Comparison of integral structures on the space of modular forms of full level
Abstract
Let and be integers and be a prime such that . One can consider two different integral structures on the space of modular forms over , one coming from arithmetic via -expansions, the other coming from geometry via integral models of modular curves. Both structures are stable under the Hecke operators; furthermore, their quotient is finite torsion. Our goal is to investigate the exponent of the annihilator of the quotient. We will apply methods due to Brian Conrad to the situation of modular forms of even weight and level over to obtain an upper bound for the exponent. We also use Klein forms to construct explicit modular forms of level whenever , allowing us to compute a lower bound which agrees with the upper bound. Hence we are able to compute the exponent precisely.
1 Introduction
Let be an integer and let be a congruence subgroup of of level i.e. a subgroup of containing the kernel, , of the usual reduction mod map
Consider the -vector space consisting of modular forms of weight and level over . We associate to each modular form in and each cusp of , the -expansion of at , denoted , which is a power series in .
We concern ourselves with -structures on the vector space i.e. -submodules of such that the natural map is an isomorphism. The space naturally has two different -structures. Define the first -structure to be
which consists of modular forms in whose -expansion at the cusp has integral coefficients. One can show the Hecke operators of preserve integrality at the cusp by explicitly computing the -expansion at under the Hecke operators (see [Kat04, §4.9.2]).
The second -structure we consider is , which consists of modular forms in whose -expansions at all cusps have integral coefficients. This structure is also stable under the Hecke operators (cf. [Con07, Theorem 1.2.2]). There is an obvious containment with quotient that is a torsion -module. Our aim is to study and determine the annihilator.
To better understand and work with , we realize as the global sections of some line bundle on , the moduli space parameterizing -level structures over . More precisely, for an integer ,[KM85, §3] considers four different moduli problem parameterizing -, -, balanced -, and -structures on elliptic curves. The definition of -structures is given in Appendix A since this will be our focus. When , the moduli problems parameterizing -, -, and balanced -structures are represented by a regular, flat two-dimensional scheme over , by [KM85, 5.5.1]. For the rest of the introduction, we let denote one of these level structures. The scheme extends to an arithmetic surface over , known as the modular curve, so that is identified with . From this, we can show that and are both finitely-generated -modules of the same rank. Thus the quotient is torsion.
Let be a prime. If does not divide the level , then the -adic valuation of the annihilator is zero (see Remark 2.8). Thus we focus on primes which divide the level and instead can work over . Let be a cusp of . Define to be the minimal -adic valuation among all the coefficients of . Then our general goal can be restated as follows: we seek to compute the smallest integer such that for all cusps of not equal to and all .
The problem of computing and bounding has a long history, which we briefly recall. Computing a bound for for arbitrary weight and level was done in [DR73, §3.19, §3.20], where they obtained an upper bound of
Their methods involved using intersection theory on and the fact that is reduced with two irreducible components.
In [Edi06], Edixhoven investigates the situation of weight and level cusp forms where . In particular, he establishes the existence of a non-zero global section of where is the relative dualizing sheaf of , is essentially111More precisely, since is not necessarily regular, [Edi06] works instead with some finite cover by appending extra level structure to . Let denote the irreducible component of containing the cusp . Then is the inverse image of under . the irreducible component containing the cusp , and is the divisor given by the sum of the supersingular points. By computing in terms of and using the inequality
Edixhoven is able to conclude which forces whenever . This agrees with the bound provided by [DR73]. In the case , Edixhoven separately concludes .
In [BDP17, Appendix B], Conrad investigates in the general situation and arbitrary weight . He begins by developing intersection theory on the regular proper Artin stack over . It is worth noting the stack in [Con07] which Conrad considers does not agree with the one in [DR73]. In particular, as defined in [DR73] is a Deligne-Mumford stack instead of merely Artin (see [Čes17] for more on this issue). We will not need to concern ourselves with this issue since we only work with the arithmetic surface which parameterizes -structures. Furthermore, Conrad’s expression for the exponent readily holds for as we show in Section 2. Let denote the matrix obtained by removing the column and row corresponding to the irreducible component containing the cusp from the intersection matrix of . Conrad provides an expression for the upper bound in terms of and the entries of . As pointed out in [ČNS23, footnote 5], this bound is incorrect for due to a typo in the values of the multiplicities of the components of used in the calculation of . In our situation, since is reduced, these multiplicities are all equal to . For general , it is not clear how to obtain a uniform description of the entries of , and this prevents Conrad from establishing explicit bounds for general .
Lastly, [ČNS23] uses an automorphic approach to bound the exponent in the situation . This approach appears quite powerful; in particular [ČNS23, Theorem 4.6] provides bounds depending on the cusp at which the q-expansion is being taken. These bounds are also shown to be sharp in a few cases via explicit computations (see [ČNS23, Example 4.8]). We will still follow the approach in [BDP17, Appendix B] as this yields an upper bound which we show is sharp in all cases except two: when the level is exactly or . Consequently, this provides an exact computation for the exponent.
