An asymptotic formula for the number of integral matrices with a fixed characteristic polynomial via orbital integrals
Abstract.
For an arbitrarily given irreducible polynomial of degree with totally real, let be the number of matrices over whose characteristic polynomial is , bounded by a positive number with respect to a certain norm. In this paper, we will provide an asymptotic formula for as in terms of the orbital integrals of . This generalizes the work of A. Eskin, S. Mozes, and N. Shah (1996) which assumed that is the ring of integers.
In addition, we will provide an asymptotic formula for , using the orbital integrals of , when is generalized to a totally real number field and when is a prime number. Here we need a mild restriction on splitness of over at -adic places of for when is unramified Galois over .
Our method is based on the strong approximation property with Brauer-Manin obstruction on a variety, the formula for orbital integrals of , and the Langlands-Shelstad fundamental lemma for .
2020 Mathematics Subject Classification:
MSC11F72, 11G35, 11R37, 11S80, 14F22, 14G12, 20G30, 20G351. Introduction
For an algebraic variety defined over , understanding its -points can be viewed as a classical number-theoretic problem, studying solutions of a Diophantine equation. More generally, one can expand objects to a variety defined over , where is the ring of integers of a number field . In this context, a Diophantine analysis investigating the distribution of has been thoroughly studied for ([Bir62], [BR95], [DRS93], and [Sch85]). Here the distribution of means the asymptotic behavior of as where
with a certain norm defined on .
We concentrate on a variety which represents the set of matrices whose characteristic polynomial is , where is an irreducible monic polynomial of degree . We aim to formulate the constant and such that as , with respect to the norm given by
(1.1) |
where embeds to and denotes the set of Archimedean places of .
This is the case investigated in [EMS96, Theorem 1.1 and Theorem 1.16] under a certain restriction. More precisely, they established the following formula
(1.2) |
under the restrictions
(1.3) |
Here, for two functions and for , we denote if . is the class number of and is the regulator of , where and is its ring of integers. is the discriminant of , is the volume of the unit ball in , and .
The main goal of this paper is to weaken the first two restrictions in (1.3). More precisely, we will describe the asymptotic formula for in Theorem 7.1 (which is summarized in Theorem 1.5) using orbital integrals of , under the following restrictions
(1.4) |
This largely generalizes (1.3); if , then the above conditions (2) and (3) are unnecessary so that we do not need the condition (2) of (1.3) (cf. Remark 7.2).
In contrast to [EMS96], our method is based on the strong approximation property with Brauer-Manin obstruction suggested in [WX16]. We will also use the formula for orbital integrals of (cf. [CKL] and [CHL]) and the Langlands-Shelstad fundamental lemma for (cf. [Ngfrm[o]–0, Theorem 1]).
1.1. Backgrounds
One of the important observations to understand the asymptotic behavior of for is introduced in [BR95], which is called Hardy-Littlewood expectation. M. Borovoi and Z. Rudnick proved in [BR95, Theorem 5.3] that a particular homogeneous space of a semisimple group, including our case when , can be represented by the integration of a certain function defined on the adelic points of . However, their description of the integrand is complicated to compute directly (cf. [BR95, Section 3.5, Theorem 5.3]).
In [WX16], D. Wei and F. Xu generalized the observation of M. Borovoi and Z. Rudnick for defined over any number field . They also gave a more handleable description of the Hardy-Littlewood expectation to approximate . They mainly used the fact that satisfies the strong approximation property with Brauer-Manin obstruction following [BD13, Theorem 0.1]. This implies that the integral points of are related to the Brauer elements (cf. [LX15, Theorem 2.10]). Combining this observation with the equidistribution property in [BO12, Theorem 1.5], they obtained Proposition 1.1.
To introduce the results in [WX16], we define the following notations;
Proposition 1.1.
As a special case of Proposition 1.1, they also deduced the same results with (1.2) under the assumptions in (1.3). They observed that a non-trivial element in does not affect the summation in (1.5). Indeed, vanishes for a ramified prime number , which always exists by the assumption . After that, they computed the local integration for a trivial element in under the assumption that is the ring of integers of .
However, this method is not applicable for general number fields since the ramified place might not exist for a number field other than (cf. [WX16, Example 6.3]). In addition, the computation of the local integration is quite a challenging problem unless is the ring of integers of (cf. Proposition 5.9).
1.2. Main results
Our strategy to obtain the asymptotic formula for is based on Proposition 1.1. We first figure out the evaluation of in (1.5), and then investigate each local integration of . According to Section 2, our method depends on the -homogeneous space (resp. -homogeneous space) structure of . By Proposition 2.1, we fix and thus we have
where (resp. ) denotes the stabilizer of under the -action (resp. the -action) on .
1.2.1. Brauer evaluation of on
The Brauer evaluation of on local points is defined by using the functoriality of the Brauer group in Definition 4.3. We interpret this functoriality through the local class field theory and obtain Proposition 1.2. To establish the well-definedness of the local evaluation of , we consider the normalized evaluation (cf. Definition 4.5). In Remark 4.4, we show that this normalization does not affect the product in (1.5).
1.2.2. Computation of each local integration
To investigate the integration of for each , we will use the more handlealbe measure defined in Section 3.2.2. Here, in the equation (1.5), Proposition 3.9 enables us to interchange the measure with .
In the case of an Archimedean place, it is enough to consider the case that is trivial by the assumption (1) in (1.4). We formulate the integration as the following proposition by using the Iwasawa decomposition on .
Proposition 1.3.
(Proposition 5.6) Suppose that and are totally real number fields. Then we have
where is the volume of the unit ball in and is the discriminant of .
In the case of a non-Archimedean place, Proposition 1.2 implies that is non-trivial only if is Galois and is a Galois field extension over . In this case, we address the following cases separately; when is ramified and when is unramified.
Proposition 1.4.
Suppose .
-
(1)
(Proposition 5.9) If is trivial, then we have
Here, the right hand side is the orbital integral of for with respect to the measure .
-
(2)
If is non-trivial (and so is a Galois field extension), then we have the following equations.
-
•
(Proposition 5.11) If is a ramified Galois extension, then we have
- •
-
•
When is trivial, the formula directly follows that of the orbital integral for . Although the orbital integral, which appears in the geometric side of the Arthur-Selberg trace formula, has been studied extensively in the literature, to obtain its closed formula is quite involved.
In this paper, for the closed formula of the orbital integral for , we will refer to the results in [CKL] for the case that and the results in [CHL] for the case that is a Bass order. Here a Bass order is an order of a number field whose fractional ideals are generated by two elements (cf. [CHL, Definition 6.1]). We note that majority of number fields contain infinitely many Bass orders. For example, any order of a number field which contains the maximal order of a subfield with degree 2 or whose discriminant is fourth-power-free in , is a Bass order.
In the case that is non-trivial and is an unramified Galois extension, we use the Langlands-Shelstad fundamental lemma in [Ngfrm[o]–0, Theorem 1]. This is the reason why we need the assumption (3) in (1.4), involving the characteristic of .
