Intersections of Hecke correspondences on the modular varieties of -elliptic sheaves
Abstract
This paper studies the intersections of Hecke correspondences on the modular varieties of -elliptic sheaves in the higher-rank setting, where is a “maximal order” in a central division algebra over a global function field . Assuming that , where is a prime distinct from the characteristic of , we express the intersection numbers of Hecke correspondences as suitable combinations of modified Hurwitz class numbers of “imaginary orders”. This result establishes a higher-rank analogue of the classical class number relation.
1 Introduction
Let denote the modular curve of full level. For each non-square , the Hecke correspondence on corresponds to a -cycle on . The geometric version of the celebrated class number relation (initiated by Kronecker [22] and extended by Hurwitz [21]) expresses the “finite part” of the intersection number (i.e. excluding the “cuspidal intersection”) as
Here, denotes the “Hurwitz class number” for . The fact that these intersection points are actually “CM” provides a conceptual foundation for the proof of this relation, as it connects the intersection in question to Eichler’s theory of optimal embeddings of imaginary quadratic orders into matrix rings. This phenomenon is also evident in the context of Hilbert modular surfaces associated with real quadratic fields, as demonstrated in the celebrated work of Hirzebruch and Zagier [19]. Since then, the arithmetic significance of intersection numbers between special cycles on modular varieties has been extensively studied (see [23] and [24]), leading to the so-called “geometric Siegel-Weil formula” with profound applications (see [27], [7] [25], [26], [14]).
The primary aim of this paper is to explore this phenomenon within the function field setting and to describe the intersection numbers of “Hecke correspondences” on the modular varieties of -elliptic sheaves in higher rank, expressed in terms of class numbers. For the case where the base field is a rational function field with a finite constant field, an analogue of the Hurwitz–Kronecker class number relation (for Drinfeld modular curves) has been derived in [40] and [37] through a detailed study of Drinfeld modular polynomials in [1], [2], and [20]. Additionally, a Hirzebruch–Zagier-type formula was established in [16], which relates the intersection numbers of Hirzebruch–Zagier-type divisors on Drinfeld–Stuhler modular surfaces to modified Hurwitz class numbers through a bridge established from a particular theta integral, which is a “Drinfeld-type” automorphic form. This paper represents our initial effort to extend these ideas to the “higher rank” setting. To streamline the discussion, we focus on the case of modular varieties of -elliptic sheaves where is a “maximal order” of a central division algebra of dimension over a global function field , with being a prime number distinct from the characteristic of .
Fix a place of the global function field , which we regard as the place at infinity. Let denote the completion of at , and let be the completion of an algebraic closure of . Let be the (coarse) moduli scheme of -elliptic sheaves defined over . For each nonzero ideal of , where is the ring of elements in that are regular away from , we define an -cycle on in (4.4) arising from the Hecke correspondences on (analogous to the classical case). In particular, the cycle represents the diagonal cycle of on . The main theorem of this paper is stated as follows:
Theorem 1.1.
(Theorem 9.1) Keep the above notation and the assumption that is a prime number distinct from the characteristic of . Then for each nonzero ideal of , the intersection number of and is equal to
where:
-
•
is the cardinality of the constant field of ;
-
•
the first sum runs through all with so that is irreducible over and the field is imaginary over (i.e. does not split in );
-
•
the second sum runs through all with ;
-
•
is the modified Hurwitz class number of with respect to in Definition A.8;
-
•
is a “volume quantity” introduced in (9.3).
Remark 1.2.
As is a division algebra, there is no “cuspidal intersection” in our situation. Also, the assumption that is a prime ensures that the components in the intersection have dimensions either or . For composite , the intersection behavior may involve components associated with moduli schemes of -elliptic sheaves, where is a non-maximal order in , and is a central simple algebra over a finite extension of , which may not be a division algebra. Furthermore, the condition that is distinct from the characteristic of ensures that the imaginary fields corresponding to the intersection points are separable over . This separability guarantees the transversality of the intersection behavior of the pullback of these cycles in suitable finite coverings, enabling the intersection number to be determined simply by counting the intersection points of the pullback cycles. Without these assumptions, the overall structure of the intersections becomes significantly more complex and technical, which is why we impose these conditions in our first attempt to study higher rank cases in this paper.
1.1 Outline of the proof of Theorem 1.1
As the (coarse) modular variety , which parametrizes -elliptic sheaves, may not be smooth over , we use Briney’s theory to study the intersections in question. To address this, we consider a finite Galois covering of , where is the modular variety of -elliptic sheaves equipped with a level -structure for a suitable nonzero ideal of . This covering ensures that is smooth over . Using the projection formula, we can reformulate the intersection number of interest in terms of the corresponding pullback cycles on as follows (see Theorem 6.1):
(1.1) |
where is the degree of the “Galois covering” over ; is the diagonal cycle of on ; is introduced in (4.3); and the sum runs through a double coset space corresponding to the Hecke correspondence associated with on .
Next, the rigid-analytic structure of these varieties allows us to establish the transversality of the intersection when and are distinct. This transversality ensures that can be computed simply by counting the number of their intersection points.
Furthermore, through a sequence of double-coset comparisons, we relate these intersection numbers to the number of optimal embeddings of the corresponding imaginary orders into (see Corollary 8.5). Additionally, the self-intersection number of is shown to equal the Euler–Poincaré characteristic of , which can be interpreted as a volume quantity via an analogue of the Gauss–Bonnet formula established by Kurihara in [28] (see Proposition 8.9).
