gorodnikoh] does not assume that 𝖦\mathsf{G} is simply connected; we will however only need this case.

3.2. Proof of Theorem 3.1

We shall prove Theorem 3.1 in several steps and start with a short overview. Note that we have a morphism

𝐆𝐆¯=SpinQ×SLk×SLnk\displaystyle\mathbf{G}\to\bar{\mathbf{G}}=\mathrm{Spin}_{Q}\times\mathrm{SL}_{k}\times\mathrm{SL}_{n-k}

given by mapping g𝐏n,kg\in\mathbf{P}_{n,k} to (π1(g),π2(g1)t)(\pi_{1}(g),\pi_{2}(g^{-1})^{t}) and SpinQ\mathrm{Spin}_{Q} to itself via the identity map (see also §2.3). The first step of the theorem establishes equidistribution of the projections to the respective homogeneous quotients for SpinQ,SLk,SLnk\mathrm{Spin}_{Q},\mathrm{SL}_{k},\mathrm{SL}_{n-k} (henceforth called ’individual equidistribution’). The second step is the analogous statement for 𝐆¯\bar{\mathbf{G}}. Note that the admissibility assumption on the sequence of subspaces LiL_{i} is used for individual equidistribution and in fact, the different conditions (1)–(3) imply the corresponding individual equidistribution statements (i.e. (1) implies equidistribution in the homogeneous quotient SpinQ(𝔸)/SpinQ()\mathchoice{\text{\raise 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Q})$}}}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})} etc.).

To vaguely outline the argument here, consider a sequence of orbits

gi𝐇Li(𝔸)SpinQ()SpinQ(𝔸)/SpinQ().\displaystyle g_{i}^{\prime}\mathbf{H}_{L_{i}}(\mathbb{A})\mathrm{Spin}_{Q}(\mathbb{Q})\subset\mathchoice{\text{\raise 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Q})$}}}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}.

As the groups 𝐇Li\mathbf{H}_{L_{i}} are maximal subgroups, the theorem of Gorodnik and Oh above implies that either the orbits are equidistributed or that there exist lattice elements δi\delta_{i} so that δi𝐇Liδi1\delta_{i}\mathbf{H}_{L_{i}}\delta_{i}^{-1} is eventually independent of ii. In the latter case, we also know that the lattice elements are up to a bounded amount in the stabilizer group; this will be shown to contradict the assumption that discQ(Li)\mathrm{disc}_{Q}(L_{i})\to\infty.

3.2.1. Applying Theorem 3.2

Consider the subgroup 𝐉=SpinQ×SLn\mathbf{J}=\mathrm{Spin}_{Q}\times\mathrm{SL}_{n}. Note that 𝐉\mathbf{J} is semisimple and simply connected so that we may apply Theorem 3.2 given a suitable sequence of subgroups.

The groups 𝐇Li\mathbf{H}_{L_{i}} are potentially not simply connected so that a little more care is needed in applying Theorem 3.2 to the orbit measures μi\mu_{i}. We fix for any ii some hi𝚫𝐇Li(𝔸)h_{i}\in\mathbf{\Delta H}_{L_{i}}(\mathbb{A}) and consider the orbit measures on gihi𝚫𝐇Li(𝔸)+𝐆()g_{i}h_{i}\mathbf{\Delta H}_{L_{i}}(\mathbb{A})^{+}\mathbf{G}(\mathbb{Q}). In view of the theorem, it suffices to show that these converge to the Haar probability measure on 𝐆(𝔸)/𝐆()\mathchoice{\text{\raise 2.15277pt\hbox{$\mathbf{G}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathbf{G}(\mathbb{Q})$}}}{\mathbf{G}(\mathbb{A})\,/\,\mathbf{G}(\mathbb{Q})}{\mathbf{G}(\mathbb{A})\,/\,\mathbf{G}(\mathbb{Q})}{\mathbf{G}(\mathbb{A})\,/\,\mathbf{G}(\mathbb{Q})}. Indeed, by disintegration the Haar measure on gi𝚫𝐇Li(𝔸)𝐆()g_{i}\mathbf{\Delta H}_{L_{i}}(\mathbb{A})\mathbf{G}(\mathbb{Q}) is the integral over the Haar measures on gihi𝚫𝐇Li(𝔸)+𝐆()g_{i}h_{i}\mathbf{\Delta H}_{L_{i}}(\mathbb{A})^{+}\mathbf{G}(\mathbb{Q}) when hih_{i} is integrated with respect to the Haar probability measure on the compact group 𝚫𝐇Li(𝔸)/𝚫𝐇Li(𝔸)+\mathbf{\Delta H}_{L_{i}}(\mathbb{A})/\mathbf{\Delta H}_{L_{i}}(\mathbb{A})^{+}. In other words, the Haar measure on gi𝚫𝐇Li(𝔸)𝐆()g_{i}\mathbf{\Delta H}_{L_{i}}(\mathbb{A})\mathbf{G}(\mathbb{Q}) is a convex combination of the Haar measures on the orbits gihi𝚫𝐇Li(𝔸)+𝐆()g_{i}h_{i}\mathbf{\Delta H}_{L_{i}}(\mathbb{A})^{+}\mathbf{G}(\mathbb{Q}). To simplify notation, we replace gig_{i} by gihig_{i}h_{i} in order to omit hih_{i}. Furthermore, we abuse notation and write μi\mu_{i} for these ”components” of the original orbit measures.

We fix a compact open subgroup WW of 𝐆(𝔸f)\mathbf{G}(\mathbb{A}_{f}) in view of (2) b) in Theorem 3.2 and an odd prime pp as in the definition of admissibility of the sequence (Li)i(L_{i})_{i}.

Let μ\mu be any weak-limit of the measures μi\mu_{i}. Note that μ\mu is a probability measure. Indeed, the pushforward of the measures μi\mu_{i} to SpinQ(𝔸)/SpinQ()\mathrm{Spin}_{Q}(\mathbb{A})/\mathrm{Spin}_{Q}(\mathbb{Q}) has to converge to a probability measure as SpinQ(𝔸)/SpinQ()\mathrm{Spin}_{Q}(\mathbb{A})/\mathrm{Spin}_{Q}(\mathbb{Q}) is compact. We let 𝐌<𝐉\mathbf{M}<\mathbf{J} be as in Theorem 3.2. As gi𝐆(𝔸)g_{i}\in\mathbf{G}(\mathbb{A}) and 𝚫𝐇Li<𝐆\mathbf{\Delta H}_{L_{i}}<\mathbf{G} for all ii, the support of the measures μi\mu_{i} is contained in 𝐆(𝔸)𝐉()𝐆(𝔸)/𝐆()\mathbf{G}(\mathbb{A})\mathbf{J}(\mathbb{Q})\simeq\mathchoice{\text{\raise 2.15277pt\hbox{$\mathbf{G}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathbf{G}(\mathbb{Q})$}}}{\mathbf{G}(\mathbb{A})\,/\,\mathbf{G}(\mathbb{Q})}{\mathbf{G}(\mathbb{A})\,/\,\mathbf{G}(\mathbb{Q})}{\mathbf{G}(\mathbb{A})\,/\,\mathbf{G}(\mathbb{Q})}. Thus 𝐌<𝐆\mathbf{M}<\mathbf{G}.

Claim.

It suffices to show that 𝐌=𝐆\mathbf{M}=\mathbf{G}.

Proof of the claim.

Suppose that 𝐌=𝐆\mathbf{M}=\mathbf{G}. Let M0=M0(W)M_{0}=M_{0}(W) be as in Theorem 3.2. Since 𝐆(𝔸)\mathbf{G}(\mathbb{A}) has no proper finite-index subgroups [BorelHomAbstr, Theorem 6.7], we have M0=𝐆(𝔸)M_{0}=\mathbf{G}(\mathbb{A}) (independently of WW). Therefore, for any WW-invariant continuous compactly-supported function ff, the integral μ(f)\mu(f) agrees with the integral against the Haar measure on 𝐆(𝔸)/𝐆()\mathbf{G}(\mathbb{A})/\mathbf{G}(\mathbb{Q}). But any continuous compactly-supported function is invariant under some compact open subgroup WW, hence the claim follows. ∎

We now focus on proving that 𝐌=𝐆\mathbf{M}=\mathbf{G}. By Theorem 3.2 there exist δi𝐆()\delta_{i}\in\mathbf{G}(\mathbb{Q}) such that δi1𝚫𝐇Liδi<𝐌\delta_{i}^{-1}\mathbf{\Delta H}_{L_{i}}\delta_{i}<\mathbf{M} for all ii0i\geq i_{0}. Furthermore, we fix g𝐆(𝔸)g\in\mathbf{G}(\mathbb{A}) as well as h^i=(hi,ΨLi(hi))𝚫𝐇Li(𝔸)+\hat{h}_{i}=(h_{i},\Psi_{L_{i}}(h_{i}))\in\mathbf{\Delta H}_{L_{i}}(\mathbb{A})^{+} as in Theorem 3.2 such that

(3.1) gih^iδig.\displaystyle g_{i}\hat{h}_{i}\delta_{i}\to g.

3.2.2. Individual equidistribution of subspaces and shapes

Consider the morphism

(3.2) 𝐆𝐆¯=SpinQ×SLk×SLnk.\displaystyle\mathbf{G}\to\bar{\mathbf{G}}=\mathrm{Spin}_{Q}\times\mathrm{SL}_{k}\times\mathrm{SL}_{n-k}.

In the following step of the proof, we show that the image 𝐌¯\bar{\mathbf{M}} of the subgroup 𝐌\mathbf{M} via (3.2) projects surjectively onto each of the factors of 𝐆¯\bar{\mathbf{G}}.

Proposition 3.3.

The morphism obtained by restricting the projection of 𝐆¯\bar{\mathbf{G}} onto any almost simple factor of 𝐆¯\bar{\mathbf{G}} to 𝐌\mathbf{M} is surjective.

Proof.

We prove the proposition for each factor separately. To ease notation, π\pi will denote the projection of 𝐆¯\bar{\mathbf{G}} onto the factor in consideration, which we extend to 𝐆\mathbf{G} by precomposition.

First factor: As π(𝚫𝐇Li)=𝐇Li\pi(\mathbf{\Delta H}_{L_{i}})=\mathbf{H}_{L_{i}} we have for each ii

π(δi)1𝐇Liπ(δi)<π(𝐌).\displaystyle\pi(\delta_{i})^{-1}\mathbf{H}_{L_{i}}\pi(\delta_{i})<\pi(\mathbf{M}).

Since 𝐇Li\mathbf{H}_{L_{i}} is a maximal subgroup of SpinQ\mathrm{Spin}_{Q} (see Proposition 2.5), there are two options: either π1(𝐌)=SpinQ\pi_{1}(\mathbf{M})=\mathrm{Spin}_{Q} or π(δi)1𝐇Liπ(δi)=π(𝐌)\pi(\delta_{i})^{-1}\mathbf{H}_{L_{i}}\pi(\delta_{i})=\pi(\mathbf{M}) for all ii0i\geq i_{0}.

Suppose the second option holds (as we are done otherwise). Setting γi=π(δiδi01)\gamma_{i}=\pi(\delta_{i}\delta_{i_{0}}^{-1}) and L=Li0L=L_{i_{0}} we have

𝐇γi.L=γi𝐇Lγi1=𝐇Li.\displaystyle\mathbf{H}_{\gamma_{i}.L}=\gamma_{i}\mathbf{H}_{L}\gamma_{i}^{-1}=\mathbf{H}_{L_{i}}.

By Proposition 2.2 we have γi.L=Li\gamma_{i}.L=L_{i} or γi.L=Li\gamma_{i}.L^{\perp}=L_{i}; changing to a subsequence and increasing i0i_{0} we may suppose that the former option holds for all ii0i\geq i_{0}. By (3.1) there exist hi𝐇Li(𝔸)h_{i}\in\mathbf{H}_{L_{i}}(\mathbb{A}) such that π(gi)hiγiπ(g)\pi(g_{i})h_{i}\gamma_{i}\to\pi(g^{\prime}) for some g𝐆(𝔸)g^{\prime}\in\mathbf{G}(\mathbb{A}). Roughly speaking, this implies that Li=hiγi.Lπ(g).LL_{i}=h_{i}\gamma_{i}.L\to\pi(g).L as p\mathbb{Q}_{p}-subspaces for any prime pp contradicting the discriminant condition. More precisely, let εie\varepsilon_{i}\to e be such that π(gi)hiγi=εiπ(g)\pi(g_{i})h_{i}\gamma_{i}=\varepsilon_{i}\pi(g^{\prime}). Then for any prime pp the local discriminant gives

discp,Q(Li)=discp,Q(hi,pγi.L)=discp,Q(εi,pπ(gp).L)\displaystyle\mathrm{disc}_{p,Q}(L_{i})=\mathrm{disc}_{p,Q}(h_{i,p}\gamma_{i}.L)=\mathrm{disc}_{p,Q}(\varepsilon_{i,p}\pi(g_{p}^{\prime}).L)

If ii is large enough such that εiSpinQ(×^)\varepsilon_{i}\in\mathrm{Spin}_{Q}(\mathbb{R}\times\widehat{\mathbb{Z}}), we have

discQ(Li)=ppνp(discp,Q(Li))=ppνp(discp,Q(π(gp).L))\displaystyle\mathrm{disc}_{Q}(L_{i})=\prod_{p}p^{\nu_{p}(\mathrm{disc}_{p,Q}(L_{i}))}=\prod_{p}p^{\nu_{p}(\mathrm{disc}_{p,Q}(\pi(g_{p}^{\prime}).L))}

which is constant, contradicting discQ(Li)\mathrm{disc}_{Q}(L_{i})\to\infty.

Second factor: The proof is very similar to the first case, so we will be brief. By maximality of special orthogonal groups (Remark 1.13) and as π(𝚫𝐇Li)=SOqLiΛLi\pi(\mathbf{\Delta H}_{L_{i}})=\mathrm{SO}_{q_{{}_{L_{i}\cap\Lambda_{L_{i}}}}} we may suppose by contradiction that for all ii0i\geq i_{0}

π(δi)1SOqLiΛLiπ(δi)=π(𝐌).\displaystyle\pi(\delta_{i})^{-1}\mathrm{SO}_{q_{{}_{L_{i}\cap\Lambda_{L_{i}}}}}\pi(\delta_{i})=\pi(\mathbf{M}).

Let us simplify notation and write qiq_{i} for the least integer multiple of qLiΛLiq_{{}_{L_{i}\cap\Lambda_{L_{i}}}} that has integer coefficients. Since LiΛLiL_{i}\cap\Lambda_{L_{i}} and Li()L_{i}(\mathbb{Z}) are commensurable with indices controlled by disc(Q)\mathrm{disc}(Q), we have disc(qi)disc(qLi)\mathrm{disc}(q_{i})\asymp\mathrm{disc}(q_{L_{i}}) and disc(q~i)disc(q~Li)\mathrm{disc}(\tilde{q}_{i})\asymp\mathrm{disc}(\tilde{q}_{L_{i}}). In particular, by our assumption disc(q~i)\mathrm{disc}(\tilde{q}_{i})\to\infty as ii\to\infty.

Set γi=π(δiδi01)SLk()\gamma_{i}=\pi(\delta_{i}\delta_{i_{0}}^{-1})\in\mathrm{SL}_{k}(\mathbb{Q}) so that

(3.3) SOγiq~i0=SOγiqi0=γiSOqLi0γi1=SOqi=SOq~i.\displaystyle\mathrm{SO}_{\gamma_{i}\tilde{q}_{i_{0}}}=\mathrm{SO}_{\gamma_{i}q_{i_{0}}}=\gamma_{i}\mathrm{SO}_{q_{{}_{L_{i_{0}}}}}\gamma_{i}^{-1}=\mathrm{SO}_{q_{i}}=\mathrm{SO}_{\tilde{q}_{i}}.

By Proposition 2.4 there exist coprime integers mi,nim_{i},n_{i} such that

miγiq~i0=niq~i.\displaystyle m_{i}\gamma_{i}\tilde{q}_{i_{0}}=n_{i}\tilde{q}_{i}.

Using (3.1) write π(gi)hiγi=εiπ(g)\pi(g_{i})h_{i}\gamma_{i}=\varepsilon_{i}\pi(g^{\prime}) for some g𝐆(𝔸)g^{\prime}\in\mathbf{G}(\mathbb{A}) and εie\varepsilon_{i}\to e. By (3.3), hi(γiq~i0)=γiq~i0h_{i}(\gamma_{i}\tilde{q}_{i_{0}})=\gamma_{i}\tilde{q}_{i_{0}}. Thus, we have for any prime pp

miεi,pπ(gp)q~i0=mihi,pγiq~i0=niq~i.\displaystyle m_{i}\varepsilon_{i,p}\pi(g_{p}^{\prime})\tilde{q}_{i_{0}}=m_{i}h_{i,p}\gamma_{i}\tilde{q}_{i_{0}}=n_{i}\tilde{q}_{i}.

The form π(gp)q~i0\pi(g_{p}^{\prime})\tilde{q}_{i_{0}} is a form over p\mathbb{Q}_{p} with trivial denominators for all but finitely many pp. Applying εi,p\varepsilon_{i,p} for large ii does not change this. Furthermore, mim_{i} needs to divide all denominators of q~i0\tilde{q}_{i_{0}} over p\mathbb{Z}_{p} for all ii as q~i\tilde{q}_{i} is primitive. Hence, mim_{i} can only assume finitely many values and by reversing roles one can argue the same for nin_{i}. For any prime pp we have

discp(q~i)=pordp(mini)discp(π(gp)q~i0).\displaystyle\mathrm{disc}_{p}(\tilde{q}_{i})=p^{\mathrm{ord}_{p}(\frac{m_{i}}{n_{i}})}\mathrm{disc}_{p}(\pi(g_{p}^{\prime})\tilde{q}_{i_{0}}).

and hence

disc(q~i)=minippordp(discp(π(gp)q~i0))\displaystyle\mathrm{disc}(\tilde{q}_{i})=\tfrac{m_{i}}{n_{i}}\prod_{p}p^{\mathrm{ord}_{p}(\mathrm{disc}_{p}(\pi(g_{p}^{\prime})\tilde{q}_{i_{0}}))}

which is a contradiction to disc(q~i)\mathrm{disc}(\tilde{q}_{i})\to\infty.

Third factor: The proof here is the same as for the second factor. We do however point out that the morphism 𝐆𝐆¯\mathbf{G}\to\bar{\mathbf{G}} was constructed to satisfy that for any h𝐇Lih\in\mathbf{H}_{L_{i}} we have π((h,ΨLi(h))=ψ2,Li(h)\pi((h,\Psi_{L_{i}}(h))=\psi_{2,L_{i}}(h) and hence π(𝚫𝐇Li)=SOqLiΛLi#\pi(\mathbf{\Delta H}_{L_{i}})=\mathrm{SO}_{q_{L_{i}^{\perp}\cap\Lambda_{L_{i}}^{\#}}}. ∎

Remark 3.4.

We recall from the beginning of this section §3.2 that the first three conditions in admissibility were used in this order for the three factors in the above proof. This has a consequence: If LiGrn,k()L_{i}\in\mathrm{Gr}_{n,k}(\mathbb{Q}) is any sequence of subspaces satisfying properties (1) and (4), then for any giSpinQ()g_{i}\in\mathrm{Spin}_{Q}(\mathbb{R}) the packets

gi𝐇Li(𝔸)SpinQ()SpinQ(𝔸)/SpinQ()\displaystyle g_{i}\mathbf{H}_{L_{i}}(\mathbb{A})\mathrm{Spin}_{Q}(\mathbb{Q})\subset\mathchoice{\text{\raise 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Q})$}}}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}

are equidistributed as ii\to\infty. This can be used to obtain equidistribution of Qn,k(D)Grn,k()\mathcal{H}^{n,k}_{Q}(D)\subset\mathrm{Gr}_{n,k}(\mathbb{R}) without any restrictions on the kk-power free part of DD (as opposed to our main theorems in the introduction).

3.2.3. Simultaneous equidistribution of subspaces and shapes

Proposition 3.3 shows that the image 𝐌¯\bar{\mathbf{M}} of 𝐌\mathbf{M} under (3.2) satisfies that the projection onto each simple factors of 𝐆¯\bar{\mathbf{G}} is surjective. We claim that this implies 𝐌¯=𝐆¯\bar{\mathbf{M}}=\bar{\mathbf{G}}.

We first show that the projection of 𝐌¯\bar{\mathbf{M}} to SLk×SLnk\mathrm{SL}_{k}\times\mathrm{SL}_{n-k} is surjective. Note that any proper subgroup of SLk×SLnk\mathrm{SL}_{k}\times\mathrm{SL}_{n-k} with surjective projections is the graph of an isomorphism SLkSLnk\mathrm{SL}_{k}\to\mathrm{SL}_{n-k}. In particular, we are done with the intermediate claim if knkk\neq n-k. Suppose that k=nkk=n-k and choose for some ii0i\geq i_{0} an element h𝐇Lih\in\mathbf{H}_{L_{i}} acting trivially on LiL_{i} but not trivially on LiL_{i}^{\perp}. The projection of gLi1ρQ(h)gLig_{L_{i}}^{-1}\rho_{Q}(h)g_{L_{i}} to the first (resp. the second) SLk\mathrm{SL}_{k} is trivial (resp. non-trivial); the projection of 𝐌¯\bar{\mathbf{M}} to SLk×SLnk\mathrm{SL}_{k}\times\mathrm{SL}_{n-k} thus contains elements of the form (e,g)(e,g) with geg\neq e. This rules out graphs under isomorphisms and concludes the intermediate claim.

Now note that 𝐌¯\bar{\mathbf{M}} projects surjectively onto SpinQ\mathrm{Spin}_{Q} and SLk×SLnk\mathrm{SL}_{k}\times\mathrm{SL}_{n-k} and that the latter two \mathbb{Q}-groups do not have isomorphic simple factors. By a similar argument as above, we deduce that 𝐌¯=𝐆¯\bar{\mathbf{M}}=\bar{\mathbf{G}}.

3.2.4. Handling the unipotent radical

We now turn to proving that 𝐌=𝐆\mathbf{M}=\mathbf{G} which concludes the proof of the theorem. By §3.2.3 we know that 𝐌\mathbf{M} surjects to 𝐆¯\bar{\mathbf{G}}. In particular, by the Levi-Malcev theorem there exists some element in the unipotent radical of 𝐏n,k\mathbf{P}_{n,k}

yC=(IkC0Ink)𝐏n,k()y_{C}=\begin{pmatrix}I_{k}&C\\ 0&I_{n-k}\end{pmatrix}\in\mathbf{P}_{n,k}(\mathbb{Q})

such that 𝐌\mathbf{M} contains SpinQ×yC𝐃n,kyC1\mathrm{Spin}_{Q}\times y_{C}\mathbf{D}_{n,k}y_{C}^{-1}. By maximality of the latter group (cf. Remark 1.13), 𝐌\mathbf{M} is either equal to 𝐆\mathbf{G} or we have

𝐌=SpinQ×yC𝐃n,kyC1.\displaystyle\mathbf{M}=\mathrm{Spin}_{Q}\times y_{C}\mathbf{D}_{n,k}y_{C}^{-1}.

Assume by contradiction the latter. The inclusion δi1𝚫𝐇Liδi𝐌\delta_{i}^{-1}\mathbf{\Delta H}_{L_{i}}\delta_{i}\subset\mathbf{M} implies that

δ2,i1gLi1ρQ(h)gLiδ2,iyC𝐃n,kyC1\displaystyle\delta_{2,i}^{-1}g_{L_{i}}^{-1}\rho_{Q}(h)g_{L_{i}}\delta_{2,i}\in y_{C}\mathbf{D}_{n,k}y_{C}^{-1}

where δ2,i\delta_{2,i} denotes the second coordinate of the element δi𝐆()=SpinQ()×𝐏n,k()\delta_{i}\in\mathbf{G}(\mathbb{Q})=\mathrm{Spin}_{Q}(\mathbb{Q})\times\mathbf{P}_{n,k}(\mathbb{Q}). Since yC𝐃n,kyC1y_{C}\mathbf{D}_{n,k}y_{C}^{-1} stabilizes two subspaces, namely yCL0=L0y_{C}L_{0}=L_{0} and L=yC({(0,,0)}×nk)L^{\prime}=y_{C}(\{(0,\ldots,0)\}\times\mathbb{Q}^{n-k}), the conjugated group gLiδi,2yC𝐃n,kyC1δi,21gLi1g_{L_{i}}\delta_{i,2}y_{C}\mathbf{D}_{n,k}y_{C}^{-1}\delta_{i,2}^{-1}g_{L_{i}}^{-1} fixes the subspaces

gLiδi,2L0=gLiL0=LiandgLiδi,2L.\displaystyle g_{L_{i}}\delta_{i,2}L_{0}=g_{L_{i}}L_{0}=L_{i}\quad\text{and}\quad g_{L_{i}}\delta_{i,2}L^{\prime}.

As 𝐇Li\mathbf{H}_{L_{i}} fixes exactly the subspaces Li,LiL_{i},L_{i}^{\perp}, we must have

(3.4) Li=gLiδi,2L\displaystyle L_{i}^{\perp}=g_{L_{i}}\delta_{i,2}L^{\prime}

for all ii. We denote by v1i,,vniv_{1}^{i},\ldots,v_{n}^{i} the columns of gLig_{L_{i}} which is a basis of ΛLi\Lambda_{L_{i}} and by w1i,,wniw_{1}^{i},\ldots,w_{n}^{i} its dual basis. Recall that wk+1i,,wniw_{k+1}^{i},\ldots,w_{n}^{i} form a basis of ΛLi#Li\Lambda_{L_{i}}^{\#}\cap L_{i}^{\perp}. By (3.4), there exists a rational number αi×\alpha_{i}\in\mathbb{Q}^{\times} such that

(3.5) αi(wk+1iwni)=gLiδi,2yC(ek+1en).\displaystyle\alpha_{i}(w^{i}_{k+1}\wedge\ldots\wedge w^{i}_{n})=g_{L_{i}}\delta_{i,2}y_{C}(e_{k+1}\wedge\ldots\wedge e_{n}).

To simplify notation, we set ηi=δi,2yC\eta_{i}=\delta_{i,2}y_{C}.

We first control the numbers αi\alpha_{i}. From (3.1) we know that there are hi𝐇Lih_{i}\in\mathbf{H}_{L_{i}} such that

g2,igLi1ρQ(hi)gLiηig\displaystyle g_{2,i}g_{L_{i}}^{-1}\rho_{Q}(h_{i})g_{L_{i}}\eta_{i}\to g^{\prime}

for some g𝐏n,k(𝔸)g^{\prime}\in\mathbf{P}_{n,k}(\mathbb{A}). For ii large enough, there exist εi𝐏n,k(×^)\varepsilon_{i}\in\mathbf{P}_{n,k}(\mathbb{R}\times\widehat{\mathbb{Z}}) with g2,igLi1ρQ(hi)gLiηi=εigg_{2,i}g_{L_{i}}^{-1}\rho_{Q}(h_{i})g_{L_{i}}\eta_{i}=\varepsilon_{i}g^{\prime}. We now fix a prime pp so that ρQ(hi,p)gLiηi=gLiεi,pgp\rho_{Q}(h_{i,p})g_{L_{i}}\eta_{i}=g_{L_{i}}\varepsilon_{i,p}g^{\prime}_{p} (as g2,i𝐆()g_{2,i}\in\mathbf{G}(\mathbb{R})). Applying ρQ(hi,p)\rho_{Q}(h_{i,p}) to (3.4) we obtain

αi(wk+1iwni)=gLiεi,pgp(ek+1en).\displaystyle\alpha_{i}(w^{i}_{k+1}\wedge\ldots\wedge w^{i}_{n})=g_{L_{i}}\varepsilon_{i,p}g^{\prime}_{p}(e_{k+1}\wedge\ldots\wedge e_{n}).

Considering that the vectors wk+1iwniw^{i}_{k+1}\wedge\ldots\wedge w^{i}_{n} and ek+1ene_{k+1}\wedge\ldots\wedge e_{n} are primitive (see e.g. [casselsGeoNum, Ch. 1, Lemma 2]) and that gLig_{L_{i}} and gpg_{p}^{\prime} have bounded denominators, this shows that the denominators and numerators of the numbers αi\alpha_{i} are bounded independently of ii.

We now compute the discriminant of the lattice spanned by wk+1i,,wniw^{i}_{k+1},\ldots,w^{i}_{n} in two ways. First, note that as wk+1i,,wniw^{i}_{k+1},\ldots,w^{i}_{n} is a basis of ΛLi#Li\Lambda_{L_{i}}^{\#}\cap L_{i}^{\perp}, the discriminant in question is equal to the discriminant of ΛLi#Li\Lambda_{L_{i}}^{\#}\cap L_{i}^{\perp} and hence discQ(Li)\asymp\mathrm{disc}_{Q}(L_{i}). For the second way, observe that by (3.5) the discriminant of the lattice spanned by wk+1i,,wniw^{i}_{k+1},\ldots,w^{i}_{n} is given by αi1\alpha_{i}^{-1} multiplied by the determinant of the matrix with entries444 One conceptual way to see this is the following: the bilinear form ,Q\langle\cdot,\cdot\rangle_{Q} induces a bilinear form ,nkQ\langle\cdot,\cdot\rangle_{\bigwedge^{n-k}Q} on the wedge-product nkn\bigwedge^{n-k}\mathbb{Q}^{n} by defining it on pure wedges through v1vnk,w1wnknkQ=det(vi,wjQ).\displaystyle\langle v_{1}\wedge\ldots\wedge v_{n-k},w_{1}\wedge\ldots\wedge w_{n-k}\rangle_{\bigwedge^{n-k}Q}=\det(\langle v_{i},w_{j}\rangle_{Q}). This definition asserts that the discriminant of a rank nkn-k lattice is the quadratic value of the wedge product of any of its bases. Equation 3.6 is then obtained by replacing one of the wedges in wk+1iwni,wk+1iwninkQ\langle w_{k+1}^{i}\wedge\ldots\wedge w_{n}^{i},w_{k+1}^{i}\wedge\ldots\wedge w_{n}^{i}\rangle_{\bigwedge^{n-k}Q} via (3.5).