1.2.3. Conclusion
Theorem 1.5 (Theorem 7.1).
Let be a number field and be an irreducible monic polynomial of degree . Let be an -scheme representing the set of matrices whose characteristic polynomial is . We define
for , where the norm is defined in (1.1).
Suppose that and are totally real, and if , we further assume that is a prime number. Then we have the following asymptotic formulas.
-
(1)
If is not Galois or ramified Galois, then
-
(2)
If is unramified Galois and splits over all -adic places for , then
Here, is formulated as follows
and we use the following notations
In Proposition 7.3 (respectively, Proposition 7.4), we will provide a closed formula for the product when or (respectively, when is a Bass order).
Remark 1.6.
According to [Yun13, Theorem 1.5], is an integer, which is a -polynomial in , whose leading term is . Since for all but finitely many places , the product is a finite product.
Remark 1.7.
In this remark, we explain how the assumptions in (1.4) affect the main steps of our argument.
-
(1)
Since we assume that is totally real, for the integration on where , we can apply the theory of -manifolds (cf. Propsition 5.6-5.8).
On the other hand, by virtue of the assumption that is totally real, Proposition 5.5 yields that the Brauer evaluations for Archimedean places are trivial. Moreover, this condition enables us to describe the stabilizer , for , in Section 3.2.2. This description affects the computations in Proposition 5.2.
-
(2)
The assumption (2) in (1.4) implies that the Brauer evaluation is non-trivial only if is an abelian extension, as explained in Proposition 5.1, which enables us to use Proposion 5.2.
This assumption also facilitates figuring out an endoscopic group associated with a local integration for , in Lemma 6.6, when is unramified.
- (3)
Organizations. We organize this paper as follows. In Section 2, we observe the structure of as a homogeneous space and investigate the related objects. In Section 3, we introduce the Tamagawa measures on homogeneous spaces and the canonical measure on which is used in our local calculations. To apply Proposition 1.3, we define the notion of Brauer evaluation in Section 4. The description of Brauer evaluation and the computation of local integrations are in Section 5-6. In Section 7, we provide the asymptotic formula for , in terms of the orbital integrals of , and we refer to the closed formula of orbital integrals of in [CKL] and [CHL] for certain cases.
Acknowledgments. We would like to thank our advisor Sungmun Cho for suggesting this problem and encouraging us, and Sug Woo Shin for helpful comments on Section 6. We also thank Seewoo Lee for meaningful discussions on the exterior product of differential forms and Fei Xu for his lecture note on Brauer-Manin obstruction, which was given at POSTECH in 2019.
1.3. Notations
-
•
Let be a number field with ring of integers , and be an algebraic closure of .
-
•
Suppose that is a commutative -algebra. We define the following schemes over .
If there is no confusion, we sometimes omit in the subscript to express schemes over .
-
•
Let be an irreducible monic polynomial of degree . We define to be the closed subscheme of representing the set of matrices whose characteristic polynomial is .
-
•
Let be the set of places and be the set of Archimedean places. For , we denote the normalized absolute value associated with by and the completion with respect to by .
-
•
For , we define the following notations
For an element , the exponential order of with respect to the maximal ideal in is written by . We then have that .
-
•
We define with ring of integers . For , we define with ring of integers .
-
•
For , let be an index set in bijection with the irreducible factors of over . For , we define the following notations
Here, we note that and .
-
•
For a finite field extension where is or , let be the discriminant ideal of .
-
–
If , then we denote by , and by the absolute value of a genearator of as an ideal in .
-
–
If , then (resp. ) denotes the exponential order (resp. the normalized absolute value) of a generator of as an ideal in .
-
–
-
•
For a field , let be the discriminant of a polynomial .
-
•
We define norms for by
where embeds to . For , we define
We note that is finite since is discrete in for each Archimendian place .
-
•
For two functions and for , we denote if .
2. Homogeneous space
2.1. Homogeneous space
In this section, we observe that is a homogeneous space of both and . We write where for . To regard as a homogeneous space of (resp. ), we fix an integral matrix whose characteristic polynomial is . For example, we can choose as the companion matrix of . We define a map
(2.1) |
We define a right -action (resp. -action) on by the conjugation , so that (resp. ) is an equivariant map under the right action of (resp. ).
Proposition 2.1.
-
(1)
is a homogeneous space of (resp. ) with respect to conjugation by (resp. ) and (resp. ) induces an following isomorphism
where (resp. ) denotes the stabilizer of under the -action (resp. the -action) on .
-
(2)
We have the following isomorphism for the stabilizer of
where we use the following notations:
Proof.
Since every square matrix over a field has a rational canonical form determined by its characteristic polynomial, for , there exists such that . Therefore, is a homogeneous space of . Similarly, we can show that is also a homogeneous space of by choosing . We then obtain the following identifications induced by and defined in (2.1),
We now prove the second statement. Since has a rational canonical form, there exists an invertible matrix such that
(2.2) |
For each , we define by
Since is a -basis of , is also a basis.
For each -algebra , we now construct an explicit bijection . Let be a free -module . With respect to the basis of , we have the following identification
Since the companion matrix in (2.2) represents the left-multiplication by with respect to the basis , we have that equals to by the change of basis.
For , in if and only if for each . We then identify the stabilizer with the set of -linear automorphisms on that commute with , under the isomorphism . By the -linearity, we also have
(2.3) |
We now define the following -linear map
where is the identity element in . The image of is contained in since, for ,
We assert that is a bijection by constructing the inverse. For , let be a matrix whose -th column consists of the coefficients of with respect to the basis . Then we have for all . Since is invertible, is contained in using the identification (2.3). Thus, defines an inverse of . One can deduce that is functorial in and this concludes that as -schemes.
Moreover, since commutes with , it follows that . ∎
2.2. Conjugacy class of and in
In this subsection, we will recall useful properties of a homogeneous space . In the previous subsection, we proved that over is identified as follows.
By [EvdGM, 4.30], the homogeneous space structure is preserved under base change and thus we have
for any place . The structures of and are described in the following lemma.
Lemma 2.2.
We have the following isomorphisms
where
Proof.
The describtion for follows from [Vos98, 3.12]. For the second description, we consider the following exact sequence of algebraic groups
(2.4) |
We claim that the norm map for an étale algebra over is the product of norms for for .
We verify the statements for on -points for each -algebra . We fix an element and define to be the left-multiplication by . Under the following composition of isomorphisms
the transfer of from to itself is also the left-multiplication by . We then have
where is the left-multiplication on by . This proves the claim and concludes the lemma. ∎
We now describe the orbits of under the -action and the -action, respectively.
Proposition 2.3.
There is only one -orbit in and thus .
Proof.
Proposition 2.4.
The set of -orbits within is in bijection with .
Proof.
We consider the following exact sequence for the homogeneous space ,
This induces the following long exact sequence
Since , it follows that the set of -orbits is in bijection with . ∎
Remark 2.5.