Let and be an integers and be a prime such that . Fix a primitive root of unity and let be a uniformizer of . We consider the situation of the modular curve over , which parameterizes -structures (see the paragraph preceding Theorem A.4, or [KM85, §9]). In this situation, the special fiber of consists of many irreducible components. We also restrict ourselves to modular forms of even weight. Applying Conrad’s method in this situation, we obtain a formula in terms of the entries of and , where is any irreducible component of the special fiber of . We make the following two significant computations:
-
•
We explicitly compute the entries of for all and .
-
•
We explicitly compute . Note that Conrad in [BDP17, Appendix B] bounds by . In our situation,
independent of , which is smaller than by a factor of . This savings leads to a marked improvement in our upper bound.
In Theorem 3.20, we calculate the entries of for all . First we describe the entries of by explicitly computing the intersection number between each irreducible component of the special fiber, using [KM85, 13.8.5], then we compute the self-intersection numbers.
We compute
where is a circulant matrix, dependent on and , is the matrix with the first row and column removed, is the matrix consisting of all 1’s, and is the degree of supersingular locus in . Applying the Woodbury Matrix Identity (see Appendix C) we obtain a formula for the entries of involving the entries of and , as well as their row sums and the total sum of all entries.
Since is circulant, we have a description of its eigenvalues and corresponding eigenvectors, allowing us to diagonalize . In turn, this allows us to explicitly compute the entries of . Consequently, using Proposition 3.13, we also obtain the entries of the inverse of . After more careful calculations, we obtain an exact expression for each entry of .
In Theorem 4.25, we compute . Using the Kodaira-Spencer isomorphism (Theorem A.12) and the adjunction formula (Theorem 4.3), we are able to identify with the relative dualizing sheaf
twisted by minus the cuspidal divisor , tensored with relative dualizing sheaf of the map induced by forgetting the -level structure.
Investigating amounts to understanding the different of the morphism , which amounts to understanding the different of
where is a codimension 1 point. Let denote the valuation of the different of in . We split our analysis of into two cases. If is a closed point of the generic fiber, then we can compute over . This amounts to understanding the ramification of the analogous map of Riemann surfaces , which is done in [Shi94, Prop. 1.37]. If is a generic point of the special fiber, then we show all contributions are the same which taken together contribute nothing to the different of . We ultimately conclude
and consequently
This identification allows us to directly compute by computing the quantities
The relative dualizing sheaf enjoys strong functoriality properties, enabling us to identify
This allows us to compute in terms of the genus of an Igusa curve, and the self-intersection number of . Combined with our calculation of the entries of , we arrive at the following upper bound.
Theorem 1.1.
Let , , and be integers and be a prime such that . The exponent of in the annihilator of
is bounded above by
For , let denote the Klein form associated to (see Definition 5.1). A lower bound for is obtained by explicitly constructing a modular form of level out of a product of Klein forms. Let be a family of integers. [EKS11, Thm. 2.6], which builds upon results in [KL81, §2.1, §2.4], provides a criterion on for when a product of Klein forms
is a nearly holomorphic modular form of level and weight . We are able to choose a family such that is a modular form of level and weight with integral -expansion at . The inclusion induces a map between the corresponding modular curves over . Pulling back under this map and taking the th power , we can view as a modular form of level and weight . By explicitly computing the -expansion of at the cusp , we obtain the following lower bound.
Theorem 1.2.
For , and , is bounded below by . Consequently, is equal to .
We are also able to obtain an upper bound in the situation of cusp forms. Although Edixhoven only considers with , his method can be adapted to yield an upper bound in our situation as well. However, this upper bound is worse than the one we obtain by a factor of , yet coincides for (see Remark 4.30).
We now briefly summarize the contents of this paper. In Section 2, we formulate our problem of computing geometrically, and the resulting formula for it in terms of the intersection theory of , and the degree of . We follow [BDP17, Appendix B] to provide an expression that calculates the exponent in the situation of . We also provide material on intersection theory for arithmetic surfaces, most of which comes from [Liu02, §8, §9].
Section 3 is devoted to describing the intersection matrix of , and then computing the entries of . In Section 4, we compute . Combined with the work in Section 3, this culminates in Theorem 4.28 where we provide an upper bound for the exponent. Lastly, we construct explicit modular forms of level in Section 5 which provides a lower bound for the exponent that agrees with the upper bound.
1.1 Notation
If is a morphism of schemes, for any -scheme we let denote the base change of along . When and , we will often abuse notation and write for . If is a scheme over a DVR, we denote the special fiber of by .
Let be an integer. We denote the “compactified” regular integral model of the modular curve of full level by over the cyclotomic integers , as presented in [KM85]. Refer to Appendix A for more on the modular curve. We denote the cuspidal locus of by or sometimes . We let denote the modular sheaf of (see the paragraph proceeding Theorem A.8). For an integer , the global sections of define the space of modular forms of weight and level .