(3.6) gLiηiej,wmiQwithj,m>k.\langle g_{L_{i}}\eta_{i}e_{j},w_{m}^{i}\rangle_{Q}\quad\text{with}\ j,m>k.

To compute this determinant, write ηiej=ajie\eta_{i}e_{j}=\sum_{\ell}a_{\ell j}^{i}e_{\ell} for all j>kj>k so that

gLiηiej=ajivi.\displaystyle g_{L_{i}}\eta_{i}e_{j}=\sum_{\ell}a_{\ell j}^{i}v_{\ell}^{i}.

Using that {wli}\{w^{i}_{l}\} are dual vectors to {vli}\{v^{i}_{l}\} we compute

gLiηiej,wmiQ=ajivi,wmiQ=amji\langle g_{L_{i}}\eta_{i}e_{j},w_{m}^{i}\rangle_{Q}=\sum_{\ell}a_{\ell j}^{i}\langle v_{\ell}^{i},w_{m}^{i}\rangle_{Q}=a_{mj}^{i}

for all m,j>km,j>k. This implies that the determinant of the matrix with entries (3.6) is equal to the determinant of the lower right block of the matrix ηi\eta_{i}. The latter being equal to one, we conclude that the discriminant of the lattice spanned by wk+1iwniw^{i}_{k+1}\wedge\ldots\wedge w^{i}_{n} is equal to αi1\alpha_{i}^{-1}.

To summarize, we have established the following identity:

discQ(ΛLi#Li)=αi1\mathrm{disc}_{Q}(\Lambda_{L_{i}}^{\#}\cap L_{i}^{\perp})=\alpha_{i}^{-1}

Since the left-hand side of this identity goes to infinity as ii\to\infty (because discQ(Li)\asymp\mathrm{disc}_{Q}(L_{i})) while the right-hand side is bounded, we have reached a contradiction. It follows that 𝐌=𝐆\mathbf{M}=\mathbf{G} and hence the proof of Theorem 3.1 is complete.

4. The dynamical version of the theorem in codimension 2

In the following, we prove the analogue of Theorem 3.1 for the case k=2k=2 and nk3n-k\geq 3 (i.e. n5n\geq 5) ignoring the unipotent radical (cf. Remark 1.12); the case nk=2n-k=2, k3k\geq 3 is completely analogous and can be deduced by passing to the orthogonal complement. Contrary to cases treated in §3, the groups whose dynamics we use are not semisimple and have non-trivial central torus (see also Remark 1.6).

Recall the following notation (for k=2k=2):

  • 𝐆¯=SpinQ×SL2×SLn2\bar{\mathbf{G}}=\mathrm{Spin}_{Q}\times\mathrm{SL}_{2}\times\mathrm{SL}_{n-2} (here the ambient group).

  • 𝚫𝐇¯L={(h,ψ1,L(h),ψ2,L(h)):h𝐇L}\mathbf{\Delta\bar{H}}_{L}=\{(h,\psi_{1,L}(h),\psi_{2,L}(h))\mathrel{\mathop{\mathchar 58\relax}}h\in\mathbf{H}_{L}\} (here the acting group) for any LGrn,k()L\in\mathrm{Gr}_{n,k}(\mathbb{Q}) where ψ1,L\psi_{1,L} resp. ψ2,L\psi_{2,L} is roughly the restriction of the action of hh to LL resp. LL^{\perp} (cf. §2.3).

  • For any LGrn,2()L\in\mathrm{Gr}_{n,2}(\mathbb{Q}) a choice of intermediate lattice nΛL(n)#\mathbb{Z}^{n}\subset\Lambda_{L}\subset(\mathbb{Z}^{n})^{\#} (also implicit in the definition of 𝚫𝐇¯L\mathbf{\Delta\bar{H}}_{L}). For simplicity, we assume here in addition that ΛLL=L()\Lambda_{L}\cap L=L(\mathbb{Z}) and ΛL#L=L()\Lambda_{L}^{\#}\cap L^{\perp}=L^{\perp}(\mathbb{Z}); such a choice will be constructed later (cf. Proposition 6.6). Again, if QQ is unimodular, ΛL=n\Lambda_{L}=\mathbb{Z}^{n} satisfies this property.

Theorem 4.1.

Let LiGrn,2()L_{i}\in\mathrm{Gr}_{n,2}(\mathbb{Q}) for i1i\geq 1 be an admissible sequence of rational subspaces and let gi𝐆¯()g_{i}\in\bar{\mathbf{G}}(\mathbb{R}) be such that gi𝚫𝐇¯Li()gi1=𝚫𝐇¯L0()g_{i}\mathbf{\Delta\bar{H}}_{L_{i}}(\mathbb{R})g_{i}^{-1}=\mathbf{\Delta\bar{H}}_{L_{0}}(\mathbb{R}). Let μi\mu_{i} be the Haar probability measure on the closed orbit

gi𝚫𝐇¯Li(𝔸)𝐆¯()𝐆¯(𝔸)/𝐆¯().\displaystyle g_{i}\mathbf{\Delta\bar{H}}_{L_{i}}(\mathbb{A})\bar{\mathbf{G}}(\mathbb{Q})\subset\mathchoice{\text{\raise 2.15277pt\hbox{$\bar{\mathbf{G}}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\bar{\mathbf{G}}(\mathbb{Q})$}}}{\bar{\mathbf{G}}(\mathbb{A})\,/\,\bar{\mathbf{G}}(\mathbb{Q})}{\bar{\mathbf{G}}(\mathbb{A})\,/\,\bar{\mathbf{G}}(\mathbb{Q})}{\bar{\mathbf{G}}(\mathbb{A})\,/\,\bar{\mathbf{G}}(\mathbb{Q})}.

Then μi\mu_{i} converges to the Haar probability measure on 𝐆¯(𝔸)/𝐆¯()\mathchoice{\text{\raise 2.15277pt\hbox{$\bar{\mathbf{G}}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\bar{\mathbf{G}}(\mathbb{Q})$}}}{\bar{\mathbf{G}}(\mathbb{A})\,/\,\bar{\mathbf{G}}(\mathbb{Q})}{\bar{\mathbf{G}}(\mathbb{A})\,/\,\bar{\mathbf{G}}(\mathbb{Q})}{\bar{\mathbf{G}}(\mathbb{A})\,/\,\bar{\mathbf{G}}(\mathbb{Q})} as ii\to\infty.

We will structure the proof somewhat differently as equidistribution in the first component turns out to be the most difficult challenge in the proof. We fix an admissible sequence of subspaces LiL_{i} and a prime pp as in the definition of admissibility.

Recall (cf. §2.1) that for any LGrn,2()L\in\mathrm{Gr}_{n,2}(\mathbb{Q}) the group 𝐇L\mathbf{H}_{L} is not semisimple but only reductive. Let us describe the center as well as the commutator subgroup of 𝐇L\mathbf{H}_{L}. Define the pointwise stabilizer subgroup

𝐇Lpt={gSpinQ:g.v=v for all vL}.\displaystyle\mathbf{H}^{\mathrm{pt}}_{L}=\{g\in\mathrm{Spin}_{Q}\mathrel{\mathop{\mathchar 58\relax}}g.v=v\text{ for all }v\in L\}.

The center of 𝐇L\mathbf{H}_{L} is equal to 𝐇Lpt\mathbf{H}^{\mathrm{pt}}_{L^{\perp}} which we denote by 𝐓L\mathbf{T}_{L} for simplicity as it is abelian in this case. The commutator subgroup of 𝐇L\mathbf{H}_{L} is the semisimple group 𝐇Lpt\mathbf{H}^{\mathrm{pt}}_{L} and 𝐇L\mathbf{H}_{L} is isogenous to 𝐇Lpt×𝐓L\mathbf{H}^{\mathrm{pt}}_{L}\times\mathbf{T}_{L} (see Remark 2.1). As in §3, one can use the measure rigidity result of Gorodnik and Oh [gorodnikoh], this time for subgroups of the form 𝐇Lpt\mathbf{H}^{\mathrm{pt}}_{L}. These are however non-maximal so that we need to put extra effort to rule out intermediate groups555Roughly speaking, the obstacle to overcome are ’short vectors’ in LL. Ellenberg and Venkatesh [localglobalEV] prove the theorem we are alluding to here assuming that LL does not contain ’short vectors’ – see also Proposition 4.7.. Here, we use an averaging procedure involving the torus 𝐓L\mathbf{T}_{L} as well as Duke’s theorem [duke88] to show that these obstructions typically do not occur.

Let us outline the structure of the proof:

  • In §4.1, we show (in Lemma 4.4) that it is sufficient to prove equidistribution in each of the factors of 𝐆¯\bar{\mathbf{G}}, that is, to show equidistribution of the projections of the packets in Theorem 4.1 to

    (4.1) SpinQ(𝔸)/SpinQ(),SL2(𝔸)/SL2(),SLn2(𝔸)/SLn2().\displaystyle\mathchoice{\text{\raise 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Q})$}}}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})},\ \mathchoice{\text{\raise 2.15277pt\hbox{$\mathrm{SL}_{2}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathrm{SL}_{2}(\mathbb{Q})$}}}{\mathrm{SL}_{2}(\mathbb{A})\,/\,\mathrm{SL}_{2}(\mathbb{Q})}{\mathrm{SL}_{2}(\mathbb{A})\,/\,\mathrm{SL}_{2}(\mathbb{Q})}{\mathrm{SL}_{2}(\mathbb{A})\,/\,\mathrm{SL}_{2}(\mathbb{Q})},\ \mathchoice{\text{\raise 2.15277pt\hbox{$\mathrm{SL}_{n-2}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathrm{SL}_{n-2}(\mathbb{Q})$}}}{\mathrm{SL}_{n-2}(\mathbb{A})\,/\,\mathrm{SL}_{n-2}(\mathbb{Q})}{\mathrm{SL}_{n-2}(\mathbb{A})\,/\,\mathrm{SL}_{n-2}(\mathbb{Q})}{\mathrm{SL}_{n-2}(\mathbb{A})\,/\,\mathrm{SL}_{n-2}(\mathbb{Q})}.

    As mentioned in Remark 1.7, we use the elementary fact that ergodic systems are disjoint from trivial systems for this reduction (see Lemma 4.2).

  • To prove equidistribution in each of the factors of 𝐆¯\bar{\mathbf{G}}, we first note that equidistribution in the third factor can be verified as in §3, Proposition 3.3. Equidistribution in the second factor turns out to be a variant of Duke’s theorem [duke88] which we discuss in §4.2.

  • Due to the difficulties described above, equidistribution in the first factor of 𝐆¯\bar{\mathbf{G}}, is the hardest to prove (cf. §4.3) and implies Theorem 4.1 by the first two items in this list. In §4.3.2, we collect a useful corollary of the above variant of Duke’s theorem which we then use in Lemma 4.10 to prove that the subspaces in the packet do not contain short vectors on average.

4.1. Reduction to individual equidistribution

As explained, we begin by reducing Theorem 4.1 to the corresponding equidistribution statement in each of the factors of 𝐆¯\bar{\mathbf{G}}. To this end, we will use the following elementary fact from abstract ergodic theory.

Lemma 4.2.

Let 𝖷1=(X,1,μ1,T1)\mathsf{X}_{1}=(X,\mathcal{B}_{1},\mu_{1},T_{1}) and 𝖷2=(X2,2,μ2,T2)\mathsf{X}_{2}=(X_{2},\mathcal{B}_{2},\mu_{2},T_{2}) be measure-preserving systems. Suppose that 𝖷1\mathsf{X}_{1} is ergodic and that 𝖷2\mathsf{X}_{2} is trivial (i.e. T2(x)=xT_{2}(x)=x for μ2\mu_{2}-almost every xX2x\in X_{2}). Then the only joining of 𝖷1\mathsf{X}_{1} and 𝖷2\mathsf{X}_{2} is μ1×μ2\mu_{1}\times\mu_{2}.

Proof.

Let ν\nu be a joining and let A1×A2X1×X2A_{1}\times A_{2}\subset X_{1}\times X_{2} be measurable. It suffices to show that ν(A1×A2)=μ1(A1)μ2(A2)\nu(A_{1}\times A_{2})=\mu_{1}(A_{1})\mu_{2}(A_{2}). By T1×T2T_{1}\times T_{2}-invariance of ν\nu, we have

ν(A1×A2)\displaystyle\nu(A_{1}\times A_{2}) =X1×X21A1(x1)1A2(x2)dν(x1,x2)\displaystyle=\int_{X_{1}\times X_{2}}1_{A_{1}}(x_{1})1_{A_{2}}(x_{2})\,\mathrm{d}\nu(x_{1},x_{2})
=1Mm=0M1X1×X21A1(T1mx1)1A2(T2mx2)dν(x1,x2)\displaystyle=\frac{1}{M}\sum_{m=0}^{M-1}\int_{X_{1}\times X_{2}}1_{A_{1}}(T_{1}^{m}x_{1})1_{A_{2}}(T_{2}^{m}x_{2})\,\mathrm{d}\nu(x_{1},x_{2})

As 𝖷1\mathsf{X}_{1} is ergodic, there is a μ1\mu_{1}-conull set B1X1B_{1}\subset X_{1} with

1Mm=0M11A1(T1m(x))μ1(A1)\displaystyle\frac{1}{M}\sum_{m=0}^{M-1}1_{A_{1}}(T_{1}^{m}(x))\to\mu_{1}(A_{1})

for every xB1x\in B_{1} by Birkhoff’s ergodic theorem. As 𝖷2\mathsf{X}_{2} is trivial, there is a μ2\mu_{2}-conull set B2B_{2} with T2(x)=xT_{2}(x)=x for all xB2x\in B_{2}. We let B=B1×B2B=B_{1}\times B_{2} and note that BB has full measure as it is the intersection of the full-measure sets B1×X2B_{1}\times X_{2} and X1×B2X_{1}\times B_{2} (we use here that ν\nu is a joining). Therefore,

ν(A1×A2)\displaystyle\nu(A_{1}\times A_{2}) =1Mm=0M1B1A1(T1mx1)1A2(T2mx2)dν(x1,x2)\displaystyle=\frac{1}{M}\sum_{m=0}^{M-1}\int_{B}1_{A_{1}}(T_{1}^{m}x_{1})1_{A_{2}}(T_{2}^{m}x_{2})\,\mathrm{d}\nu(x_{1},x_{2})
=1Mm=0M1B1A1(T1mx1)1A2(x2)dν(x1,x2)\displaystyle=\frac{1}{M}\sum_{m=0}^{M-1}\int_{B}1_{A_{1}}(T_{1}^{m}x_{1})1_{A_{2}}(x_{2})\,\mathrm{d}\nu(x_{1},x_{2})
=B1Mm=0M11A1(T1mx1)1A2(x2)dν(x1,x2)\displaystyle=\int_{B}\frac{1}{M}\sum_{m=0}^{M-1}1_{A_{1}}(T_{1}^{m}x_{1})1_{A_{2}}(x_{2})\,\mathrm{d}\nu(x_{1},x_{2})
Bμ1(A1)1A2(x2)dν(x1,x2)=μ1(A1)μ2(A2)\displaystyle\to\int_{B}\mu_{1}(A_{1})1_{A_{2}}(x_{2})\,\mathrm{d}\nu(x_{1},x_{2})=\mu_{1}(A_{1})\mu_{2}(A_{2})

as claimed. ∎

We aim to apply Lemma 4.2 to any weak-limit μ\mu of the measures in Theorem 4.1. Thus, we need to establish some invariance of the latter. Let pp be as in the definition of admissibility.

Lemma 4.3.

There exists gGLn(p)g\in\mathrm{GL}_{n}(\mathbb{Q}_{p}) with the following property: Let LGrn,2(p)L\in\mathrm{Gr}_{n,2}(\mathbb{Q}_{p}) be the subspace spanned by the first two columns of gg. Then μ\mu is invariant under the subgroup of 𝚫𝐇¯L(p)𝐆¯(p)\mathbf{\Delta\bar{H}}_{L}(\mathbb{Q}_{p})\subset\bar{\mathbf{G}}(\mathbb{Q}_{p}) where

𝚫𝐇¯L={(h,π1(g1ρQ(h)g),π2(g1ρQ(h1)g)t):h𝐇L}.\displaystyle\mathbf{\Delta\bar{H}}_{L}=\{(h,\pi_{1}(g^{-1}\rho_{Q}(h)g),\pi_{2}(g^{-1}\rho_{Q}(h^{-1})g)^{t})\mathrel{\mathop{\mathchar 58\relax}}h\in\mathbf{H}_{L}\}.

Moreover, the p\mathbb{Q}_{p}-group 𝚫𝐇¯L\mathbf{\Delta\bar{H}}_{L} is strongly isotropic.

Proof.

First of all, we prove that there exists a compact subset KGLn(p)K\subset\mathrm{GL}_{n}(\mathbb{Q}_{p}) such that gLiKg_{L_{i}}\in K for all ii\in\mathbb{N}. Recall that gLig_{L_{i}} consists of a basis of an intermediate lattice nΛLi(n)#\mathbb{Z}^{n}\subseteq\Lambda_{L_{i}}\subseteq(\mathbb{Z}^{n})^{\#} (cf. §2.3). The set KK of elements gGLn(p)g\in\mathrm{GL}_{n}(\mathbb{Q}_{p}) with pngpn(pn)#\mathbb{Z}_{p}^{n}\subset g\mathbb{Z}_{p}^{n}\subset(\mathbb{Z}_{p}^{n})^{\#} is compact (in fact, it consists of finitely many cosets modulo GLn(p)\mathrm{GL}_{n}(\mathbb{Z}_{p}) on the right).

By compactness of KK we may assume (by passing to a subsequence) that the sequence (gLi)i(g_{L_{i}})_{i\in\mathbb{N}} converges to some gKg\in K. Let LL denote the p\mathbb{Q}_{p}-plane spanned by the first two columns of gg. Note that μ\mu is 𝚫𝐇¯L(p)\mathbf{\Delta\bar{H}}_{L}(\mathbb{Q}_{p})-invariant because each μi\mu_{i} is 𝚫𝐇¯Li(p)\mathbf{\Delta\bar{H}}_{L_{i}}(\mathbb{Q}_{p})-invariant. Therefore, we are left to show that LL is non-degenerate and 𝚫𝐇¯L(p)\mathbf{\Delta\bar{H}}_{L}(\mathbb{Q}_{p}) is strongly isotropic.

We first observe that LL and LL^{\perp} are non-degenerate. Indeed, since gLigg_{L_{i}}\rightarrow g, there exist p\mathbb{Z}_{p}-bases of the subspaces LiL_{i} which converge towards a basis of LL. Taking discriminants of LiL_{i} and LL with respect to these bases, we obtain

discp,Q(Li)discp,Q(L).\displaystyle\mathrm{disc}_{p,Q}(L_{i})\rightarrow\mathrm{disc}_{p,Q}(L).

Since p×/(p×)2\mathbb{Z}_{p}^{\times}/(\mathbb{Z}_{p}^{\times})^{2} is discrete, discp,Q(Li)\mathrm{disc}_{p,Q}(L_{i}) is eventually constant and therefore discp,Q(L)=discp,Q(Li)\mathrm{disc}_{p,Q}(L)=\mathrm{disc}_{p,Q}(L_{i}) for ii large enough; non-degeneracy of LL follows. In particular, LL^{\perp} is non-degenerate.

We may now use Corollary 2.8 to show 𝚫𝐇¯L\mathbf{\Delta\bar{H}}_{L} or equivalently 𝐇L\mathbf{H}_{L} is strongly isotropic. Since 𝐇Li\mathbf{H}_{L_{i}} is strongly isotropic at pp, the quadratic spaces (Q|Li,Li)(Q|_{L_{i}},L_{i}) and (Q|Li,Li)(Q|_{L_{i}^{\perp}},L_{i}^{\perp}) are isotropic over p\mathbb{Q}_{p}. By isotropy of the spaces (Q|Li,Li)(Q|_{L_{i}},L_{i}), we have a sequence of non-zero primitive vectors viLi(p)v_{i}\in L_{i}(\mathbb{Z}_{p}) such that Q(vi)=0Q(v_{i})=0 (after multiplying with denominators). By compactness of pnppn\mathbb{Z}_{p}^{n}\setminus p\mathbb{Z}_{p}^{n}, the sequence viv_{i} admits a limit vpnppnv\in\mathbb{Z}_{p}^{n}\setminus p\mathbb{Z}_{p}^{n} after passing to a subsequence. This limit clearly satisfies vL(p)v\in L(\mathbb{Z}_{p}) and Q(v)=0Q(v)=0, so (Q|L,L)(Q|_{L},L) is isotropic. An identical argument proves that (Q|L,L)(Q|_{L^{\perp}},L^{\perp}) is also isotropic, which proves (cf. Corollary 2.8) that 𝐇L\mathbf{H}_{L} is a strongly isotropic group. The proof is complete. ∎

Recall that ψ1,L,ψ2,L\psi_{1,L},\psi_{2,L} denote the epimorphisms 𝐇LiSOqLi\mathbf{H}_{L_{i}}\to\mathrm{SO}_{q_{L_{i}}}, 𝐇LiSOqLi\mathbf{H}_{L_{i}}\to\mathrm{SO}_{q_{L_{i}^{\perp}}} respectively.

Lemma 4.4.

Suppose that individual equidistribution holds i.e. that

  1. (1)

    gi,1𝐇Li(𝔸)SpinQ()g_{i,1}\mathbf{H}_{L_{i}}(\mathbb{A})\mathrm{Spin}_{Q}(\mathbb{Q}) is equidistributed in SpinQ(𝔸)/SpinQ()\mathchoice{\text{\raise 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Q})$}}}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})},

  2. (2)

    gi,2ψ1,L(𝐇Li(𝔸))SL2()g_{i,2}\psi_{1,L}(\mathbf{H}_{L_{i}}(\mathbb{A}))\mathrm{SL}_{2}(\mathbb{Q}) is equidistributed in SL2(𝔸)/SL2()\mathchoice{\text{\raise 2.15277pt\hbox{$\mathrm{SL}_{2}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathrm{SL}_{2}(\mathbb{Q})$}}}{\mathrm{SL}_{2}(\mathbb{A})\,/\,\mathrm{SL}_{2}(\mathbb{Q})}{\mathrm{SL}_{2}(\mathbb{A})\,/\,\mathrm{SL}_{2}(\mathbb{Q})}{\mathrm{SL}_{2}(\mathbb{A})\,/\,\mathrm{SL}_{2}(\mathbb{Q})}, and

  3. (3)

    gi,3ψ2,L(𝐇Li(𝔸))SLn2()g_{i,3}\psi_{2,L}(\mathbf{H}_{L_{i}}(\mathbb{A}))\mathrm{SL}_{n-2}(\mathbb{Q}) is equidistributed in SLn2(𝔸)/SLn2()\mathchoice{\text{\raise 2.15277pt\hbox{$\mathrm{SL}_{n-2}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathrm{SL}_{n-2}(\mathbb{Q})$}}}{\mathrm{SL}_{n-2}(\mathbb{A})\,/\,\mathrm{SL}_{n-2}(\mathbb{Q})}{\mathrm{SL}_{n-2}(\mathbb{A})\,/\,\mathrm{SL}_{n-2}(\mathbb{Q})}{\mathrm{SL}_{n-2}(\mathbb{A})\,/\,\mathrm{SL}_{n-2}(\mathbb{Q})}.

Then Theorem 4.1 holds.

Proof.

Let μ\mu be a weak-limit and choose LL as in Lemma 4.3. By assumption, μ\mu is a joining with respect to the Haar measures on each factor. We proceed in two steps and apply Lemma 4.2 once in each step.

For the first step, we choose h𝐇L(p)h\in\mathbf{H}_{L}(\mathbb{Q}_{p}) which acts trivially on LL but non-trivially on LL^{\perp}. As 𝐇L(p)\mathbf{H}_{L}(\mathbb{Q}_{p}) is strongly isotropic we can choose hh so that it is unipotent and not contained in any normal subgroup of SpinQ(p)\mathrm{Spin}_{Q}(\mathbb{Q}_{p}). Since SpinQ\mathrm{Spin}_{Q} is simply connected and SpinQ(p)\mathrm{Spin}_{Q}(\mathbb{Q}_{p}) is isotropic, SpinQ\mathrm{Spin}_{Q} has strong approximation with respect to {p}\{p\} (see for example [platonov, Thm. 7.12]). In particular, SpinQ(p)\mathrm{Spin}_{Q}(\mathbb{Q}_{p}) acts ergodically on X1=SpinQ(𝔸)/SpinQ()X_{1}=\mathchoice{\text{\raise 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Q})$}}}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})} with respect to the Haar measure on X1X_{1}. By Mautner’s phenomenon (see [MT2, §2] for this instance), hh also acts ergodically. Embedding hh diagonally (using the embedding in Lemma 4.3), we can apply Lemma 4.2 for X1X_{1} as above and X2=SL2(𝔸)/SL2()X_{2}=\mathchoice{\text{\raise 2.15277pt\hbox{$\mathrm{SL}_{2}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathrm{SL}_{2}(\mathbb{Q})$}}}{\mathrm{SL}_{2}(\mathbb{A})\,/\,\mathrm{SL}_{2}(\mathbb{Q})}{\mathrm{SL}_{2}(\mathbb{A})\,/\,\mathrm{SL}_{2}(\mathbb{Q})}{\mathrm{SL}_{2}(\mathbb{A})\,/\,\mathrm{SL}_{2}(\mathbb{Q})} and obtain that the pushforward of μ\mu to X1×X2X_{1}\times X_{2} is the Haar measure.

For the second step, we proceed similarly. Choose h𝐇L(p)h\in\mathbf{H}_{L}(\mathbb{Q}_{p}) which acts trivially on LL^{\perp} but non-trivially on LL. One checks that hh acts ergodically on X1×X2X_{1}\times X_{2} (via π2(g1ρQ(h1)g)t\pi_{2}(g^{-1}\rho_{Q}(h^{-1})g)^{t} on the second factor cf. Lemma 4.3). Applying Lemma 4.2 again for X1×X2X_{1}\times X_{2} and for X3=SLn2(𝔸)/SLn2()X_{3}=\mathchoice{\text{\raise 2.15277pt\hbox{$\mathrm{SL}_{n-2}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathrm{SL}_{n-2}(\mathbb{Q})$}}}{\mathrm{SL}_{n-2}(\mathbb{A})\,/\,\mathrm{SL}_{n-2}(\mathbb{Q})}{\mathrm{SL}_{n-2}(\mathbb{A})\,/\,\mathrm{SL}_{n-2}(\mathbb{Q})}{\mathrm{SL}_{n-2}(\mathbb{A})\,/\,\mathrm{SL}_{n-2}(\mathbb{Q})} we obtain the claim. ∎

We prove the conditions of Lemma 4.4 in an order which is potentially peculiar at first sight. The third assertion can be proven exactly as in §3 by applying [gorodnikoh] (see Proposition 3.3) so we omit it here.

4.2. Individual equidistribution in the second factor

The aim of this second section is to prove the second assertion of Lemma 4.4. As we shall see, it follows from Duke’s theorem [duke88] and its generalizations – see e.g. [Dukeforcubic, harcosmichelII]. Note that

gi,2ψ1,L(𝐇Li(𝔸))SL2()gi,2SOqLi(𝔸)SL2().\displaystyle g_{i,2}\psi_{1,L}(\mathbf{H}_{L_{i}}(\mathbb{A}))\mathrm{SL}_{2}(\mathbb{Q})\subset g_{i,2}\mathrm{SO}_{q_{L_{i}}}(\mathbb{A})\mathrm{SL}_{2}(\mathbb{Q}).

While the right-hand side is equidistributed by Duke’s theorem (specifically for instance by [Dukeforcubic, Thm. 4.6] or – as we assume a splitting condition – by [W-linnik]), one needs to verify that the left-hand side has sufficiently large ’volume’.

Proposition 4.5.

For LGrn,2()L\in\mathrm{Gr}_{n,2}(\mathbb{Q}) and any field KK of characteristic zero the image ψ1,L(𝐇L(K))\psi_{1,L}(\mathbf{H}_{L}(K)) contains the group of squares in the abelian group SOqL(K)\mathrm{SO}_{q_{L}}(K).

Proof.

The proof is surprisingly involved. Observe first that ψ1,L(𝐇L(K))\psi_{1,L}(\mathbf{H}_{L}(K)) contains ψ1,L(𝐓L(K))\psi_{1,L}(\mathbf{T}_{L}(K)) which we now identify as the set of squares in SOqL(K)\mathrm{SO}_{q_{L}}(K).

We identify the torus 𝐓L\mathbf{T}_{L} in terms of the Clifford algebra. Denote by 𝒞\mathcal{C} resp. 𝒞0\mathcal{C}^{0} the Clifford algebra of QQ resp. the even Clifford algebra of QQ. Let v1,v2v_{1},v_{2} be an orthogonal basis of LL and complete it into an orthogonal basis of n\mathbb{Q}^{n}. Consider X=v1v2(𝒞0)×X=v_{1}v_{2}\in(\mathcal{C}^{0})^{\times} (LL is non-degenerate) which satisfies the relations

(4.2) Xvi=viX,Xv1=Q(v1)v2=v1X,Xv2=v2X\displaystyle Xv_{i}=v_{i}X,\ Xv_{1}=-Q(v_{1})v_{2}=-v_{1}X,\ Xv_{2}=-v_{2}X

for all i>2i>2. Moreover, X2=Q(v1)Q(v2)×X^{2}=-Q(v_{1})Q(v_{2})\in\mathbb{Q}^{\times}. Denote by σ\sigma the standard involution on 𝒞\mathcal{C}. Then σ(X)=v2v1=X\sigma(X)=v_{2}v_{1}=-X.