Definition 2.6.
We define to be the set of representatives of each -orbits in . For each , we define to be the element in corresponding to the orbit according to Proposition 2.4.
Remark 2.7.
By Proposition 2.4, we can describe as the following disjoint union.
where is the stabilizer of under the -action.
Remark 2.8.
We remark that each -orbit is open in . For each , we define the equivariant map over , by . By the same argument in Lemma 2.2, one can deduce that the stabilizer is isomorphic to and it is smooth over . Thus, is also smooth over by [EvdGM, Corollary 4.33] and is an open map by [Poo17, Proposition 3.5.73]. Therefore, the image of via , that is the -orbit of , is open in .
3. Measures on
In this section, we introduce the Tamagawa measure on , which is used to explain the Hardy-Littlewood expectation in Proposition 1.5. The Tamagawa measure is induced from a gauge form on (cf. Definition 3.1), and so we provide various concepts about volume forms.
In Section 3.1, we define the Tamagawa measure on in the general setting: is a homogeneous space over where and are assumed to be connected reductive -groups. In Section 3.2, we return to in Section 2.1, and describe the Tamgawa measure on using another handleable measure , which is defined in Section 3.2.2.
3.1. Tamagawa Measure on a homogeneous space
Let be a homogeneous space of where and are connected reductive -groups. To define the Tamagawa measure on , we will use a top-degree volume form on that is "compatible" with non-vanishing invariant top-degree volume forms on and (cf. Definition 3.2). In our case where , we will prove that the Tamagawa measure on is unique. We mainly follow concepts and backgrounds in [Wei82, Section 2] and [BR95, Section 1].
3.1.1. Gauge forms
Definition 3.1.
For a geometrically irreducible non-singular algebraic variety over a field , a gauge form on is a nowhere-zero and regular differential form in where denotes a cotangent sheaf of .
From now on, we will simply denote by . We assume that has a -point whose stabilizer is . This gives the canonical map from into ,
(3.1) |
which induces an isomorphism . Using this identification, we introduce the notion of algebraically matching for gauge forms on homogeneous spaces.
Definition 3.2 ([Wei82, Section 2.4, p24]).
Let , , and be an invariant gauge form on , an invariant gauge form on , and a -invariant gauge form on , respectively. Let be the pullback of along the morphism in (3.1), and let a differential form on be a lifting of such that, for every , induces on the form for . Then the form is a gauge form which is independent of the choice of a lifting . We say that , , and match together algebraically, if and we denote it by .
We note that and are unimodular since they are connected reductive -groups. As shown in [Wei82, Corollary of Theorem 2.2.2], and admit translation-invariant gauge forms. Moreover, admits a -invariant gauge form by [BR95, Section 1.4]. Following the arguments in [Wei82, p. 24], by multiplying a constant in on , we can assume that .
For a gauge form on a variety over , let denote the measure on induced from the canonical Haar measure on , as in [Wei82, Section 2.2.1], for each . We then have the following proposition in the sentence right below [BR95, Lemma 1.6.4].
Proposition 3.3.
Let , and be a translation-invariant gauge form on , a translation-invariant gauge form on , and -invariant gauge form on , respectively, such that . Then we have
(3.2) |
Here we note that is an open set in . For triples of measures satisfying (3.2), we say that they match together topologically.
3.1.2. Tamagawa measure
From the previous section, we choose a -invariant gauge form on . Using this gauge form, we define the Tamagawa measure on as follows.
Definition 3.4.
We define the Tamagawa measure on with respect to to be
Remark 3.5.
In general, the volume of a compact set with respect to the non-Archimedean part of the Tamagawa measure does not converge. To resolve this problem, Ono introduced the convergence factors using the Artin L-function in [Ono65]. However, in the case of in Section 2.1, the product already converges and thus we use the notion for the Tamagawa measure without convergence factors.
The definition of the Tamagawa measure might depend on the choice of a -invariant gauge form . However, the following proposition yields that the Tamgawa measure on , in Section 2.1, is uniquely defined regardless of the choice of a gauge form.
Proposition 3.6.
The Tamagawa measure on is independent of the choice of a gauge form .
Proof.
We claim that any gauge form on is unique up to a constant factor in . A gauge form defines an isomorphism of locally free -modules, by mapping . For another gauge form on , defines a nowhere-zero regular automorphism on which corresponds to an element in the unit group of the coordinate ring . Thus, we have for some . Since is a homogeneous space of and , we have by [BR95, Lemma 1.5.1].
For , we have , and the product formula yields that . Therefore, the Tamagawa measure is independent of the choice of a gauge form on . ∎
3.2. Measures on for
We now return to our main task. We note that where both and are connected reductive groups. The Tamagawa measure used in (1.5) is therefore given by , which is independent of the choice of a gauge form on by Proposition 3.6. To clarify the subsequent statements in this section, we fix a translation-invariant gauge form on , a translation-invariant gauge form on , and let be a -invariant gauge form on such that .
To compute each integration in (1.5), we need an explicit description of the measure on . To address this, we define a translation-invariant gauge form on such that the induced measure is suitable to investigate the integration of for each . After that, we will establish a relation between the Tamagawa measure and .
3.2.1. The standard integral model for for
To define a translation-invariant volume form on for , we need a precise description of an integral model for .
For , we recall that the following description of given in Lemma 2.2,
For each , the component has an integral model, given by the smooth group scheme over (cf. [KP23, B.3.(2)]). Thus, has a smooth integral structure , given by
This is refered to the standard integral model for and we note that is the maximal compact open subgroup of . Moreover, the identity component of the Néron-Raynaud model (cf. [BLR90, Chapter 6]) for coincides with the standard integral model by the description right below [Bit11, Remark 1.2].
3.2.2. Volume forms on
To define , we construct the translation-invariant volume forms on and on . Since our main computation on Archimdean places requires the assumption (1) in (1.4), we assume that both and are totally real.
For an Archimedean place , by the assumption, we have and splits completely over for all . In this case we define the volume forms on and as follows.
-
(1)
On , we define to be a translation-invariant volume form in which generates the free -module , following [GG99, Section 9]. We note that a translation-invariant volume form on is a form of
for some . To generate , the constant must be . We note that both choices induce the same measure on .
-
(2)
Since and are totally real, we have
On , we define to be a pullback of a translation-invariant volume form in which generates the free -module , following [GG99, Section 9]. We note that a translation-invariant volume form on is a form of
for some . To generate , the constant must be . We note that both choices induce the same measure on .
On the other hand, for a non-Archimedean place , we define the volume forms on and as follows.
- (1)
- (2)
In both two cases, for , we define on to be a translation-invariant volume form such that . Then by Proposition 3.3, defines a measure on which matches topologically together with and .
For , comparing with the quotient measure induced from the canonically normalized Haar measures on and on , we have the following relation.
Lemma 3.7.
Suppose that . Let be the quotient measure on , which is defined via the isomorphism , where
We have
Lemma 3.8.