Starting in §2.2 and onward, we will usually consider modular forms of level where is a prime number such that and . We will also be working over the DVR which has uniformizer .
2 The exponent
2.1 Formulation of the exponent
Let denote the “compactified” regular integral model of the modular curve of full level and let denote the modular sheaf over following the notation in §1.1. We will consider two different sub -modules of the space of modular forms over . The first structure is , the space of modular forms over (see Definition A.10), which has the following description in terms of -expansions.
Lemma 2.1.
We have
Proof.
This comes from the -expansion principle, as stated in Proposition A.17. ∎
Let be a -algebra contained in . The second structure is defined as
which are modular forms over whose -expansion at the cusp has coefficients in .
Lemma 2.2.
The usual map
is an isomorphism.
Proof.
By Proposition A.11, we have
In particular, the coefficients of any -expansion of a modular form over have bounded denominators. Therefore we can write the -expansion of any at as for some and integer , i.e. . Then is the image of . ∎
Definition 2.3.
Let be a module over a ring . Then annihilator of is the ideal
The annihilator of an element is the ideal
The next proposition showcases some properties of our two -modules necessary to discuss the annihilator of the quotient .
Proposition 2.4.
-
a.
Both and are finitely generated -modules of the same rank.
-
b.
We have
Proof.
Since is projective over , is a finitely generated -module by [Liu02, Theorem 5.3.2]. Since is a submodule of a finitely generated module over a noetherian ring, namely of , it is also finitely generated.
Now we show the second claim. By Proposition A.11, we have
while, by Lemma 2.2, we have
Since modular forms over are determined by their -expansion at (see [Kat73, 1.6.2]), we get an equality
as desired. In particular, we have an equality of finite dimensional vector spaces
which shows both our -modules have the same rank. ∎
Our description of in Lemma 2.1 shows that it is contained in . Having established that and have the same rank, the quotient
is a torsion -module. Our goal will be to investigate the annihilator of this quotient.
By localizing, we can focus our attention on investigating a single prime in the annihilator. The following lemma, which is [Sta21, Tag080S], shows how the annihilator behaves under flat base change.
Lemma 2.5.
Let be a flat ring map. Let be an -module and . Then
for any . If is finite over , then
Corollary 2.6.
Let be a finitely generated torsion module over a Dedekind domain and let be a prime in . The exponent of appearing in is equal to the exponent of appearing in .
Proof.
We factor the annihilator as a product of distinct primes . By Lemma 2.5, we have
Hence the exponent of appearing in is precisely . ∎
Proposition 2.7.
Let be a prime lying over the prime . The exponent of appearing in the annihilator of and are the same.
Proof.
For convenience, we let
and
Consider the exact sequence Since is flat over , the above sequence remains exact after tensoring with . Thus we have a commutative diagram
By Proposition A.11, and are isomorphisms; consequently is an isomorphism. By Corollary 2.6, the exponent of appearing in agrees with that of . ∎
Remark 2.8.
If , then is invertible in so by Proposition 2.4 we have
Hence the -adic valuation of the annihilator of is trivial in this case. Therefore we restrict our attention to primes dividing the level.
Let and be integers and let be a prime such that . Let be a uniformizer of . We seek to compute the smallest integer such that
2.2 Geometric interpretation of the exponent
For convenience, we let . We will formulate an explicit, geometric description of the exponent by first providing an algebraic description coming from -expansions.
Let be a cusp of , and let be a non-zero modular form. Let denote the -expansion of at the cusp . Define to be the minimal -adic valuation among all the coefficients of i.e.
Note that this minimum exists since the denominators of the coefficients are bounded (see proof of Lemma 2.2). Thus, by Lemma 2.1, we may describe
Consequently, we seek to compute the smallest integer such that for all cusps of not equal to and all .
Next we will provide a geometric interpretation of which does not rely on -expansions. Let denote the generic point of an irreducible component of the special fiber . The stalk is a DVR so it has a valuation which we denote by . Viewing inside the stalk , we can write where is the canonical generator of (see Definition A.16). We define
Indeed, is a valuation, independent of the choice of local generator. The following result is stated in [DR73, Théoréme 3.10(ii)].
Proposition 2.9.
Let be a cusp and let be an irreducible component of with generic point on which lies. For any non-zero , we have . Furthermore, is a uniformizer of .
Proof.
Since , the stalk is a localization of . Furthermore, the map is injective since all stalks are regular local rings so are integral domains in particular. The induced map on completions is also injective by [Sta21, Tag00MB, Tag07N9].
We will show is also a uniformizer of . Consider the exact sequence
where is the maximal ideal of and is the residue field of which is of characteristic . Since when , we have in so in . Therefore . It remains to show is a maximal ideal in . Consider the quotient . Since corresponds to a minimal prime and is reduced, the stalk is a field. Therefore is a maximal ideal in hence is a uniformizer in .
Let be non-zero. Using the local generator of (see Definition A.16), we write where is the -expansion of at . Write where . The map
is the same as the map , which is also a localization map as .