Finally, by applying Eichler’s theory, we express these intersection numbers in terms of modified Hurwitz class numbers, and complete the proof of Theorem 1.1.
Remark 1.3.
The recent groundbreaking works of Feng, Yun and Zhang in [11] and [12] establish a “higher” Siegel–Weil formula in the function field setting, showing a function field analogue of the Kudla–Rapoport conjecture. They express the non-singular Fourier coefficients of all central derivatives of Siegel–Eisenstein series on unitary groups in terms of the degrees of corresponding special cycles on the moduli stack of unitary shtukas. In contrast, our result takes a complementary perspective in the case where the group is (an inner form of ): we write the intersection numbers of specific cycles on the modular varieties of -elliptic sheaves as combinations of the class numbers of imaginary orders, which are natural arithmetic invariants.
Notably, when , we would be able to adapt the method in [16] and [17], utilizing the Weil representation, to explicitly construct an automorphic form whose Fourier coefficients correspond to the intersection numbers in question. Although this method does not appear to generalize to , we believe that a deeper connection with automorphic forms on persists and remains to be fully understood. This intriguing connection will be investigated in future work.
1.2 Content
This paper is organized as follows. Section 2 introduces the basic notation and preliminaries. In Section 3, we recall the definition of -elliptic sheaves and discuss level structures. The moduli spaces of -elliptic sheaves and the Hecke correspondences are presented in Section 4. Section 5 reviews Briney’s extension of intersection theory, leading to the derivation of equality (1.1) in Section 6.
The transversality of the non-self-intersection is verified in Section 7 from a rigid-analytic perspective. In Section 8, we perform several technical double-coset comparisons to connect the intersection points to optimal embeddings of imaginary orders and to express the self-intersection of in terms of a volume quantity. Finally, Section 9 connects the intersections in question with the modified Hurwitz class numbers and completes the proof of Theorem 1.1. For completeness, a brief review of Eichler’s theory of optimal embeddings is included in Appendix A.
Acknowledgments
Part of this work was carried out during the first author’s visit to the National Center for Theoretical Sciences (NCTS). The authors sincerely thank the NCTS for its generous support of this collaboration. The first author is supported by Academia Sinica Investigator Grant AS-IA-112-M01 and NSTC grant 113-2115-M-001-001. The second author is supported by the NSTC grant 109-2115-M-007-017-MY5, 113-2628-M-007-003, and the NCTS.
2 Notation
Let denote the finite field with elements. Let be a geometrically connected smooth projective curve over a finite field with function field . We denote the set of closed points of by , identifying with the set of places of . Fix one closed point of , referred to the place at infinity. Let denote the ring of regular functions outside of where is the structure sheaf of .
Given , the completion of at is denoted by , and let be the maximal compact subring of . We define the adele ring of to be the restricted product of the completions at the places of :
and the ring of finite adeles as the restricted product of the completions at the finite places:
The maximal compact subring of is .
Let be a central division algebra over of dimension . We say that is ramified at a place if is not isomorphic to the matrix algebra . Let denote the set of ramified places of . Here we assume that . Let be a locally free sheaf of -algebras such that the stalk of at the generic point of is the division algebra , i.e, . Throughout this paper, we always assume that is a maximal -order in . For a place of , we define
where (resp. ) is the stalk of (resp. ) at . We put
In particular, our assumption implies that , the matrix ring with entries in . Finally, let be the global sections of away from . Then
3 -elliptic sheaves
In this section, we recall the definition and some basic facts about -elliptic sheaves. For more details on -elliptic sheaves, please refer to [29].
3.1 Defintion of -elliptic sheaves
First, we introduce some notations. For an -scheme , we denote the Frobenius endomorphism on by
which is defined as the identity on points and as the -power map on the structure sheaf. Note that for an -scheme , one can form the fiber product over . We will assume this is understood over and denote the fiber product simply by . Let
i.e., acts on as the identity and on as the Frobenius endomorphism .
Definition 3.1.
Let be a scheme over . A -elliptic sheaf over is a pair where is a morphism and is a sequence . Here each is a locally free -modules of rank with a right -action which is -linear and the restriction of to the scalars is same as the -action. And the morphisms
are injective -linear morphisms which are compatible with the -action such that the following diagram commutes
Moreover, for each the following conditions hold:
-
1.
(periodicity) Put We have
where
-
2.
The is supported on and it is locally free -module of rank .
-
3.
The has support on the graph of and is locally free of rank over .
The morphism in the definition of a -elliptic sheaf is called the characteristic of the -elliptic sheaf. For simplicity, we denote the -elliptic sheaf by if no confusion arises.
A morphism between two -elliptic sheaves and over is a sequence of morphisms of locally free -modules which are compatible with -action and with the morphisms ’s and ’s.
Remark 3.2.
Let be a -elliptic sheaf over . By the condition (2) in Definition 3.1, is independent of . Moreover, if for a ring , it becomes an -module where the -action comes from the morphism . This module is referred to as a Drinfeld-Stuhler -module. The concept was first implicitly introduced in [29, Section 3], with further studies available in [30].