We have the following equations which are independent of the choice of gauge forms and , respectively,
Proof.
For or , by [GG99, Corollary 7.3 and Proposition 9.3] we have that
which is independent of the choice of a gauge form where is the global conductor of the motive of (cf. [GG99, Equation (9.1)]).
-
(1)
Since is split over , we take a gauge form to generate the Chevalley differetials over . Then by the proof of [GG99, Proposition 9.3], we have for all and so we have .
- (2)
∎
Then Proposition 3.3 and Lemma 3.8 directly yield the following relation between two global measures and .
Proposition 3.9.
We have
4. Brauer groups and evaluations
In this section, we will explain various notions, used in the observation of D. Wei and F. Xu in Proposition 1.1. For the reader’s convenience, we restate this proposition as follows.
Proposition 4.1.
We mainly follow [Poo17, Section 8] to define the Brauer group and the evaluation of each element in the Brauer group.
Definition 4.2.
For any -scheme , the Brauer group is defined by
where stands for the étale sheaf associated with the multiplicative group over (cf. [Poo17, Proposition 6.3.19]). To ease the notation, we denote by for affine schemes.
Since gives a contravariant functor from to , each element of induces an evaluation on the set of -points , for any -algebra , as follows.
Definition 4.3.
Let be a -scheme and be a -algebra.
-
(1)
For any and , we define the Brauer evaluation to be the image of under the morphism .
-
(2)
In particular, for a local field , [Poo17, Theorem 1.5.34] yields that the invariant map is injective. In this case, for any and , we define the Brauer evaluation to be the image of under the following composition,
Remark 4.4.
We note that in Section 2.1 has a -point . Since is a -scheme, the structure morphism induces a morphism . Let be the quotient group of by the image of under this induced morphism. We remark that the summation in (4.1) is well-defined. More precisely, each infinite product over is independent of the choice of representatives of .
Since the Brauer evaluation is a group homomorphism, it suffices to show that the evaluation is constant on and the product of these constant values over is trivial. For a fixed embedding , the composition is independenet of . This concludes that is constant on by Definition 4.3.(2). Therefore it suffices to show that is trivial for , and this follows from [Poo17, Proposition 8.2.2].
For in Section 2.1, we define the normalized Brauer evaluation as follows to remove the vagueness.
Definition 4.5.
For each , we define the normalized Brauer evaluation on to be
where and is the point in chosen in Section 2.1.
Remark 4.4 indicates that the evaluation on if and belong to the same coset in . We also have that for
(4.2) | ||||
by [Poo17, Proposition 8.2.2] since is in . Therefore, we can use the normalized Brauer evaluation on (4.1).
Remark 4.6.
If is trivial in , then the normalized Brauer evaluation is trivial on by Remark 4.4. However, the normalized Brauer evaluation can be trivial even if is non-trivial in .
5. The asymptotic formula for
From now on, throughout the remaining part of this paper, we will work on the following assumptions in (1.4),
We mainly use the result in Proposition 4.1. According to the formula (4.1), we will investigate the Brauer evaluation , using the local class field theory, in Section 5.1. Based on this observation, we will investigate the integration for each place throughout Section 5.2- 5.3 and Section 6.
5.1. Evaluation of
In Definition 4.3, we described a concrete definition of the Brauer evaluation using the functor and the invariant map. Alternatively, we can explicitly formulate the Brauer evaluation using local class field theory in our case, where is a homogeneous space of . The key fact is that is a semisimple and simply connected algebraic group (cf. [CTX09, Section 2]). We divide into cases based on whether is Galois or not.
5.1.1. The case that is not a Galois extension
If is not Galois, the lemma below states that is trivial. Thus, by Remark 4.6, we do not need to consider any nontrivial Brauer evaluation in this case.
Proposition 5.1.
If is of prime degree and not Galois, then is trivial.
Proof.
Let be the Galois closure of , with and as the corresponding Galois groups. By the proof of [WX16, Theorem 6.1] and [CTX09, Proposition 2.10], we have
Suppose that there is a non-trivial element . Let be the image of via the above isomorphism, and let be the fixed field of . We claim that . Then is Galois over since is normal in , which contradicts the hypothesis. Since vanishes on , we have . The non-trivial yields that is not equal to . We then have that since is a prime number, which implies that . ∎
5.1.2. The case that is a Galois extension
Since we assume is a prime number, the field extension is an abelian extension if it is Galois. For any non-Archimedean place , if is a field, we can explictly describe the normalized Brauer evaluation.
Proposition 5.2.
Suppose that is an abelian extension. Let . Then there is an isomorphism
In particular, if is a field for , then the normalized Brauer evaluation for is given by
for . Here, is the Artin reciprocity isomorphism (cf. [Ser79, Chapter XI]) and such that .
Proof.
The first isomorphism, in , is proved in [CTX09, Proposition 2.10]. On the other hand, the isomorphism in is proved in [WX16, Theorem 6.1]. The same arguments apply for where , and it satisfies the following commutative diagram,
Therefore it suffices to consider the evaluation for the base change .
The Brauer evaluation of a homogeneous space is induced from the following commutative diagram, as shown in [CTX09, Proposition 2.9-2.10],
(5.1) |
Here, is the subgroup of the Brauer group consisting of the elements whose evaluation vanishes at . The pairing in the first row is the Brauer evaluation for , and the second row is the cup product on the group cohomology. We will explain the first vertical map later. The second vertical map is defined by
First, we explain how to operate the cup product in the second row of (5.1). Since the cup product is compatible with inflation homomorphisms, and both and split over , we consider the cup product on the group cohomology for -modules as follows,
We claim that the following diagram commutes, allowing us to use the cup product in the second row.
(5.2) |
Here, the right vertical map is obtained from the canonical pairing. and are connecting homomorphisms derived from the following exact sequences of -modules,
(5.3) |
One can show that two exact sequences in (5.3) split by choosing splitting homomorphisms and , respectively. We have that is surjective since the proof of Proposition 2.3 yields that is trivial. Since , we have
for and , by Shapiro’s lemma, and thus is an isomorphism. From the exact sequences in (5.3), we consider the diagram obtained by applying at the first row, and by applying at the second row as follows,
The first vertical map is defined by the canonical pairing of the character modules. Since and preserve split exactness, both two rows are also split exact. Thus, there is an induced central map that such that the diagram commutes, through the splitting homomorphisms. Let and be the connecting homomorphisms from the above two. Then the diagram below commutes.
For and , it follows from [Ser79, Proposition 5, VIII] that and . By the commutativity of and , and coincide in , so that the diagram (5.2) commutes.
Now we describe the map in the commutative diagram (5.1) explicitly. By Proposition 2.1, we have
(5.4) |
To identify two in the diagram, we choose for such that . Thus, represents the image of in . For , we can choose such that , which implies that . Let be the image of under
We note that the evaluation of equals the normalized evaluation . By the proof of [WX16, Theorem 6.1] that , the image of under
coincides with . Here, is a connecting homomorphism derived from the exact sequence
Therefore, the evaluation of in (5.1) is
by [Ser79, Proposition 2, XI]. ∎
Corollary 5.3.