By [KM85, 10.9.1(2)], is smooth at the cusps so, in particular, is a domain. Therefore is injective so the map on completions
is injective. Since , we can conclude i.e. is a unit in . Thus coincides with , the valuation in .
∎
With this geometric description of the valuation, we can reformulate our integrality condition as follows. Let be the irreducible components of the special fiber where contains the cusp . Let denote the multiplicity of . We desire to find the smallest integer such that for all and all .
By Proposition 2.9, is a uniformizer of , where is the generic point of . Hence
Since is reduced, . Hence the condition is equivalent to . Therefore we have the following expression for the exponent:
(1) |
which illustrates we can investigate by trying to understand the quantities .
2.3 Intersection theory on arithmetic surfaces
In this section we briefly recall some important intersection theory facts from [Liu02, §9]. By Theorem A.8, the modular curve is an arithmetic surface (see Definition A.7). We begin with a more general situation with an arithmetic surface over a Dedekind domain and a line bundle on .
Since is regular, we have, by [Liu02, 7.2.16], an isomorphism between Cartier divisors and Weil divisors given by
where . Furthermore, this map respects principal divisors and effective divisors. Let be a closed subscheme of and let be the generic points of . The Weil divisor associated to is given by
Let be an irreducible Weil divisor of . According to [Liu02, 8.3.4], is either an irreducible component of a closed fiber, or the closure of a closed point of the generic fiber .
Definition 2.10.
If is an irreducible component of a closed fiber, or equivalently is a point, then is called vertical. If is instead the closure of a closed point of , or equivalently , then is called horizontal.
Recall a Weil divisor is a formal -linear sum of irreducible closed subsets of codimension 1. In general, a Weil divisor is horizontal (resp. vertical) if all its irreducible components are horizontal (resp. vertical).
Let and be two effective divisors of with no common irreducible component. As in [Liu02, §9.1.1], we define the (local) intersection number of and at a point by
By definition, we immediately have that is symmetric and bilinear. If , then .
We will now establish the intersection number between a general divisor of and a vertical divisor of a fixed closed fiber. For a fixed closed point , let denote the set of divisors of with support in ; such divisors are vertical divisors. A -basis of consists of all irreducible components of .
Theorem 2.11.
Let be a closed point. Then there exists a unique bilinear map of -modules
such that:
-
a.
If and have no common component, then
where the sum is over all closed points.
-
b.
restricted to is symmetric.
-
c.
if are linearly equivalent.
-
d.
If , then . .
-
e.
If is principal, then .
Proof.
This is proved in [Liu02, Theorem 9.1.12] besides part (e), which we prove. By [Liu02, Corollary 9.1.10], there exists a principal divisor such that and have no common component. Thus we may assume and have no common component. By (a), it suffices to show if is principal, then the local intersection number is zero. By definition,
Since is principal, . Thus
∎ |
Definition 2.12.
Let be an arithmetic surface, and let be closed. For any in , we call the intersection number as defined in Theorem 2.4. More generally, if is a vertical divisor, we define
where the sum is over all closed points of , which is a -cycle on . We call the self-intersection numberof . If is only concentrated at a single point , as in the case when is the spectrum of a DVR, we identify with the integer .
Proposition 2.13.
Let be an arithmetic surface and fix a closed point . Let be the irreducible components of . We have an equality of Weil divisors
where is the Weil divisor associated to the Cartier divisor .
Proof.
The first equality follows from the definition of . Let and be uniformizers where is the generic point of . Let denote the normalized valuation of associated to and let denote the generic point of .
In , we can write for some . Then
Note that
∎ |
Lastly we state some results on the intersection between a horizontal divisor and a closed fiber which won’t be used until §4.4. The following is [Liu02, Proposition 9.1.30].
Proposition 2.14.
Let be an arithmetic surface. Let be the generic point of and a closed point. Then for any closed point , we have
where is the Zariski closure of in , endowed with the reduced closed subscheme structure.
Corollary 2.15.
Let be an arithmetic surface and let be a -rational point. Then is reduced to a single point and .
In particular, intersects exactly one irreducible component of . Moreover, has multiplicity and .
Proof.
Since is -rational, by Proposition 2.14 we have . Let denote the irreducible components of with multiplicities respectively. Then
Since and are effective divisors with no common components, which forces for some . Hence . Moreover, so for i.e. does not intersect . ∎
2.4 A more explicit description of the exponent
In this section, we will provide a more explicit description of the exponent by relating the quantities to intersection numbers and the degree of a line bundle, due to Conrad in [BDP17, Appendix B]. Let be an arithmetic surface over a DVR . Let be a line bundle on and let denote the special fiber of .
Definition 2.16.
Let be a non-zero global section and let be a trivializing open cover of so that for some generator . We can write for some . Then the system is an effective Cartier divisor of , which we denote by or simply if the line bundle is clear from context.