It follows directly from (4.2) that for all a,bKa,b\in K the element t=a+bXt=a+bX satisfies tvi=vittv_{i}=v_{i}t for i>2i>2. Also,

tv1σ(t)\displaystyle tv_{1}\sigma(t) =(a+bX)v1(abX)=a2v1+abXv1abv1Xb2Xv1X\displaystyle=(a+bX)v_{1}(a-bX)=a^{2}v_{1}+abXv_{1}-abv_{1}X-b^{2}Xv_{1}X
=a2v12abQ(v1)v2b2Q(v1)Q(v2)v1L\displaystyle=a^{2}v_{1}-2abQ(v_{1})v_{2}-b^{2}Q(v_{1})Q(v_{2})v_{1}\in L

and similarly for v2v_{2}. Therefore, t𝐓Lt\in\mathbf{T}_{L} if and only if

σ(t)t=(abX)(a+bX)=a2b2X2=a2b2Q(v1)Q(v2)=1.\displaystyle\sigma(t)t=(a-bX)(a+bX)=a^{2}-b^{2}X^{2}=a^{2}-b^{2}Q(v_{1})Q(v_{2})=1.

We set

F=(Q(v1)Q(v2))=(discQ(L))\displaystyle F=\mathbb{Q}(-Q(v_{1})Q(v_{2}))=\mathbb{Q}(-\mathrm{disc}_{Q}(L))

and embed FF into 𝒞0\mathcal{C}^{0} via discQ(L)X\sqrt{-\mathrm{disc}_{Q}(L)}\mapsto X. The non-trivial Galois automorphism on FF is then given by σ|F\sigma|_{F}. To summarize, we obtain

𝐓L(K)={tFK:σ(t)t=1}.\displaystyle\mathbf{T}_{L}(K)=\{t\in F\otimes K\mathrel{\mathop{\mathchar 58\relax}}\sigma(t)t=1\}.

Also, recall that the special Clifford group surjects onto SOQ\mathrm{SO}_{Q} so that one may show quite analogously

SOqL(K)=(FK)×/K×.\displaystyle\mathrm{SO}_{q_{L}}(K)=(F\otimes K)^{\times}/K^{\times}.

The proposition then follows from Hilbert’s theorem 90 as in the proof of [W-linnik, Lemma 7.2]. ∎

Corollary 4.6.

The orbits

gi,2ψ1,L(𝐇Li(𝔸))SL2()SL2(𝔸)/SL2()\displaystyle g_{i,2}\psi_{1,L}(\mathbf{H}_{L_{i}}(\mathbb{A}))\mathrm{SL}_{2}(\mathbb{Q})\subset\mathchoice{\text{\raise 2.15277pt\hbox{$\mathrm{SL}_{2}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathrm{SL}_{2}(\mathbb{Q})$}}}{\mathrm{SL}_{2}(\mathbb{A})\,/\,\mathrm{SL}_{2}(\mathbb{Q})}{\mathrm{SL}_{2}(\mathbb{A})\,/\,\mathrm{SL}_{2}(\mathbb{Q})}{\mathrm{SL}_{2}(\mathbb{A})\,/\,\mathrm{SL}_{2}(\mathbb{Q})}

equidistribute as ii\to\infty.

Proof.

We deduce the corollary from existing literature and Proposition 4.5. We first claim that as ii\to\infty the sets

(4.3) gi,2SOqLi(^)ψ1,L(𝐇Li(𝔸))SL2()\displaystyle g_{i,2}\mathrm{SO}_{q_{L_{i}}}(\widehat{\mathbb{Z}})\psi_{1,L}(\mathbf{H}_{L_{i}}(\mathbb{A}))\mathrm{SL}_{2}(\mathbb{Q})

are equidistributed. By Proposition 4.5 the abelian group SOqLi(^)ψ1,L(𝐇Li(𝔸))\mathrm{SO}_{q_{L_{i}}}(\widehat{\mathbb{Z}})\psi_{1,L}(\mathbf{H}_{L_{i}}(\mathbb{A})) contains the group SOqLi(^)SOqLi(𝔸)2\mathrm{SO}_{q_{{}_{L_{i}}}}(\widehat{\mathbb{Z}})\mathrm{SO}_{q_{{}_{L_{i}}}}(\mathbb{A})^{2} where SOqLi(𝔸)2\mathrm{SO}_{q_{{}_{L_{i}}}}(\mathbb{A})^{2} denotes the group of squares.

The orbit (4.3) is then a union of suborbits of the same form associated to these subgroups. Any sequence of such suborbits are equidistributed, for instance, by [harcosmichelII] as the volume is of size666Since the 22-torsion of the Picard group of the order of discriminant discQ(Li)\mathrm{disc}_{Q}(L_{i}) has size discQ(Li)o(1)\mathrm{disc}_{Q}(L_{i})^{o(1)} (see e.g. [Cassels, p. 342]), the squares form a subgroup of size discQ(Li)12+o(1)\mathrm{disc}_{Q}(L_{i})^{\frac{1}{2}+o(1)}. discQ(Li)12+o(1)\mathrm{disc}_{Q}(L_{i})^{\frac{1}{2}+o(1)}. We note that the result in [harcosmichelII] allows for smaller volumes (where the exponent 12\frac{1}{2} can be replaced by 12η\frac{1}{2}-\eta for some not too large η>0\eta>0). In the case needed here one can also apply Linnik’s ergodic method as we assume a splitting condition at a fixed prime – see [W-linnik, §7]. By averaging, the claim in (4.3) follows. The corollary is implied by (4.3) and ergodicity of the Haar measure on SL2(𝔸)/SL2()\mathrm{SL}_{2}(\mathbb{A})/\mathrm{SL}_{2}(\mathbb{Q}) under any diagonal flow. ∎

4.3. Individual equidistribution in the first factor

In view of the discussion in §4.2 and Lemma 4.4, it suffices to show equidistribution of the packets

gi,1𝐇Li(𝔸)SpinQ()SpinQ(𝔸)/SpinQ()\displaystyle g_{i,1}\mathbf{H}_{L_{i}}(\mathbb{A})\mathrm{Spin}_{Q}(\mathbb{Q})\subset\mathchoice{\text{\raise 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Q})$}}}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}

to prove Theorem 4.1. We proceed in several steps.

4.3.1. An equidistribution theorem for the pointwise stabilizers

We first establish the following proposition which shows that either orbits of the pointwise stabilizer are equidistributed or there is some arithmetic obstruction.

Proposition 4.7.

Let (Li)i(L_{i})_{i} be a sequence of 22-dimensional rational subspaces such that there exists a prime pp for which 𝐇Lipt(p)\mathbf{H}^{\mathrm{pt}}_{L_{i}}(\mathbb{Q}_{p}) is strongly isotropic for all ii. Let gi𝐆()g_{i}\in\mathbf{G}(\mathbb{R}) and assume that discQ(Li)\mathrm{disc}_{Q}(L_{i})\to\infty as ii\to\infty. Then one of the following statements is true:

  1. (1)

    The packets gi𝐇Lipt(𝔸)SpinQ()g_{i}\mathbf{H}^{\mathrm{pt}}_{L_{i}}(\mathbb{A})\mathrm{Spin}_{Q}(\mathbb{Q}) are equidistributed in SpinQ(𝔸)/SpinQ()\mathchoice{\text{\raise 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Q})$}}}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})} as ii\to\infty.

  2. (2)

    There exists a rational vector vn{0}v\in\mathbb{Q}^{n}\setminus\{0\} and lattice elements δiSpinQ()\delta_{i}\in\mathrm{Spin}_{Q}(\mathbb{Q}) such that

    v=iδi1.Li().\displaystyle\mathbb{Q}v=\bigcap_{i}\delta_{i}^{-1}.L_{i}(\mathbb{Q}).

    The lattice elements additionally satisfy that there exist hi𝐇Lipt(𝔸)h_{i}\in\mathbf{H}^{\mathrm{pt}}_{L_{i}}(\mathbb{A}) such that the sequence gihiδig_{i}h_{i}\delta_{i} is convergent as ii\to\infty.

Proof.

We prove the proposition in exactly the same way we proved the first case in Proposition 3.3; we will thus be rather brief. Let δiSpinQ()\delta_{i}\in\mathrm{Spin}_{Q}(\mathbb{Q}) and a connected \mathbb{Q}-group 𝐌<𝐆¯\mathbf{M}<\bar{\mathbf{G}} be as in Theorem 3.2. In particular,

δi1𝐇Liptδi<𝐌\displaystyle\delta_{i}^{-1}\mathbf{H}^{\mathrm{pt}}_{L_{i}}\delta_{i}<\mathbf{M}

and it suffices for equidistribution to verify that 𝐌=SpinQ\mathbf{M}=\mathrm{Spin}_{Q}. One can see that 𝐌\mathbf{M} strictly contains δi1𝐇Liptδi\delta_{i}^{-1}\mathbf{H}^{\mathrm{pt}}_{L_{i}}\delta_{i} for all ii by using discQ(Li)\mathrm{disc}_{Q}(L_{i})\to\infty and repeating the proof of the first case in Proposition 3.3.

Contrary to the case treated in Proposition 3.3 the groups 𝐇Lipt\mathbf{H}^{\mathrm{pt}}_{L_{i}} are non-maximal. The intermediate groups can however be understood explicitly: they are of the form 𝐇Wpt\mathbf{H}^{\mathrm{pt}}_{W} where WW is a rational line contained in δi1.Li\delta_{i}^{-1}.L_{i} for all ii. For a proof of this fact we refer to [localglobalEV, Prop. 4], see also the arXiv version of the same paper where the authors give an elementary proof in the case n27n-2\geq 7. This concludes the proof of the proposition. ∎

Corollary 4.8.

Let the notation and the assumptions be as in Proposition 4.7 and suppose that the second case holds. Then

minwLi(){0}Q(w)=minw2{0}qLi(w)\displaystyle\min_{w\in L_{i}(\mathbb{Z})\setminus\{0\}}Q(w)=\min_{w\in\mathbb{Z}^{2}\setminus\{0\}}q_{L_{i}}(w)

is bounded as ii\to\infty.

Proof.

Let vnv\in\mathbb{Q}^{n} be as in Proposition 4.7 and suppose without loss of generality that vv is integral and primitive. Suppose also that gihiδigSpinQ(𝔸)g_{i}h_{i}\delta_{i}\to g^{\prime}\in\mathrm{Spin}_{Q}(\mathbb{A}) and write gihiδi=εigg_{i}h_{i}\delta_{i}=\varepsilon_{i}g^{\prime} where εie\varepsilon_{i}\to e. Let i0i_{0} be large enough so that εi𝐆¯(×^)\varepsilon_{i}\in\bar{\mathbf{G}}(\mathbb{R}\times\widehat{\mathbb{Z}}) for all ii0i\geq i_{0} and let NN\in\mathbb{N} be the smallest integer such that NgpNg^{\prime}_{p} is integral for all primes pp.

We claim that vi:=Nδi.vLi()v_{i}\mathrel{\mathop{\mathchar 58\relax}}=N\delta_{i}.v\in L_{i}(\mathbb{Z}). To see this, first note that viLi()v_{i}\in L_{i}(\mathbb{Q}). Furthermore, for any prime pp the vector viv_{i} is contained in L(p)L(\mathbb{Z}_{p}). Indeed, hi,p𝐇Lipt(p)h_{i,p}\in\mathbf{H}^{\mathrm{pt}}_{L_{i}}(\mathbb{Q}_{p}) necessarily fixes viv_{i} and, as gi,p=eg_{i,p}=e,

vi=hi,p.vi=Nhi,pδi.v=Nεi,pgp.vpn.\displaystyle v_{i}=h_{i,p}.v_{i}=Nh_{i,p}\delta_{i}.v=N\varepsilon_{i,p}g_{p}.v\in\mathbb{Z}_{p}^{n}.

This proves the claim and hence the corollary as Q(Nδi.v)=N2Q(v)Q(N\delta_{i}.v)=N^{2}Q(v). ∎

4.3.2. A corollary of equidistribution in the second factor

In the following we would like to give an estimate on the measure of the set of points in gi,1𝐓Li(𝔸)SpinQ()g_{i,1}\mathbf{T}_{L_{i}}(\mathbb{A})\mathrm{Spin}_{Q}(\mathbb{Q}) whose associated point in SL2(𝔸)/SL2()\mathchoice{\text{\raise 2.15277pt\hbox{$\mathrm{SL}_{2}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathrm{SL}_{2}(\mathbb{Q})$}}}{\mathrm{SL}_{2}(\mathbb{A})\,/\,\mathrm{SL}_{2}(\mathbb{Q})}{\mathrm{SL}_{2}(\mathbb{A})\,/\,\mathrm{SL}_{2}(\mathbb{Q})}{\mathrm{SL}_{2}(\mathbb{A})\,/\,\mathrm{SL}_{2}(\mathbb{Q})} is ’close’ to the cusp. This will allow to ’wash out’ the effect of the obstructions in Proposition 4.7 on average across the full stabilizer group. To obtain said estimate, we introduce a height function that suits our needs.

Let 𝒮2\mathcal{S}_{2} be the space of positive definite real binary quadratic forms up to similarity777Two positive definite binary real quadratic forms Q1,Q2Q_{1},Q_{2} are similar if there exist λ>0\lambda>0 and gGL2()g\in\mathrm{GL}_{2}(\mathbb{Z}) with Q2()=λQ1(g)Q_{2}(\cdot)=\lambda Q_{1}(g\cdot). Note that the space 𝒮2\mathcal{S}_{2} will be discussed in more detail in §6.2. and write [q][q] for the similarity class of a binary form qq. We define for ε>0\varepsilon>0

𝒮2(ε)={[q]𝒮2:minw2{0}q(w)>εdisc(q)}.\displaystyle\mathcal{S}_{2}(\varepsilon)=\big{\{}[q]\in\mathcal{S}_{2}\mathrel{\mathop{\mathchar 58\relax}}\min_{w\in\mathbb{Z}^{2}\setminus\{0\}}q(w)>\varepsilon\sqrt{\mathrm{disc}(q)}\big{\}}.

Note that the condition is independent of the choice of representative of [q][q].

By Mahler’s compactness criterion [Mahlercompactness], these are compact subsets of 𝒮2\mathcal{S}_{2} and any compact subset is contained in 𝒮2(ε)\mathcal{S}_{2}(\varepsilon) for some ε>0\varepsilon>0. Furthermore, one can show that the Haar measure of 𝒮2𝒮2(ε)\mathcal{S}_{2}\setminus\mathcal{S}_{2}(\varepsilon) is ε\ll\varepsilon by direct integration of the hyperbolic area measure on that region.

We define KεSL2(𝔸)/SL2()K_{\varepsilon}\subset\mathchoice{\text{\raise 2.15277pt\hbox{$\mathrm{SL}_{2}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathrm{SL}_{2}(\mathbb{Q})$}}}{\mathrm{SL}_{2}(\mathbb{A})\,/\,\mathrm{SL}_{2}(\mathbb{Q})}{\mathrm{SL}_{2}(\mathbb{A})\,/\,\mathrm{SL}_{2}(\mathbb{Q})}{\mathrm{SL}_{2}(\mathbb{A})\,/\,\mathrm{SL}_{2}(\mathbb{Q})} to be the preimage of 𝒮2(ε)\mathcal{S}_{2}(\varepsilon) under the composition

SL2(𝔸)/SL2()SL2()/SL2()PGL2()/PGL2()𝒮2.\displaystyle\mathchoice{\text{\raise 2.15277pt\hbox{$\mathrm{SL}_{2}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathrm{SL}_{2}(\mathbb{Q})$}}}{\mathrm{SL}_{2}(\mathbb{A})\,/\,\mathrm{SL}_{2}(\mathbb{Q})}{\mathrm{SL}_{2}(\mathbb{A})\,/\,\mathrm{SL}_{2}(\mathbb{Q})}{\mathrm{SL}_{2}(\mathbb{A})\,/\,\mathrm{SL}_{2}(\mathbb{Q})}\to\mathchoice{\text{\raise 2.15277pt\hbox{$\mathrm{SL}_{2}(\mathbb{R})$}\big{/}\lower 2.15277pt\hbox{$\mathrm{SL}_{2}(\mathbb{Z})$}}}{\mathrm{SL}_{2}(\mathbb{R})\,/\,\mathrm{SL}_{2}(\mathbb{Z})}{\mathrm{SL}_{2}(\mathbb{R})\,/\,\mathrm{SL}_{2}(\mathbb{Z})}{\mathrm{SL}_{2}(\mathbb{R})\,/\,\mathrm{SL}_{2}(\mathbb{Z})}\to\mathchoice{\text{\raise 2.15277pt\hbox{$\mathrm{PGL}_{2}(\mathbb{R})$}\big{/}\lower 2.15277pt\hbox{$\mathrm{PGL}_{2}(\mathbb{Z})$}}}{\mathrm{PGL}_{2}(\mathbb{R})\,/\,\mathrm{PGL}_{2}(\mathbb{Z})}{\mathrm{PGL}_{2}(\mathbb{R})\,/\,\mathrm{PGL}_{2}(\mathbb{Z})}{\mathrm{PGL}_{2}(\mathbb{R})\,/\,\mathrm{PGL}_{2}(\mathbb{Z})}\to\mathcal{S}_{2}.

By the previous discussion, this is a compact set whose complement has Haar measure ε\ll\varepsilon. For xSL2(𝔸)/SL2()x\in\mathchoice{\text{\raise 2.15277pt\hbox{$\mathrm{SL}_{2}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathrm{SL}_{2}(\mathbb{Q})$}}}{\mathrm{SL}_{2}(\mathbb{A})\,/\,\mathrm{SL}_{2}(\mathbb{Q})}{\mathrm{SL}_{2}(\mathbb{A})\,/\,\mathrm{SL}_{2}(\mathbb{Q})}{\mathrm{SL}_{2}(\mathbb{A})\,/\,\mathrm{SL}_{2}(\mathbb{Q})} we shall call the supremum over all ε>0\varepsilon>0 with xKεx\in K_{\varepsilon} the minimal quadratic value for xx.

The following is a direct corollary of equidistribution in the second factor.

Corollary 4.9.

For any ε(0,1)\varepsilon\in(0,1) there exists i01i_{0}\geq 1 so that the measure of the set of points t𝐓Li(𝔸)/𝐓Li()t\in\mathchoice{\text{\raise 2.15277pt\hbox{$\mathbf{T}_{L_{i}}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathbf{T}_{L_{i}}(\mathbb{Q})$}}}{\mathbf{T}_{L_{i}}(\mathbb{A})\,/\,\mathbf{T}_{L_{i}}(\mathbb{Q})}{\mathbf{T}_{L_{i}}(\mathbb{A})\,/\,\mathbf{T}_{L_{i}}(\mathbb{Q})}{\mathbf{T}_{L_{i}}(\mathbb{A})\,/\,\mathbf{T}_{L_{i}}(\mathbb{Q})} for which g2ψ1,Li(t)SL2()Kεg_{2}\psi_{1,L_{i}}(t)\mathrm{SL}_{2}(\mathbb{Q})\not\in K_{\varepsilon} is ε\ll\varepsilon for all ii0i\geq i_{0}.

4.3.3. Using the shape in the subspace

In the following we identify the minimal quadratic value for the points on the orbits in the context of proving Theorem 4.1. As SpinQ(×^)\mathrm{Spin}_{Q}(\mathbb{R}\times\widehat{\mathbb{Z}}) is a compact open subgroup, it has finitely many orbits on SpinQ(𝔸)/SpinQ()\mathrm{Spin}_{Q}(\mathbb{A})/\mathrm{Spin}_{Q}(\mathbb{Q}) (these correspond to the spin genus of the quadratic form QQ). We choose a finite set of representatives SpinQ(𝔸f)\mathcal{R}\subset\mathrm{Spin}_{Q}(\mathbb{A}_{f}) such that

(4.4) SpinQ(𝔸)/SpinQ()=𝗋SpinQ(×^)𝗋SpinQ().\displaystyle\mathrm{Spin}_{Q}(\mathbb{A})/\mathrm{Spin}_{Q}(\mathbb{Q})=\bigsqcup_{\mathsf{r}\in\mathcal{R}}\mathrm{Spin}_{Q}(\mathbb{R}\times\widehat{\mathbb{Z}})\mathsf{r}\,\mathrm{Spin}_{Q}(\mathbb{Q}).

Note that in SL2\mathrm{SL}_{2} (or SLn2\mathrm{SL}_{n-2}) any gSL2(𝔸)g\in\mathrm{SL}_{2}(\mathbb{A}) can be written as g=bγg=b\gamma where bSL2(×^)b\in\mathrm{SL}_{2}(\mathbb{R}\times\widehat{\mathbb{Z}}) and γSL2()\gamma\in\mathrm{SL}_{2}(\mathbb{Q}).

Lemma 4.10.

Let h𝚫𝐇¯Li(𝔸)h\in\mathbf{\Delta\bar{H}}_{L_{i}}(\mathbb{A}) and write hγ=b𝗋h\gamma=b\mathsf{r} for some γ𝐆¯()\gamma\in\bar{\mathbf{G}}(\mathbb{Q}), b𝐆¯(×^)b\in\bar{\mathbf{G}}(\mathbb{R}\times\widehat{\mathbb{Z}}), and 𝗋\mathsf{r}\in\mathcal{R}. Then γ11.Li\gamma_{1}^{-1}.L_{i} is a rational subspace of discriminant D\asymp D. Furthermore, the minimum

minw2{0}qγ11.Li(w)discQ(γ11.Li)\displaystyle\min_{w\in\mathbb{Z}^{2}\setminus\{0\}}\frac{q_{\gamma_{1}^{-1}.L_{i}}(w)}{\sqrt{\mathrm{disc}_{Q}(\gamma_{1}^{-1}.L_{i})}}

is comparable to the minimal quadratic value for gi,2ψ1,L(h)SL2()g_{i,2}\psi_{1,L}(h)\mathrm{SL}_{2}(\mathbb{Q}).

Note that a lemma in this spirit will later on be used to deduce the main theorems from their dynamical counterparts (cf. Proposition 7.1). The statement here is more technical in nature (as it needs to treat different genera) and the reader is encouraged to return to the proof after reading Proposition 7.1. We note that such a treatment has appeared in different context in the literature [localglobalEV, ALMW-elliptic].

Proof.

The ingredients for this proof are all contained in the proof of Proposition 7.1; so we will be brief. Write L=LiL=L_{i} for simplicity. Note that h1,pγ1=b1,p𝗋ph_{1,p}\gamma_{1}=b_{1,p}\mathsf{r}_{p} and hence

discp,Q(γ11.L)\displaystyle\mathrm{disc}_{p,Q}(\gamma_{1}^{-1}.L) =discp,Q(γ11h1,p1.L)=discp,Q(𝗋p1b1,p.L)\displaystyle=\mathrm{disc}_{p,Q}(\gamma_{1}^{-1}h_{1,p}^{-1}.L)=\mathrm{disc}_{p,Q}(\mathsf{r}_{p}^{-1}b_{1,p}.L)
𝗋discp,Q(b1,p.L)=discp,Q(L).\displaystyle\asymp_{\mathsf{r}}\mathrm{disc}_{p,Q}(b_{1,p}.L)=\mathrm{disc}_{p,Q}(L).

As the discriminant is a product of the local discriminants (1.5), this proves the first claim.

For the second claim, we let L=γ11.LL^{\prime}=\gamma_{1}^{-1}.L and first consider m=gL1ρQ(γ11)gLγ2GLn()m=g_{L^{\prime}}^{-1}\rho_{Q}(\gamma_{1}^{-1})g_{L}\gamma_{2}\in\mathrm{GL}_{n}(\mathbb{Q}). Observe that

mL0=gL1ρQ(γ11)gLL0=gL1ρQ(γ11)L=gL1L=L0.\displaystyle mL_{0}=g_{L^{\prime}}^{-1}\rho_{Q}(\gamma_{1}^{-1})g_{L}L_{0}=g_{L^{\prime}}^{-1}\rho_{Q}(\gamma_{1}^{-1})L=g_{L^{\prime}}^{-1}L^{\prime}=L_{0}.

As we will now see, mm is ’almost integral’ and invertible. For this, compute

ρQ(γ11)gLγ2=ρQ(γ11h1,p1)gLh2,pγ2=ρQ(𝗋p1b1,p1)gLb2,p.\displaystyle\rho_{Q}(\gamma_{1}^{-1})g_{L}\gamma_{2}=\rho_{Q}(\gamma_{1}^{-1}h_{1,p}^{-1})g_{L}h_{2,p}\gamma_{2}=\rho_{Q}(\mathsf{r}_{p}^{-1}b_{1,p}^{-1})g_{L}b_{2,p}.

This implies that there exists some NN\in\mathbb{N} independent of LL such that NρQ(γ11)gLγ2N\rho_{Q}(\gamma_{1}^{-1})g_{L}\gamma_{2} and NρQ(γ11)gL1γ2N\rho_{Q}(\gamma_{1}^{-1})g_{L}^{-1}\gamma_{2} are integral. Recall that disc(Q)gL,disc(Q)gL1\mathrm{disc}(Q)g_{L^{\prime}},\mathrm{disc}(Q)g_{L^{\prime}}^{-1} are integral so that Ndisc(Q)mN\mathrm{disc}(Q)m and Ndisc(Q)m1N\mathrm{disc}(Q)m^{-1} are integral. This discussion implies that for any two positive definite real binary quadratic forms q,qq,q^{\prime} with the property that π1(m)q\pi_{1}(m)q and qq^{\prime} are similar we have

minw2{0}q(w)disc(q)minw2{0}q(w)disc(q).\displaystyle\min_{w\in\mathbb{Z}^{2}\setminus\{0\}}\frac{q(w)}{\sqrt{\mathrm{disc}(q)}}\asymp\min_{w\in\mathbb{Z}^{2}\setminus\{0\}}\frac{q^{\prime}(w)}{\sqrt{\mathrm{disc}(q^{\prime})}}.

Here, recall that GL2()\mathrm{GL}_{2}(\mathbb{R}) acts on binary forms via gq(x)=q(gtx)gq(x)=q(g^{t}x).

Now note

[qL]=[Q(gL)]=[Q(ρQ(γ1)gL)]\displaystyle[q_{L^{\prime}}]=[Q(g_{L}^{\prime}\cdot)]=[Q(\rho_{Q}(\gamma_{1})g_{L^{\prime}}\cdot)]

while the similarity class belonging to g2ψ1,L(t)SL2()g_{2}\psi_{1,L}(t)\mathrm{SL}_{2}(\mathbb{Q}) is

[γ21qL]=[Q(gLγ2)]=[Q(ρQ(γ1)gLm)]\displaystyle[\gamma_{2}^{-1}q_{L}]=[Q(g_{L}\gamma_{2}\cdot)]=[Q(\rho_{Q}(\gamma_{1})g_{L^{\prime}}m\cdot)]

The claim follows. ∎

4.3.4. Proof of Theorem 4.1

As explained, it now suffices to prove that the packets for LiL_{i}

gi,1𝐇Li(𝔸)SpinQ()SpinQ(𝔸)/SpinQ()\displaystyle g_{i,1}\mathbf{H}_{L_{i}}(\mathbb{A})\mathrm{Spin}_{Q}(\mathbb{Q})\subset\mathchoice{\text{\raise 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Q})$}}}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}

equidistribute as discQ(Li)\mathrm{disc}_{Q}(L_{i})\to\infty. Similarly to the situation in the proof of Theorem 3.1, we need to circumvent the problem that 𝐇L\mathbf{H}_{L} for LGrn,2()L\in\mathrm{Gr}_{n,2}(\mathbb{Q}) is not exactly isomorphic to 𝐇Lpt×𝐓L\mathbf{H}^{\mathrm{pt}}_{L}\times\mathbf{T}_{L} – see Remark 2.1 for a more careful discussion. Denote by 𝐇L(𝔸)\mathbf{H}_{L}(\mathbb{A})^{\star} the image of 𝐇Lpt(𝔸)×𝐓L(𝔸)𝐇L(𝔸)\mathbf{H}^{\mathrm{pt}}_{L}(\mathbb{A})\times\mathbf{T}_{L}(\mathbb{A})\to\mathbf{H}_{L}(\mathbb{A}); this is a normal subgroup of 𝐇L(𝔸)\mathbf{H}_{L}(\mathbb{A}) with the property that KL:=𝐇L(𝔸)/𝐇L(𝔸)K_{L}\mathrel{\mathop{\mathchar 58\relax}}=\mathchoice{\text{\raise 2.15277pt\hbox{$\mathbf{H}_{L}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathbf{H}_{L}(\mathbb{A})^{\star}$}}}{\mathbf{H}_{L}(\mathbb{A})\,/\,\mathbf{H}_{L}(\mathbb{A})^{\star}}{\mathbf{H}_{L}(\mathbb{A})\,/\,\mathbf{H}_{L}(\mathbb{A})^{\star}}{\mathbf{H}_{L}(\mathbb{A})\,/\,\mathbf{H}_{L}(\mathbb{A})^{\star}} is compact and abelian. By an argument as in the beginning of the proof of Theorem 3.1 it suffices to show that for any kiKLik_{i}\in K_{L_{i}} the orbits

gi,1ki𝐇Li(𝔸)SpinQ()SpinQ(𝔸)/SpinQ()\displaystyle g_{i,1}k_{i}\mathbf{H}_{L_{i}}(\mathbb{A})^{\star}\mathrm{Spin}_{Q}(\mathbb{Q})\subset\mathchoice{\text{\raise 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Q})$}}}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}

are equidistributed as ii\to\infty. We let μi\mu_{i} be the Haar measure on the ii-th such orbit and let

μi,1=m𝐇Lipt(𝔸)SpinQ(),μi,2=m𝐓Li(𝔸)SpinQ()\displaystyle\mu_{i,1}=m_{\mathbf{H}^{\mathrm{pt}}_{L_{i}}(\mathbb{A})\mathrm{Spin}_{Q}(\mathbb{Q})},\quad\mu_{i,2}=m_{\mathbf{T}_{L_{i}}(\mathbb{A})\mathrm{Spin}_{Q}(\mathbb{Q})}

be the Haar measure on the closed orbits of 𝐇Lipt(𝔸)\mathbf{H}^{\mathrm{pt}}_{L_{i}}(\mathbb{A}) resp. 𝐓Li(𝔸)\mathbf{T}_{L_{i}}(\mathbb{A}). Then we have for any function fCc(SpinQ(𝔸)/SpinQ())f\in C_{c}(\mathchoice{\text{\raise 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Q})$}}}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})})

(4.5) fdμi=f(gi,1kiht)dμ1,i(h)dμi,2(t).\displaystyle\int f\,\mathrm{d}\mu_{i}=\int\int f(g_{i,1}k_{i}ht)\,\mathrm{d}\mu_{1,i}(h)\,\mathrm{d}\mu_{i,2}(t).