Suppose that is an abelian extension. For , if is a field, then the normalized Brauer evaluation on is constant on each -orbits in , for any .
Proof.
For on the same -orbit, we have for some . Then, by Proposition 5.2, we have for any . ∎
Corollary 5.4.
If is a Galois extension of prime degree and is also a field extension, then for a non-trivial , the evaluation on is non-trivial.
Proof.
For the case that is not a field, the normalized Brauer evaluation is always trivial.
Proposition 5.5.
If is a Galois extension of prime degree and is not a field, then on .
Proof.
Since is a Galois extension of of prime degree, if is not a field, then splits completely over . By [CTX09, Propositoin 2.10], we have
By Lemma 2.2, is the kernel of the multiplication map , which is isomorphic to . Since splits over , is also splits over . We conclude that is trivial. If is contained in the image of , then for , is defined by the image of via
Since is constant on , is trivial on . ∎
5.2. Computation on Archimedean places
We now investigate the integration in (4.1) for Archimedean places. Since we assume that and are totally real number fields, splits completely over for any Archimedean place . Thus, by Proposition 5.5, the Brauer evaluation is trivial. It suffices to consider the volume of defined in (4.1), with respect to .
Proposition 5.6.
Suppose that and are totally real number fields. Then we have
where is the volume of the unit ball in .
Proof.
We fix an Archimedean place . Since and are totally real, we have and splits completely over . Let be the distinct roots of , and be the diagonal matrix. We denote by the stabilizer of under the conjugation of . Since the roots are distinct, the stabilizer is the set of diagonal matrices in .
To obtain the volume , we define a test function where denotes
Here, represents the unique RQ-decomposition of , where is an upper triangular matrix with positive diagonal entries, and is an orthogonal matrix in .
We note that
(5.5) |
As explained in Section 3.2.2, we have . Comparing this with , we have by . Thus, the pullback of under the isomorphism induces the Haar measure on . Therefore, through the identification (5.5), the triple of measures match together topologically by Proposition 3.3, we have
(5.6) |
First, we express the inner integration in the right hand side of (5.6) in terms of . We fix an element , with the unique RQ-decomposition . For any , the condition is equivalent that and . Thus, we have
(5.7) |
Next, we compute the left hand side of (5.6) by applying the Iwasawa decomposition
(5.8) |
where
By decomposing , we have
where represents the set of diagonal matrices with entries in and . Indeed, for , we have .
We decompose the Haar measure following [Kna02, Section 8] with respect to the decomposition (5.8). Let and be Haar measures on and defined as follows,
(5.9) |
Then [Kna02, Proposition 8.43] implies that there exists the Haar measure on such that
(5.10) |
To obtain the volume of with respect to , it suffices to compute each factor in the right hand side of the following equation,
By definition of , we have . The formulas in [EMS96, p. 282] yield that
Lemma 5.7, provided below, proves that
Lemma 5.7.
Let be the Haar measure on satisfying (5.10). Then we have
Proof.
To obtain the volume of , we introduce an alternative Haar measure on using another decomposition
(5.11) |
where denotes the set of signature matrices, and , , and are as defined in (5.8). Like defining in (5.10), we should fix the other Haar measures on , , and . We define on and on as (5.9).
We choose a Haar measure on such that
(5.12) |
where denotes the sum of positive roots in the root system of (cf. [Kna02, Proposition 8.45]). Here, denotes a Haar measure on the subgroup of consisting of unipotent lower triangular matrices. We claim that is the counting measure on so that . To express the Haar measure in terms of , , and , we apply Lemma 5.8, provided below, which describe the Jacobian of the LDU-decomposition. Let be the diagonal matrix, and (resp. ) be a unipotent lower (resp. upper) triangular matrix with entries for (resp. ). By a change of variables, we obtain
The last equality follows from the root system of , for ,
The decomposition (5.12) then implies that each element in has measure 1.
[Kna02, Proposition 8.44] implies that there exists a Haar measure on such that
(5.13) |
By [Kna02, Proposition 8.43], exchanging and in (5.10) yields
Combining this with (5.13), we have . [Kna02, Proposition 8.46] determines the normalization of ,
where is the diagonal matrix in the decomposition of with respect to (5.11) and the second equality follows from [Vos98, Theorem 1, 14.10]. Therefore, we conclude that
∎
Lemma 5.8.
Let
defined by
where (resp. ) is the unipotent lower (resp. upper) triangular matrix with the entries for (resp. ), and . Then the absolute value of the Jacobian of is .
Proof.
We use an induction on . For and , we have the Jacobian of is 1. Suppose that the lemma holds for . We denote the blocks of , , and by
Then, we have
Since the entries of are independent of , , and for , the Jacobian matrix equals . This completes the proof by the inductive assumption. ∎
5.3. Computation on non-Archimedean places
Now we consider the integration on for . In the case that the evaluation is trivial, we find that the integration is associated with the orbital integral for (cf. Proposition 5.9). Otherwise, is a Galois field extension by Proposition 5.1 and Proposition 5.5. In this case, we will address the two cases separately; when is ramified and when is unramified.
In particular, when is unramified, we use the Langlands-Shelstad fundamental lemma for , proven by Ngô in [Ngfrm[o]–0]. Because of the complexity and the various involved concepts, we will provide it in Section 6.
5.3.1. The case that the evaluation is trivial
By Proposition 5.1 and Proposition 5.5, the evaluation is trivial if the following cases hold.
Otherwise, is a Galois field extension and is non-trivial, then the evaluation is non-trivial by Proposition 5.4.
Proposition 5.9.
For , we identify with so that . Then, we have the following equation.
Here, the right hand side is the orbital integral of for with respect to the measure (See [Yun13, Section 1.3] for the definition of the orbital integral for with respect to the measure ).
Proof.
In general, obtaining the closed formula of the orbital integral for is a challenging problem, and the difficulty increases as the rank grows. In certain cases, including , we will summarize the closed formula of the orbital integrals for in Proposition 7.3-7.4.
Remark 5.10.
Suppose that is a irreducible polynomial such that (cf. assumption (2) in (1.3)). By [Yun13, Theorem 1.5], we have
(5.14) |
where the measure is defined in Lemma 3.7. Indeed, according to loc. cit. the orbtal integral of for with respect to is an integer formulated by a -polynomial in whose leading term is . Here is the -module length between and . Therefore the assumption yields that and thus (5.14) holds.
5.3.2. The case that the evaluation is non-trivial and is ramified
Proposition 5.11.
Suppose that is a ramified Galois extension of prime degree . For whose evaluation is non-trivial, we have
Proof.