We can equivalently define as a Weil divisor. Let be a prime divisor of with generic point . Then is a DVR with valuation which we denote by . We can write the image of in the stalk of as for some and generator . Define which agrees with our valuation defined in Section 2.2 for an irreducible component of and is independent of choice of . According to [Sta21, Tag02SE], the Weil divisor associated to is equal to
where the sum is over prime divisors of . Decomposing the divisor into its horizontal and vertical components, we can write
(2) |
where is some effective horizontal divisor and the sum is over the irreducible components of .
Proposition 2.17.
Let be a non-zero global section. Then .
Proof.
Decomposing into its horizontal and vertical components, as in (2), we get
Using the equation
(3) |
we will provide an explicit expression for the quantities for where is the multiplicity of .
Let denote the intersection matrix of . Since is an arithmetic surface, [Liu02, 9.1.23] (see also [Lan88, III, 3.4]) says is negative semi-definite and moreover the kernel of is one-dimensional whenever is connected. Multiplication by induces an exact sequence
Let which is a non-zero vector in . Note that
where the last equality is due to being principal. Therefore we can write . The following lemma describes the image of .
Lemma 2.18.
We have
Proof.
This is shown in the paragraph preceding [BDP17, Remark B.2.3.1]. We will provide the proof here for convenience. Let denote the space of vectors such that . Let so for some . Then
so
Therefore . Since and are both of dimension , we get . ∎
In order to isolate the terms appearing in (3), we would ideally invert the matrix . In light of being not invertible, we will instead invert an submatrix of to obtain an expression for each , excluding , which suffices for our purpose.
Let and let denote the projection map
Consider the restriction . Given any , we let
which forces by Lemma 2.18. Hence is surjective. Since and are both of dimension , the map is an isomorphism of vector spaces.
Next we consider the restriction . If where , then for some . Coordinatewise, this means
Comparing the first coordinates and noting for each , we must have . Hence so is injective and therefore an isomorphism. Define as the composition of our isomorphisms
After identifying by forgetting the first coordinate, we can identify as the lower right submatrix of . The following is [BDP17, Proposition B.2.3.2].
Proposition 2.19.
For any and such that , we have
(4) |
We now apply Proposition 2.19 to our specific situation involving the quantities .
Remark 2.20.
We make a quick remark about the term that appears in the following theorem. Recall the equation for the exponent in (1). We claim that
Indeed, the maximum on the left hand side must occur at some with . Otherwise if with , , and , so
Hence so the maximum must occur over such .
Furthermore, the differences are visibly invariant under arbitrary scaling of , so we have
In particular, this shows the exponent can be computed using Theorem 2.21 below, which provides a formula for the differences in terms of geometric data.
Theorem 2.21.
Let and be non-zero and let denote the irreducible components of . We have
(5) |
where is the entry of and is the matrix obtained by removing the first row and column of the intersection matrix of .
Proof.
We will write out equation (4) entry-wise in the generality of Proposition 2.19, with an arithmetic surface, and a line bundle on . Then we apply our results to the situation of the modular curve.
Let and such that . Then equation 4 gives
Recall is the lower right submatrix of the intersection matrix of . Write where is the -entry of the inverse of . Then we have
Therefore
Let be non-zero and let . The th coordinate of is precisely
by equation (3). Thus we have
(6) |
Now we take to be our modular curve and to be the modular sheaf. Recall is reduced so for each . Then equation (6) becomes
as desired. ∎
Thus Theorem 2.21 expresses the quantities in terms of , , and the entries of the inverse of .
3 Intersection matrix
Throughout this chapter, we will use the following notation:
-
•
and will be integers and will be a prime such that
-
•
will denote the residue field of where where is the order of in .
-
•
will denote the modular curve .
3.1 Intersection matrix for the modular curve
In this section, we will obtain an explicit description of the intersection matrix of . First we recall the description of the irreducible components of the special fiber as in Theorem A.25.
Theorem 3.1.
The special fiber of is the disjoint union, with crossings at the supersingular points of , of the exotic Igusa curves over , indexed by
Furthermore, is reduced.
We will identify an irreducible component of by it’s index . By [KM85, 13.8.5], the local intersection number at a supersingular point between two distinct irreducible components and is precisely
(7) |
Remark 3.2.
In [KM85, 13.8.5], we require to be a -rational supersingular point. However, when , the supersingular points may not be -rational. Indeed, [KM85, p. 96], shows that the supersingular points are -rational. However, we can instead compute the intersection numbers by étale base change. Consider the ring . By [Ser79, IV, §4, Prop 16 & 17], is the ring of integers of the local field so is a DVR. Since has order in , the residue field of is . The map is unramified by [Ser79, IV, §4, Prop. 16], noting that , and flat because is torsion-free over the DVR (see [Sta21, Tag0539]). Therefore is étale.
Let and be two irreducible components of . Let and let be a supersingular point of which maps to . By [Liu02, 9.1.5, 9.1.6], we have
Note that [Liu02, 9.2.15] uses a desingularization of i.e. a proper birational morphism where is regular. Since and are both DVRs, they are both regular, excellent, and noetherian. Therefore is an arithmetic surface (by Proposition A.14), so we can take .