In the following, we identify kik_{i} with a representative in a fixed bounded region of 𝐇Li(𝔸)\mathbf{H}_{L_{i}}(\mathbb{A}).

For a fixed ti𝐓Li(𝔸)t_{i}\in\mathbf{T}_{L_{i}}(\mathbb{A}), the inner integral is the integral over the orbit

gi,1ki𝐇Lipt(𝔸)tiSpinQ()=gi,1kiti𝐇Lipt(𝔸)SpinQ().\displaystyle g_{i,1}k_{i}\mathbf{H}^{\mathrm{pt}}_{L_{i}}(\mathbb{A})t_{i}\mathrm{Spin}_{Q}(\mathbb{Q})=g_{i,1}k_{i}t_{i}\mathbf{H}^{\mathrm{pt}}_{L_{i}}(\mathbb{A})\mathrm{Spin}_{Q}(\mathbb{Q}).

Writing tiγi=bi𝗋t_{i}\gamma_{i}=b_{i}\mathsf{r} as in (4.4), we see that

gi,1kiti𝐇Lipt(𝔸)SpinQ()\displaystyle g_{i,1}k_{i}t_{i}\mathbf{H}^{\mathrm{pt}}_{L_{i}}(\mathbb{A})\mathrm{Spin}_{Q}(\mathbb{Q}) =gi,1kibi𝗋γi1𝐇Lipt(𝔸)SpinQ()\displaystyle=g_{i,1}k_{i}b_{i}\mathsf{r}\gamma_{i}^{-1}\mathbf{H}^{\mathrm{pt}}_{L_{i}}(\mathbb{A})\mathrm{Spin}_{Q}(\mathbb{Q})
=gi,1kibi𝗋𝐇γi1.Lipt(𝔸)SpinQ()\displaystyle=g_{i,1}k_{i}b_{i}\mathsf{r}\mathbf{H}^{\mathrm{pt}}_{\gamma_{i}^{-1}.L_{i}}(\mathbb{A})\mathrm{Spin}_{Q}(\mathbb{Q})

which is equidistributed if and only if 𝐇γi1.Lipt(𝔸)SpinQ()\mathbf{H}^{\mathrm{pt}}_{\gamma_{i}^{-1}.L_{i}}(\mathbb{A})\mathrm{Spin}_{Q}(\mathbb{Q}) is equidistributed (as gi,1kibig_{i,1}k_{i}b_{i} is bounded). By Proposition 4.7 and its corollary it suffices to show that the minimal non-zero value of qγi1.Liq_{\gamma_{i}^{-1}.L_{i}} goes to infinity. This minimum is comparable to the minimal quadratic value for gi,2ψ1,L(ti)SL2()g_{i,2}\psi_{1,L}(t_{i})\mathrm{SL}_{2}(\mathbb{Q}) by Lemma 4.10.

Motivated by this observation, we define for ε>0\varepsilon>0

i(ε)={t𝐓Li():gi,2ψ1,L(t)SL2()Kε}𝐓Li(𝔸)/𝐓Li()\displaystyle\mathcal{B}_{i}(\varepsilon)=\{t\mathbf{T}_{L_{i}}(\mathbb{Q})\mathrel{\mathop{\mathchar 58\relax}}g_{i,2}\psi_{1,L}(t)\mathrm{SL}_{2}(\mathbb{Q})\in K_{\varepsilon}\}\subset\mathchoice{\text{\raise 2.15277pt\hbox{$\mathbf{T}_{L_{i}}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathbf{T}_{L_{i}}(\mathbb{Q})$}}}{\mathbf{T}_{L_{i}}(\mathbb{A})\,/\,\mathbf{T}_{L_{i}}(\mathbb{Q})}{\mathbf{T}_{L_{i}}(\mathbb{A})\,/\,\mathbf{T}_{L_{i}}(\mathbb{Q})}{\mathbf{T}_{L_{i}}(\mathbb{A})\,/\,\mathbf{T}_{L_{i}}(\mathbb{Q})}

so that the complement of i(ε)\mathcal{B}_{i}(\varepsilon) has μi,2\mu_{i,2}-measure ε\ll\varepsilon for all ii large enough (depending on ε\varepsilon) by Corollary 4.9. In view of (4.5), this implies that

fdμi=1μi,2(i(ε))i(ε)f(gi,1kiht)dμ1,i(h)dμi,2(t)+𝒪(ε)\displaystyle\int f\,\mathrm{d}\mu_{i}=\tfrac{1}{\mu_{i,2}(\mathcal{B}_{i}(\varepsilon))}\int_{\mathcal{B}_{i}(\varepsilon)}\int f(g_{i,1}k_{i}ht)\,\mathrm{d}\mu_{1,i}(h)\,\mathrm{d}\mu_{i,2}(t)+\mathcal{O}(\varepsilon)

By the previous paragraph, the orbits gi,1ki𝐇Lipt(𝔸)tiSpinQ()g_{i,1}k_{i}\mathbf{H}^{\mathrm{pt}}_{L_{i}}(\mathbb{A})t_{i}\mathrm{Spin}_{Q}(\mathbb{Q}) are equidistributed for any sequence tii(ε)t_{i}\in\mathcal{B}_{i}(\varepsilon). The integral on the right-hand side is a convex combination of such orbital integrals and hence must converge to the integral of ff over the Haar measure. Letting μ\mu be any weak-limit of the measures μi\mu_{i} we obtain

fdμ=fdmSpinQ(𝔸)/SpinQ()+𝒪(ε).\displaystyle\int f\,\mathrm{d}\mu=\int f\,\mathrm{d}m_{\mathchoice{\text{\raise 1.50694pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{A})$}\big{/}\lower 1.50694pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Q})$}}}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}{\mathrm{Spin}_{Q}(\mathbb{A})\,/\,\mathrm{Spin}_{Q}(\mathbb{Q})}}+\mathcal{O}(\varepsilon).

As ε\varepsilon is arbitrary, this implies the claim.

Part II

From equidistribution of orbits to the main theorems


\@afterheading

For the contents of this part we refer the reader to the overview of this article in §1.3.

5. Discriminants and glue groups

Recall that QQ is a positive definite integral quadratic form on n\mathbb{Q}^{n} and ,Q\langle\cdot,\cdot\rangle_{Q} is its symmetric bilinear form. By integrality we mean that ,Q\langle\cdot,\cdot\rangle_{Q} takes integer values on n×n\mathbb{Z}^{n}\times\mathbb{Z}^{n}. The goal of this section is to prove the following proposition:

Proposition 5.1.

For any subspace LnL\subset\mathbb{Q}^{n} there exist two positive divisors m1,m2m_{1},m_{2} of disc(Q)\mathrm{disc}(Q) with

discQ(L)=m1m2discQ(L).\displaystyle\mathrm{disc}_{Q}(L^{\perp})=\frac{m_{1}}{m_{2}}\mathrm{disc}_{Q}(L).

In particular,

1disc(Q)discQ(L)discQ(L)disc(Q)discQ(L).\frac{1}{\mathrm{disc}(Q)}\mathrm{disc}_{Q}(L)\leq\mathrm{disc}_{Q}(L^{\perp})\leq\mathrm{disc}(Q)\mathrm{disc}_{Q}(L).

To that end, we will use the notion of glue groups defined in §5.1 and, in fact, prove a significantly finer statement in Proposition 5.4 below.

5.1. Definitions

For any \mathbb{Z}-lattice Γn\Gamma\subset\mathbb{Q}^{n} we define the dual lattice

Γ#={xΓ:x,yQ for all yΓ}.\displaystyle\Gamma^{\#}=\{x\in\Gamma\otimes\mathbb{Q}\mathrel{\mathop{\mathchar 58\relax}}\langle x,y\rangle_{Q}\in\mathbb{Z}\text{ for all }y\in\Gamma\}.

If Γn\Gamma\subset\mathbb{Z}^{n} (or more generally if ,Q\langle\cdot,\cdot\rangle_{Q} takes integral values on Γ×Γ\Gamma\times\Gamma), the dual lattice Γ#\Gamma^{\#} contains Γ\Gamma. Note that if Γ1Γ2\Gamma_{1}\subset\Gamma_{2} are any two \mathbb{Z}-lattices then Γ1#Γ2#\Gamma_{1}^{\#}\supset\Gamma_{2}^{\#}.

For the purposes of this section, a very useful classical tool is the so-called glue-group, which one could see as a concept generalizing the discriminant. We introduce only what is needed here; for better context we refer to [conwaysloane, McMullenglue] (in particular, we do not introduce the fractional form). We define the glue group of a rational subspace LL (or of the lattice L()L(\mathbb{Z})) as

𝒢(L)=L()#/L().\displaystyle\mathcal{G}(L)=L(\mathbb{Z})^{\#}/L(\mathbb{Z}).

Note that L()#L(\mathbb{Z})^{\#} contains L()L(\mathbb{Z}) by integrality. The glue-group is a finite abelian group whose cardinality is exactly the discriminant (see e.g. [kitaoka, §5.1]). We remark that the glue group may be computed from local data – this is made explicit in §B.1 of the appendix.

Remark 5.2.

For each discriminant DD, one may consider the collection of subspaces LGrn,k()L\in\mathrm{Gr}_{n,k}(\mathbb{Q}) with discriminant DD and glue group a fixed abelian group of order DD. In principle, the results of the current article should carry over to prove equidistribution of these subspaces together with their shapes (cf. [2in4]). However, it is not clear when one expects such collections to be non-empty, even when QQ is the sum of squares.

5.2. The glue group of the orthogonal complement

We study the relation between the glue group of a subspace and that of its orthogonal complement. Any subspace LnL\subset\mathbb{Q}^{n} contains various lattices which are (potentially) of interest and are natural:

  • the intersections L()=L()nL(\mathbb{Z})=L(\mathbb{Q})\cap\mathbb{Z}^{n} and L()(n)#L(\mathbb{Q})\cap(\mathbb{Z}^{n})^{\#},

  • the dual lattice L()#L(\mathbb{Z})^{\#}, and

  • the projection lattices πL(n)\pi_{L}(\mathbb{Z}^{n}) and πL((n)#)\pi_{L}((\mathbb{Z}^{n})^{\#}) where πL:nL\pi_{L}\mathrel{\mathop{\mathchar 58\relax}}\mathbb{Q}^{n}\to L denotes the orthogonal projection.

Lemma 5.3 (Elementary properties).

The following relations between the aforementioned lattices hold:

  1. (i)

    L()#=πL((n)#)L(\mathbb{Z})^{\#}=\pi_{L}((\mathbb{Z}^{n})^{\#}) and (L(n)#)#=πL(n)(L\cap(\mathbb{Z}^{n})^{\#})^{\#}=\pi_{L}(\mathbb{Z}^{n}).

  2. (ii)

    (L(n)#)/L()L()#/πL(n)(L\cap(\mathbb{Z}^{n})^{\#})/L(\mathbb{Z})\simeq L(\mathbb{Z})^{\#}/\pi_{L}(\mathbb{Z}^{n}).

Proof.

We prove i first. Since the proofs of the two assertions in i are similar, we only detail the first. Let v1,,vkv_{1},\ldots,v_{k} be a \mathbb{Z}-basis of L()L(\mathbb{Z}). Moreover, let w1,,wkLw_{1},\ldots,w_{k}\in L be the dual basis to v1,,vkv_{1},\ldots,v_{k}. Extend v1,,vkv_{1},\ldots,v_{k} to a basis v1,,vnv_{1},\ldots,v_{n} of n\mathbb{Z}^{n} and consider y1,,yny_{1},\ldots,y_{n} the dual basis to v1,,vnv_{1},\ldots,v_{n}. Then πL(yi)=wi\pi_{L}(y_{i})=w_{i} for any iki\leq k as

πL(yi),vjQ=yi,vjQ=δij\displaystyle\langle\pi_{L}(y_{i}),v_{j}\rangle_{Q}=\langle y_{i},v_{j}\rangle_{Q}=\delta_{ij}

whenever jkj\leq k. Moreover, yiLy_{i}\in L^{\perp} for i>ki>k by construction. Thus,

πL((n)#)=πL(span(y1,,yn))=span(w1,,wk)=L()#\pi_{L}((\mathbb{Z}^{n})^{\#})=\pi_{L}(\mathrm{span}_{\mathbb{Z}}(y_{1},\ldots,y_{n}))=\mathrm{span}_{\mathbb{Z}}(w_{1},\ldots,w_{k})=L(\mathbb{Z})^{\#}

as claimed. The proof of the second equality is analogous.

For ii, note that for any two lattices Λ1Λ2\Lambda_{1}\subset\Lambda_{2} in LL one has

(5.1) Λ2/Λ1Λ1#/Λ2#,\displaystyle\Lambda_{2}/\Lambda_{1}\simeq\Lambda_{1}^{\#}/\Lambda_{2}^{\#},

so ii follows from i. To construct such an isomorphism one proceeds as follows. Fix a basis v1,,vnv_{1},\ldots,v_{n} of Λ2\Lambda_{2} such that d1v1,,dnvnd_{1}v_{1},\ldots,d_{n}v_{n} is a basis888Such a basis is sometimes called ’adapted basis’ (in geometry of numbers). The existence can be easily seen using Smith’s normal form. of Λ1\Lambda_{1} with did_{i}\in\mathbb{Z} and let w1,,wnw_{1},\ldots,w_{n} be the dual basis to v1,,vnv_{1},\ldots,v_{n}. Then, the map

f:Λ2Λ1#,vidi1wif\mathrel{\mathop{\mathchar 58\relax}}\Lambda_{2}\rightarrow\Lambda_{1}^{\#},\ v_{i}\mapsto d_{i}^{-1}w_{i}

induces the desired isomorphism. ∎

Proposition 5.4.

We have an isomorphism

πL(n)/L()πL(n)/L().\pi_{L}(\mathbb{Z}^{n})/L(\mathbb{Z})\rightarrow\pi_{L^{\perp}}(\mathbb{Z}^{n})/L^{\perp}(\mathbb{Z}).

When QQ is unimodular, i.e. disc(Q)=1\mathrm{disc}(Q)=1, this together with Lemma 5.3 shows that the glue-groups of LL and LL^{\perp} are isomorphic. Indeed, in this case (n)#=n(\mathbb{Z}^{n})^{\#}=\mathbb{Z}^{n} and hence πL(n)=L()#\pi_{L}(\mathbb{Z}^{n})=L(\mathbb{Z})^{\#}. In particular, LL and LL^{\perp} have the same discriminant. When QQ is not unimodular, the proposition gives an isomorphism between subgroups of the respective glue-groups.

Proof.

We define a map ff from πL(n)\pi_{L}(\mathbb{Z}^{n}) to πL(n)/L()\pi_{L^{\perp}}(\mathbb{Z}^{n})/L^{\perp}(\mathbb{Z}) as follows. For xπL(n)x\in\pi_{L}(\mathbb{Z}^{n}) choose a lift x^n\hat{x}\in\mathbb{Z}^{n} of xx for the projection πL\pi_{L} and define

f(x)=πL(x^)+L().\displaystyle f(x)=\pi_{L^{\perp}}(\hat{x})+L^{\perp}(\mathbb{Z}).

Note that ff is well-defined since if x^,y^n\hat{x},\hat{y}\in\mathbb{Z}^{n} are two lifts of xπL(n)x\in\pi_{L}(\mathbb{Z}^{n}), then x^y^L()\hat{x}-\hat{y}\in L^{\perp}(\mathbb{Z}) which implies πL(x^)+L()=πL(y^)+L()\pi_{L^{\perp}}(\hat{x})+L^{\perp}(\mathbb{Z})=\pi_{L^{\perp}}(\hat{y})+L^{\perp}(\mathbb{Z}).

We show that ker(f)=L()\ker(f)=L(\mathbb{Z}). Obviously, L()ker(f)L(\mathbb{Z})\subset\ker(f) since for any xL()x\in L(\mathbb{Z}) we can choose xx itself as lift. On the other hand, if xker(f)x\in\ker(f) there is a lift x^n\hat{x}\in\mathbb{Z}^{n} of xx for πL\pi_{L} such that πL(x^)L()\pi_{L^{\perp}}(\hat{x})\in L^{\perp}(\mathbb{Z}). In particular,

x=πL(x^)=πL(x^)πL(πL(x^))=πL(x^πL(x^))=x^πL(x^)L().x=\pi_{L}(\hat{x})=\pi_{L}(\hat{x})-\pi_{L}(\pi_{L^{\perp}}(\hat{x}))=\pi_{L}(\hat{x}-\pi_{L^{\perp}}(\hat{x}))=\hat{x}-\pi_{L^{\perp}}(\hat{x})\in L(\mathbb{Z}).

We deduce that ker(f)L()\ker(f)\subset L(\mathbb{Z}) and hence equality. This proves the proposition. ∎

Proof of Proposition 5.1.

By Proposition 5.4

discQ(L)=|𝒢(L)|\displaystyle\mathrm{disc}_{Q}(L)=|\mathcal{G}(L)| =|L()#/πL(n)||πL(n)/L()|\displaystyle=|L(\mathbb{Z})^{\#}/\pi_{L}(\mathbb{Z}^{n})|\cdot|\pi_{L}(\mathbb{Z}^{n})/L(\mathbb{Z})|
=|L()#/πL(n)||πL(n)/L()|\displaystyle=|L(\mathbb{Z})^{\#}/\pi_{L}(\mathbb{Z}^{n})|\cdot|\pi_{L^{\perp}}(\mathbb{Z}^{n})/L^{\perp}(\mathbb{Z})|
=|L()#/πL(n)||L()#/πL(n)||𝒢(L)|.\displaystyle=\frac{|L(\mathbb{Z})^{\#}/\pi_{L}(\mathbb{Z}^{n})|}{|L^{\perp}(\mathbb{Z})^{\#}/\pi_{L^{\perp}}(\mathbb{Z}^{n})|}|\mathcal{G}(L^{\perp})|.

Using Lemma 5.3 note that the finite group L()#/πL(n)=πL((n)#)/πL(n)L(\mathbb{Z})^{\#}/\pi_{L}(\mathbb{Z}^{n})=\pi_{L}((\mathbb{Z}^{n})^{\#})/\pi_{L}(\mathbb{Z}^{n}) is a quotient of (n)#/n(\mathbb{Z}^{n})^{\#}/\mathbb{Z}^{n} and hence |L()#/πL(n)||L(\mathbb{Z})^{\#}/\pi_{L}(\mathbb{Z}^{n})| is a divisor of disc(Q)=|(n)#/n|\mathrm{disc}(Q)=|(\mathbb{Z}^{n})^{\#}/\mathbb{Z}^{n}|. As the analogous statement holds for LL^{\perp}, the proposition follows. ∎

Remark 5.5.

When disc(Q)=1\mathrm{disc}(Q)=1, Proposition 5.4 states that 𝒢(L)𝒢(L)\mathcal{G}(L)\simeq\mathcal{G}(L^{\perp}). Apart from the discriminants of LL and LL^{\perp} being the same, this includes information about the local coefficients of the quadratic forms on LL and LL^{\perp}. This is exploited e.g. in Proposition B.6. When k=nkk=n-k, one can ask whether this implies that Q|L()Q|_{L(\mathbb{Z})} and Q|L()Q|_{L^{\perp}(\mathbb{Z})} are in the same genus.

6. Moduli spaces

In this section we study the moduli space 𝒴\mathcal{Y} of basis extensions which was introduced in §1.1 consisting of (certain) homothety classes [L,Λ][L,\Lambda] where LL is a kk-dimensional subspace, Λ\Lambda is a full rank lattice in n\mathbb{R}^{n} and LΛL\cap\Lambda is a lattice in LL. We also discuss a slight refinement of Theorem 1.11 (Theorem 6.9 below) and see how it implies Theorem 1.4.

6.1. Oriented subspaces

For the purposes of proving the main theorems from their dynamical analogues, it is convenient to work with subspaces with an orientation. In fact, the main theorems may be refined to include orientation.

Oriented subspaces of dimension kk form an affine variety Grn,k+\mathrm{Gr}^{+}_{n,k} (defined over \mathbb{Q}) with a morphism (of algebraic varieties) Grn,k+Grn,k\mathrm{Gr}^{+}_{n,k}\to\mathrm{Gr}_{n,k} where the preimage of any point consists of two points corresponding to two choices of orientation.

Remark 6.1.

To construct Grn,k+\mathrm{Gr}^{+}_{n,k} explicitly, observe that the positive definite form QQ induces a rational form discQ\mathrm{disc}_{Q} on the exterior product kn\bigwedge^{k}\mathbb{Q}^{n} via

discQ(v1vk)=det(v1,v1Qv1,vkQvk,v1Qvk,vkQ).\displaystyle\mathrm{disc}_{Q}(v_{1}\wedge\ldots\wedge v_{k})=\det\begin{pmatrix}\langle v_{1},v_{1}\rangle_{Q}&\cdots&\langle v_{1},v_{k}\rangle_{Q}\\ \vdots&&\vdots\\ \langle v_{k},v_{1}\rangle_{Q}&\cdots&\langle v_{k},v_{k}\rangle_{Q}\end{pmatrix}.

Note that this merely extends the previous definition of discriminant. The variety Grn,k+\mathrm{Gr}^{+}_{n,k} is then as the subvariety of the variety of pure wedges 𝒫\mathcal{P} satisfying the additional equation discQ(v1vk)=1\mathrm{disc}_{Q}(v_{1}\wedge\ldots\wedge v_{k})=1. Note that rational subspaces with an orientation do not correspond to rational points of Grn,k+\mathrm{Gr}^{+}_{n,k} but rather to primitive integer points of the variety of pure wedges 𝒫\mathcal{P}. In that sense, it is often more natural to work with 𝒫\mathcal{P} instead of Grn,k+\mathrm{Gr}^{+}_{n,k}.

The orthogonal group SOQ\mathrm{SO}_{Q} (and hence also SpinQ\mathrm{Spin}_{Q}) acts on oriented subspaces. For an oriented rational subspace LL the stabilizer group in SpinQ\mathrm{Spin}_{Q} under this action is exactly equal to the stabilizer group 𝐇L\mathbf{H}_{L} defined in §2.1. Moreover, the action of SpinQ()\mathrm{Spin}_{Q}(\mathbb{R}) on Grn,k+()\mathrm{Gr}^{+}_{n,k}(\mathbb{R}) is transitive (as the action of SOQ()\mathrm{SO}_{Q}(\mathbb{R}) is).

Remark 6.2 (Orientation on the orthogonal complement).

For any oriented kk-dimensional subspace LL the orthogonal complement inherits an orientation: if v1,,vkv_{1},\ldots,v_{k} is an oriented basis of LL then a basis vk+1,,vnv_{k+1},\ldots,v_{n} of LL^{\perp} is oriented if det(v1,,vn)>0\det(v_{1},\ldots,v_{n})>0. The orthogonal complement yields an isomorphism Grn,k+Grn,nk+\mathrm{Gr}^{+}_{n,k}\to\mathrm{Gr}^{+}_{n,n-k} which is explicitly realizable in Plücker coordinates at least when disc(Q)=1\mathrm{disc}(Q)=1 [Schmidt-heights, §1].

6.2. Quotients of homogeneous spaces

6.2.1. The moduli space of oriented basis extensions

We extend the definition of the moduli space of basis extensions to include orientation. Consider the pairs (L,Λ)(L,\Lambda) where LL is an oriented subspace, Λn\Lambda\subset\mathbb{R}^{n} is a full rank lattice, and LΛL\cap\Lambda is a lattice in LL. Two such pairs (L,Λ),(L,Λ)(L,\Lambda),(L^{\prime},\Lambda^{\prime}) are equivalent if L=LL=L^{\prime} (including orientation) and if there exists gGLn()g\in\mathrm{GL}_{n}(\mathbb{R}) which acts by positive scalar multiplication of LL and LL^{\perp} such that gΛ=Λg\Lambda=\Lambda^{\prime}. The moduli space of oriented basis extensions 𝒴+\mathcal{Y}^{+} is defined to be the set of such equivalence classes [L,Λ][L,\Lambda]. There exists a natural map 𝒴+𝒴\mathcal{Y}^{+}\to\mathcal{Y} (simply by forgetting orientation).

We begin by realizing 𝒴+\mathcal{Y}^{+} as a double quotient of a Lie group. We use the following notation:

  • The groups 𝐏n,k\mathbf{P}_{n,k} and 𝐆\mathbf{G} as defined in §1.4.4:

    𝐏n,k\displaystyle\mathbf{P}_{n,k} ={(AB0D)SLn:det(A)=det(D)=1},\displaystyle=\Big{\{}\begin{pmatrix}A&B\\ 0&D\end{pmatrix}\in\mathrm{SL}_{n}\mathrel{\mathop{\mathchar 58\relax}}\det(A)=\det(D)=1\Big{\}},
    𝐆\displaystyle\mathbf{G} =SpinQ×𝐏n,k.\displaystyle=\mathrm{Spin}_{Q}\times\mathbf{P}_{n,k}.
  • The reference subspace L0L_{0} spanned by the first kk standard basis vectors (1.6) as well as the ’standardization’ ηQ\eta_{Q} defined in (1.3). Note that L0L_{0} is oriented using the standard basis.

  • For any oriented subspace LnL\subset\mathbb{Q}^{n} we let 𝐇L<SpinQ\mathbf{H}_{L}<\mathrm{Spin}_{Q} be the stabilizer group of LL.

  • The subgroup 𝐇L0<SpinQ\mathbf{H}_{L_{0}}<\mathrm{Spin}_{Q} maps to a subgroup of 𝐏n,k\mathbf{P}_{n,k} under the (spin) isogeny ρQ\rho_{Q}; we again denote by 𝚫𝐇L0<𝐆\mathbf{\Delta H}_{L_{0}}<\mathbf{G} the diagonally embedded group (this agrees with the definition in §2.3 with the choice of the standard basis).

Lemma 6.3.

There is an identification

𝒴+𝚫𝐇L0()\𝐆()/ ​​𝐏n,k().\displaystyle\mathcal{Y}^{+}\simeq\mathchoice{\text{\lower 2.15277pt\hbox{$\mathbf{\Delta H}_{L_{0}}(\mathbb{R})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathbf{G}(\mathbb{R})$\!}\big{/} \lower 2.15277pt\hbox{\!\!$\mathbf{P}_{n,k}(\mathbb{Z})$}}}{\mathbf{\Delta H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathbf{G}(\mathbb{R})\,/\,\mathbf{P}_{n,k}(\mathbb{Z})}{\mathbf{\Delta H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathbf{G}(\mathbb{R})\,/\,\mathbf{P}_{n,k}(\mathbb{Z})}{\mathbf{\Delta H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathbf{G}(\mathbb{R})\,/\,\mathbf{P}_{n,k}(\mathbb{Z})}.

By Lemma 6.3, we may pull back the Haar quotient probability measure on the right-hand side to 𝒴+\mathcal{Y}^{+} (and by push-forward on 𝒴\mathcal{Y}).

Proof.

The above identification runs as follows. If (g1,g2)𝐆()(g_{1},g_{2})\in\mathbf{G}(\mathbb{R}) is given, we set L=ρQ(g11)g2L0()=g11.L0()L=\rho_{Q}(g_{1}^{-1})g_{2}L_{0}(\mathbb{R})=g_{1}^{-1}.L_{0}(\mathbb{R}) and Λ=ρQ(g11)g2n\Lambda=\rho_{Q}(g_{1}^{-1})g_{2}\mathbb{Z}^{n}. Clearly, Λ\Lambda intersects LL in the lattice ρQ(g11)g2L0()\rho_{Q}(g_{1}^{-1})g_{2}L_{0}(\mathbb{Z}). As any element of 𝐏n,k()\mathbf{P}_{n,k}(\mathbb{Z}) stabilizes L0()L_{0}(\mathbb{R}) and n\mathbb{Z}^{n} and as 𝚫𝐇L0()\mathbf{\Delta H}_{L_{0}}(\mathbb{R}) is diagonally embedded, we obtain a well-defined map

𝚫𝐇L0()\𝐆()/ ​​𝐏n,k()𝒴.\mathchoice{\text{\lower 2.15277pt\hbox{$\mathbf{\Delta H}_{L_{0}}(\mathbb{R})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathbf{G}(\mathbb{R})$\!}\big{/} \lower 2.15277pt\hbox{\!\!$\mathbf{P}_{n,k}(\mathbb{Z})$}}}{\mathbf{\Delta H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathbf{G}(\mathbb{R})\,/\,\mathbf{P}_{n,k}(\mathbb{Z})}{\mathbf{\Delta H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathbf{G}(\mathbb{R})\,/\,\mathbf{P}_{n,k}(\mathbb{Z})}{\mathbf{\Delta H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathbf{G}(\mathbb{R})\,/\,\mathbf{P}_{n,k}(\mathbb{Z})}\rightarrow\mathcal{Y}.