[Neu99, (1.7) Proposition, Chapter 5] states that is ramified if and only if there exists an element such that . Therefore the image of under the Artin reciprocity isomorphism , defined in Proposition 5.2, is non-trivial in . On the other hand, by the assumption that is a prime number, we have that . Then the kernel of is trivial, where is defined in Proposition 5.2. We then have that .
Let . Since is a -invariant measure and is stable under the -action, we have
by Proposition 5.2. Therefore, we have
∎
6. Computation on non-Archimedean places: the case that the evaluation is non-trivial and is unramified
By Proposition 5.3, the integration of on turns to be the weighted summation for each -orbit. In the case that is unramified and is non-trivial, we will show that the enumeration of this summation follows from the Langlands-Shelstad fundamental lemma, which is proven in [Ngfrm[o]–0].
We first collect the necessary materials to state the Langlands-Shelstad fundamental lemma for in Section 6.1. Summing up our observations, we will provide the conclusion in Section 6.2. Throughout this section, we fix such that is an unramified Galois field extension.
6.1. Measures on each -orbit in
In the Langlands-Shelstad fundamental lemma, the measure defined in Lemma 6.1 is used, whereas our integration is defined with respect to the measure . Therefore, it is necessary to describe the difference between two measures and on , for each .
To address this, we will define a measure on each , which behaves as a bridge between two measures and (cf. Lemma 6.1-6.2). The construction of this measure will follow the method used in Section 3.2.2.
6.1.1. Integral models for when is an unramified field extension
As in Subsection 3.2.1, we first construct an integral model for . Since Lemma 2.2 is independent of the choice of , we have the following description of ,
By [Vos98, Section 10.5], the standard integral model for is given by which is the kernel of the following morphism,
Since is unramified, [Vos98, Theorem 2, 10.3] yields that is an -torus, with which is smooth over and has a connected fiber over each point in . Then by [Poo17, Theorem 4.3.7], these properties descend to as well and thus is also smooth over and its fiber over each point in is connected. Indeed, is a fpqc morphism. Moreover, the identity component of the Néron-Raynaud model (cf. [BLR90, Chapter 10]) for coincides with the standard integral model by [Bit11, Corollary 1.4].
6.1.2. Volume forms on
We construct translation-invariant volume forms on , on , and the corresponding -invariant volume form on as follows.
- (1)
- (2)
-
(3)
On , we define to be a -invariant volume form such that . Then by Proposition 3.3, defines a measure on which matches topologically together with and .
Comparing with the quotient measure induced from the canonically normalized Haar measures on and on , we have the following relation.
Lemma 6.1.
Let be the quotient measure on , which is defined via the isomorphism , where
We have
On the other hand, the measure coincides with the restriction of on , by the following lemma.
Lemma 6.2.
We have
Proof.
In this proof, to ease the notations, we will omit the subscript for each scheme over if there is no confusion. For the centralizer of in , we define following Section 3.2.2. Since there exists such that , we have and . Therefore and are the same measure on by Lemma 3.7. Here we note that is embedded in as the kernel of . We consider the following commutative diagram defined over ,
(6.1) |
Here the first horizontal morphism is smooth over since is unramified. On the other hand, the second horizontal morphism is also smooth over since its differential is surjective.
We first claim that there exists a translation-invariant gauge form on satisfying the following two equations, up to multiplication of a unit in ,
Since generates an -module , its lifting on (cf. Definition 3.2), which is in , does not vanish under the modulo reduction. Since generates an -module and is smooth over , a translation-invariant volume form on , such that , has good reduction (mod ). Since is smooth over , a translation-invariant gauge form such that also has good reduction (mod ) by the same argument. By [Gro97, p. 293], differs from up to multiplication of a unit in . This yields the claim.
Now we denote a lifting of a volume form on or in the sense of Definition 3.2, along the embedding into , by . Since , we have
where represents the mapping defined by . The above claim yields that
(6.2) |
where is equal to a lifting of , along . Since the right hand side of (6.2) is independent of the choice of a lifting (cf. Definition 3.2), we choose it to be the exterior product of liftings of and on . By the commutative diagram (6.1), is a lifting of along . We then have
Therefore the -form vanishes under the exterior product with . Since is of degree 1, there exists an -differential form on such that
Restricting this equation on , by plugging with , we then have the following equation up to sign
where represents the mapping defined by . This yields that on by Proposition 3.3. ∎
6.2. Computation using the Langlands-Shelstad fundamental lemma
The Langlands-Shelstad fundamental lemma, stated in [Ngfrm[o]–0, Theorem 1], represents the equation between two orbital integrals encoded by an endoscopic data (cf. [Hal05, Section 2-3] and [Kot84, Section 7.1]). We find that the evaluation on is related to an endoscopic data for , and one side of the fundamental lemma for coincides with the integration of on .
In Section 6.2.1, we will provide the explicit relation between the evaluation and an endoscopic data for , and demonstrate that is an endoscopic group associated with the endoscopic data derived from . This connection provides the information on the other side of the fundamental lemma for . Using this argument, in Section 6.2.2, we will formulate the integration of on .
6.2.1. Endoscopic data and endoscopic group for
Since is a split group over , we use the definition for the endoscopic data and the endoscopic group, in terms of the root data following [Hal05, Section 2-3].
Definition 6.3 ([Hal05, Section 2]).
Let be a reductive group over . If is a quasi-split group which splits over an unramified extension over , then is said to be an unramified reductive group.
An unramified reductive group is classified by the following root data
where
In general, is obtained from the action on induced from the Frobenius automorphism of on the maximally split Cartan subgroup in . Here, is the maximal unramified extension over , in a fixed algebraically closure of containing .
Definition 6.4 ([Kot84, Section 7.1] and [Hal05, Section 3]).
Let be an unramified reductive group over with the root data . is an endoscopic group of if it is an unramified reductive group over whose classifying data has the form
The data for is subject to the constraints that there exists an element and a Weyl group element such that , , and .
Now, we return to our context, plugging and . The following two lemmas explain an endoscopic data of associated with the evaluation and the corresponding endoscopic group.
Lemma 6.5.
Proof.
By Proposition 2.4 and Corollary 5.3, the morphism is well-defined and independent of the choice of . For such that , we have
by Proposition 5.2, where the notations and are defined in loc. cit. and is defined in the diagram (5.4). Since the morphisms , , , and are group homomorphisms, the above equation concludes the desired result. ∎
By the Tate-Nakayama isomorphism (cf. [Lab08, Theorem 6.5.1]) for , we can identify the character of as an element in . Since by Section 6.1.1, we have
where . Here, the action of on , induced from that on , is given by cyclic permutations of the factors of . On the other hand, by [Lab08, Proposition 6.5.2], we have the following isomorphism
where
Since is a trivial morphism in , the following composition is surjective
Hence, a character on assigns an element in .
Lemma 6.6.
Under the above identifiaction of as an element in , an unramified reductive group is an endoscopic group of , associated with (cf. Definition 6.4).
Proof.