From equation (7), the local intersection number doesn’t depend on the supersingular point. Let (resp. ) denote the supersingular locus of (resp. ). By [KM85, 12.7.2], the map is totally ramified at the supersingular points so . Therefore to obtain the global intersection number we multiply by . We will now compute the local intersection number between each pair of irreducible components of , beginning with the case of distinct pairs.
Proposition 3.3.
Let denote the -adic valuation normalized so and let be a supersingular point. We have
for distinct and distinct.
Proof.
Since each group homomorphism corresponding to an irreducible component is a surjective, we know . By definition,
Therefore
Similarly, since
we have
By equation (7), we have
We will now compute by considering the following three cases.
Case 1: We have
Note that
is invertible in . Therefore is a basis for hence
We conclude .
Case 2: We have
For any , consider the injective group homomorphism
given by multiplication by . This gives us a filtration
which is exhaustive and separated. Thus for any non-zero , there exists a unique smallest such that . We let denote the quantity . We claim for some . Indeed, if is not a unit of , then so for some . Hence , contradicting the minimality of .
We will show the sequence of abelian groups
(8) |
is exact. Exactness at the second and fourth term are clear. Let . Then
so . On the other hand, let so . Then
Since , we have . Hence or equivalently . Therefore allowing us to conclude exactness at the third term.
Let . Consider the sequence of abelian groups
(9) |
where and
One can show exactness similar to the above sequence (8). Indeed, showing is clear. Conversely, if , then and consequently . Writing , we similarly conclude so . Using exactness of (9), we get
We conclude .
Case 3: Lastly, we have
Arguing as above by creating a sequence similar to (9) and replacing (resp. ) with (resp. ), we get
We conclude
.
To finish our calculation of intersection numbers, we will now compute each self-intersection. First we introduce the following lemma.
Lemma 3.4.
Let . We have
Proof.
We will group the index based on its -adic valuation and then sum over . Observe that there are precisely -many elements in with -adic valuation equal to . Thus
as desired. ∎
Proposition 3.5.
For any , the self-intersection number is
Proof.
By [Liu02, Proposition 9.1.21], we have
where is the multiplicity of for any irreducible component . Since is reduced, .
We will now describe the intersection matrix by specifying four blocks which make up . We label the columns (and by symmetry the rows) of in the following order:
so that the entry of is equal to the intersection number between the th row label and th column label. Since is a common factor among each entry of , we will describe the matrix to simplify exposition.
Let (resp. ) denote the submatrix of corresponding to the column and row labels of the form (resp. ). We let and denote the remaining two submatrices of so that
We also let denote the matrix to highlight the dependence on . By convention, we define to be the matrix consisting of the entry . The matrices and take on a special form.
Definition 3.6.
An circulant matrix is of the form
where each column is equal to the previous column shifted downward by 1, looping around as appropriate.
Refer to Appendix B for a discussion on circulant matrices, including an explicit description of the eigenvalues, eigenvectors, and inverse in terms of the entries of the matrix and roots of unity.
Proposition 3.7.
-
a.
The entries of and are all equal to 1.
-
b.
The matrix is a circulant matrix whose first entry of the first column is . For , the entry in the first column is equal to .
-
c.
The matrix is a circulant matrix equal to .
Proof.
Proposition 3.3 immediately tells us the entries of and are all 1. Next we describe the first column of . The first entry is equal to the local self-intersection number of , which is by Proposition 3.5. Using Proposition 3.3, for , the entry of the first column is equal to the local intersection number
Now we show is circulant. Recall the th column of corresponds to the label for . The th entry in the th column is equal to
Note for any nonzero , if (mod ), then . Therefore the quantity
for remains unchanged if we take modulo . We conclude
which says the th column is equal to the first column with every entry shifted downward by , looping around as appropriate. Hence is a circulant matrix.
Lastly, we show . Suppose . By Proposition 3.3, the entry of for is equal to
which is equal to the entry of multiplied by . When , the entry of is while the entry of is
Thus .
When , is a matrix consisting of the entry . By convention so in the case . ∎
Example.
The intersection matrix for is the matrix
where the entries in bold comprise and the entries in italic comprise . The intersection matrix for is the matrix
3.2 Inverting
Recall our goal is to invert the matrix obtained by removing the first row and column of . For a general matrix , let denote the matrix obtained by removing the first row and column of and let denote the matrix whose entries are all 1. Using Proposition 3.7, we have the following description of :
We will use the following identity to invert .
Proposition 3.8 (Woodbury Matrix Identity).
Let be an invertible matrix, an invertible matrix where , an matrix, and a matrix. Then
Refer to Appendix C for a discussion on using the Woodbury Matrix Identity to compute the inverse of in the situation that is a block diagonal matrix with two blocks, is the identity, and both and consists of 0’s and 1’s which we specify later. Note that the inverse of a block diagonal matrix is obtained by inverting each block. Thus computing amounts to computing and which we will do in this subsection.