Let denote the category whose objects are the -elliptic sheaves over and whose morphisms are isomorphisms of -elliptic sheaves. If is a morphism of -schemes, then taking pullback of a -elliptic sheaf over gives us a -elliptic sheaf over . This defines a fibered category
over the category of -schemes . Moreover this gives us a stack, denoted by , with respect to fppf-topology. The proof follows as a special case of [36, Proposition 2.12].
3.2 -elliptic sheaves with level structures
Now we will define level structures on -elliptic sheaves. First, we want to recall the following definition:
Definition 3.3.
Let and be two schemes, and let , be the natural projections. Given a locally free -module and a locally free -module , we define the external tensor product of and as follows:
which is naturally a locally free -module.
Let be a finite closed subscheme of . Then, the restrictions are all isomorphic by the morphisms ’s (cf. Remark 3.2). Therefore it is independent of the chosen , and we will denote this restriction simply by . We write for where is the structure sheaf of , and let be the ring of global sections of .
Definition 3.4.
Let be an -scheme and be a -elliptic sheaf over . Let be a finite closed subscheme of . A level -structure on is an isomorphism of -modules
which is compatible with -action and the Frobenius endeomorphism on .
Remark 3.5.
Throughout the paper, when we talk about level -structures, is always a finite closed subscheme of .
Let denote the category whose objects are -elliptic sheaves over with level -structure and whose morphisms are morphisms of -elliptic sheaves that respect the level -structure. As before, by using pullbacks one can define a fibered category over , which is a stack with respect to fppf-topology. For a detailed proof we refer to [36, Proposition 2.19].
Now, let be a -elliptic sheaf over with level -structure and let denote the category of schemes over . One can define -invariant elements functor
by . We then have the following:
Theorem 3.6.
Remark 3.7.
([29, (4.8), page 240]) For connected , the set of -level structures is a torsor over the unit group . More precisely, Let be an -scheme, be two finite closed subschemes with . Then, the morphism
which associates a level -structure to its restriction gives us a -torsor over , i.e, the finite group acts on the set of level -structures transitively and freely.
4 Moduli schemes of -elliptic sheaves and Hecke correspondences
We shall introduce the moduli schemes of -elliptic sheaves and the cycles associated with the Hecke correspondences.
4.1 Moduli schemes of -elliptic sheaves
In the previous section, we defined the stack of -elliptic sheaves with level -structures. In fact, is an algebraic stack in the sense of Deligne-Mumford (cf. [9]).
Theorem 4.1.
([29, Theorem 4.1 and Theorem 5.1])
The stack is an algebraic stack in the sense of Deligne-Mumford which is smooth of relative dimension over . Moreover, if , it is actually a quasi-projective scheme.
Remark 4.2.
Let be a finite closed subscheme of such that . There exists a nonzero ideal of associated with this closed subscheme. Conversely, if we are given a nonzero ideal of , we get a corresponding finite closed subscheme of away from .
Definition 4.3.
Let be a nonzero ideal of with for every and be the finite closed subscheme of associated with . We denote the representing scheme of by .
Moreover, different from Drinfeld modular varieties, we have the following theorem:
Remark 4.5.
We want to note that in [29], it is assumed that the characteristic of -elliptic sheaf is away from . However, both theorems above hold in general. We refer to [36, Section 5 and Section 6] for the proof in general case. Moreover, the base change of to is actually a projective scheme (see [29, p. 218]).
We denote by the formal completion of the scheme along the fiber over , which is a formal scheme over .
Let denote the Drinfeld symmetric space:
It has a rigid analytic stucture, and thus becomes a rigid space in the sense of Raynaud. By taking a formal completion over the fiber over , we obtain a formal scheme ([3, Remark 4.3.1]). For more on the Drinfeld symmetric space, we refer to [10], [8] and [33].
Recall that . Set and
Then the following theorem holds:
Theorem 4.6.
One has an isomorphism of formal schemes
Let be an open compact subgroup of . We want to look at the quotient closer. Let , denote the representatives of the double coset space and let
be the stabilizer of with respect to -action. In other words,
so . Put , which is a discrete subgroup of .
Proposition 4.7.
Keep notations as above. We have
Proof.
Write as the finite disjoint union . Then
Hence we can identify with .
∎
By construction, the generic fibre over of is the rigid space . Therefore, together with the Proposition 4.7, we have the following proposition:
Proposition 4.8.
Let be a nonzero ideal of and take . We have
4.2 Hecke correspondences and associated cycles
In this section we will give a brief summary of Hecke correspondences on the modular variety of -elliptic sheaves. For details we refer to [29, Section 7].
Let be a finite closed subscheme of such that . Define
For an -scheme , a section of over consists of triples as in the Definition 3.1 with the following ‘level structure’, i.e, a -linear isomorphism
where .
There is a right action of on , extending the action of (cf. [29, (7.4)]). Let be an open compact subgroup and fix . Then, we get a correspondence over :
(4.1) |
where the morphisms and are induced, respectively, by the inclusions
Given a nonzero ideal of with for every , recall that we let
For every , define
We also put
Let be the scheme corresponding to . Then the correspondence (4.1) gives us the following:
|
In particular, similar to Proposition 4.8, we may identify
Then the two morphisms can be realized as follows: for every and , let and be the representing double coset in and in , respectively. Then
Therefore we have a morphism
which is realized by
(4.2) |
for every .