The injectivity of this map is clear from the definition of 𝚫𝐇L0()\mathbf{\Delta H}_{L_{0}}(\mathbb{R}), so let us argue for the surjectivity.

Let [L,Λ]𝒴[L,\Lambda]\in\mathcal{Y}. By choosing the representative correctly, we may assume that Λ\Lambda as well as LΛL\cap\Lambda are unimodular. Choose g1SpinQ()g_{1}\in\mathrm{Spin}_{Q}(\mathbb{R}) such that g1.L=L0g_{1}.L=L_{0}. Then L0()L_{0}(\mathbb{R}) is g1.Λg_{1}.\Lambda-rational. Pick a basis v1,,vkv_{1},\ldots,v_{k} of g1.ΛL0()g_{1}.\Lambda\cap L_{0}(\mathbb{R}) and complete it into a basis v1,,vnv_{1},\ldots,v_{n} of g1.Λg_{1}.\Lambda. Set

g2=(v1vn){gSLn():gL0()=L0()}.\displaystyle g_{2}=(v_{1}\mid\ldots\mid v_{n})\in\{g\in\mathrm{SL}_{n}(\mathbb{R})\mathrel{\mathop{\mathchar 58\relax}}gL_{0}(\mathbb{R})=L_{0}(\mathbb{R})\}.

As g1.ΛL0()g_{1}.\Lambda\cap L_{0}(\mathbb{R}) is unimodular, we have that g2𝐏n,k()g_{2}\in\mathbf{P}_{n,k}(\mathbb{R}). Under these choices we have ρQ(g11)g2L0()=L\rho_{Q}(g_{1}^{-1})g_{2}L_{0}(\mathbb{R})=L and ρQ(g11)g2n=Λ\rho_{Q}(g_{1}^{-1})g_{2}\mathbb{Z}^{n}=\Lambda; surjectivity follows. ∎

Remark 6.4 (Action of SpinQ()\mathrm{Spin}_{Q}(\mathbb{Z})).

Note that SpinQ()\mathrm{Spin}_{Q}(\mathbb{Z}) acts on 𝒴+\mathcal{Y}^{+} via g[L,Λ]=[g.L,g.Λ]g[L,\Lambda]=[g.L,g.\Lambda]. In view of the identification in Lemma 6.3 (and its proof) this action of SpinQ()\mathrm{Spin}_{Q}(\mathbb{Z}) corresponds to the SpinQ()\mathrm{Spin}_{Q}(\mathbb{Z})-action from the right on the double quotient 𝚫𝐇L0()\𝐆()/𝐏n,k()\mathchoice{\text{\lower 2.15277pt\hbox{$\mathbf{\Delta H}_{L_{0}}(\mathbb{R})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathbf{G}(\mathbb{R})$\!}\big{/} \lower 2.15277pt\hbox{\!\!$\mathbf{P}_{n,k}(\mathbb{Z})$}}}{\mathbf{\Delta H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathbf{G}(\mathbb{R})\,/\,\mathbf{P}_{n,k}(\mathbb{Z})}{\mathbf{\Delta H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathbf{G}(\mathbb{R})\,/\,\mathbf{P}_{n,k}(\mathbb{Z})}{\mathbf{\Delta H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathbf{G}(\mathbb{R})\,/\,\mathbf{P}_{n,k}(\mathbb{Z})}. In particular,

SpinQ()\𝒴+𝚫𝐇L0()\𝐆()/ ​​𝐆().\displaystyle\mathchoice{\text{\lower 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Z})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{Y}^{+}$}}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{Y}^{+}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{Y}^{+}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{Y}^{+}}\simeq\mathchoice{\text{\lower 2.15277pt\hbox{$\mathbf{\Delta H}_{L_{0}}(\mathbb{R})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathbf{G}(\mathbb{R})$\!}\big{/} \lower 2.15277pt\hbox{\!\!$\mathbf{G}(\mathbb{Z})$}}}{\mathbf{\Delta H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathbf{G}(\mathbb{R})\,/\,\mathbf{G}(\mathbb{Z})}{\mathbf{\Delta H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathbf{G}(\mathbb{R})\,/\,\mathbf{G}(\mathbb{Z})}{\mathbf{\Delta H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathbf{G}(\mathbb{R})\,/\,\mathbf{G}(\mathbb{Z})}.

Recall from the introduction that 𝒮k\mathcal{S}_{k} is the space of positive definite real quadratic forms in kk variables up to similarity. Here, we say that two forms q,qq,q^{\prime} in kk-variables are equivalent if there is gGLk()g\in\mathrm{GL}_{k}(\mathbb{Z}) such that gq=qgq=q^{\prime} and similar if qq is equivalent to a multiple of qq^{\prime}. We may identify 𝒮k\mathcal{S}_{k} with

(6.1) Ok()\PGLk()/ ​​PGLk().\displaystyle\mathchoice{\text{\lower 2.15277pt\hbox{$\mathrm{O}_{k}(\mathbb{R})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathrm{PGL}_{k}(\mathbb{R})$\!}\big{/} \lower 2.15277pt\hbox{\!\!$\mathrm{PGL}_{k}(\mathbb{Z})$}}}{\mathrm{O}_{k}(\mathbb{R})\,\backslash\,\mathrm{PGL}_{k}(\mathbb{R})\,/\,\mathrm{PGL}_{k}(\mathbb{Z})}{\mathrm{O}_{k}(\mathbb{R})\,\backslash\,\mathrm{PGL}_{k}(\mathbb{R})\,/\,\mathrm{PGL}_{k}(\mathbb{Z})}{\mathrm{O}_{k}(\mathbb{R})\,\backslash\,\mathrm{PGL}_{k}(\mathbb{R})\,/\,\mathrm{PGL}_{k}(\mathbb{Z})}.

Indeed, to any point Ok()gPGLk()\mathrm{O}_{k}(\mathbb{R})g\mathrm{PGL}_{k}(\mathbb{Z}) one associates the similarity class of the form represented by gtgg^{t}g. Conversely, given the similarity class of a form qq and a matrix representation MM of qq one can write M=gtgM=g^{t}g for some gGLk()g\in\mathrm{GL}_{k}(\mathbb{R}). Another way to view the quotient in (6.1) is as the space of lattices in k\mathbb{R}^{k} up to isometries and homothety. For a lattice Γk\Gamma\subset\mathbb{R}^{k}, we denote by Γ\langle\Gamma\rangle its equivalence class. The map

(6.2) Γ[Q0|Γ]\langle\Gamma\rangle\mapsto[Q_{0}|_{\Gamma}]

is the desired bijection. In words, the class of lattices Γ\langle\Gamma\rangle is associated to the similarity class of the standard form Q0Q_{0} represented in a basis of the lattice Γ\Gamma.

Note that we have a map [L,Λ]𝒴[Q|LΛ]𝒮k[L,\Lambda]\in\mathcal{Y}\mapsto[Q|_{L\cap\Lambda}]\in\mathcal{S}_{k} already alluded to in the introduction. It is natural to ask what equivalence class of lattices corresponds to the similarity class (or shape) [Q|LΛ][Q|_{L\cap\Lambda}] from the introduction under the identification (6.2). To answer this question, choose a rotation kLSOQ()k_{L}\in\mathrm{SO}_{Q}(\mathbb{R}) with kLL()=L0()k_{L}L(\mathbb{R})=L_{0}(\mathbb{R}). Apply ηQ\eta_{Q} to the lattice kL(LΛ)L0()k_{L}(L\cap\Lambda)\subset L_{0}(\mathbb{R}). Recall that ηQ\eta_{Q} was chosen in §1.4.1 to preserve L0()L_{0}(\mathbb{R}) so that ηQkL(LΛ)L0()\eta_{Q}k_{L}(L\cap\Lambda)\subset L_{0}(\mathbb{R}). Since

Q0|ηQkL(LΛ)Q|LΛ,\displaystyle Q_{0}|_{\eta_{Q}k_{L}(L\cap\Lambda)}\simeq Q|_{L\cap\Lambda},

the equivalence class of the lattice ηQkL(LΛ)\eta_{Q}k_{L}(L\cap\Lambda) corresponds to the similarity class or shape [Q|LΛ][Q|_{L\cap\Lambda}]. As we did in the introduction, we will also write [LΛ][L\cap\Lambda] for that shape.

Lemma 6.5.

There is a surjective map

𝒴+Grn,k()×𝒮k×𝒮nk\displaystyle\mathcal{Y}^{+}\to\mathrm{Gr}_{n,k}(\mathbb{R})\times\mathcal{S}_{k}\times\mathcal{S}_{n-k}

given explicitly by [L,Λ](L,[LΛ],[LΛ#])[L,\Lambda]\mapsto(L,[L\cap\Lambda],[L^{\perp}\cap\Lambda^{\#}]). Moreover, the pushforward of the Haar (quotient) probability measure is the Haar probability measure on the target.

Proof.

Recall that 𝐇L0\mathbf{H}^{\prime}_{L_{0}} is the stabilizer of L0L_{0} in SOQ\mathrm{SO}_{Q}. Over \mathbb{R}, we have 𝐇L0()=ρQ(𝐇L0())\mathbf{H}^{\prime}_{L_{0}}(\mathbb{R})=\rho_{Q}(\mathbf{H}_{L_{0}}(\mathbb{R})). Consider the (surjective) composition

𝒴+\displaystyle\mathcal{Y}^{+} 𝚫𝐇L0()\𝐆()/ ​​𝐏n,k()\displaystyle\to\mathchoice{\text{\lower 2.15277pt\hbox{$\mathbf{\Delta H}_{L_{0}}(\mathbb{R})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathbf{G}(\mathbb{R})$\!}\big{/} \lower 2.15277pt\hbox{\!\!$\mathbf{P}_{n,k}(\mathbb{Z})$}}}{\mathbf{\Delta H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathbf{G}(\mathbb{R})\,/\,\mathbf{P}_{n,k}(\mathbb{Z})}{\mathbf{\Delta H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathbf{G}(\mathbb{R})\,/\,\mathbf{P}_{n,k}(\mathbb{Z})}{\mathbf{\Delta H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathbf{G}(\mathbb{R})\,/\,\mathbf{P}_{n,k}(\mathbb{Z})}
𝐇L0()\SpinQ()×𝐇L0())\𝐏n,k()/ ​​𝐏n,k()\displaystyle\to\mathchoice{\text{\lower 2.15277pt\hbox{$\mathbf{H}_{L_{0}}(\mathbb{R})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{R})$}}}{\mathbf{H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathrm{Spin}_{Q}(\mathbb{R})}{\mathbf{H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathrm{Spin}_{Q}(\mathbb{R})}{\mathbf{H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathrm{Spin}_{Q}(\mathbb{R})}\times\mathchoice{\text{\lower 2.15277pt\hbox{$\mathbf{H}^{\prime}_{L_{0}}(\mathbb{R}))$}\big{\backslash}\raise 2.15277pt\hbox{$\mathbf{P}_{n,k}(\mathbb{R})$\!}\big{/} \lower 2.15277pt\hbox{\!\!$\mathbf{P}_{n,k}(\mathbb{Z})$}}}{\mathbf{H}^{\prime}_{L_{0}}(\mathbb{R}))\,\backslash\,\mathbf{P}_{n,k}(\mathbb{R})\,/\,\mathbf{P}_{n,k}(\mathbb{Z})}{\mathbf{H}^{\prime}_{L_{0}}(\mathbb{R}))\,\backslash\,\mathbf{P}_{n,k}(\mathbb{R})\,/\,\mathbf{P}_{n,k}(\mathbb{Z})}{\mathbf{H}^{\prime}_{L_{0}}(\mathbb{R}))\,\backslash\,\mathbf{P}_{n,k}(\mathbb{R})\,/\,\mathbf{P}_{n,k}(\mathbb{Z})}
𝐇L0()\SpinQ()×ηQ𝐇L0()ηQ1\𝐏n,k()/ ​​𝐏n,k()\displaystyle\to\mathchoice{\text{\lower 2.15277pt\hbox{$\mathbf{H}_{L_{0}}(\mathbb{R})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{R})$}}}{\mathbf{H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathrm{Spin}_{Q}(\mathbb{R})}{\mathbf{H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathrm{Spin}_{Q}(\mathbb{R})}{\mathbf{H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathrm{Spin}_{Q}(\mathbb{R})}\times\mathchoice{\text{\lower 2.15277pt\hbox{$\eta_{Q}\mathbf{H}^{\prime}_{L_{0}}(\mathbb{R})\eta_{Q}^{-1}$}\big{\backslash}\raise 2.15277pt\hbox{$\mathbf{P}_{n,k}(\mathbb{R})$\!}\big{/} \lower 2.15277pt\hbox{\!\!$\mathbf{P}_{n,k}(\mathbb{Z})$}}}{\eta_{Q}\mathbf{H}^{\prime}_{L_{0}}(\mathbb{R})\eta_{Q}^{-1}\,\backslash\,\mathbf{P}_{n,k}(\mathbb{R})\,/\,\mathbf{P}_{n,k}(\mathbb{Z})}{\eta_{Q}\mathbf{H}^{\prime}_{L_{0}}(\mathbb{R})\eta_{Q}^{-1}\,\backslash\,\mathbf{P}_{n,k}(\mathbb{R})\,/\,\mathbf{P}_{n,k}(\mathbb{Z})}{\eta_{Q}\mathbf{H}^{\prime}_{L_{0}}(\mathbb{R})\eta_{Q}^{-1}\,\backslash\,\mathbf{P}_{n,k}(\mathbb{R})\,/\,\mathbf{P}_{n,k}(\mathbb{Z})}

where the first map is the identification in Lemma 6.3, the second map is the quotient map and the third map is multiplication by ηQ\eta_{Q} in the second factor. First, observe that 𝐇L0()\SpinQ()\mathchoice{\text{\lower 2.15277pt\hbox{$\mathbf{H}_{L_{0}}(\mathbb{R})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{R})$}}}{\mathbf{H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathrm{Spin}_{Q}(\mathbb{R})}{\mathbf{H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathrm{Spin}_{Q}(\mathbb{R})}{\mathbf{H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathrm{Spin}_{Q}(\mathbb{R})} is identified with Grn,k+()\mathrm{Gr}^{+}_{n,k}(\mathbb{R}) via 𝐇L0()g0g01.L0()\mathbf{H}_{L_{0}}(\mathbb{R})g_{0}\mapsto g_{0}^{-1}.L_{0}(\mathbb{R}). Note also that ηQ𝐇L0())ηQ1\eta_{Q}\mathbf{H}^{\prime}_{L_{0}}(\mathbb{R}))\eta_{Q}^{-1} is equal to the group SOk()×SOnk()\mathrm{SO}_{k}(\mathbb{R})\times\mathrm{SO}_{n-k}(\mathbb{R}) embedded block-diagonally. We apply projections onto the blocks (π1,π2\pi_{1},\pi_{2} defined in §1.4.4) as well as inverse-transpose in the second block to obtain a surjective map

ηQ𝐇L0()ηQ1\𝐏n,k()/ ​​𝐏n,k()𝒮k×𝒮nk.\displaystyle\mathchoice{\text{\lower 2.15277pt\hbox{$\eta_{Q}\mathbf{H}^{\prime}_{L_{0}}(\mathbb{R})\eta_{Q}^{-1}$}\big{\backslash}\raise 2.15277pt\hbox{$\mathbf{P}_{n,k}(\mathbb{R})$\!}\big{/} \lower 2.15277pt\hbox{\!\!$\mathbf{P}_{n,k}(\mathbb{Z})$}}}{\eta_{Q}\mathbf{H}^{\prime}_{L_{0}}(\mathbb{R})\eta_{Q}^{-1}\,\backslash\,\mathbf{P}_{n,k}(\mathbb{R})\,/\,\mathbf{P}_{n,k}(\mathbb{Z})}{\eta_{Q}\mathbf{H}^{\prime}_{L_{0}}(\mathbb{R})\eta_{Q}^{-1}\,\backslash\,\mathbf{P}_{n,k}(\mathbb{R})\,/\,\mathbf{P}_{n,k}(\mathbb{Z})}{\eta_{Q}\mathbf{H}^{\prime}_{L_{0}}(\mathbb{R})\eta_{Q}^{-1}\,\backslash\,\mathbf{P}_{n,k}(\mathbb{R})\,/\,\mathbf{P}_{n,k}(\mathbb{Z})}\to\mathcal{S}_{k}\times\mathcal{S}_{n-k}.

Overall, we have a surjection ϕ:𝒴+Grn,k()×𝒮k×𝒮nk\phi\mathrel{\mathop{\mathchar 58\relax}}\mathcal{Y}^{+}\to\mathrm{Gr}_{n,k}(\mathbb{R})\times\mathcal{S}_{k}\times\mathcal{S}_{n-k}.

It remains to verify that this surjection is the map from the lemma. Let [L,Λ]𝒴+[L,\Lambda]\in\mathcal{Y}^{+} and let (g1,g2)𝐆()(g_{1},g_{2})\in\mathbf{G}(\mathbb{R}) be a representative of its double coset in Lemma 6.3. It is clear from the proof of Lemma 6.3 that ϕ([L,Λ])1=g11.L0()=L()\phi([L,\Lambda])_{1}=g_{1}^{-1}.L_{0}(\mathbb{R})=L(\mathbb{R}). For the second component, note that using g11.L0()=L()g_{1}^{-1}.L_{0}(\mathbb{R})=L(\mathbb{R})

[Q|LΛ]=[Q0|ηQρQ(g1)(LΛ))]=[Q0|ηQ(L0g2n)]=[Q0|π1(ηQg2)k]=ϕ([L,Λ])2.\displaystyle[Q|_{L\cap\Lambda}]=[Q_{0}|_{\eta_{Q}\rho_{Q}(g_{1})(L\cap\Lambda)})]=[Q_{0}|_{\eta_{Q}(L_{0}\cap g_{2}\mathbb{Z}^{n})}]=[Q_{0}|_{\pi_{1}(\eta_{Q}g_{2})\mathbb{Z}^{k}}]=\phi([L,\Lambda])_{2}.

For the third component, we first observe that L()=g11.L0()L^{\perp}(\mathbb{R})=g_{1}^{-1}.L_{0}(\mathbb{R})^{\perp} as well as Λ#=ρQ(g11)(g21)tn\Lambda^{\#}=\rho_{Q}(g_{1}^{-1})(g_{2}^{-1})^{t}\mathbb{Z}^{n}. Hence,

[Q|LΛ#]\displaystyle[Q|_{L^{\perp}\cap\Lambda^{\#}}] =[Q0|ηQρQ(g1)(LΛ#)]=[Q0|ηQ(L0(g21)tn)]=[Q0|π2(ηQ(g21)t)k]\displaystyle=[Q_{0}|_{\eta_{Q}\rho_{Q}(g_{1})(L^{\perp}\cap\Lambda^{\#})}]=[Q_{0}|_{\eta_{Q}(L_{0}^{\perp}\cap(g_{2}^{-1})^{t}\mathbb{Z}^{n}})]=[Q_{0}|_{\pi_{2}(\eta_{Q}(g_{2}^{-1})^{t})\mathbb{Z}^{k}}]
=ϕ([L,Λ])3\displaystyle=\phi([L,\Lambda])_{3}

which concludes the lemma. ∎

6.3. A construction of an intermediate lattice

As was already observed in Remark 1.10, equidistribution of the tuples [L,n][L,\mathbb{Z}^{n}] for LQn,k(D)L\in\mathcal{H}^{n,k}_{Q}(D) (Conjecture 1.9) does not necessarily imply equidistribution of the tuples (L,[L()],[L()])(L,[L(\mathbb{Z})],[L^{\perp}(\mathbb{Z})]) when QQ is not unimodular (Conjecture 1.1). Indeed, one can see from Lemma 6.5 that it implies equidistribution of the tuples (L,[L()],[L(n)#])(L,[L(\mathbb{Z})],[L^{\perp}\cap(\mathbb{Z}^{n})^{\#}]) for LQn,k(D)L\in\mathcal{H}^{n,k}_{Q}(D). Here, we construct for every LL a full rank sublattice ΛLn\Lambda_{L}\subset\mathbb{Q}^{n} so that equidistribution of the tuples [L,ΛL][L,\Lambda_{L}] does have this desired implication. For any subspace LnL\subset\mathbb{Q}^{n} write πL\pi_{L} for the orthogonal projection onto LL.

Proposition 6.6.

For any subspace LGrn,k()L\in\mathrm{Gr}_{n,k}(\mathbb{Q}) there exists a full rank \mathbb{Z}-lattice ΛLn\Lambda_{L}\subset\mathbb{Q}^{n} with the following properties:

  1. (1)

    nΛL(n)#\mathbb{Z}^{n}\subset\Lambda_{L}\subset(\mathbb{Z}^{n})^{\#}

  2. (2)

    We have

    LΛL=L(),πL(ΛL)=L()#,andL()=ΛL#L.\displaystyle L\cap\Lambda_{L}=L(\mathbb{Z}),\quad\pi_{L^{\perp}}(\Lambda_{L})=L^{\perp}(\mathbb{Z})^{\#},\quad\text{and}\quad L^{\perp}(\mathbb{Z})=\Lambda_{L}^{\#}\cap L^{\perp}.
  3. (3)

    Suppose that LL^{\prime} satisfies that there are γSpinQ()\gamma\in\mathrm{Spin}_{Q}(\mathbb{Q}) and kpSpinQ(p)k_{p}\in\mathrm{Spin}_{Q}(\mathbb{Z}_{p}) for every prime pp such that γ.L=L\gamma.L=L^{\prime} and kp.L(p)=L(p)k_{p}.L(\mathbb{Z}_{p})=L^{\prime}(\mathbb{Z}_{p}). Then

    ΛL=pkp.(ΛLp)n.\displaystyle\Lambda_{L^{\prime}}=\bigcap_{p}k_{p}.(\Lambda_{L}\otimes\mathbb{Z}_{p})\cap\mathbb{Q}^{n}.

We remark that if QQ is unimodular, one may simply take ΛL=n\Lambda_{L}=\mathbb{Z}^{n}. For QQ not unimodular, this choice generally satisfies (1) and (3) but not necessarily (2).

Remark 6.7 (Equivalence relation).

We write LLL\sim L^{\prime} for rational subspaces L,LL,L^{\prime} of dimension kk if there are γSpinQ()\gamma\in\mathrm{Spin}_{Q}(\mathbb{Q}) and kpSpinQ(p)k_{p}\in\mathrm{Spin}_{Q}(\mathbb{Z}_{p}) for every prime pp such that γ.L=L\gamma.L=L^{\prime} and kp.L(p)=L(p)k_{p}.L(\mathbb{Z}_{p})=L^{\prime}(\mathbb{Z}_{p}). This defines an equivalence relation. As L,LL,L^{\prime} are locally rotated into each other, they have the same discriminant (see Equation (1.5)).

Proof of Proposition 6.6.

In view of Remark 6.7 and the required property in (3) we first observe that if LL^{\prime} is equivalent to LL and LL satisfies (1),(2) then LL^{\prime} also does so. We may hence split Grn,k()\mathrm{Gr}_{n,k}(\mathbb{Q}) into equivalence classes, choose a representative LL in each equivalence class, and construct ΛL\Lambda_{L} with the properties in (1) and (2) ignoring (3).

So let LGrn,k()L\in\mathrm{Gr}_{n,k}(\mathbb{Q}) be such a representative. Choose a basis v1,,vkv_{1},\ldots,v_{k} of L()L(\mathbb{Z}). We consider the \mathbb{Z}-module (n)#/L()\mathchoice{\text{\raise 2.15277pt\hbox{$(\mathbb{Z}^{n})^{\#}$}\big{/}\lower 2.15277pt\hbox{$L(\mathbb{Z})$}}}{(\mathbb{Z}^{n})^{\#}\,/\,L(\mathbb{Z})}{(\mathbb{Z}^{n})^{\#}\,/\,L(\mathbb{Z})}{(\mathbb{Z}^{n})^{\#}\,/\,L(\mathbb{Z})} which fits into the following exact sequence

(6.3) 0L(n)#/L()(n)#/L()(n)#/L(n)#0.\displaystyle 0\to\mathchoice{\text{\raise 2.15277pt\hbox{$L\cap(\mathbb{Z}^{n})^{\#}$}\big{/}\lower 2.15277pt\hbox{$L(\mathbb{Z})$}}}{L\cap(\mathbb{Z}^{n})^{\#}\,/\,L(\mathbb{Z})}{L\cap(\mathbb{Z}^{n})^{\#}\,/\,L(\mathbb{Z})}{L\cap(\mathbb{Z}^{n})^{\#}\,/\,L(\mathbb{Z})}\to\mathchoice{\text{\raise 2.15277pt\hbox{$(\mathbb{Z}^{n})^{\#}$}\big{/}\lower 2.15277pt\hbox{$L(\mathbb{Z})$}}}{(\mathbb{Z}^{n})^{\#}\,/\,L(\mathbb{Z})}{(\mathbb{Z}^{n})^{\#}\,/\,L(\mathbb{Z})}{(\mathbb{Z}^{n})^{\#}\,/\,L(\mathbb{Z})}\to\mathchoice{\text{\raise 2.15277pt\hbox{$(\mathbb{Z}^{n})^{\#}$}\big{/}\lower 2.15277pt\hbox{$L\cap(\mathbb{Z}^{n})^{\#}$}}}{(\mathbb{Z}^{n})^{\#}\,/\,L\cap(\mathbb{Z}^{n})^{\#}}{(\mathbb{Z}^{n})^{\#}\,/\,L\cap(\mathbb{Z}^{n})^{\#}}{(\mathbb{Z}^{n})^{\#}\,/\,L\cap(\mathbb{Z}^{n})^{\#}}\to 0.

As L(n)#L\cap(\mathbb{Z}^{n})^{\#} is primitive999A sublattice Γ\Gamma of a lattice Λn\Lambda\subset\mathbb{Q}^{n} is primitive if it is not strictly contained in any sublattice of the same rank. in (n)#(\mathbb{Z}^{n})^{\#}, the module on the very right is free of rank nkn-k. We choose a basis of it as well as representatives vk+1,,vn(n)#v_{k+1},\ldots,v_{n}\in(\mathbb{Z}^{n})^{\#} of these basis elements. Define

ΛL=v1++vn.\displaystyle\Lambda_{L}=\mathbb{Z}v_{1}+\ldots+\mathbb{Z}v_{n}.

It is not very hard to see that this lattice contains n\mathbb{Z}^{n} and is contained in (n)#(\mathbb{Z}^{n})^{\#} so that (1) is satisfied.

Suppose that

v=iαiviLΛL.\displaystyle v=\sum_{i}\alpha_{i}v_{i}\in L\cap\Lambda_{L}.

This implies that i>kαiviL\sum_{i>k}\alpha_{i}v_{i}\in L and so i>kαivi=0\sum_{i>k}\alpha_{i}v_{i}=0 by linear independence. The identity LΛL=L()L\cap\Lambda_{L}=L(\mathbb{Z}) follows.

By Lemma 5.3, the projection πL:(n)#L()#\pi_{L^{\perp}}\mathrel{\mathop{\mathchar 58\relax}}(\mathbb{Z}^{n})^{\#}\to L^{\perp}(\mathbb{Z})^{\#} is surjective. Clearly, the kernel is L(n)#L\cap(\mathbb{Z}^{n})^{\#} and hence by construction of ΛL\Lambda_{L} we have πL(ΛL)=πL((n)#)=L()#\pi_{L^{\perp}}(\Lambda_{L})=\pi_{L^{\perp}}((\mathbb{Z}^{n})^{\#})=L^{\perp}(\mathbb{Z})^{\#}.

It remains to prove the last identity. As ΛL#n\Lambda_{L}^{\#}\supset\mathbb{Z}^{n} we have ΛL#LL()\Lambda_{L}^{\#}\cap L^{\perp}\supset L^{\perp}(\mathbb{Z}) so it suffices to show that

L()#=πL(ΛL)(ΛL#L)#.\displaystyle L^{\perp}(\mathbb{Z})^{\#}=\pi_{L^{\perp}}(\Lambda_{L})\subset(\Lambda_{L}^{\#}\cap L^{\perp})^{\#}.

For v=πL(v)πL(ΛL)v=\pi_{L^{\perp}}(v^{\prime})\in\pi_{L^{\perp}}(\Lambda_{L}) and wLΛL#w\in L^{\perp}\cap\Lambda_{L}^{\#} we have v,w=v,w\langle v,w\rangle=\langle v^{\prime},w\rangle\in\mathbb{Z} proving the remaining claim. ∎

Remark 6.8.

Observe that ΛL\Lambda_{L} constructed above depends on the choice of basis for the free module (n)#/L(n)#\mathchoice{\text{\raise 2.15277pt\hbox{$(\mathbb{Z}^{n})^{\#}$}\big{/}\lower 2.15277pt\hbox{$L\cap(\mathbb{Z}^{n})^{\#}$}}}{(\mathbb{Z}^{n})^{\#}\,/\,L\cap(\mathbb{Z}^{n})^{\#}}{(\mathbb{Z}^{n})^{\#}\,/\,L\cap(\mathbb{Z}^{n})^{\#}}{(\mathbb{Z}^{n})^{\#}\,/\,L\cap(\mathbb{Z}^{n})^{\#}} which forms the ’free part’ of (n)#/L()\mathchoice{\text{\raise 2.15277pt\hbox{$(\mathbb{Z}^{n})^{\#}$}\big{/}\lower 2.15277pt\hbox{$L(\mathbb{Z})$}}}{(\mathbb{Z}^{n})^{\#}\,/\,L(\mathbb{Z})}{(\mathbb{Z}^{n})^{\#}\,/\,L(\mathbb{Z})}{(\mathbb{Z}^{n})^{\#}\,/\,L(\mathbb{Z})} in the sense of (6.3). But the short exact sequence (6.3) does not split in general so that the basis elements have no canonical lifts to (n)#/L()\mathchoice{\text{\raise 2.15277pt\hbox{$(\mathbb{Z}^{n})^{\#}$}\big{/}\lower 2.15277pt\hbox{$L(\mathbb{Z})$}}}{(\mathbb{Z}^{n})^{\#}\,/\,L(\mathbb{Z})}{(\mathbb{Z}^{n})^{\#}\,/\,L(\mathbb{Z})}{(\mathbb{Z}^{n})^{\#}\,/\,L(\mathbb{Z})}; different choices yield different lattices ΛL\Lambda_{L}. This dependency is inconsequential as the set of lattices Λ\Lambda with nΛ(n)#\mathbb{Z}^{n}\subset\Lambda\subset(\mathbb{Z}^{n})^{\#} is finite.