Let be the action on induced from the Frobenius automorphism of on . We claim that there exists an element of Weyl group for such that and . Since is split over , we have . The Weyl group is isomorphic to the symmetric group which acts on by permuting the factors. Thus we can find such that . Indeed, acts on by a cyclic permuation of the factors. On the other hand, since , we have .
We now verify that
is the root data classifying an unramified reductive group , where and is its dual set. Since the evaluation is non-trivial, is also non-trivial by Lemma 6.5. Moreover since and is a prime number, the induced morphism in from , via the above identification, maps to a non-trivial -th root of unity in where . Indeed, represents a non-trivial element in . For any coroot where , we then have
(6.3) |
where such that . Here, the last equality in (6.3) follows from the fact that . Since , we then have
and the dual set is also empty. In conculusion, the root data corresponds to . ∎
6.2.2. Computation
We first introduce one lemma which will be used in our main formulation.
Lemma 6.7.
For , let be the closed subscheme of representing the set of matrices whose characteristic polynomial is .
-
(1)
Then, is a homogeneous space of (resp. ), in the sence of Section 2.1 with a point , and we have the following isomorphism
where (resp. ) is the stabilizer of under the -action (resp. the -action) on .
-
(2)
There is an isomorphism such that and the following equation holds
Proof.
Since the map is an isomorphism which is compatible with conjugation of (resp. ), we obtain the first argument from the homogeneous space structure of with a point , in Proposition 2.1. This directly yields the following commutative diagram.
Here, . Therefore the measure transports to the measure along the isomorphism , according to Section 3.2.2.
On the other hand, applying the functor on the isomorphism (cf. Definition 4.2), we have the induced isomorphism . We then have the following commutative diagram for ,
For , this yields that where . By Definition 4.5, we have
for . Since the measure and the evaluation on are compatible with the measure and the evaluation on along the isomorphism , we conclude that
∎
Proposition 6.8.
Suppose that is an unramified Galois extension of prime degree . For whose evaluation is non-trivial, we have
where .
Proof.
Since , we can always choose for such that . Accordingly, we can take to be as well. Since the discriminant is invaraint under the constant translation on and , we may and do assume that with by Lemma 6.7.
By Lemma 5.3 and Lemma 6.5, we have the folowing equation where which corresponds to (cf. Definition 2.6),
(6.4) |
with the identification in Remark 2.7. By the similar argument in the proof of Proposition 5.9 and the assumption that , we have
Combining the above two equations, we have
(6.5) |
for .
To compute the right hand side of (6.5), we apply the formula in [Ngfrm[o]–0, Theorem 1] by plugging , and according to Lemma 6.6. We note that the condition is necessary to apply this formula. For the measure on such that , we have
(6.6) |
We explain the notations in the following (1)-(3), and further compute the right hand side of (6.6) in (2)-(3).
- (1)
-
(2)
We denote the Lie algebra of by , and correspondingly the Lie algebra of by . Let be an element in , whose stable conjugacy class transfers to via the map given by [Ngfrm[o]–0, Section 1.9]. By [Ngfrm[o]–0, Lemma 1.4.3 and Lemma 1.9.2], the centralizer of in coincides with . Then the stable orbital integral in the right hand side of (6.6) is formulated as follows
where . Indeed, the adjoint action of on is trivial and so . Therefore we have
-
(3)
In the right hand side of (6.6), and are discriminant functions of and defined in [Ngfrm[o]–0, Section 1.10], respectively. Since is torus, is trivial and so
On the other hand, by [Gor22, (20) and Example 3.6], we have
7. Main Result
We provide our main theorem on the asymptotic formula for based on the results in Section 5-6. We note that the product which appears in the following theorem is a finite product by Remark 1.6.
Theorem 7.1.
Let be a number field and be an irreducible monic polynomial of degree . Let be an -scheme representing the set of matrices whose characteristic polynomial is . We define
for , where the norm is defined in (1.1).
Suppose that and are totally real, and if , we further assume that is a prime number. Then we have the following asymptotic formulas.
-
(1)
If is not Galois or ramified Galois, then
-
(2)
If is unramified Galois and splits over all -adic places for , then
Here, is formulated as follows
where we use the following notations,
Proof.
We recall the formula in Proposition 4.1, by the equation (4.2), we have
(7.1) |
By Proposition 3.9, we have
For Archimedean place , the assumption yields that is trivial. By Proposition 5.6, we have
We denote the -module length between and by . Then by [CKL, Proposition 2.5], we have
where the last equality follows from [Yun13, Section 4.1] and [Bou03, Corollary 1 of Prosition 11 in Chapter 4.6]. Therefore we have
We now compute the integration of on with respect to for each . For , we note that the only case (1) is possible since every extension of is ramified.
-
(1)
We claim that the summand in (7.1) for a non-trivial element vanishes. Therefore, it suffices to obtain the volume of with respect to by Remark 4.6. In the case that and is not Galois, is trivial by Proposition 5.1. On the other hand, if and is ramified, then Corollary 5.4 and Proposition 5.11 yields the claim. In the case that , without the assumption that is a prime number, the claim also holds by the proof of [WX16, Theorem 6.1].
-
(2)
In the case that is trivial, the computation of the integration of on is exactly same with that of (1), and so we will omit this. For a non-trivial element , we address the following two cases separately; splits completely over and is unramified over .
-
•
If splits completely over , then the evaluation is trivial by Proposition 5.5. We then apply the computation in (1),
Here, by the following argument. Since is a linear polynomial for each where over , the orbital integral for associated with is equal to 1. In addition, the -module length between and equals to 0. By [Yun13, Corollary 4.10], we then have
- •
To sum up, for a non-trivial , we have
Proposition 5.2 yields that . Since is a cyclic group of prime order , we have nontrivial elements in . In conclusion, we have
-
•
∎
Remark 7.2.
In the case that and , we refer to the results in [CKL], for the closed formula of .
Proposition 7.3.
-
(1)
For , the product of orbital integrals is given as follows
-
(2)
For , the product of orbital integrals is given as follows
Here,
where
Proof.
-
(1)
By [CKL, Remark 5.7], we have
In order to write the above in a uniform way, we describe explicitly as follows.
where in the second case, is a quadratic field extension over . Plugging it into the above formula for , we obtain the desired formula.
-
(2)
By [CKL, Remark 6.7], we have
-
(a)
If is an irreducible monic polynomial , then we have
Here, we note that cannot be of the form (cf. [CKL, Proposition 6.2]).
-
(b)
If over where is an irreducible monic quadratic polynomial, we have
-
(c)
If splits into linear terms completely over , we have
On the other hand, by [Yun13, Section 4.1], we have . Since for a linear factor of , we verify that
In order to write the above in a uniform way, we describe explicitly as follows.
where is a cubic field extension over in the first case, and is a quadratic field extension over in the third case. Plugging it into the above formula for , we obtain the desired formula.