By abuse of notations, we still denote by its base change to . Recall that it is known that has constant dimension over (cf. Theorem 4.1). Let be the -cycle of associated with itself. We define
(4.3) |
which is an -cycle of . In particular, we denote
Moreover, let be a nonzero ideal of . We define the following associated cycle on :
(4.4) |
where the sum is taken over double cosets with the condition . Here Nr denotes the natural extension of the reduced norm of to .
The main goal of this paper is to study the intersection cycle , and to connect the intersection number with the class numbers of “imaginary orders”.
5 Briney’s Theory
In this section we briefly recall Briney’s theory of intersections of quotients of algebraic varieties in [4]. Here we assume all varieties are (absolutely) irreducible and abstract as in [4].
5.1 Quotient varieties by finite groups
Fix an algebraically closed field . We will focus on quotients of an algebraic variety . Let be a finite subgroup of the group of all automorphisms of . Then one can look at the quotient of by . We assume that is an algebraic variety, i.e, every -orbit is contained in an affine open set in ([34, Section 1.4]). Then the natural projection map is a proper morphism. Moreover it is a covering map. If all are defined over and is a field of definition of , we say that is defined over .
Let and be as in the previous paragraph. We say they are equivalent if there are surjective isomorphisms and such that for all and for all . In this case, gives us isomorphism of quotients such that .
Now, let and be varieties such that there are morphisms and . Let and be finite subgroups of the automorphism groups of and , respectively. Then, each morphism splits through the associated quotient:
We say and are equivalent if and are equivalent and if . In other words every triangle in the following diagram commutes:
Now, we can define a quotient structure.
Definition 5.1.
A quotient structure on the variety is the equivalence classes of the ‘real’ quotients. More precisely, a quotient structure is the equivalence classes of the triples such that is an isomorphism ([4, Definition 2.1]). In this case we say is a quotient variety of by and is denoted by .
Let be an algebraic variety and be a quotient variety of by as above. For a subvariety , a typical component of will be denoted by . In this case we say lies over . For each , we define the following two groups:
-
1.
The splitting group of is defined as
-
2.
The inertia group of is the group
For any subvariety of the quotient we associate certain numerical characters where denotes a typical component of :
-
1.
The degree of over is defined as
-
2.
The degree
is called separable degree of over .
-
3.
We define the inseparable degree of over as
-
4.
Define
which is analogous to the ramification index for a valuation ring.
Remark 5.2.
Let denote the number of components of and . We have the relation
for any and any lying over .
Note that the components of are permuted transitively by the elements of ([5, Ch. V, Proposition 3 on p. 189]), so that the above characters depend only on and not on the choice of .
5.2 Intersection multiplicity and Projection formula
We recall the following theorem:
Theorem 5.3.
By this theorem we can talk about the intersection of smooth (quasi-) projective varieties. Briney’s theory enables us to extend the intersection theory of finite quotients of smooth (quasi-) projective varieties to not necessarily smooth ones by allowing rational multiplicities. First we define covering cycles.
Let be (quasi-)projective (not necessarily smooth) varieties over . And, let be the projection map as defined in the Definition 5.1. For any subvariety , let be distinct irreducible components of . Set
We call the covering cycle of .
By the projection map , we get the following mappings on subvarieties
and
One can extend the mappings and linearly to the cycles.
Remark 5.4.
For every cycle on we have , where . This relation is a special case of projection formula (cf. Proposition 5.7).
Let and be the corresponding covering cycles. Let be a proper component of on .
We can compute the intersection multiplicity by carrying over the multiplicities given on via the quotient map .
Definition 5.5.
([4, Definition 3.2])
The intersection multiplicity of and at on is
where is any component of and the multiplicity is on .
Proposition 5.7.
([4, Proposition 3.3])
Let be cycles on and be cycle on .
-
1.
If is defined on then is defined on . And in this case, we have
-
2.
(Projection formula) is defined on iff is defined on . And in this case, we have
Remark 5.8.
Suppose that is projective and smooth over . Given two cycles on with middle dimension which have transversal intersection, the intersection number is the counting number of their intersection points. By the moving lemma, this intersection number can be extended to every two cycles on with middle dimension. Therefore for every two cycle on with middle dimension, take a cycle on so that for some . We may define the intersection number via the above projection formula:
6 Projection Formula for Hecke cycles
Let be a nonzero proper ideal of such that for every and for every . The main result of this section is the following theorem:
Theorem 6.1.
where the sum is taken over with .
The rest of the section is devoted to the proof this theorem.
6.1 Projection Formula
In order to study the intersection , we will be working over product of these varieties. More precisely we have
which can be written as follows: we define as the tuple where correspond to first component of and correspond to the second component, i.e., we can write
Similarly for and we can write
Recall the definition of :
sending an element to , i.e, we have a morphism
In particular, gives a morphism
where is determined by and . Moreover, let be the projection to the first component (which projects to ). Then coincides with the original covering map. So, we have the following relations between these objects:
For we define as the intersection of the cycle on and the -th component of :
Similarly, define
For and define
Let and . One can see by definition that . By the map
one can define the map . We denote by for and the morphism on the irreducible components , i.e,
We want to remark that in Section 5, the numerical components are defined for irreducible varieties. Let be a cycle of . Write . Let be a component of . We define
and
In particular, one has
Unlike Briney, we used the upper script and to point out the separable and inseparable degrees to avoid confusion with the irreducible component indices .