Before we compute the eigenvalues of , we will need the following technical lemma.
Lemma 3.9.
Let and be integers. We have
Proof.
By the geometric partial sum formula, we have
Next we consider the sum
Therefore
Lemma 3.10.
The eigenvalues of are and
for .
Proof.
By Lemma B.2, the eigenvalues of an circulant matrix whose first column has entries are given by
For , the eigenvalues are therefore given by
for .
Assume . When , we have
Since for , we have
As , we have so . Using the geometric series partial sum formula, we get
Lastly we handle the case. We compute the sum appearing in the expression for by breaking it up according to the value of . We have
We can rewrite the index in each sum as where and . Re-indexing, with , we get
Since for , we have
(10) |
Let . We will simplify (10) using Lemma 3.9 with , and . For ease of exposition, we split into two different cases depending on and will consequently obtain our desired expression for .
Corollary 3.11.
is invertible.
Proof.
By Lemma 3.10, all the eigenvalues of are nonzero hence is invertible. ∎
Now that we know the eigenvalues of , we can use Proposition B.4 to compute the inverse of .
Proposition 3.12.
Let denote the -entry of . We have
Proof.
By Lemma B.4, the -entry of is equal to
where are the eigenvalues of as in Lemma 3.10. Continuing,
We split into two cases, breaking down the sum in a similar manner as in the proof of Lemma 3.10. We have
Case 1: Suppose . Then
Thus the -entry of is
Case 2: Suppose and let . Then this situation resembles that of Equation (10) which we have already computed.
Thus the -entry of is
as desired. ∎
3.3 Inverting
Having calculated the entries of , we can calculate the entries of using the following result. We will provide a sketch of the proof. A full proof can be found in [JCP16, Theorem 2.2].
Proposition 3.13.
Let be an invertible matrix and let . Let and let denote the matrix obtain from by removing the row and column. Then the -entry of is given by
for with and .
Proof.
Write . Let denote the column of after removing the component and let denote the row of after removing the component. Then one can verify
where is the identity matrix. Using the Sherman-Morrison formula, which is a special case of Proposition 3.8, to calculate , we get
Therefore the entry of is
∎ |
Proposition 3.14.
Let and let . We have
3.4 Inverting
Recall in Section 3.2 we wrote where
The matrix is rank 2 and can be written as where is the identity matrix, is the matrix whose first and last column are the same as those of , and is the matrix
where the first entries of the first row of are all 1 with the remaining entries are all 0 and the first entries of the second row are 0 while the remaining entries are all 1. Note that .
In the more general situation where is an invertible block diagonal matrix with
an explicit formula for the entries of is provided in Proposition C.1. We state it here for convenience:
Proposition 3.15.
Let where and are the given matrices above. Let denote the -entry of . We have
where is the sum of all entries in the first block in and is the sum of all entries in the second block in .
To obtain a closed formula for the entries , we will therefore need to calculate the row and column sums of and . Note that both these matrices are symmetric so it suffices to compute, say, the row sums. The following lemma will be used when computing these sums.
Lemma 3.16.
We have
Proof.
The number of positive integers with valuation is precisely . Hence
Using the identity
with and , we obtain
∎ |
Corollary 3.17.
Let . We have
Proof.
Write and . We first compute the row and column sums of and the quantity , the sum of all the entries of .
Lemma 3.18.
Consider the matrix . We have
for all .
Proof.
Since is circulant, the inverse is circulant by Corollary B.5. Therefore all the row sums are the same. Furthermore, by Lemma 3.10, is an eigenvalue of with corresponding eigenvector (see Lemma B.2).
Note that the entries of are precisely the row sums of . Hence the row sums of are all . Observe that
so the row sums of are all . Consequently the row sums of are all . The matrix has rows so
as desired. ∎
Next we compute the row sums for which is substantially more tedious than Lemma 3.18, noting that fails to be circulant in general.
Lemma 3.19.
Consider the matrix and let . Fix a row . We have
and
Proof.
Using Proposition 3.14, we compute
(11) | ||||
(12) | ||||
(13) |
We will now compute the sums above, starting with (12).
Using Corollary 3.17, we have
Now we compute (13). Note that the third expression in (11) is what would be the term in the sum. We can combine them and compute instead
Combining everything together, the th row sum is
Lastly, we compute :
∎ |
We will now provide an explicit description of the entries of .
Theorem 3.20.
Let denote the -entry of . We have
Proof.
We will break into four cases, using Proposition 3.15 to calculate the along with our results in Lemma 3.18 and Lemma 3.19.
Case 1. Suppose . Then
If , then is equal to
If , then is equal to
Case 2. Suppose and . Then
Case 3. Suppose and . Then
Case 4. Suppose . If , then
If , then
∎ |
The following corollary will be useful when we find an upper bound for the exponent in Theorem 4.28.
Corollary 3.21.
Let denote the -entry of . Then each is negative.
Proof.