6.4. A refinement of Theorem 1.11

We now present a refinement of Theorem 1.11 which is necessary in order to deduce the desired equidistribution theorem of shapes (i.e. Theorem 1.4).

Theorem 6.9.

Let k3k\geq 3 with knkk\leq n-k and let pp be a prime with p2disc(Q)p\nmid 2\mathrm{disc}(Q). Let LGrn,k()ΛLL\in\mathrm{Gr}_{n,k}(\mathbb{Q})\mapsto\Lambda_{L} satisfy conditions (1) and (3) from Proposition 6.6. Suppose that DiD_{i}\in\mathbb{N} is a sequence of integers with Di[k]D_{i}^{[k]}\to\infty, Qn,k(Di)\mathcal{H}_{Q}^{n,k}(D_{i})\neq\emptyset as well pDip\nmid D_{i} if k{3,4}k\in\{3,4\}. Then the sets

(6.4) {([L,ΛL]:Ln oriented ,discQ(L)=Di,dim(L)=k}\displaystyle\{([L,\Lambda_{L}]\mathrel{\mathop{\mathchar 58\relax}}L\subset\mathbb{Q}^{n}\text{ oriented },\mathrm{disc}_{Q}(L)=D_{i},\dim(L)=k\}

equidistribute in 𝒴+\mathcal{Y}^{+} as ii\to\infty

We observe that the special case ΛL=n\Lambda_{L}=\mathbb{Z}^{n} for every LGrn,k()L\in\mathrm{Gr}_{n,k}(\mathbb{Q}) in Theorem 6.9 implies Theorem 1.11 after projection 𝒴+𝒴\mathcal{Y}^{+}\to\mathcal{Y}.

Proof of Theorem 1.4 from Theorem 6.9 when k3k\geq 3.

Let ΛL\Lambda_{L} for LGrn,k()L\in\mathrm{Gr}_{n,k}(\mathbb{Q}) be defined as in Proposition 6.6. Let pp be a prime and Di1D_{i}\geq 1 be a sequence of discriminants as in Theorem 1.4. Then Theorem 6.9 is applicable and the sets in (6.4) are equidistributed in 𝒴+\mathcal{Y}^{+} when ii\to\infty. By construction of ΛL\Lambda_{L}, the image of these sets under the map in Lemma 6.5 is exactly

{(L,[L()],[L()]):LQn,k(Di)}.\displaystyle\{(L,[L(\mathbb{Z})],[L^{\perp}(\mathbb{Z})])\mathrel{\mathop{\mathchar 58\relax}}L\in\mathcal{H}_{Q}^{n,k}(D_{i})\}.

These images are equidistributed with respect to the pushforward measure, which is the Haar probability measure on Grn,k()×𝒮k×𝒮nk\mathrm{Gr}_{n,k}(\mathbb{R})\times\mathcal{S}_{k}\times\mathcal{S}_{n-k}. ∎

Remark 6.10 (Theorem 1.4 for oriented subspaces).

Let 𝒳k\mathcal{X}_{k} be the space of positive definite real quadratic forms in kk variables up to proper similarity. Observe that the shape of an oriented kk-dimensional subspace makes sense as a point in 𝒳k\mathcal{X}_{k}. Very much related to this is the fact that the proof of Lemma 6.5 actually establishes a surjective map 𝒴+Grn,k+()×𝒳k×𝒳nk\mathcal{Y}^{+}\to\mathrm{Gr}^{+}_{n,k}(\mathbb{R})\times\mathcal{X}_{k}\times\mathcal{X}_{n-k}. Theorem 1.4 may thus be generalized to this latter space. For k=1k=1, this oriented version already appears in the works [AES-dim3, AES-higherdim].

7. Proof of the main theorems from the dynamical versions

The aim of this section is to prove Theorem 6.9 and Theorem 1.4 for k=2k=2. We remark that any possible future upgrades to the dynamical versions (in regard to the congruence conditions at fixed primes) imply the analogous upgrades to the arithmetic versions.

7.1. Notation

We recall and introduce here some notation used throughout this Section 7. In the following, LnL\subset\mathbb{Q}^{n} is an arbitrary kk-dimensional oriented subspace unless specified otherwise:

  • 𝒴+\mathcal{Y}^{+} is the moduli space of oriented basis extensions defined in §6.2.1 (see also §1.1). Recall that SpinQ()\mathrm{Spin}_{Q}(\mathbb{Z}) acts on 𝒴+\mathcal{Y}^{+} via g[L,Λ]=[g.L,g.Λ]g[L,\Lambda]=[g.L,g.\Lambda]. Moreover, by Lemma 6.3 and the subsequent Remark 6.4, we have

    (7.1) 𝒴+\displaystyle\mathcal{Y}^{+} 𝚫𝐇L0()\𝐆()/ ​​𝐏n,k(),\displaystyle\simeq\mathchoice{\text{\lower 2.15277pt\hbox{$\mathbf{\Delta H}_{L_{0}}(\mathbb{R})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathbf{G}(\mathbb{R})$\!}\big{/} \lower 2.15277pt\hbox{\!\!$\mathbf{P}_{n,k}(\mathbb{Z})$}}}{\mathbf{\Delta H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathbf{G}(\mathbb{R})\,/\,\mathbf{P}_{n,k}(\mathbb{Z})}{\mathbf{\Delta H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathbf{G}(\mathbb{R})\,/\,\mathbf{P}_{n,k}(\mathbb{Z})}{\mathbf{\Delta H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathbf{G}(\mathbb{R})\,/\,\mathbf{P}_{n,k}(\mathbb{Z})},
    (7.2) SpinQ()\mathrm{Spin}_{Q}(\mathbb{Z})\𝒴+\mathcal{Y}^{+} 𝚫𝐇L0()\𝐆()/ ​​𝐆()\displaystyle\simeq\mathchoice{\text{\lower 2.15277pt\hbox{$\mathbf{\Delta H}_{L_{0}}(\mathbb{R})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathbf{G}(\mathbb{R})$\!}\big{/} \lower 2.15277pt\hbox{\!\!$\mathbf{G}(\mathbb{Z})$}}}{\mathbf{\Delta H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathbf{G}(\mathbb{R})\,/\,\mathbf{G}(\mathbb{Z})}{\mathbf{\Delta H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathbf{G}(\mathbb{R})\,/\,\mathbf{G}(\mathbb{Z})}{\mathbf{\Delta H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathbf{G}(\mathbb{R})\,/\,\mathbf{G}(\mathbb{Z})}

    where L0=k×{(0,,0)}nL_{0}=\mathbb{Q}^{k}\times\{(0,\ldots,0)\}\subset\mathbb{Q}^{n} is the fixed reference subspace (cf. (1.6)) and 𝐆=SpinQ×𝐏n,k\mathbf{G}=\mathrm{Spin}_{Q}\times\mathbf{P}_{n,k} (cf. 1.4.4).

  • The subgroup 𝐇L<SpinQ\mathbf{H}_{L}<\mathrm{Spin}_{Q} is the identity component of the stabilizer group of LL (cf. §2.1 and see also §6.1).

  • We fix a full-rank lattice nΛL(n)#\mathbb{Z}^{n}\subset\Lambda_{L}\subset(\mathbb{Z}^{n})^{\#} satisfying (1) and (3) in Proposition 6.6. The reader is encouraged to keep in mind the case disc(Q)=1\mathrm{disc}(Q)=1 where one may take ΛL=n\Lambda_{L}=\mathbb{Z}^{n} for all LL.

  • We fix an oriented basis of ΛL\Lambda_{L} where the first kk vectors are an oriented basis of LΛLL\cap\Lambda_{L}. Let gLGLn()g_{L}\in\mathrm{GL}_{n}(\mathbb{Q}) be the element whose columns consist of this basis.

  • The subgroup 𝚫𝐇L<𝐆\mathbf{\Delta H}_{L}<\mathbf{G} is defined as in §2.3 using the basis in gLg_{L}.

  • For any [L,Λ]𝒴+[L,\Lambda]\in\mathcal{Y}^{+} (where LL is not necessarily rational) we write to shorten notation [L,Λ][L,\Lambda]_{\star} for the equivalence class SpinQ()[L,Λ]SpinQ()\𝒴+\mathrm{Spin}_{Q}(\mathbb{Z})[L,\Lambda]\in\mathchoice{\text{\lower 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Z})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{Y}^{+}$}}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{Y}^{+}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{Y}^{+}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{Y}^{+}}.

  • Let 𝗌L𝐆()\mathsf{s}_{L}\in\mathbf{G}(\mathbb{R}) be the representative of the double coset of [L,ΛL][L,\Lambda_{L}] defined using gLg_{L} (see also the proof of Lemma 6.3).

  • For any DD\in\mathbb{N} with Qn,k(D)\mathcal{H}^{n,k}_{Q}(D)\neq\emptyset we consider the finite set Qn,k(D)𝒴+\mathcal{R}^{n,k}_{Q}(D)\subset\mathcal{Y}^{+} consisting of classes [L,ΛL][L,\Lambda_{L}] where LL runs over all oriented kk-dimensional subspaces LnL\subset\mathbb{Q}^{n} with discQ(L)=D\mathrm{disc}_{Q}(L)=D – see also (6.4). The action of SpinQ()\mathrm{Spin}_{Q}(\mathbb{Z}) on 𝒴+\mathcal{Y}^{+} leaves Qn,k(D)\mathcal{R}^{n,k}_{Q}(D) invariant.

7.2. Outline of the proof

Let 𝒰=𝐆(×^)𝐆()𝐆(𝔸)/𝐆()\mathcal{U}=\mathbf{G}(\mathbb{R}\times\widehat{\mathbb{Z}})\mathbf{G}(\mathbb{Q})\subset\mathchoice{\text{\raise 2.15277pt\hbox{$\mathbf{G}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathbf{G}(\mathbb{Q})$}}}{\mathbf{G}(\mathbb{A})\,/\,\mathbf{G}(\mathbb{Q})}{\mathbf{G}(\mathbb{A})\,/\,\mathbf{G}(\mathbb{Q})}{\mathbf{G}(\mathbb{A})\,/\,\mathbf{G}(\mathbb{Q})} be the principal genus101010The genera (i.e. orbits of 𝐆()×𝐆(^)\mathbf{G}(\mathbb{R})\times\mathbf{G}(\widehat{\mathbb{Z}})) correspond to classes in the spinor genus of QQ. Recall here that if QQ is the sum of squares in 8\leq 8 variables, then the spinor genus consists of one class (cf. [Cassels, p.232]) and hence 𝒰=𝐆(𝔸)\𝐆()\mathcal{U}=\mathchoice{\text{\lower 2.15277pt\hbox{$\mathbf{G}(\mathbb{A})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathbf{G}(\mathbb{Q})$}}}{\mathbf{G}(\mathbb{A})\,\backslash\,\mathbf{G}(\mathbb{Q})}{\mathbf{G}(\mathbb{A})\,\backslash\,\mathbf{G}(\mathbb{Q})}{\mathbf{G}(\mathbb{A})\,\backslash\,\mathbf{G}(\mathbb{Q})}.. There is a natural map

(7.3) 𝐆(𝔸)/𝐆()𝒰SpinQ()\𝒴+\displaystyle\mathchoice{\text{\raise 2.15277pt\hbox{$\mathbf{G}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathbf{G}(\mathbb{Q})$}}}{\mathbf{G}(\mathbb{A})\,/\,\mathbf{G}(\mathbb{Q})}{\mathbf{G}(\mathbb{A})\,/\,\mathbf{G}(\mathbb{Q})}{\mathbf{G}(\mathbb{A})\,/\,\mathbf{G}(\mathbb{Q})}\supset\mathcal{U}\to\mathchoice{\text{\lower 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Z})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{Y}^{+}$}}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{Y}^{+}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{Y}^{+}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{Y}^{+}}

given by taking the quotient on the left of 𝐆(𝔸)/𝐆()\mathbf{G}(\mathbb{A})/\mathbf{G}(\mathbb{Q}) by the maximal compact open subgroup 𝐆(^)\mathbf{G}(\widehat{\mathbb{Z}}) and 𝚫𝐇L0()\mathbf{\Delta H}_{L_{0}}(\mathbb{R}). Consider now an oriented subspace LL of discriminant DD and the orbit 𝗌L𝚫𝐇L(𝔸)𝐆()\mathsf{s}_{L}\mathbf{\Delta H}_{L}(\mathbb{A})\mathbf{G}(\mathbb{Q}). For any LQn,k(D)L\in\mathcal{H}^{n,k}_{Q}(D), the image of the intersection of 𝗌L𝚫𝐇L(𝔸)𝐆()\mathsf{s}_{L}\mathbf{\Delta H}_{L}(\mathbb{A})\mathbf{G}(\mathbb{Q}) with 𝒰\mathcal{U} under (7.3) is a subset of the collection SpinQ()\Qn,k(D)\mathchoice{\text{\lower 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Z})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{R}^{n,k}_{Q}(D)$}}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{R}^{n,k}_{Q}(D)}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{R}^{n,k}_{Q}(D)}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{R}^{n,k}_{Q}(D)} and contains [L,ΛL][L,\Lambda_{L}] – see Proposition 7.1. In other words, we have a commutative diagram

𝗌L𝚫𝐇L(𝔸)𝐆()𝒰{\mathsf{s}_{L}\mathbf{\Delta H}_{L}(\mathbb{A})\mathbf{G}(\mathbb{Q})\cap\mathcal{U}}𝒰{\mathcal{U}}SpinQ()\Qn,k(D){\mathchoice{\text{\lower 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Z})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{R}^{n,k}_{Q}(D)$}}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{R}^{n,k}_{Q}(D)}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{R}^{n,k}_{Q}(D)}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{R}^{n,k}_{Q}(D)}}SpinQ()\𝒴+.{\mathchoice{\text{\lower 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Z})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{Y}^{+}$}}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{Y}^{+}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{Y}^{+}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{Y}^{+}}.}

Assuming here k3k\geq 3, the intersection 𝗌L𝚫𝐇L(𝔸)𝐆()𝒰\mathsf{s}_{L}\mathbf{\Delta H}_{L}(\mathbb{A})\mathbf{G}(\mathbb{Q})\cap\mathcal{U} is equidistributed in 𝒰\mathcal{U} with respect to the normalized restriction of the Haar measure (along any sequence of admissible subspaces). This immediately implies equidistribution of the pushforwards under the map in (7.3).

It remains to compare the pushforward of the Haar measure on the orbit to the measure on SpinQ()\Qn,k(D)\mathchoice{\text{\lower 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Z})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{R}^{n,k}_{Q}(D)$}}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{R}^{n,k}_{Q}(D)}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{R}^{n,k}_{Q}(D)}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{R}^{n,k}_{Q}(D)} induced by the normalized counting measure on Qn,k(D)\mathcal{R}^{n,k}_{Q}(D). (This technical argument constitutes a large part of this section §7.) To this end, we first note that the projection P(L)\mathrm{P}(L) of 𝗌L𝚫𝐇L(𝔸)𝐆()𝒰\mathsf{s}_{L}\mathbf{\Delta H}_{L}(\mathbb{A})\mathbf{G}(\mathbb{Q})\cap\mathcal{U} is not surjective but SpinQ()\Qn,k(D)\mathchoice{\text{\lower 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Z})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{R}^{n,k}_{Q}(D)$}}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{R}^{n,k}_{Q}(D)}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{R}^{n,k}_{Q}(D)}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{R}^{n,k}_{Q}(D)} may be decomposed into such images for different subspaces LL – see Remark 7.2. Thus, it is enough to determine the weights of individual points in P(L)\mathrm{P}(L) – see Lemmas 7.3 and 7.4.

7.3. Generating integer points from the packet

As a first step towards the proof of Theorem 6.9, we illustrate a general technique for generating points in Qn,k(D)\mathcal{R}^{n,k}_{Q}(D) from a given point in Qn,k(D)\mathcal{R}^{n,k}_{Q}(D). This kind of idea appears in many recent or less recent articles in the literature – see for example [platonov, Thm. 8.2], [localglobalEV], [AES-dim3], [AES-higherdim], and [2in4].

For g𝐆=SpinQ×𝐏n,kg\in\mathbf{G}=\mathrm{Spin}_{Q}\times\mathbf{P}_{n,k} we write g=(g1,g2)g=(g_{1},g_{2}) where g1g_{1} is the first (resp. g2g_{2} is the second) coordinate of gg. Consider the open subset (principal genus)

𝒰=𝐆(×^)𝐆()𝐆(𝔸)/𝐆()\displaystyle\mathcal{U}=\mathbf{G}(\mathbb{R}\times\widehat{\mathbb{Z}})\mathbf{G}(\mathbb{Q})\subset\mathchoice{\text{\raise 2.15277pt\hbox{$\mathbf{G}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathbf{G}(\mathbb{Q})$}}}{\mathbf{G}(\mathbb{A})\,/\,\mathbf{G}(\mathbb{Q})}{\mathbf{G}(\mathbb{A})\,/\,\mathbf{G}(\mathbb{Q})}{\mathbf{G}(\mathbb{A})\,/\,\mathbf{G}(\mathbb{Q})}

On 𝒰\mathcal{U}, there is a projection map

(7.4) Φ:𝒰𝐆()/𝐆()𝚫𝐇L0()\𝐆()/ ​​𝐆()SpinQ()\𝒴+\Phi\mathrel{\mathop{\mathchar 58\relax}}\mathcal{U}\to\mathchoice{\text{\raise 2.15277pt\hbox{$\mathbf{G}(\mathbb{R})$}\big{/}\lower 2.15277pt\hbox{$\mathbf{G}(\mathbb{Z})$}}}{\mathbf{G}(\mathbb{R})\,/\,\mathbf{G}(\mathbb{Z})}{\mathbf{G}(\mathbb{R})\,/\,\mathbf{G}(\mathbb{Z})}{\mathbf{G}(\mathbb{R})\,/\,\mathbf{G}(\mathbb{Z})}\to\mathchoice{\text{\lower 2.15277pt\hbox{$\mathbf{\Delta H}_{L_{0}}(\mathbb{R})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathbf{G}(\mathbb{R})$\!}\big{/} \lower 2.15277pt\hbox{\!\!$\mathbf{G}(\mathbb{Z})$}}}{\mathbf{\Delta H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathbf{G}(\mathbb{R})\,/\,\mathbf{G}(\mathbb{Z})}{\mathbf{\Delta H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathbf{G}(\mathbb{R})\,/\,\mathbf{G}(\mathbb{Z})}{\mathbf{\Delta H}_{L_{0}}(\mathbb{R})\,\backslash\,\mathbf{G}(\mathbb{R})\,/\,\mathbf{G}(\mathbb{Z})}\simeq\mathchoice{\text{\lower 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Z})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{Y}^{+}$}}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{Y}^{+}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{Y}^{+}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{Y}^{+}}

where the first map takes for any point x𝒰x\in\mathcal{U} a representative in 𝐆(×^)\mathbf{G}(\mathbb{R}\times\widehat{\mathbb{Z}}) and projects onto the real component. Note that the first map is clearly 𝐆()\mathbf{G}(\mathbb{R})-equivariant. For LGrn,k()L\in\mathrm{Gr}_{n,k}(\mathbb{Q}) we define

(7.5) P(L):=Φ(𝗌L𝚫𝐇L(𝔸)𝐆()𝒰).\displaystyle\mathrm{P}(L)\mathrel{\mathop{\mathchar 58\relax}}=\Phi(\mathsf{s}_{L}\mathbf{\Delta H}_{L}(\mathbb{A})\mathbf{G}(\mathbb{Q})\cap\mathcal{U}).
Proposition 7.1.

For any oriented kk-dimensional subspace LnL\subset\mathbb{Q}^{n} of discriminant DD we have

P(L)SpinQ()Qn,k(D).\mathrm{P}(L)\subset\mathrm{Spin}_{Q}(\mathbb{Z})\setminus\mathcal{R}^{n,k}_{Q}(D).
Proof.

Fix a coset b𝐆()𝚫𝐇L(𝔸)𝐆()𝒰b\mathbf{G}(\mathbb{Q})\in\mathbf{\Delta H}_{L}(\mathbb{A})\mathbf{G}(\mathbb{Q})\cap\mathcal{U} and a representative b=(b1,b2)𝐆(×^)b=(b_{1},b_{2})\in\mathbf{G}(\mathbb{R}\times\widehat{\mathbb{Z}}). By definition of Φ\Phi

Φ(𝗌Lb𝐆())=𝚫𝐇L0()𝗌Lb𝐆().\displaystyle\Phi(\mathsf{s}_{L}b\mathbf{G}(\mathbb{Q}))=\mathbf{\Delta H}_{L_{0}}(\mathbb{R})\mathsf{s}_{L}b_{\infty}\mathbf{G}(\mathbb{Z}).

Note that since b𝐆()𝚫𝐇L(𝔸)𝐆()b\mathbf{G}(\mathbb{Q})\in\mathbf{\Delta H}_{L}(\mathbb{A})\mathbf{G}(\mathbb{Q}) there exists h𝚫𝐇L(𝔸)h\in\mathbf{\Delta H}_{L}(\mathbb{A}) and γ𝐆()\gamma\in\mathbf{G}(\mathbb{Q}) such that b=hγb=h\gamma. By definition of 𝚫𝐇L\mathbf{\Delta H}_{L} we have h2=gL1ρQ(h1)gLh_{2}=g_{L}^{-1}\rho_{Q}(h_{1})g_{L}. We first show that the point in SpinQ()\𝒴+\mathchoice{\text{\lower 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Z})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{Y}^{+}$}}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{Y}^{+}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{Y}^{+}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{Y}^{+}} corresponding to Φ(𝗌Lb𝐆())\Phi(\mathsf{s}_{L}b\mathbf{G}(\mathbb{Q})) lies above a rational subspace under the natural map 𝒴Grn,k+()\mathcal{Y}\to\mathrm{Gr}^{+}_{n,k}(\mathbb{R}). Note that by definition of the maps in (7.1) the subspace attached to Φ(𝗌Lb𝐆())\Phi(\mathsf{s}_{L}b\mathbf{G}(\mathbb{Q})) is ρQ(b1,1)ρL1L0=b1,1.L\rho_{Q}(b_{1,\infty}^{-1})\rho_{L}^{-1}L_{0}=b_{1,\infty}^{-1}.L. But

(7.6) b1,1.L=γ11h1,1.L=γ11.Ln.\displaystyle b_{1,\infty}^{-1}.L=\gamma_{1}^{-1}h_{1,\infty}^{-1}.L=\gamma_{1}^{-1}.L\subset\mathbb{Q}^{n}.

Next, we show that γ11.L\gamma_{1}^{-1}.L has discriminant DD. To this end, note that by an analogous argument as in (7.6) for a prime pp we have b1,p1.L=γ11.Lb_{1,p}^{-1}.L=\gamma_{1}^{-1}.L so that

discp,Q(L)=discp,Q(b1,p1.L)=discp,Q(γ11.L)\displaystyle\mathrm{disc}_{p,Q}(L)=\mathrm{disc}_{p,Q}(b_{1,p}^{-1}.L)=\mathrm{disc}_{p,Q}(\gamma_{1}^{-1}.L)

where we used that b1,pSpinQ(p)b_{1,p}\in\mathrm{Spin}_{Q}(\mathbb{Z}_{p}) preserves the local discriminant at pp. Thus, discQ(γ11.L)=D\mathrm{disc}_{Q}(\gamma_{1}^{-1}.L)=D by (1.5).

It remains to show that Φ(𝗌Lb𝐆())\Phi(\mathsf{s}_{L}b\mathbf{G}(\mathbb{Q})) corresponds to [γ11.L,Λγ11.L][\gamma_{1}^{-1}.L,\Lambda_{\gamma_{1}^{-1}.L}]_{\star}. For this, notice first that under (7.1)

Φ(𝗌Lb𝐆())=[γ11.L,ρQ(b1,1)gLb2,n]\displaystyle\Phi(\mathsf{s}_{L}b\mathbf{G}(\mathbb{Q}))=[\gamma_{1}^{-1}.L,\rho_{Q}(b_{1,\infty}^{-1})g_{L}b_{2,\infty}\mathbb{Z}^{n}]_{\star}

by definition of the equivalence relation. Now,

ρQ(b1,1)gLb2,=ρQ(γ11h11)gLh2γ2=ρQ(γ11)gLγ2.\displaystyle\rho_{Q}(b_{1,\infty}^{-1})g_{L}b_{2,\infty}=\rho_{Q}(\gamma_{1}^{-1}h_{1}^{-1})g_{L}h_{2}\gamma_{2}=\rho_{Q}(\gamma_{1}^{-1})g_{L}\gamma_{2}.

Quite analogously, we have ρQ(γ11)gLγ2=ρQ(b1,p1)gLb2,p\rho_{Q}(\gamma_{1}^{-1})g_{L}\gamma_{2}=\rho_{Q}(b_{1,p}^{-1})g_{L}b_{2,p} so that

ρQ(γ11)gLγ2pn=ρQ(b1,p1)gLpn=b1,p1.(ΛLp).\displaystyle\rho_{Q}(\gamma_{1}^{-1})g_{L}\gamma_{2}\mathbb{Z}_{p}^{n}=\rho_{Q}(b_{1,p}^{-1})g_{L}\mathbb{Z}_{p}^{n}=b_{1,p}^{-1}.(\Lambda_{L}\otimes\mathbb{Z}_{p}).

This shows that

ρQ(γ11)gLγ2n=p(ρQ(γ11)gLγ2pn)n=pb1,p1.(ΛLp)n=Λγ11.L\displaystyle\rho_{Q}(\gamma_{1}^{-1})g_{L}\gamma_{2}\mathbb{Z}^{n}=\bigcap_{p}(\rho_{Q}(\gamma_{1}^{-1})g_{L}\gamma_{2}\mathbb{Z}_{p}^{n})\cap\mathbb{Q}^{n}=\bigcap_{p}b_{1,p}^{-1}.(\Lambda_{L}\otimes\mathbb{Z}_{p})\cap\mathbb{Q}^{n}=\Lambda_{\gamma_{1}^{-1}.L}

by the third property of ΛL\Lambda_{L} in Proposition 6.6. This shows that

Φ(𝗌Lb𝐆())=[γ11.L,Λγ11.L]\displaystyle\Phi(\mathsf{s}_{L}b\mathbf{G}(\mathbb{Q}))=[\gamma_{1}^{-1}.L,\Lambda_{\gamma_{1}^{-1}.L}]_{\star}

and hence the proposition follows. ∎

Remark 7.2 (Equivalence class induced by packets).

Note that for any two L,LL,L^{\prime} of discriminant DD the sets P(L),P(L)\mathrm{P}(L),\mathrm{P}(L^{\prime}) are either equal or disjoint. Indeed, these sets are equivalence classes for an equivalence relation which is implicitly stated in the proof of Proposition 7.1; see also Remark 6.7.

We analyze the fibers of the map Φ\Phi when restricted to the piece of the homogeneous set 𝗌L𝚫𝐇L(𝔸)𝐆()\mathsf{s}_{L}\mathbf{\Delta H}_{L}(\mathbb{A})\mathbf{G}(\mathbb{Q}) in the open set 𝒰\mathcal{U}. We set for any LGrn,k()L\in\mathrm{Gr}_{n,k}(\mathbb{Q})

ΔHLcpt={h𝚫𝐇L(𝔸):h1𝐇L(×^)}.\displaystyle\Delta H^{\mathrm{cpt}}_{L}=\{h\in\mathbf{\Delta H}_{L}(\mathbb{A})\mathrel{\mathop{\mathchar 58\relax}}h_{1}\in\mathbf{H}_{L}(\mathbb{R}\times\widehat{\mathbb{Z}})\}.

We remark that ΔHLcpt\Delta H^{\mathrm{cpt}}_{L} is not equal to 𝚫𝐇L(×^)\mathbf{\Delta H}_{L}(\mathbb{R}\times\widehat{\mathbb{Z}}) as gLg_{L} can have denominators (cf. (2.2)).

Lemma 7.3.

Let x,y𝚫𝐇L(𝔸)𝐆()𝒰x,y\in\mathbf{\Delta H}_{L}(\mathbb{A})\mathbf{G}(\mathbb{Q})\cap\mathcal{U}. Then

Φ(𝗌Lx)=Φ(𝗌Ly)yΔHLcptx.\Phi(\mathsf{s}_{L}x)=\Phi(\mathsf{s}_{L}y)\iff y\in\Delta H^{\mathrm{cpt}}_{L}x.
Proof.

We fix representatives bx𝐆(×^)b^{x}\in\mathbf{G}(\mathbb{R}\times\widehat{\mathbb{Z}}) of xx and by𝐆(×^)b^{y}\in\mathbf{G}(\mathbb{R}\times\widehat{\mathbb{Z}}) of yy. Moreover, we write bx=hxγxb^{x}=h^{x}\gamma^{x} and by=hyγyb^{y}=h^{y}\gamma^{y} with hx,hy𝚫𝐇L(𝔸)h^{x},h^{y}\in\mathbf{\Delta H}_{L}(\mathbb{A}) and γx,γy𝐆()\gamma^{x},\gamma^{y}\in\mathbf{G}(\mathbb{Q}). The direction ”\Leftarrow” is straightforward to verify; we leave it to the reader.