Note that in our case, we have the “Galois group” and so
In particular:
let be the Galois group associated with the irreducible component , of the covering . Then is a subgroup of . Also:
Lemma 6.2.
We have .
Proof.
Given represented by for , one has that for every in ,
As this equality holds for every , we must get , which means that . Hence we get .
Moreover, the fact that follows from the fact that is a Galois covering. Therefore,
∎
Now, assume that for every , which implies that as . Put
We can rewrite as
where . Note that is nonempty if and only if and . Thus we do not need to take all pairs into account. More precisely, for every open compact subgroup of , let . Then
where . Moreover:
Lemma 6.3.
We have
Proof.
By definition one has
where is the sum of the irreducible components of . Note that the Galois group of the covering map can be identified with , and one checks that
∎
Remark 6.4.
It is known that the self-intersection number is equal to , the “Euler–Poincaré characteristic of ”. We refer the reader to Section 8.2 for further discussion and the precise formula for .
Proof of Theorem 6.1. From the previous discussion with the projection formula, we then get that
Therefore the result holds.
In the next section, we will determine the transversality of the intersection when .
7 Transversal intersection
We start with recalling the following lemma:
Lemma 7.1.
([15, Chapter 3, Exercise 6.7])
Let be a field, let be a -scheme, and let and be closed subschemes of and let be their schematic intersection. Let be an -valued point and assume that , , and are smooth at over of relative dimension and , respectively. The following assertions are equivalent.
-
1.
is smooth of relative dimension .
-
2.
.
Here is the tangent space of at for , regarding as subspaces of , the tangent space of at .
If these equivalent conditions are satisfied, we say that and intersect transversally.
Recall that we defined a morphism
for every in Section 4.2. Put
As and are both smooth of dimension in (which has dimension ), showing the transversality of the intersection reduces to verify that the intersection of the corresponding “tangent spaces” at each intersection point is trivial.
To proceed,
we will now consider our objects as rigid analytic spaces, and study the corresponding tangent spaces via their rigid analytic uniformization.
Recall the following identification in Proposition 4.8:
Put for every . Then
Similarly, there exist elements and arithmetic subgroups such that
where
Here acts diagonally on as well. Let . Since the covering is étale and the intersection behavior is a local property, it is sufficient to lift to a point (still denoted by by abuse of notation) in for some nonconstant .
Lemma 7.2.
Suppose is a prime number distinct from the characteristic of . Given such that , suppose there exists . Let be the tangent space of at , and be the tangent space of at (regarded as a subspace of via the inclusion ). Then .
Proof.
For simplicity, we may identify with
Given
write
Regarding as functions in , let be the partial derivative of with respect to , i.e,
(7.1) |
Then for every , the tangent space is spanned by the vectors
In particular, the tangent space is spanned by the vectors
Suppose that , i.e. . Write
Then implies that for , and
(7.2) |
On the other hand, given , we may write with , and the condition implies that
By (7.1) we get
which is equivalent to
(7.3) |
By suitable row operations on the following matrix
we can see that .
Now, if , i.e. , then . As the field is separable of degree over under our assumption, the eigenspace of corresponding to the eigenvalue (with multiplicity one) is spanned by . Hence implies that the nonzero column vector lies in the range of , and also in the null space of by (7.2). However, the diagonalizability of in assures that the intersection of the range of and the null space of must be trivial, which is a contradiction. Therefore , i.e. the intersection of and is trivial. ∎
Consequently, we have that:
Corollary 7.3.
Suppose is a prime number distinct from the characteristic of . When , the intersection is transversal.
8 Counting the intersection numbers of Hecke correspondences
From the previous section, we know that when is a prime number distinct from the characteristic of , the intersection of and is always transversal when they are distinct. In this case, the intersection number is simply equal to the cardinality of their intersection points. In what follows, we shall count the intersection points in question by using “optimal embeddings.”
We first release our condition on , (i.e. is just a positive integer). Recall the following uniformization of :
Every -valued point of corresponds to a class in . Note that the morphism gives an isomorphism between and . Then:
Lemma 8.1.
The set can be identified with
Proof.
Let . From the unifomization of , we may identify with a class . Then (resp. ) is the class in corresponding to (resp. ). Suppose , which is equivalent to
This means that there exists and so that and , which says that
Conversely, suppose there exists so that and . Write where . Then
which implies that . This completes the proof. ∎
Set
which is equipped with a left action of and a right action of defined below: for every , , and ,
By Lemma 8.1, we have a natural surjective map from to sending to (the point corresponding to) . Moreover:
Lemma 8.2.
The above map induces a bijection between and .
Proof.
Given , suppose
which implies that there exists and so that and . Thus
Put , which satisfies that and .
Note that by [30, Lemma 4.6], one has that is a subfield of which is imaginary with respect to (i.e. is non-split in ). Hence
This says in particular that is a unit in , the integral closure of in , and . Since is imaginary and the degree divides , the units of must be contained in a finite field with elements. The condition then implies that , whence
Therefore and represent the same class in , and the proof is complete. ∎
Corollary 8.3.
Given , one has that if and only if .
Proof.