-
(a)
∎
In the case that an -order is a Bass order, we refer to the results in [CHL], for the closed formula of . A Bass order is an order of a number field whose fractional ideals are generated by two elements (cf. [CHL, Definition 6.1]). For example, any order of a number field which contains the maximal order of a subfield with degree 2 or whose discriminant is fourth-power-free in , is a Bass order.
To state the closed formula of when is a Bass order, we define to be the set of maximal ideals of and define the following notations for ;
Let and be subsets of such that
We denote the residue field of by and the cardinality of by . For , we denote the residue field of by .
Proposition 7.4.
Suppose that is a Bass order. The product of orbital integrals is given as follows
where .
Proof.
To ease the notation, we denote by . For or respectively, we define
By [CHL, Definition 2.3], we have the following identification
The isomorphism in [CHL, Remark B.(1)] yields that and thus . Since the map from to is surjective, we then have
(7.2) |
For , by [CHL, Theorem 3.7.(1)], we have
For , by [CHL, Theorem 6.11], we have
Plugging these closed formula for into (7.2), we have the desired result. ∎
References
- [BD13] Mikhail Borovoi and Cyril Demarche. Manin obstruction to strong approximation for homogeneous spaces. Comment. Math. Helv., 88(1):1–54, 2013.
- [Bir62] B. J. Birch. Forms in many variables. Proc. Roy. Soc. London Ser. A, 265:245–263, 1961/62.
- [Bit11] Rony A. Bitan. The discriminant of an algebraic torus. J. Number Theory, 131(9):1657–1671, 2011.
- [BLR90] Siegfried Bosch, Werner Lütkebohmert, and Michel Raynaud. Néron models, volume 21. Springer Science & Business Media, 1990.
- [BO12] Yves Benoist and Hee Oh. Effective equidistribution of -integral points on symmetric varieties. Ann. Inst. Fourier (Grenoble), 62(5):1889–1942, 2012.
- [Bou03] Nicolas Bourbaki. Algebra II. Chapters 4–7. Elements of Mathematics (Berlin). Springer-Verlag, Berlin, french edition, 2003.
- [BR95] Mikhail Borovoi and Zeév Rudnick. Hardy-Littlewood varieties and semisimple groups. Invent. Math., 119(1):37–66, 1995.
- [CHL] Sungmun Cho, Jungtaek Hong, and Yuchan Lee. Orbital integrals and ideal calss monoids for a bass order. arXiv:2408.16199v2.
- [CKL] Sungmun Cho, Taeyeoup Kang, and Yuchan Lee. Orbital integrals for classical lie algebras and smooth integral models. arXiv:2411.16054.
- [CTX09] Jean-Louis Colliot-Thélène and Fei Xu. Brauer-Manin obstruction for integral points of homogeneous spaces and representation by integral quadratic forms. Compos. Math., 145(2):309–363, 2009. With an appendix by Dasheng Wei and Xu.
- [DRS93] W. Duke, Z. Rudnick, and P. Sarnak. Density of integer points on affine homogeneous varieties. Duke Math. J., 71(1):143–179, 1993.
- [EMS96] Alex Eskin, Shahar Mozes, and Nimish Shah. Unipotent flows and counting lattice points on homogeneous varieties. Ann. of Math. (2), 143(2):253–299, 1996.
- [EvdGM] Bas Edixhoven, Gerard van der Geer, and Ben Moonen. Abelian varieties, preprint available at http://van-der-geer.nl/~gerard/AV.pdf.
- [GG99] Benedict H. Gross and Wee Teck Gan. Haar measure and the Artin conductor. Trans. Amer. Math. Soc., 351(4):1691–1704, 1999.
- [Gor22] Julia Gordon. Orbital integrals and normalizations of measures. arXiv preprint arXiv:2205.02391, 2022.
- [Gro97] Benedict H. Gross. On the motive of a reductive group. Invent. Math., 130(2):287–313, 1997.
- [Hal05] Thomas C. Hales. A statement of the fundamental lemma. In Harmonic analysis, the trace formula, and Shimura varieties, volume 4 of Clay Math. Proc., pages 643–658. Amer. Math. Soc., Providence, RI, 2005.
- [Kna02] Anthony W. Knapp. Lie groups beyond an introduction, volume 140 of Progress in Mathematics. Birkhäuser Boston, Inc., Boston, MA, second edition, 2002.
- [Kot84] Robert E. Kottwitz. Stable trace formula: cuspidal tempered terms. Duke Math. J., 51(3):611–650, 1984.
- [KP23] Tasho Kaletha and Gopal Prasad. Bruhat-Tits theory—a new approach, volume 44 of New Mathematical Monographs. Cambridge University Press, Cambridge, 2023.
- [Lab08] Jean-Pierre Labesse. Introduction to endoscopy. In Representation theory of real reductive Lie groups, volume 472 of Contemp. Math., pages 175–213. Amer. Math. Soc., Providence, RI, 2008.
- [Lee21] Jungin Lee. Counting algebraic tori over by artin conductor. arXiv preprint arXiv:2104.02855, 2021.
- [LX15] Qing Liu and Fei Xu. Very strong approximation for certain algebraic varieties. Math. Ann., 363(3-4):701–731, 2015.
- [Neu99] Jürgen Neukirch. Algebraic number theory, volume 322 of Grundlehren der mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 1999. Translated from the 1992 German original and with a note by Norbert Schappacher, With a foreword by G. Harder.
- [Ngfrm[o]–0] Bao Châu Ngô. Le lemme fondamental pour les algèbres de Lie. Publ. Math. Inst. Hautes Études Sci., (111):1–169, 2010.
- [Ono65] Takashi Ono. On the relative theory of Tamagawa numbers. Ann. of Math. (2), 82:88–111, 1965.
- [Poo17] Bjorn Poonen. Rational points on varieties, volume 186 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2017.
- [Sch85] Wolfgang M. Schmidt. The density of integer points on homogeneous varieties. Acta Math., 154(3-4):243–296, 1985.
- [Ser79] Jean-Pierre Serre. Local fields, volume 67 of Graduate Texts in Mathematics. Springer-Verlag, New York-Berlin, 1979. Translated from the French by Marvin Jay Greenberg.
- [Vos98] V. E. Voskresenskiĭ. Algebraic groups and their birational invariants, volume 179 of Translations of Mathematical Monographs. American Mathematical Society, Providence, RI, 1998. Translated from the Russian manuscript by Boris Kunyavski [Boris È. Kunyavskiĭ].
- [Wei82] André Weil. Adeles and algebraic groups, volume 23 of Progress in Mathematics. Birkhäuser, Boston, MA, 1982. With appendices by M. Demazure and Takashi Ono.
- [WX16] Dasheng Wei and Fei Xu. Counting integral points in certain homogeneous spaces. J. Algebra, 448:350–398, 2016.
- [Yun13] Zhiwei Yun. Orbital integrals and Dedekind zeta functions. In The legacy of Srinivasa Ramanujan, volume 20 of Ramanujan Math. Soc. Lect. Notes Ser., pages 399–420. Ramanujan Math. Soc., Mysore, 2013.