Assume that Φ(𝗌Lx)=Φ(𝗌Ly)\Phi(\mathsf{s}_{L}x)=\Phi(\mathsf{s}_{L}y). We recall from Proposition 7.1 and its proof that

Φ(𝗌Lx)=[(γ1x)1.L,Λ(γ1x)1.L]\displaystyle\Phi(\mathsf{s}_{L}x)=[(\gamma_{1}^{x})^{-1}.L,\Lambda_{(\gamma_{1}^{x})^{-1}.L}]_{\star}

and similarly for Φ(𝗌Ly)\Phi(\mathsf{s}_{L}y). By assumption, we have that there exists ηSpinQ()\eta\in\mathrm{Spin}_{Q}(\mathbb{Z}) such that η(γ1x)1.L=(γ1y)1.L\eta(\gamma_{1}^{x})^{-1}.L=(\gamma_{1}^{y})^{-1}.L. Therefore, γ1yη(γ1x)1𝐇L()\gamma_{1}^{y}\eta(\gamma_{1}^{x})^{-1}\in\mathbf{H}_{L}(\mathbb{Q}) and we obtain that

SpinQ(×^)b1xη(b1y)1=h1xγ1xη(γ1y)1(h1y)1𝐇L(𝔸).\displaystyle\mathrm{Spin}_{Q}(\mathbb{R}\times\widehat{\mathbb{Z}})\ni b_{1}^{x}\eta(b_{1}^{y})^{-1}=h_{1}^{x}\gamma_{1}^{x}\eta(\gamma_{1}^{y})^{-1}(h_{1}^{y})^{-1}\in\mathbf{H}_{L}(\mathbb{A}).

The element h=(h1,gL1ρQ(h1)gL)ΔHLcpth=(h_{1},g_{L}^{-1}\rho_{Q}(h_{1})g_{L})\in\Delta H^{\mathrm{cpt}}_{L} corresponding to h1=b1xη(b1y)1𝐇L(×^)h_{1}=b_{1}^{x}\eta(b_{1}^{y})^{-1}\in\mathbf{H}_{L}(\mathbb{R}\times\widehat{\mathbb{Z}}) satisfies hy=xhy=x. To see this, note that

hy=hby𝐆()=(b1xη(b1y)1b1ySpinQ(),gL1ρQ(b1xη(b1y)1)gLb2y𝐏n,k()).\displaystyle hy=hb^{y}\mathbf{G}(\mathbb{Q})=(b_{1}^{x}\eta(b_{1}^{y})^{-1}b_{1}^{y}\mathrm{Spin}_{Q}(\mathbb{Q}),g_{L}^{-1}\rho_{Q}(b_{1}^{x}\eta(b_{1}^{y})^{-1})g_{L}b_{2}^{y}\mathbf{P}_{n,k}(\mathbb{Q})).

For the first component we have b1xη(b1y)1b1ySpinQ()=b1xSpinQ()b_{1}^{x}\eta(b_{1}^{y})^{-1}b_{1}^{y}\mathrm{Spin}_{Q}(\mathbb{Q})=b_{1}^{x}\mathrm{Spin}_{Q}(\mathbb{Q}) because ηSpinQ()\eta\in\mathrm{Spin}_{Q}(\mathbb{Z}). For the second component, we first recall that

b2y=h2yγ2y=gL1ρQ(t1y)gLγ2yandb1xη(b1y)1=h1xγ1xη(γ1y)1(t1y)1.\displaystyle b_{2}^{y}=h_{2}^{y}\gamma_{2}^{y}=g_{L}^{-1}\rho_{Q}(t_{1}^{y})g_{L}\gamma_{2}^{y}\quad\text{and}\quad b_{1}^{x}\eta(b_{1}^{y})^{-1}=h_{1}^{x}\gamma_{1}^{x}\eta(\gamma_{1}^{y})^{-1}(t_{1}^{y})^{-1}.

We may therefore rewrite

gL1ρQ(b1xη(b1y)1)gLb2y𝐏n,k()=gL1ρQ(h1xγ1xη(γ1y)1)gLγ2y𝐏n,k().\displaystyle g_{L}^{-1}\rho_{Q}(b_{1}^{x}\eta(b_{1}^{y})^{-1})g_{L}b_{2}^{y}\mathbf{P}_{n,k}(\mathbb{Q})=g_{L}^{-1}\rho_{Q}(h_{1}^{x}\gamma_{1}^{x}\eta(\gamma_{1}^{y})^{-1})g_{L}\gamma_{2}^{y}\mathbf{P}_{n,k}(\mathbb{Q}).

Using that γ2y𝐏n,k()\gamma_{2}^{y}\in\mathbf{P}_{n,k}(\mathbb{Q}) and h2x=gL1ρQ(h1x)gLh_{2}^{x}=g_{L}^{-1}\rho_{Q}(h_{1}^{x})g_{L} we obtain:

gL1ρQ(h1xγ1xη(γ1y)1)gLγ2y𝐏n,k()=h2xgL1ρQ(γ1xη(γ1y)1)gL𝐏n,k().\displaystyle g_{L}^{-1}\rho_{Q}(h_{1}^{x}\gamma_{1}^{x}\eta(\gamma_{1}^{y})^{-1})g_{L}\gamma_{2}^{y}\mathbf{P}_{n,k}(\mathbb{Q})=h_{2}^{x}g_{L}^{-1}\rho_{Q}(\gamma_{1}^{x}\eta(\gamma_{1}^{y})^{-1})g_{L}\mathbf{P}_{n,k}(\mathbb{Q}).

Finally, gL1ρQ(γ1xη(γ1y)1)gL𝐏n,k()g_{L}^{-1}\rho_{Q}(\gamma_{1}^{x}\eta(\gamma_{1}^{y})^{-1})g_{L}\in\mathbf{P}_{n,k}(\mathbb{Q}) because γ1xη(γ1y)1\gamma_{1}^{x}\eta(\gamma_{1}^{y})^{-1} stabilizes LL, thus

h2xgL1ρQ(γ1xη(γ1y)1)gL𝐏n,k()=h2x𝐏n,k()=b2x𝐏n,k().\displaystyle h_{2}^{x}g_{L}^{-1}\rho_{Q}(\gamma_{1}^{x}\eta(\gamma_{1}^{y})^{-1})g_{L}\mathbf{P}_{n,k}(\mathbb{Q})=h_{2}^{x}\mathbf{P}_{n,k}(\mathbb{Q})=b_{2}^{x}\mathbf{P}_{n,k}(\mathbb{Q}).

It follows that hx=yhx=y and the proof is complete. ∎

7.4. The correct weights

Let μL\mu_{L} be the Haar probability measure on the orbit 𝗌L𝚫𝐇L(𝔸)𝐆()𝐆(𝔸)/𝐆()\mathsf{s}_{L}\mathbf{\Delta H}_{L}(\mathbb{A})\mathbf{G}(\mathbb{Q})\subset\mathbf{G}(\mathbb{A})/\mathbf{G}(\mathbb{Q}) and let μL|𝒰\mu_{L}|_{\mathcal{U}} be the normalized restriction111111Note that the normalized restriction is well-defined (i.e. μL(𝒰)0\mu_{L}(\mathcal{U})\neq 0) as the intersection 𝗌L𝚫𝐇L(𝔸)𝐆()𝒰\mathsf{s}_{L}\mathbf{\Delta H}_{L}(\mathbb{A})\mathbf{G}(\mathbb{Q})\cap\mathcal{U} contains 𝗌L(𝚫𝐇L(𝔸)𝐆(×^))𝐆()\mathsf{s}_{L}(\mathbf{\Delta H}_{L}(\mathbb{A})\cap\mathbf{G}(\mathbb{R}\times\widehat{\mathbb{Z}}))\mathbf{G}(\mathbb{Q}) which is open in 𝗌L𝚫𝐇L(𝔸)𝐆()\mathsf{s}_{L}\mathbf{\Delta H}_{L}(\mathbb{A})\mathbf{G}(\mathbb{Q}). to 𝒰\mathcal{U}.

We compute the measure of a fiber through any point x𝒰x\in\mathcal{U} in the packet.

Lemma 7.4.

Let x𝚫𝐇L(𝔸)𝐆()𝒰x\in\mathbf{\Delta H}_{L}(\mathbb{A})\mathbf{G}(\mathbb{Q})\cap\mathcal{U} and write Φ(𝗌Lx)=[L^,ΛL^]\Phi(\mathsf{s}_{L}x)=[\hat{L},\Lambda_{\hat{L}}]_{\star}. Then

(7.7) μL|𝒰(𝗌LΔHLcptx)=([L,ΛL]P(L)|𝐇L^()||𝐇L()|)1.\displaystyle\mu_{L}|_{\mathcal{U}}\big{(}\mathsf{s}_{L}\Delta H^{\mathrm{cpt}}_{L}x\big{)}=\Big{(}\sum_{[L^{\prime},\Lambda_{L^{\prime}}]_{\star}\in\mathrm{P}(L)}\frac{|\mathbf{H}_{\hat{L}}(\mathbb{Z})|}{|\mathbf{H}_{L^{\prime}}(\mathbb{Z})|}\Big{)}^{-1}.
Proof.

We must trace through a normalization: let mm be the Haar measure on 𝚫𝐇L(𝔸)\mathbf{\Delta H}_{L}(\mathbb{A}) induced by requiring that μL\mu_{L} is a probability measure and let C1=m(ΔHLcpt)C_{1}=m(\Delta H^{\mathrm{cpt}}_{L}). Then

(7.8) μL(𝗌LΔHLcptx)=C1|StabΔHLcpt(x)|.\displaystyle\mu_{L}\big{(}\mathsf{s}_{L}\Delta H^{\mathrm{cpt}}_{L}x\big{)}=\frac{C_{1}}{|\mathrm{Stab}_{\Delta H^{\mathrm{cpt}}_{L}}(x)|}.

We compute the stabilizer. Write x=b𝐆()x=b\mathbf{G}(\mathbb{Q}) for some b𝐆(×^)b\in\mathbf{G}(\mathbb{R}\times\widehat{\mathbb{Z}}) and observe

(7.9) StabΔHLcpt(x)=bStabΔHL^cpt(𝐆())b1\displaystyle\mathrm{Stab}_{\Delta H^{\mathrm{cpt}}_{L}}(x)=b\mathrm{Stab}_{\Delta H^{\mathrm{cpt}}_{\hat{L}}}(\mathbf{G}(\mathbb{Q}))b^{-1}

as L^=b1,1.L\hat{L}=b^{-1}_{1,\infty}.L. The intersection ΔHL^cpt𝐆()\Delta H^{\mathrm{cpt}}_{\hat{L}}\cap\mathbf{G}(\mathbb{Q}) consists of rational elements gg of 𝚫𝐇L^()\mathbf{\Delta H}_{\hat{L}}(\mathbb{Q}) whose first component g1g_{1} is in SpinQ(×^)\mathrm{Spin}_{Q}(\mathbb{R}\times\widehat{\mathbb{Z}}). Equivalently, it is the subgroup of 𝚫𝐇L^()\mathbf{\Delta H}_{\hat{L}}(\mathbb{Q}) of elements gg with g1SpinQ()g_{1}\in\mathrm{Spin}_{Q}(\mathbb{Z}) which is clearly isomorphic to 𝐇L^()\mathbf{H}_{\hat{L}}(\mathbb{Z}). In particular,

|StabΔHLcpt(x)|=|𝐇L^()|.\displaystyle|\mathrm{Stab}_{\Delta H^{\mathrm{cpt}}_{L}}(x)|=|\mathbf{H}_{\hat{L}}(\mathbb{Z})|.

We now use that the one-to-one correspondence between P(L)\mathrm{P}(L) and ΔHLcpt\Delta H^{\mathrm{cpt}}_{L}-orbits in 𝚫𝐇L(𝔸)𝐆()𝒰\mathbf{\Delta H}_{L}(\mathbb{A})\mathbf{G}(\mathbb{Q})\cap\mathcal{U} (Lemma 7.3). By summing (7.7) over all such orbits we obtain

μL(𝒰)=[L,ΛL]P(L)C1|𝐇L()|\displaystyle\mu_{L}(\mathcal{U})=\sum_{[L^{\prime},\Lambda_{L^{\prime}}]_{\star}\in\mathrm{P}(L)}\frac{C_{1}}{|\mathbf{H}_{L^{\prime}}(\mathbb{Z})|}

which determines C1C_{1}. This concludes the lemma as by (7.8) and (7.9)

μL|𝒰(𝗌LΔHLcptx)=C1μL(𝒰)1|𝐇L^()|1.\displaystyle\mu_{L}|_{\mathcal{U}}\big{(}\mathsf{s}_{L}\Delta H^{\mathrm{cpt}}_{L}x\big{)}=C_{1}\mu_{L}(\mathcal{U})^{-1}|\mathbf{H}_{\hat{L}}(\mathbb{Z})|^{-1}.

7.4.1. Measures on SpinQ()\Qn,k(D)\mathchoice{\text{\lower 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Z})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{R}^{n,k}_{Q}(D)$}}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{R}^{n,k}_{Q}(D)}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{R}^{n,k}_{Q}(D)}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{R}^{n,k}_{Q}(D)}

We have different measures on the set of cosets SpinQ()\Qn,k(D)\mathchoice{\text{\lower 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Z})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{R}^{n,k}_{Q}(D)$}}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{R}^{n,k}_{Q}(D)}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{R}^{n,k}_{Q}(D)}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{R}^{n,k}_{Q}(D)}:

  • νD\nu_{D} is the pushforward of the normalized sum of Dirac measures on Qn,k(D)\mathcal{R}^{n,k}_{Q}(D).

  • For any LnL\subset\mathbb{Q}^{n} oriented kk-dimensional with discQ(L)=D\mathrm{disc}_{Q}(L)=D the measure νP(L)\nu_{\mathrm{P}(L)} is the pushforward of μL|𝒰\mu_{L}|_{\mathcal{U}} under the map Φ\Phi defined in (7.4). Here, the collection P(L)\mathrm{P}(L) is defined in (7.5).

We claim that νD\nu_{D} is a convex combination of the measures νP(L)\nu_{\mathrm{P}(L)} for LL varying with discriminant DD. The weights of the above measures may be computed explicitly. Beginning with the former, note that the mass νD\nu_{D} gives to a point [L^,ΛL^]SpinQ()\Qn,k(D)[\hat{L},\Lambda_{\hat{L}}]_{\star}\in\mathchoice{\text{\lower 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Z})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{R}^{n,k}_{Q}(D)$}}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{R}^{n,k}_{Q}(D)}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{R}^{n,k}_{Q}(D)}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{R}^{n,k}_{Q}(D)} is, up to a fixed scalar multiple, the number of preimages of [L^,ΛL^][\hat{L},\Lambda_{\hat{L}}]_{\star} under the quotient map Qn,k(D)SpinQ()\Qn,k(D)\mathcal{R}^{n,k}_{Q}(D)\rightarrow\mathchoice{\text{\lower 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Z})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{R}^{n,k}_{Q}(D)$}}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{R}^{n,k}_{Q}(D)}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{R}^{n,k}_{Q}(D)}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{R}^{n,k}_{Q}(D)}. In other words, it is a constant times

#{[g.L^,Λg.L^]:gSpinQ()}=#{g.L^:gSpinQ()}=|SpinQ()||𝐇L^()|.\displaystyle\#\{[g.\hat{L},\Lambda_{g.\hat{L}}]\mathrel{\mathop{\mathchar 58\relax}}g\in\mathrm{Spin}_{Q}(\mathbb{Z})\}=\#\{g.\hat{L}\mathrel{\mathop{\mathchar 58\relax}}g\in\mathrm{Spin}_{Q}(\mathbb{Z})\}=\frac{|\mathrm{Spin}_{Q}(\mathbb{Z})|}{|\mathbf{H}_{\hat{L}}(\mathbb{Z})|}.

By the same argument as in Lemma 7.4, we have (as |SpinQ()||\mathrm{Spin}_{Q}(\mathbb{Z})| cancels out)

(7.10) νD([L^,ΛL^])=([L,ΛL]SpinQ()\Qn,k(D)1|𝐇L()|)11|𝐇L^()|\displaystyle\nu_{D}\big{(}[\hat{L},\Lambda_{\hat{L}}]_{\star}\big{)}=\Big{(}\sum_{[L^{\prime},\Lambda_{L^{\prime}}]_{\star}\in\mathchoice{\text{\lower 1.50694pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Z})$}\big{\backslash}\raise 1.50694pt\hbox{$\mathcal{R}^{n,k}_{Q}(D)$}}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{R}^{n,k}_{Q}(D)}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{R}^{n,k}_{Q}(D)}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{R}^{n,k}_{Q}(D)}}\frac{1}{|\mathbf{H}_{L^{\prime}}(\mathbb{Z})|}\Big{)}^{-1}\frac{1}{|\mathbf{H}_{\hat{L}}(\mathbb{Z})|}

On the other hand, the measure νP(L)\nu_{\mathrm{P}(L)} satisfies for any [L^,ΛL^]P(L)[\hat{L},\Lambda_{\hat{L}}]_{\star}\in\mathrm{P}(L)

(7.11) νP(L)([L^,ΛL^])=([L,ΛL]P(L)1|𝐇L()|)11|𝐇L^()|\displaystyle\nu_{\mathrm{P}(L)}\big{(}[\hat{L},\Lambda_{\hat{L}}]_{\star}\big{)}=\Big{(}\sum_{[L^{\prime},\Lambda_{L^{\prime}}]_{\star}\in\mathrm{P}(L)}\frac{1}{|\mathbf{H}_{L^{\prime}}(\mathbb{Z})|}\Big{)}^{-1}\frac{1}{|\mathbf{H}_{\hat{L}}(\mathbb{Z})|}

by Lemma 7.4.

Thus, the relative weights the measures νD\nu_{D} and νP(L)\nu_{\mathrm{P}(L)} assign agree. It follows from Remark 7.2 and (7.11) and (7.10) that νD\nu_{D} is a convex combination of the measures νP(L)\nu_{\mathrm{P}(L)} as claimed.

7.5. Conclusion

We now prove the remaining theorems. We proved in §6.4 that Theorem 6.9 implies Theorem 1.4 when k>2k>2 and Theorem 1.11. So it is left to prove Theorem 6.9 and Theorem 1.4 when k=2k=2.

Proof of Theorem 6.9.

The key insight is that νDi\nu_{D_{i}} is a convex combination of measures that are equidistributed along any sequence of admissible subspaces. The assumption of DiD_{i} to be kk-power free implies admissibility.

Let pp be an odd prime not dividing disc(Q)\mathrm{disc}(Q) and let DiD_{i}\to\infty be a sequence of integers as in the assumptions of the theorem for the prime pp. We first claim that any sequence LiQn,k(Di)L_{i}\in\mathcal{H}_{Q}^{n,k}(D_{i}) is admissible (cf. §3). Observe first that Condition (1) is automatic. Also, the assumption Di[k]D_{i}^{[k]}\to\infty implies Condition (2). By Proposition 5.1 and nkkn-k\geq k we have

disc(Li)[nk]disc(Li)[k]QDi[k]\displaystyle\mathrm{disc}(L_{i}^{\perp})^{[n-k]}\geq\mathrm{disc}(L_{i}^{\perp})^{[k]}\asymp_{Q}D_{i}^{[k]}

proving Condition (3). Lastly, Condition (4) follows from Propositions 5.1 and 2.9 (where the former implies pdiscQ(L)p\nmid\mathrm{disc}_{Q}(L^{\perp})).

For any sequence LiL_{i} as above together with an additional given orientation the measures νP(Li)\nu_{\mathrm{P}(L_{i})} equidistribute to the Haar measure on SpinQ()\𝒴+\mathchoice{\text{\lower 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Z})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{Y}^{+}$}}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{Y}^{+}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{Y}^{+}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{Y}^{+}}. Indeed, by admissibility the measures μLi\mu_{L_{i}} converge to the Haar measure μ\mu on 𝐆(𝔸)/𝐆()\mathchoice{\text{\raise 2.15277pt\hbox{$\mathbf{G}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathbf{G}(\mathbb{Q})$}}}{\mathbf{G}(\mathbb{A})\,/\,\mathbf{G}(\mathbb{Q})}{\mathbf{G}(\mathbb{A})\,/\,\mathbf{G}(\mathbb{Q})}{\mathbf{G}(\mathbb{A})\,/\,\mathbf{G}(\mathbb{Q})} by Theorem 3.1. In particular, as 𝒰\mathcal{U} is compact open we have μLi|𝒰μ|𝒰\mu_{L_{i}}|_{\mathcal{U}}\to\mu|_{\mathcal{U}}. Taking the pushforward under Φ\Phi yields νP(Li)ν\nu_{\mathrm{P}(L_{i})}\to\nu where ν\nu is the Haar measure on SpinQ()\𝒴+\mathchoice{\text{\lower 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Z})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{Y}^{+}$}}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{Y}^{+}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{Y}^{+}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{Y}^{+}}.

The fact that νDi\nu_{D_{i}} is a convex combination of the measures νP(Li)\nu_{\mathrm{P}(L_{i})} finally implies Theorem 6.9. ∎

Proof of Theorem 1.4 for k=2k=2.

Let 𝒰¯\bar{\mathcal{U}} be the principal genus of 𝐆¯(𝔸)/𝐆¯()\mathchoice{\text{\raise 2.15277pt\hbox{$\bar{\mathbf{G}}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\bar{\mathbf{G}}(\mathbb{Q})$}}}{\bar{\mathbf{G}}(\mathbb{A})\,/\,\bar{\mathbf{G}}(\mathbb{Q})}{\bar{\mathbf{G}}(\mathbb{A})\,/\,\bar{\mathbf{G}}(\mathbb{Q})}{\bar{\mathbf{G}}(\mathbb{A})\,/\,\bar{\mathbf{G}}(\mathbb{Q})}. The following diagram commutes by construction:

𝐆(𝔸)/𝐆()𝒰{\mathchoice{\text{\raise 2.15277pt\hbox{$\mathbf{G}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\mathbf{G}(\mathbb{Q})$}}}{\mathbf{G}(\mathbb{A})\,/\,\mathbf{G}(\mathbb{Q})}{\mathbf{G}(\mathbb{A})\,/\,\mathbf{G}(\mathbb{Q})}{\mathbf{G}(\mathbb{A})\,/\,\mathbf{G}(\mathbb{Q})}\supset\mathcal{U}}𝒰¯𝐆¯(𝔸)/𝐆¯(){\bar{\mathcal{U}}\subset\mathchoice{\text{\raise 2.15277pt\hbox{$\bar{\mathbf{G}}(\mathbb{A})$}\big{/}\lower 2.15277pt\hbox{$\bar{\mathbf{G}}(\mathbb{Q})$}}}{\bar{\mathbf{G}}(\mathbb{A})\,/\,\bar{\mathbf{G}}(\mathbb{Q})}{\bar{\mathbf{G}}(\mathbb{A})\,/\,\bar{\mathbf{G}}(\mathbb{Q})}{\bar{\mathbf{G}}(\mathbb{A})\,/\,\bar{\mathbf{G}}(\mathbb{Q})}}SpinQ()\𝒴+{\mathchoice{\text{\lower 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Z})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathcal{Y}^{+}$}}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{Y}^{+}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{Y}^{+}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathcal{Y}^{+}}}SpinQ()\Grn,2()×𝒮2×𝒮n2.{\mathchoice{\text{\lower 2.15277pt\hbox{$\mathrm{Spin}_{Q}(\mathbb{Z})$}\big{\backslash}\raise 2.15277pt\hbox{$\mathrm{Gr}_{n,2}(\mathbb{R})$}}}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathrm{Gr}_{n,2}(\mathbb{R})}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathrm{Gr}_{n,2}(\mathbb{R})}{\mathrm{Spin}_{Q}(\mathbb{Z})\,\backslash\,\mathrm{Gr}_{n,2}(\mathbb{R})}\times\mathcal{S}_{2}\times\mathcal{S}_{n-2}.}

By Theorem 4.1, the images of 𝗌Li𝚫𝐇Li(𝔸)𝐆()𝒰\mathsf{s}_{L_{i}}\mathbf{\Delta H}_{L_{i}}(\mathbb{A})\mathbf{G}(\mathbb{Q})\cap\mathcal{U} in 𝒰¯\bar{\mathcal{U}} along any admissible sequence of subspaces LiL_{i} are equidistributed. On the other hand, by the above commutative diagram these images are given by the images of P(Li)\mathrm{P}(L_{i}) under the bottom map. The rest of the argument is analogous to the case k>2k>2. ∎

Appendix A Non-emptiness for the sum of squares

In this section, we discuss non-emptiness conditions for the set Qn,k(D)\mathcal{H}^{n,k}_{Q}(D) when QQ is the sum of squares. To simplify notation, we write n,k(D)\mathcal{H}^{n,k}(D). Note that we have a bijection

Ln,k(D)Ln,nk(D)\displaystyle L\in\mathcal{H}^{n,k}(D)\mapsto L^{\perp}\in\mathcal{H}^{n,n-k}(D)

as QQ is unimodular (see Proposition 5.4 and its corollary). In view of our goal, we will thus assume that knkk\leq n-k throughout. We will also suppose that nk2n-k\geq 2.

The question of when n,k(D)\mathcal{H}^{n,k}(D) is non-empty is a very classical problem in number theory, particularly if k=1k=1. Here, note that n,1(D)\mathcal{H}^{n,1}(D) is non-empty if and only if there exists a primitive vector vnv\in\mathbb{Z}^{n} with Q(v)=DQ(v)=D (i.e. DD is primitively represented as a sum of nn squares).

  • For n=3n=3, Legendre proved, assuming the existence of infinitely many primes in arithmetic progression, that 3,1(D)\mathcal{H}^{3,1}(D) is non-empty if and only if D0,4,7mod8D\not\equiv 0,4,7\mod 8. A complete proof was later given by Gauss [gauss]; we shall nevertheless refer to this result as Legendre’s three squares theorem.

  • For n=4n=4, Lagrange’s four squares theorem states that 4,1(D)\mathcal{H}^{4,1}(D) is non-empty if and only if D0mod8D\not\equiv 0\mod 8.

  • For n5n\geq 5, we have 5,1(D)\mathcal{H}^{5,1}(D)\neq\emptyset for all DD\in\mathbb{N} as one can see from Lagrange’s four square theorem. Indeed, if D0mod8D\not\equiv 0\mod 8 the integer DD is primitively represented as a sum of four squares and hence also of nn squares (by adding zeros). If D0mod8D\equiv 0\mod 8 one can primitively represent D1D-1 as sum of four squares which yields a primitive representation of DD as sum of five squares.

When k=2k=2, this question has been studied by Mordell [mordell1, mordell2] and Ko [ko]. In [2in4], the first and last named authors have shown together with Einsiedler that

(A.1) 4,2(D)D0,7,12,15mod16.\displaystyle\mathcal{H}^{4,2}(D)\neq\emptyset\iff D\not\equiv 0,7,12,15\mod 16.

This concludes all cases with n{3,4}n\in\{3,4\}. In this appendix we show by completely elementary methods the following.

Proposition A.1.

Suppose that n5n\geq 5. Then n,k(D)\mathcal{H}^{n,k}(D) is non-empty.

We first claim that it suffices to show that 5,2(D)\mathcal{H}^{5,2}(D) is non-empty. For this, observe that there exists for any (n,k)(n,k) injective maps

(A.2) n,k(D)n+1,k(D),n,k(D)n+1,k+1(D).\displaystyle\mathcal{H}^{n,k}(D)\hookrightarrow\mathcal{H}^{n+1,k}(D),\quad\mathcal{H}^{n,k}(D)\hookrightarrow\mathcal{H}^{n+1,k+1}(D).

The first map is given by viewing Ln,k(D)L\in\mathcal{H}^{n,k}(D) as a subspace of n+1\mathbb{Q}^{n+1} via nn×{0}n+1\mathbb{Q}^{n}\to\mathbb{Q}^{n}\times\{0\}\subset\mathbb{Q}^{n+1}. The second map associates to L=v1vkn,k(D)L=\mathbb{Q}v_{1}\oplus\ldots\mathbb{Q}v_{k}\in\mathcal{H}^{n,k}(D) the subspace (v1,0)(vk,0)en+1n+1,k+1(D)\mathbb{Q}(v_{1},0)\oplus\ldots\oplus\mathbb{Q}(v_{k},0)\oplus\mathbb{Q}e_{n+1}\in\mathcal{H}^{n+1,k+1}(D). In particular, Proposition A.1 for (n,k)=(5,2)(n,k)=(5,2) implies Proposition A.1 for (n,k)=(6,2),(6,3)(n,k)=(6,2),(6,3). One then proceeds inductively to verify the claim.

A.1. A construction of Schmidt

Though it is not, strictly speaking, necessary, we introduce here a conceptual construction of Schmidt [Schmidt-count] that captures what can be done with inductive arguments as in (A.2). As before, we identify n\mathbb{Q}^{n} with a subspace of n+1\mathbb{Q}^{n+1} via nn×{0}\mathbb{Q}^{n}\simeq\mathbb{Q}^{n}\times\{0\}. Given any LGrn+1,k()L\in\mathrm{Gr}_{n+1,k}(\mathbb{Q}) we have that either the intersection LnL\cap\mathbb{Q}^{n} is (k1)(k-1)-dimensional or LL is contained in n\mathbb{Q}^{n}. In particular, we can write

n+1,k(D)=n,k(D)ndn+1,k(D)\displaystyle\mathcal{H}^{n+1,k}(D)=\mathcal{H}^{n,k}(D)\sqcup\mathcal{H}_{\mathrm{nd}}^{n+1,k}(D)

where ndn+1,k(D)\mathcal{H}_{\mathrm{nd}}^{n+1,k}(D) denotes the subspaces Ln+1,k(D)L\in\mathcal{H}^{n+1,k}(D) for which LnL\not\subset\mathbb{Q}^{n}. We also let Grn+1,knd()\mathrm{Gr}_{n+1,k}^{\mathrm{nd}}(\mathbb{Q}) be the subspaces LGrn+1,k()L\in\mathrm{Gr}_{n+1,k}(\mathbb{Q}) for which LnL\not\subset\mathbb{Q}^{n}. Here, ’nd\mathrm{nd}’ stands for ’non-degenerate’.