It is clear that when . Conversely, suppose . By Lemma 8.2, we have that for every , there exists such that and . When taking with algebraically independent coordinates over , we must have that , whence as desired. ∎
8.1 The case when
Let
Then we have a natural map from defined by for every triple . On the other hand, the fiber of each in under can be identified with the set of fix points of on , which is always non-empty. Therefore is surjective. Moreover, the following lemma holds:
Lemma 8.4.
The map
induced by is surjective. Moreover, suppose and is a prime number. For each class , the cardinality of is resp. if is separable resp. purely inseparable over .
Proof.
The surjectivity is clear. Suppose and is a prime number. Let . By assumption one has that is a maximal subfield of which is imaginary. Thus there are (resp. ) fixed point(s) of on , say (resp. ), corresponding to distinct eigenvalues (resp. the unique eigenvalue) of . Hence
It suffices to verify that represents distinct classes in the space when is separable. To show this, suppose there exists and such that
Since is a maximal subfield of , the centralizer of is itself. Hence and
By our assumption on , we can apply the same argument in the proof of Lemma 8.2 to get , which implies and . ∎
Corollary 8.5.
Suppose and is a prime number distinct from the characteristic of . The cardinality of is equal to
Remark 8.6.
Given , the map from to sending to for every is bijective. Thus we have the following induced bijection
8.2 Self-intersection of
When (i.e. by Corollary 8.3), the self-intersection number of in is equal to the Euler–Poincaré characteristic of , denoted by , which is the degree of the top Chern class of the tangent bundle of (see [13, Example 8.1.12]). Recall the following identification
where for every . Identifying with , we may regard as a discrete and co-compact torsion free subgroup of . Therefore by Kurihara’s analogue of Hirzebruch proportionality (see [28, Theorem 2.2.8]), we obtain:
(8.1) |
Here is the Euler–Poincaré measure on introduced by Serre (see [34]), i.e.
and is the cardinality of the residue field at . Extending to the Haar measure on satisfying that
the equation (8.1) leads to the following Gauss–Bonnet-type formula:
Proposition 8.7.
Remark 8.8.
(1) We may rewrite the above formula as the following form:
where for every .
(2) Let be the Tamagawa measure on introduced in [39, §3.2]. It is known that (see [39, Theorem 3.2.1])
and (cf. [39, §3.1, p. 32] and [38, §5])
where is the cardinality of the residue field of for each , and is the zeta function of , i.e.
Therefore we get
(8.2) | |||||
Consequently, Proposition 8.7 implies the following formula.
Proposition 8.9.
The self-intersection number of in is equal to
Remark 8.10.
When is a polynomial ring over , the formula for appears in [31, equation (6.10)].
8.3 Projection to the full-level case
Now, we assume to be a prime number distinct from the characteristic of . Let be a non-zero ideal of . Set
where
Then (as well as and ) is equipped with a left action of and a right action of defined by
Choose a nonzero ideal of so that for every and for every . For each with , one has . Put
We point out that either or is empty, and is non-empty if and only if . Moreover, let be the “prime-to-Ram” part of , i.e. is the ideal of so that
the following lemma holds.
Lemma 8.11.
Taking , we may decompose into
(8.4) |
Consequently, we have the bijection:
Proof.
Given , we may take and get . Thus it remains to show that the right hand side of the equality (8.4) is indeed a disjoint union.
Given with , suppose is non-empty, i.e. there exists . Then , which implies that . Hence
and the desired decomposition in (8.4) follows from:
∎
Next, we look closely into for or . We may decompose into
where is the centralizer of in . Applying Eichler’s theory of optimal embeddings of imaginary -orders into in Appendix A, we know that is finite (see Remark A.7). Hence
Finally, by Lemma 8.11, we obtain that
where runs through the double cosets in with . Therefore
(8.5) | |||||
Recall that is non-empty if and only if , which is equivalent to by Corollary 8.3. For and , we put
By Corollary 8.5, the equation (8.1), and the projection formula in Example 6.1, we obtain the following.
Proposition 8.12.
Suppose is a prime number different from the characteristic of . Let be a nonzero ideal of . We have that
Proof.
In the next section, we shall express the right hand side of the above equality in terms of class numbers of imaginary fields, and derive an Kronecker-Hurwitz-type class number relation.
9 Class number relation
Keep the notation as in the last section. Suppose is non-empty, which implies that is a principal ideal of . In this case, for each , the condition is equivalent to
where is a maximal -order of . This implies that is integral over . Suppose . Let
be the minimal polynomial of over . Since is imaginary over of degree , the polynomial remains irreducible over . Hence every root of in has the same absolute value. Notice that , which gives us that for . Therefore the possibility of the minimal polynomial is finite for running through all pairs in .
On the other hand, given , let
Suppose that:
-
(i)
;
-
(ii)
is an imaginary field extension over ;
-
(iii)
there exists a -algebra embedding .
Put , which is an -order in , and is the coset in represented by . For each satisfying that
set . Then lies in .
Let
which is invariant by left multiplication of and right multiplication of . For each with , by the Noether-Skolem theorem, there exists , which is unique up to left mulltiplication of , such that . Then the condition is equivalent to . Hence we have a well-defined bijective map
(9.1) |
which sends to when .
Now, for an -order of with , put
Then
(9.2) |
Note that for , one has that
Therefore by the equation (8.5) and Proposition A.9, we obtain that
where is the modified Hurwitz class number with respect to introduced in Definition A.8.
Finally, set
(9.3) |
As is in bijection with , which is finite, by Proposition 8.9 and Remark 8.8 (1) one gets
Therefore Proposition 8.12 leads to the following class number relation:
Theorem 9.1.
Let be a prime number distinct from the characteristic of . For each nonzero ideal of , we have
Remark 9.2.
Suppose and the characteristic of is odd. Given , we may denote by if is an imaginary quadratic field extension over . For such , we let
Then Theorem 9.1 can be reformulate as the following: for each nonzero ideal of , we have
Furthermore, set
By employing the “Weil representation” as illustrated in [16, Sec. 3] (or alternatively, by combining the work of [6] with the relations among Hurwitz class numbers established in [17, Sec. 2]), we are able to construct a “harmonic” theta series on whose nonzero (resp. constant) Fourier coefficients come from the intersection numbers for nonzero ideal of (resp. ). Although the approach in [16] seems not applicable for , we believe that, after further work, this phenomenon remains valid.
Appendix A Eichler’s theory of optimal embeddings
Recall that is a division algebra over with . Here we review Eichler’s theory of optimal embeddings from an -order of an imaginary extension over into the division algebra . For simpliciy, we always assume that is a prime number in this section, which is sufficient for our purpose.
A.1 Local theory
Fix a place of with , let
As is a prime number, either is division or . Thus there always exists an embedding from into when either or is a field (see [35, Theorem 1.1 (1)]). Fix an embedding . For each -order in , let
We shall determine the finiteness of the double coset space
where , the -component of .
Lemma A.1.
Suppose is division and is a field. Then
Proof.
When is division and is a field, one knows that is the unique maximal -order in , for which . Moreover, every element in which is integral over is contained in . Therefore for every ,
Thus if is not maximal in .
When , we get that . Hence
This completes the proof. ∎
Lemma A.2.
Suppose . Then the cardinality of is finite and positive. In particular, if is maximal in , then
Proof.
Under the isomorphism , we may identify with without loss of generality. Let , which is equipped with a left module structure over . An -lattice is a free -submodule of which is invariant under the multiplication by , and is optimal if . Two -lattices and are isomorphic if there exists so that , and
Then , and we have a bijection
which is induced by sending to the optimal -lattice for every . Thus it suffices to count the isomorphism classes of optimal -lattices in .
Notice that when is maximal in , every -lattice in must be free of rank one over . Hence there is only one isomorhism class. In general, for each optimal -lattice in , consider , which is an -lattice. On the other hand, there exists such that (the smallest choice of ‘’ refers to the “conductor” of ), whence . Since has finite cardinality, we obtain that the isomorphism classes of -lattices in is finite (and bounded by ). ∎
Remark A.3.
Suppose . Let be the cardinality of the isomorphism classes of optimal -lattices in . Then the above proof shows that is finite positive, and for almost all . In particular,
A.2 Global theory and modified Hurwitz class numbers
First, the condition for the existence of a -algebra embedding from into is determined by the following.
Theorem A.4.
(See [32, Proposition A.1]) Let be a central simple algebra of degree over and is a field extension of degree over . Then the local-global principle holds, i.e. can be embedded into if and only if can be embedded into for every place of .
Now, let be a division algebra over and is a prime number. Suppose an embedding exists, which corresponds to embeddings for all finite places of . Let be an -order of . Put
Here is defined in the beginning of Section 9. We shall express the cardinality of the double coset space as “modified class number” of .
For each finite place of , let . For , one has that if and only if for every . Define . Then Lemma A.2 ensures that
On the other hand, let be the class number of , i.e.
We have the following.
Lemma A.5.
Under the canonical surjective map
the cardinality of each fiber is equal to .
Proof.
Given , the fiber of can be identified with the double coset space
and the condition of lying in implies that
Therefore the result follows. ∎
Consequently, let be the following modified class number (with respect to ):
where is the introduced in Lemma A.1 when is division and Remark A.3 when . We then conclude that:
Proposition A.6.
The double coset space is a finite set, and its cardinality is equal to .
Proof.
Remark A.7.
Definition A.8.
Given satisfying that is imaginary, the modified Hurwitz class number of (with respect to ) is given by
We finally arrive at:
Proposition A.9.