We now associate to LGrn+1,knd()L\in\mathrm{Gr}_{n+1,k}^{\mathrm{nd}}(\mathbb{Q}) three quantities. Let L=LnL^{\prime}=L\cap\mathbb{Q}^{n}. Furthermore, note that the projection of L()L(\mathbb{Z}) onto the xn+1x_{n+1}-axis consists of multiples of some vector (0,,0,hL)(0,\ldots,0,h_{L}) where hLh_{L}\in\mathbb{N}. As (0,,0,hL)(0,\ldots,0,h_{L}) comes from projection of L()L(\mathbb{Z}), there exists some vector (uL,hL)L()(u_{L},h_{L})\in L(\mathbb{Z}). We define vLv_{L} to be the projection of uLu_{L} onto the orthogonal complement of LL^{\prime} inside n\mathbb{Q}^{n}.

Proposition A.2 ([Schmidt-count, §5]).

The following properties hold:

  1. (i)

    For any LGrn+1,knd()L\in\mathrm{Gr}_{n+1,k}^{\mathrm{nd}}(\mathbb{Q}) the pair (hL,vL)(h_{L},v_{L}) is relatively prime in the following sense: there is no integer d>1d>1 such that d1hLd^{-1}h_{L}\in\mathbb{N} and d1vLπL(n1)d^{-1}v_{L}\in\pi_{L^{\prime\perp}}(\mathbb{Z}^{n-1}).

  2. (ii)

    Let (h,L¯,v)(h,\bar{L},v) be any triplet with hh\in\mathbb{N}, L¯Grn,k1()\bar{L}\in\mathrm{Gr}_{n,k-1}(\mathbb{Q}) and vπL¯(n1)v\in\pi_{\bar{L}}(\mathbb{Z}^{n-1}) such that (hL,vL)(h_{L},v_{L}) is relatively prime. Then there exists a unique LGrn+1,knd()L\in\mathrm{Gr}_{n+1,k}^{\mathrm{nd}}(\mathbb{Q}) with (h,L¯,v)=(hL,L,vL)(h,\bar{L},v)=(h_{L},L^{\prime},v_{L}).

  3. (iii)

    We have

    disc(L)=disc(L)(hL2+Q(vL)).\mathrm{disc}(L)=\mathrm{disc}(L^{\prime})(h_{L}^{2}+Q(v_{L})).

We remark that the construction in (ii) is quite explicit: If un1u\in\mathbb{Z}^{n-1} satisfies πL¯(u)=v\pi_{\bar{L}}(u)=v, one defines LL to be the span of L¯\bar{L} and the vector (u,h)(u,h).

To illustrate this construction we show the direction in (A.1) that we will need for Proposition A.1.

Lemma A.3.

If DD\in\mathbb{N} satisfies D0,7,12,15mod16D\not\equiv 0,7,12,15\mod 16, then 4,2(D)\mathcal{H}^{4,2}(D) is non-empty.

Proof.

By Legendre’s three squares theorem and (A.2), we have

D0,4,7mod84,2(D).\displaystyle D\not\equiv 0,4,7\mod 8\implies\mathcal{H}^{4,2}(D)\neq\emptyset.

So suppose that DD is congruent to 4,84,8 modulo 1616. In view of Proposition A.2, we let LL^{\prime} be the line through (1,1,0)(1,-1,0) so that disc(L)=2\mathrm{disc}(L^{\prime})=2. Thus, it remains to find relatively prime hh\in\mathbb{N} and vπL(3)v\in\pi_{L^{\prime}}(\mathbb{Z}^{3}) with D2=h2+Q(v)\frac{D}{2}=h^{2}+Q(v). Note that

πL(3)=e1+e22+e3\displaystyle\pi_{L^{\prime}}(\mathbb{Z}^{3})=\mathbb{Z}\frac{e_{1}+e_{2}}{2}+\mathbb{Z}e_{3}

so that we may choose v=ae1+e22+be3v=a\frac{e_{1}+e_{2}}{2}+be_{3} for a,ba,b\in\mathbb{Z}. Hence, we need to find a solution to

D2=h2+a24+a24+b2=h2+a22+b2\displaystyle\frac{D}{2}=h^{2}+\frac{a^{2}}{4}+\frac{a^{2}}{4}+b^{2}=h^{2}+\frac{a^{2}}{2}+b^{2}

such that (h,a,b)(h,a,b) is primitive.

Equivalently, this corresponds to finding a primitive representation of DD by the ternary form x12+2x22+2x32x_{1}^{2}+2x_{2}^{2}+2x_{3}^{2}. This is again a very classical problem and has been settled by Dickson [Dickson]; as the argument is very short and elementary we give it here. Note that D4\frac{D}{4} is congruent to 11 or 22 modulo 44 and hence there is (x,y,z)3(x,y,z)\in\mathbb{Z}^{3} primitive with x2+y2+z2=D4x^{2}+y^{2}+z^{2}=\frac{D}{4}. As D41,2mod4\frac{D}{4}\equiv 1,2\mod 4 at least one and at most two of the integers x,y,zx,y,z must be even. Suppose without loss of generality that xx is even and yy is odd. One checks that

D=2(x+y)2+2(xy)2+(2z)2\displaystyle D=2(x+y)^{2}+2(x-y)^{2}+(2z)^{2}

and observing that (x+y,xy,2z)(x+y,x-y,2z) is primitive as x+yx+y is odd, the claim follows in this case. ∎

Proof of Proposition A.1.

As explained, it suffices to consider the case (n,k)=(5,2)(n,k)=(5,2). In view of Lagrange’s four squares theorem and (A.2), we may suppose that D0mod8D\equiv 0\mod 8. Moreover, we can assume that D0,7,12,15mod16D\equiv 0,7,12,15\mod 16 by (A.2) and Lemma A.3. To summarize, we only need to consider the case D0mod16D\equiv 0\mod 16.

We again employ the technique in Proposition A.2. Consider the subspace L4L^{\prime}\subset\mathbb{Q}^{4} spanned by the vector (1,1,0,0)(1,-1,0,0) which has discriminant 22. Then

πL(4)=e1+e22+e3+e4\displaystyle\pi_{L^{\prime}}(\mathbb{Z}^{4})=\mathbb{Z}\frac{e_{1}+e_{2}}{2}+\mathbb{Z}e_{3}+\mathbb{Z}e_{4}

and as in the proof of Lemma A.3 we need to find a primitive representation (h,a,b,c)(h,a,b,c) of D2\frac{D}{2} as

D2=h2+a22+b2+c2.\displaystyle\frac{D}{2}=h^{2}+\frac{a^{2}}{2}+b^{2}+c^{2}.

Setting a=2a=2 and observing that D226mod8\frac{D}{2}-2\equiv 6\mod 8 the claim follows from Legendre’s three squares theorem. ∎

Appendix B More results around discriminants and induced forms

The contents of this section of the appendix are of elementary nature and complement the results in §5.

B.1. Local glue groups

In this section we briefly explain how to compute the glue group in terms of local data. This is largely in analogy to the local formula for the discriminant (1.5). Define for any prime pp

𝒢p(L)=L(p)#/L(p)\displaystyle\mathcal{G}_{p}(L)=L(\mathbb{Z}_{p})^{\#}/L(\mathbb{Z}_{p})

where we recall that L(p)=L(p)pnL(\mathbb{Z}_{p})=L(\mathbb{Q}_{p})\cap\mathbb{Z}_{p}^{n} and

L(p)#={vL(p):v,wp}.\displaystyle L(\mathbb{Z}_{p})^{\#}=\{v\in L(\mathbb{Q}_{p})\mathrel{\mathop{\mathchar 58\relax}}\langle v,w\rangle\in\mathbb{Z}_{p}\}.

Observe that 𝒢p(L)\mathcal{G}_{p}(L) is trivial for all but finitely many pp. Indeed, 𝒢p(L)\mathcal{G}_{p}(L) is trivial if LL is pp-unimodular for an odd prime pp, i.e. pdiscQ(L)p\nmid\mathrm{disc}_{Q}(L) (see also Remark B.2 for a much finer statement). Also, it is easy to adapt Lemma 5.3 and Proposition 5.4 to their local analogues. Here, we prove the following:

Lemma B.1.

We have

(B.1) 𝒢(L)p𝒢p(L).\displaystyle\mathcal{G}(L)\simeq\prod_{p}\mathcal{G}_{p}(L).

Taking cardinalities, (B.1) encodes the (obvious) local product formula for discriminants (1.5).

Proof.

The image of the natural inclusion L()L(p)L(\mathbb{Z})\hookrightarrow L(\mathbb{Z}_{p}) is dense for every pp. In particular, the image of L()#L(\mathbb{Z})^{\#} under L()L(p)L(\mathbb{Q})\hookrightarrow L(\mathbb{Q}_{p}) lies in L(p)#L(\mathbb{Z}_{p})^{\#} and is dense therein. We obtain a homomorphism ι:𝒢(L)p𝒢p(L)\iota\mathrel{\mathop{\mathchar 58\relax}}\mathcal{G}(L)\to\prod_{p}\mathcal{G}_{p}(L). We prove that ι\iota is the desired isomorphism. Let (vi)i(v_{i})_{i} be an integral basis of L()L(\mathbb{Z}).

Let v+L()v+L(\mathbb{Z}) be in the kernel of ι\iota. Then vL(p)v\in L(\mathbb{Z}_{p}) for every pp or, equivalently, the coordinates of vv in the \mathbb{Z}-basis (vi)i(v_{i})_{i} of L()L(\mathbb{Z}) have no denominators in pp for every pp. Hence vL()v\in L(\mathbb{Z}) and ι\iota is injective.

As 𝒢p(L)\mathcal{G}_{p}(L) is trivial for all but finitely many pp, it suffices to find for any vL(p)#v\in L(\mathbb{Z}_{p})^{\#} an element wL()#w\in L(\mathbb{Z})^{\#} with w+L(p)=v+L(p)w+L(\mathbb{Z}_{p})=v+L(\mathbb{Z}_{p}) and wL(q)w\in L(\mathbb{Z}_{q}) for any qpq\neq p. Let vL(p)#v\in L(\mathbb{Z}_{p})^{\#} and write v=iαiviv=\sum_{i}\alpha_{i}v_{i} where αip\alpha_{i}\in\mathbb{Q}_{p}. For every ii let βi[1p]\beta_{i}\in\mathbb{Z}[\frac{1}{p}] be such that αiβi+p\alpha_{i}\in\beta_{i}+\mathbb{Z}_{p} and set w=iβiviL()w=\sum_{i}\beta_{i}v_{i}\in L(\mathbb{Q}) as well as u=wvL(p)u=w-v\in L(\mathbb{Z}_{p}). Then clearly for every ii

w,vi=v,vi+u,vip,\displaystyle\langle w,v_{i}\rangle=\langle v,v_{i}\rangle+\langle u,v_{i}\rangle\in\mathbb{Z}_{p},

that is, wL(p)#w\in L(\mathbb{Z}_{p})^{\#}, and w,vi[1p]\langle w,v_{i}\rangle\in\mathbb{Z}[\frac{1}{p}]. But p[1p]=\mathbb{Z}_{p}\cap\mathbb{Z}[\frac{1}{p}]=\mathbb{Z} and hence wL()#w\in L(\mathbb{Z})^{\#}. Observe also that by construction wL(q)w\in L(\mathbb{Z}_{q}) for every prime qpq\neq p. Hence ι\iota is surjective. ∎

Remark B.2.

The isomorphism in (B.1) is particularly handy when one tries to explicitly compute glue-groups. Indeed, recall that for any odd prime pp an integral quadratic form qq over p\mathbb{Z}_{p} is diagonalizable [Cassels, Ch. 8]. For

q(x1,,xk)=α1p1x12++αkpkxk2\displaystyle q(x_{1},\ldots,x_{k})=\alpha_{1}p^{\ell_{1}}x_{1}^{2}+\ldots+\alpha_{k}p^{\ell_{k}}x_{k}^{2}

with units αip×\alpha_{i}\in\mathbb{Z}_{p}^{\times} and i0\ell_{i}\geq 0, the glue-group is

/p1××/pk.\displaystyle\mathbb{Z}/p^{\ell_{1}}\mathbb{Z}\times\ldots\times\mathbb{Z}/p^{\ell_{k}}\mathbb{Z}.

For p=2p=2 an integral quadratic form qq need not be diagonalizable over 2\mathbb{Z}_{2}. However, by [Cassels, Lemma 4.1] we may write qq as a (direct) sum of forms of the following types in distinct variables:

(B.2) 2αx12,2(2x1x2)and2(2x12+2x1x2+2x22)2^{\ell}\alpha x_{1}^{2},\quad 2^{\ell}(2x_{1}x_{2})\quad\text{and}\quad 2^{\ell}(2x_{1}^{2}+2x_{1}x_{2}+2x_{2}^{2})

with 0\ell\geq 0 and α2×\alpha\in\mathbb{Z}_{2}^{\times}. An elementary computation leads to observing that the glue groups of the quadratic forms in (B.2) are respectively:

(B.3) /2/2×/2and/2×/2.\mathbb{Z}/2^{\ell}\mathbb{Z}\quad\mathbb{Z}/2^{\ell}\mathbb{Z}\times\mathbb{Z}/2^{\ell}\mathbb{Z}\quad\text{and}\quad\mathbb{Z}/2^{\ell}\mathbb{Z}\times\mathbb{Z}/2^{\ell}\mathbb{Z}.

It follows that the glue group has essentially the same structure as in the case of pp odd. More precisely, assume that

q(x1,,xk)=q1++qmq(x_{1},\ldots,x_{k})=q_{1}+\cdots+q_{m}

where the qiq_{i}’s are forms as in (B.2) with exponents =i\ell=\ell_{i} satisfying 1m\ell_{1}\leq\ldots\leq\ell_{m}, Then the glue group is a product of groups as in (B.3) with exponents 1m\ell_{1}\leq\ldots\leq\ell_{m}.

B.2. Indices of projected lattices

For any subspace LnL\subset\mathbb{Q}^{n} we denote the index of L()L(\mathbb{Z}) in L(n)#L\cap(\mathbb{Z}^{n})^{\#} by i(L)i(L). Then the proof of Proposition 5.1 and Lemma 5.3 shows that

discQ(L)=i(L)i(L)discQ(L).\displaystyle\mathrm{disc}_{Q}(L^{\perp})=\frac{i(L^{\perp})}{i(L)}\mathrm{disc}_{Q}(L).

The following proposition establishes a fundamental relation between the indices for LL and LL^{\perp}.

Proposition B.3.

Let LnL\subset\mathbb{Q}^{n} be a subspace. The sequence

0(L(n)#)/L()(n)#/nL()#/πL(n)0.0\rightarrow(L^{\perp}\cap(\mathbb{Z}^{n})^{\#})/L^{\perp}(\mathbb{Z})\rightarrow(\mathbb{Z}^{n})^{\#}/\mathbb{Z}^{n}\rightarrow L(\mathbb{Z})^{\#}/\pi_{L}(\mathbb{Z}^{n})\rightarrow 0.

obtained by inclusion and projection is exact. In particular,

i(L)i(L)=disc(Q).\displaystyle i(L)i(L^{\perp})=\mathrm{disc}(Q).

Similarly, for any prime pp

[L(p)(pn)#:L(p)][L(p)(pn)#:L(p)]=pνp(disc(Q)).\displaystyle[L(\mathbb{Q}_{p})\cap(\mathbb{Z}_{p}^{n})^{\#}\mathrel{\mathop{\mathchar 58\relax}}L(\mathbb{Z}_{p})]\cdot[L^{\perp}(\mathbb{Q}_{p})\cap(\mathbb{Z}_{p}^{n})^{\#}\mathrel{\mathop{\mathchar 58\relax}}L^{\perp}(\mathbb{Z}_{p})]=p^{\nu_{p}(\mathrm{disc}(Q))}.
Proof.

By Lemma 5.3, the orthogonal projection πL\pi_{L} defines a surjective morphism

f:(n)#L()#/πL(n).f\mathrel{\mathop{\mathchar 58\relax}}(\mathbb{Z}^{n})^{\#}\rightarrow L(\mathbb{Z})^{\#}/\pi_{L}(\mathbb{Z}^{n}).

The kernel of this morphism can be described by:

(B.4) ker(f)={v(n)#:there existswnsuch thatvwL}.\ker(f)=\{v\in(\mathbb{Z}^{n})^{\#}\mathrel{\mathop{\mathchar 58\relax}}\text{there exists}\ w\in\mathbb{Z}^{n}\ \text{such that}\ v-w\in L^{\perp}\}.

Clearly L(n)#ker(f)L^{\perp}\cap(\mathbb{Z}^{n})^{\#}\subset\ker(f). We claim that the inclusion of L(n)#L^{\perp}\cap(\mathbb{Z}^{n})^{\#} into ker(f)\ker(f) induces an isomorphism:

L(n)#/L()ker(f)/n.L^{\perp}\cap(\mathbb{Z}^{n})^{\#}/L^{\perp}(\mathbb{Z})\rightarrow\ker(f)/\mathbb{Z}^{n}.

The fact that the map L(n)#ker(f)/nL^{\perp}\cap(\mathbb{Z}^{n})^{\#}\rightarrow\ker(f)/\mathbb{Z}^{n} induced by the inclusion is surjective follows immediately from the characterization of ker(f)\ker(f) in (B.4). Since the kernel of this map is clearly L()L^{\perp}(\mathbb{Z}), the claim is proven. It follows that

0L(n)#/L()(n)#/nL()#/πL(n)0.0\rightarrow L^{\perp}\cap(\mathbb{Z}^{n})^{\#}/L^{\perp}(\mathbb{Z})\rightarrow(\mathbb{Z}^{n})^{\#}/\mathbb{Z}^{n}\rightarrow L(\mathbb{Z})^{\#}/\pi_{L}(\mathbb{Z}^{n})\rightarrow 0.

is a short exact sequence. The local analogue follows similarly. ∎

Remark B.4.

It would be interesting to see statistical results regarding these indices. To give a concrete example, suppose disc(Q)=2\mathrm{disc}(Q)=2. Then clearly i(L){1,2}i(L)\in\{1,2\} for any subspace LL and one can ask what the proportion of subspaces LL with i(L)=1i(L)=1 (or i(L)=2i(L^{\perp})=2) is. If k=nkk=n-k, Proposition B.3 shows that the number of subspaces with i(L)=1i(L)=1 and i(L)=2i(L)=2 is the same.

B.3. Primitive forms

Here, we study to what extent the induced forms qL,qLq_{L},q_{L^{\perp}} (defined in §1.4.2 up to equivalence) for a given subspace LGrn,k()L\in\mathrm{Gr}_{n,k}(\mathbb{Q}) need to be primitive. For instance, we establish that for k<nkk<n-k the form qLq_{L^{\perp}} needs to be essentially primitive (while qLq_{L} does not). First, observe that indeed the form qLq_{L} need not be primitive:

Example B.5.

Let n6n\geq 6, let (ei)i=1n(e_{i})_{i=1}^{n} denote the standard basis vectors of n\mathbb{Q}^{n} and suppose that Q=Q0Q=Q_{0} is the standard positive definite form. Let (v1,v2)2(v_{1},v_{2})\in\mathbb{Z}^{2} be a primitive vector. Then the integer lattice in the subspace

L=span{v1e1+v2e2,v1e3+v2e4,v1e5+v2e6}\displaystyle L=\mathrm{span}_{\mathbb{Q}}\{v_{1}e_{1}+v_{2}e_{2},v_{1}e_{3}+v_{2}e_{4},v_{1}e_{5}+v_{2}e_{6}\}

is spanned by v1e1+v2e2,v1e3+v2e4,v1e5+v2e6v_{1}e_{1}+v_{2}e_{2},v_{1}e_{3}+v_{2}e_{4},v_{1}e_{5}+v_{2}e_{6} which are orthogonal vectors. In this basis,

qL(x1,x2,x3)=(v12+v22)(x12+x22+x32)\displaystyle q_{L}(x_{1},x_{2},x_{3})=(v_{1}^{2}+v_{2}^{2})(x_{1}^{2}+x_{2}^{2}+x_{3}^{2})

which is a highly non-primitive form. Similarly, L()L^{\perp}(\mathbb{Z}) is spanned by the integer vectors v2e11v1e2,v2e3v1e4,v2e5v1e6,e7,,env_{2}e_{1}-1v_{1}e_{2},v_{2}e_{3}-v_{1}e_{4},v_{2}e_{5}-v_{1}e_{6},e_{7},\ldots,e_{n} and hence in this basis

qL(x1,,xn3)=(v12+v22)(x12+x22+x32)+x42++xn32\displaystyle q_{L^{\perp}}(x_{1},\ldots,x_{n-3})=(v_{1}^{2}+v_{2}^{2})(x_{1}^{2}+x_{2}^{2}+x_{3}^{2})+x_{4}^{2}+\ldots+x_{n-3}^{2}

In particular, qLq_{L^{\perp}} is primitive if n>3n>3 and otherwise gcd(qL)=gcd(qL)\gcd(q_{L^{\perp}})=\gcd(q_{L}) (as qL=qLq_{L^{\perp}}=q_{L} in this specific example). This type of behavior is generally true as established below. For more examples we also refer to [2in4, Example 2.4].

Proposition B.6.

Let LGrn,k()L\in\mathrm{Gr}_{n,k}(\mathbb{Q}). If k>nkk>n-k, gcd(qL)\gcd(q_{L}) divides disc(Q)\mathrm{disc}(Q) and

disc(q~L)QdiscQ(L).\displaystyle\mathrm{disc}(\tilde{q}_{L})\asymp_{Q}\mathrm{disc}_{Q}(L).

Conversely, if k<nkk<n-k, gcd(qL)\gcd(q_{L^{\perp}}) divides disc(Q)\mathrm{disc}(Q) and disc(q~L)QdiscQ(L)\mathrm{disc}(\tilde{q}_{L^{\perp}})\asymp_{Q}\mathrm{disc}_{Q}(L).

Moreover, if k=nkk=n-k we have gcd(qL)Qgcd(qL)\gcd(q_{L})\asymp_{Q}\gcd(q_{L^{\perp}}) and

disc(q~L)Qdisc(q~L).\displaystyle\mathrm{disc}(\tilde{q}_{L})\asymp_{Q}\mathrm{disc}(\tilde{q}_{L^{\perp}}).

For convenience of the reader, we provide two proofs of the first claim in the proposition; the second uses glue groups and generalizes to k=nkk=n-k.

First proof for knkk\neq n-k.

Fix a basis v1,,vkv_{1},\ldots,v_{k} of L()L(\mathbb{Z}) and complete it into a basis v1,,vnv_{1},\ldots,v_{n} of n\mathbb{Z}^{n}. Let v1,,vnv_{1}^{*},\ldots,v_{n}^{*} be its dual basis. Since k>nkk>n-k, without loss of generality we may assume v1span(vk+1,,vn)v_{1}\in\mathrm{span}_{\mathbb{R}}(v_{k+1},\ldots,v_{n})^{\perp}. Note that v1(n)#v_{1}^{*}\in(\mathbb{Z}^{n})^{\#} and so disc(Q)v1n\mathrm{disc}(Q)v_{1}^{*}\in\mathbb{Z}^{n}. In particular, we may write

disc(Q)v1=snasvswithas.\mathrm{disc}(Q)v_{1}^{*}=\sum_{s\leq n}a_{s}v_{s}\ \text{with}\ a_{s}\in\mathbb{Z}.

By our choice of v1v_{1} we have

disc(Q)=disc(Q)v1,v1Q=skasvs,v1Q\mathrm{disc}(Q)=\langle\mathrm{disc}(Q)v_{1}^{*},v_{1}\rangle_{Q}=\sum_{s\leq k}a_{s}\langle v_{s},v_{1}\rangle_{Q}

and the first claim follows as gcd(qL)\gcd(q_{L}) divides the right-hand side. ∎

Proof.

Given a prime pp we write ordp(qL)\mathrm{ord}_{p}(q_{L}) for the largest integer mm with pmgcd(qL)p^{m}\mid\gcd(q_{L}). Note that ordp(qL)\mathrm{ord}_{p}(q_{L}) can be extracted from the glue-group of LL whenever pgcd(qL)p\mid\gcd(q_{L}) – see Remark B.2.

To begin the proof, fix pp and note that aL:=ordp(qL)a_{L}\mathrel{\mathop{\mathchar 58\relax}}=\mathrm{ord}_{p}(q_{L}) can be characterized as follows: it is the smallest integer mm so that there exists a primitive vector vL(p)#v\in L(\mathbb{Z}_{p})^{\#} with pmvL(p)p^{m}v\in L(\mathbb{Z}_{p}). To see this, first assume pp is an odd prime. Then, as in Remark B.2 (after possibly changing the basis), we may write

qL(x1,,xk)=α1p1x12++αkpkxk2\displaystyle q_{L}(x_{1},\ldots,x_{k})=\alpha_{1}p^{\ell_{1}}x_{1}^{2}+\ldots+\alpha_{k}p^{\ell_{k}}x_{k}^{2}

with 12k\ell_{1}\leq\ell_{2}\leq\ldots\leq\ell_{k}. If vv is a vector as above, the expression for the glue-group in Remark B.2 as well as primitivity imply that m1m\geq\ell_{1}. Conversely, it is easy to see that the first vector vv in the above (implicit) choice of basis of L(p)L(\mathbb{Z}_{p}) satisfies p1vL(p)#p^{-\ell_{1}}v\in L(\mathbb{Z}_{p})^{\#} and is primitive. For p=2p=2 the proof above can be adapted using Remark B.2.

Define aLa_{L}^{\prime} as the smallest integer mm so that there exists a primitive vector vπL(pn)v^{\prime}\in\pi_{L}(\mathbb{Z}_{p}^{n}) with pmvL(p)p^{m}v^{\prime}\in L(\mathbb{Z}_{p}). We argue that aLaLa^{\prime}_{L}\leq a_{L}. Let vv be as in the above definition of aLa_{L}. Then, there exists an integer iaLi\leq a_{L} such that pivπL(pn)p^{i}v\in\pi_{L}(\mathbb{Z}_{p}^{n}) and pivp^{i}v is primitive in πL(pn)\pi_{L}(\mathbb{Z}_{p}^{n}). For this integer ii set v:=pivv^{\prime}\mathrel{\mathop{\mathchar 58\relax}}=p^{i}v and observe that paLiv=paLvL(p)p^{a_{L}-i}v^{\prime}=p^{a_{L}}v\in L(\mathbb{Z}_{p}). Therefore, aLaLiaLa_{L}^{\prime}\leq a_{L}-i\leq a_{L} as claimed. In analogous fashion, one argues that aLaL+ordp(ip(L))a_{L}\leq a_{L}^{\prime}+\mathrm{ord}_{p}(i_{p}(L)) so that the following inequalities hold:

aLaLaL+ordp(ip(L)).\displaystyle a_{L}^{\prime}\leq a_{L}\leq a_{L}^{\prime}+\mathrm{ord}_{p}(i_{p}(L)).

Suppose now that k>nkk>n-k. Applying Proposition 5.4 we see that there exists vπL(pn)v^{\prime}\in\pi_{L}(\mathbb{Z}_{p}^{n}) primitive with vL(p)v^{\prime}\in L(\mathbb{Z}_{p}). Indeed, as πL(pn)/L(p)\pi_{L^{\perp}}(\mathbb{Z}_{p}^{n})/L^{\perp}(\mathbb{Z}_{p}) is a product of at most kk non-trivial cyclic groups, the same is true for πL(pn)/L(p)\pi_{L}(\mathbb{Z}_{p}^{n})/L(\mathbb{Z}_{p}) implying the claim. Therefore, aL=0a_{L}^{\prime}=0 and hence aLordp(ip(L))a_{L}\leq\mathrm{ord}_{p}(i_{p}(L)). This shows that gcd(qL)i(L)\gcd(q_{L})\mid i(L) which proves a sharpened version of the first part of the proposition (cf. Proposition B.3).

Suppose now that k=nkk=n-k. We show first that aL=aLa_{L}^{\prime}=a_{L^{\perp}}^{\prime}. If aL=0a_{L}^{\prime}=0, πL(pn)/L(p)\pi_{L}(\mathbb{Z}_{p}^{n})/L(\mathbb{Z}_{p}) is a product of at most k1k-1 cyclic groups and hence the same is true for πL(pn)/L(p)\pi_{L^{\perp}}(\mathbb{Z}_{p}^{n})/L^{\perp}(\mathbb{Z}_{p}) by Proposition 5.4. This implies that aL=0a_{L^{\perp}}^{\prime}=0. If aL0a_{L}^{\prime}\neq 0, the number aLa_{L}^{\prime} is exactly the smallest order of a non-trivial element in πL(pn)/L(p)\pi_{L}(\mathbb{Z}_{p}^{n})/L(\mathbb{Z}_{p}). Applying the same for LL^{\perp} yields aL=aLa_{L}^{\prime}=a_{L^{\perp}}^{\prime} in all cases. In particular,

aLaL+ordp(ip(L))aL+ordp(ip(L)).\displaystyle a_{L}\leq a_{L^{\perp}}^{\prime}+\mathrm{ord}_{p}(i_{p}(L))\leq a_{L^{\perp}}+\mathrm{ord}_{p}(i_{p}(L)).

Varying the prime pp we obtain that

gcd(qL)gcd(qL)i(L)\displaystyle\gcd(q_{L})\mid\gcd(q_{L^{\perp}})i(L)

and conversely. This finishes the proof of the proposition. ∎

References