References
- [1] S. Bae “On the modular equation for Drinfeld modules of rank 2” In J. Number Theory 42, 1992, pp. 123–133
- [2] S. Bae “On the coefficients of the Drinfeld modular equation” In J. Number Theory 66, 1997, pp. 85–101
- [3] Stuhler U. Blum A. “Drinfeld modules and elliptic sheaves” In Vector Bundles on Curves — New Directions: Lectures given at the 3rd Session of the Centro Internazionale Matematico Estivo (C.I.M.E.) held in Cetraro (Cosenza), Italy, June 19–27, 1995 Springer Berlin Heidelberg, 1997, pp. 110–188
- [4] R.. Briney “Intersection theory on quotients of algebraic varieties” In Amer. J. of Math 84, 1962, pp. 217–238
- [5] Claude Chevalley “Les classes d’équivalence rationnelle, I” talk:2 In Séminaire Claude Chevalley 3 Secrétariat mathématique, 1958
- [6] C.-Y. Chuang, T.-F. Lee, F.-T. Wei and J. Yu “Brandt matrices and theta series over global function fields” In Mem. Amer. Math. Soc. 237.1117, 2015
- [7] J.. Cogdell “Arithmetic cycles on Picard modular surfaces and modular forms of nebentypus” In J. Reine Angew. Math. 357, 1985, pp. 115–137
- [8] Husemöller D. Deligne P. “Survey of Drinfeld modules” In Contemporary Math. 67, 1987, pp. 25–91
- [9] Mumford D. Deligne P. “The Irreducibility of the Space of Curves of Given Genus” In Publ. Math. IHES 36, 1969, pp. 75–109
- [10] V.G. Drinfeld “Elliptic modules” In Math. USSR Sbornik23, 561–592, 1974
- [11] T. Feng, Z. Yun and W. Zhang “Higher Siegel–Weil formula for unitary groups: the non-singular terms” In preprint arXiv:2103.11514
- [12] T. Feng, Z. Yun and W. Zhang “Higher theta series for unitary groups over function fields” In preprint arXiv:2110.07001
- [13] W. Fulton “Intersection Theory, Second edition” Springer New York, 1998, pp. VIII\bibrangessep470
- [14] J. Funke “Heegner divisors and nonholomorphic modular forms” In Compositio math. 133, 2002, pp. 289–321
- [15] Wedhorn T. Görtz U. “Algebraic Geometry”, Advanced Lectures in Mathematics Vieweg+Teubner Verlag Wiesbaden, 2010, pp. IV\bibrangessep615
- [16] J.-W. Guo and F.-T. Wei “On class number relations and intersections over function fields” In Doc. math. 27, 2022, pp. 1321–1368
- [17] J.-W. Guo and F.-T. Wei “On CM masses over global function fields” In Math. Z. 303.29, 2023
- [18] R. Hartshorne “Algebraic Geometry”, Graduate Texts in Mathematics Springer New York, 1977, pp. XVI\bibrangessep496
- [19] D. Hirzebruch “Intersection numbers of curves on Hilbert modular surfaces and modular forms of nebentypus” In Invent. math. 36, 1976, pp. 57–113
- [20] L.-C. Hsia “On the coefficients of modular polynomials for Drinfeld modules” In J. Number Theory 72, 1998, pp. 236–256
- [21] A. Hurwitz “Über Relationen zwischen Klassenzahlen binärer quadratischer Formen von negativer Determinante” In Math. Ann. 25, 1885, pp. 157–196
- [22] L. Kronecker “Über die Anzahl der verschiedenen Klassen quadratischer Formen von negativer Determinante” In J. Reine Angew. Math. 57, 1860, pp. 248–255
- [23] J.. Kudla “Intersection numbers of cycles on locally symmetric spaces and Fourier coefficients of holomorphic modular forms in several complex variables” In Publications mathématiques de l’I.H.É.S. 71, 1990, pp. 121–172
- [24] J.. Kudla “The theta correspondence and harmonic forms. I” In Math. Ann. 274, 1986, pp. 353–378
- [25] J.. Kudla “The theta correspondence and harmonic forms. II” In Math. Ann. 277, 1987, pp. 267–317
- [26] S.. Kudla “Algebraic cycles on Shimura varieties of orthogonal type” In Duke math. 86.1, 1997, pp. 39–78
- [27] S.. Kudla “Intersection numbers for quotients of the complex -ball and Hilbert modular forms” In Invent. math. 47.2, 1978, pp. 189–208
- [28] A. Kurihara “Construction of p-adic unit balls and the Hirtzbruch proportionality” In Amer. J. math 102.3, 1980, pp. 565–648
- [29] Stuhler U. Laumon G. “-elliptic sheaves and the Langlands correspondence” In Invent. Math. 113, 1993, pp. 217–338
- [30] M. Papikian “Drinfeld–Stuhler modules” In Res Math Sci 5, 2018
- [31] M. Papikian “Modular varieties of -elliptic sheaves and the Weil–Deligne bound” In J. Reine Angew. Math. 626, 2009, pp. 115–134
- [32] G. Prasad and A.. Rapinchuk “Computation of the metaplectic kernel” In Publ. Math. IHES 84, 1996, pp. 91–187
- [33] Stuhler U. Schneider P. “The cohomology of p-adic symmetric spaces” In Invent. Math. 105, 1991, pp. 47–122
- [34] J.-P. Serre “Groupes algébriques et corps de classes”, Actualités Scientifiques et Industrielles 1264 Hermann (Paris), 1959
- [35] S.-C. Shih, T.-C. Yang and C.-F. Yu “Embeddings of fields into simple algebras over global fields” In Asian J. Math 18.2, 2014, pp. 365–386
- [36] Ö. Ülkem “Uniformization of generalized -elliptic sheaves, PhD thesis” Heidelberg University, 2020 URL: http://www.ub.uni-heidelberg.de/archiv/29531
- [37] J. Wang “On class number relations over function fields” In J. Number Theory 69, 1998, pp. 181–196
- [38] F.-T. Wei and C.-F. Yu “Mass formula of division algebras over global function fields” In J. Number Theory 132.3, 2012, pp. 1170–1184
- [39] A. Weil “Adeles and algebraic groups (with appendices by M. Demazure and T. Ono)” 23, Progress in Mathematics Birkhäuser, Boston, Mass., 1982
- [40] J.-K. Yu “A class number relation over function fields” In J. Number Theory 54, 1995, pp. 318–340