Based on our result in Theorem 3.20, we will show the third case is negative i.e. we will show
(14) |
is negative for . The other cases are either clearly negative or are essentially the same as this case.
Since the largest value attains is , the largest value expression (14) attains is
which is always negative. ∎
4 Computing the degree of the modular sheaf
Throughout this chapter, we will keep the notation of §3. Unless otherwise stated, we let . We will compute the degree of restricted to an irreducible component of and ultimately compute an upper bound for the exponent .
4.1 Decomposing the Modular Sheaf
We will make use of the Kodaira-Spencer isomorphism, as stated in Theorem A.12, which we restate here for convenience.
Theorem 4.1.
Let be a noetherian, regular, excellent -algebra containing . The Kodaira-Spencer isomorphism on extends to an isomorphism on
We will need the following definition, which we take from [Liu02, 6.4.18].
Definition 4.2.
Let be a proper morphism of relative dimension . A relative (-)dualizing sheaf for is a quasi-coherent sheaf on , endowed with a homomorphism of -modules
such that for any quasi-coherent sheaf on , the natural bilinear map
induces an isomorphism
By [Liu02, 6.4.19], uniqueness of is automatic once we have existence. If is locally noetherian and is a projective morphism with fibers of dimension , then as remarked in [Liu02, 6.4.30], the relative -dualizing sheaf exists. Furthermore, by [Liu02, Theorem 6.4.32], the relative dualizing sheaf is isomorphic to the canonical sheaf (see [Liu02, 6.4.7]) whenever is a flat projective l.c.i. and is locally noetherian. When is smooth, the canonical sheaf coincides with the sheaf of Kahler differentials . We will use the following result, which is in [Liu02, Theorem 6.4.9] known as the adjunction formula, to eventually relate with the relative dualizing sheaf of and of .
Theorem 4.3.
Let and be quasi-projective l.c.i.s. We have a canonical isomorphism of canonical sheaves
We cannot directly apply the Kodaira-Spencer isomorphism to our modular curve since and consequently the level , is not invertible in . Instead, we will apply it to the modular curve over since the level is invertible in . Consider
(15) |
where is the projection map and is the structural morphism. For convenience, we let denote the base change .
According to [Liu02, 6.3.18], if is a morphism of finite type of regular locally noetherian schemes, then is an l.c.i. Therefore the maps and are l.c.i.s. By Theorem A.8, and are projective. Hence by [KM85, 3.3.32(e)], is projective. Applying the adjunction formula to (15), we have
Combining this with the Kodaira-Spencer isomorphism applied to , we get
(16) |
We will later show in Lemma 4.18
which is where the sheaf appears in (16). This will then allow us to identify with , the relative dualizing sheaf twisted by the cuspidal divisor. Thus, computing amounts to computing . Our first step will be to investigate .
4.2 The relative dualizing sheaf
In this section, our goal will be to better understand the relative dualizing sheaf . Since is invertible, we have for some divisor of . As we will see, will be the divisor associated to the different of the morphism
We begin by discussing the trace map which generalizes the usual notion over a finite extension of fields. Let be a finite, flat map of noetherian rings. According to [Sta21, Tag0BSY], is a finite locally free -module and so we can consider the trace of the -linear map given by multiplication by . This gives us an -linear map . The following definition is from [Sta21, Tag0BW0].
Definition 4.4.
Let be a ring map and let , the total ring of fractions of (see [Sta21, 02C5], note when is a domain, coincides with the field of fractions), and . We say the Dedekind different is defined if is noetherian, is finite and maps any non-zerodivisor of to a non-zerodivisor of , and is étale. In this situation, is finite flat. Let
We define the Dedekind different of to be the inverse of :
viewed as a sub -module of .
Remark 4.5.
Let be a Dedekind domain, , a finite separable extension of , and the integral closure of in . In this situation, [Ser79, §4.3] defines the different in the same manner as we have done. Since is normal and noetherian, by [Sta21, Tag032L], is finite. Furthermore, and so indeed, the Dedekind different is defined for . We record a few useful facts for calculating the different in this situation.
Proposition 4.6.
Let be a Dedekind domain, , a finite separable extension of , and the integral closure of in .
-
a.
Let be a non-zero prime of such that the corresponding residue extension is separable and let denote the ramification index of . Then the exponent of in the different is greater than or equal to with equality precisely when is tamely ramified.
-
b.
Suppose for each prime of , the corresponding residue extension is separable. The annihilator of the -module of Kahler differentials is equal to .
Lemma 4.7.
Suppose the Dedekind different is defined for . Let be a multiplicatively closed subset such that the Dedekind different is defined for . Then as -modules.
Proof.
First we show . By definition,
Now
Thus can be identified with
Lastly, we show . Let so . Then hence .
For the other inclusion, we first note that is finitely generated since is noetherian. Let denote the generators of over . Let so . Then for each there exists such that . Let . Then so . Therefore so .
In conclusion,
Let be a proper222More generally, one can define the different of a locally quasi-finite morphism of locally noetherian schemes, as in [Sta21, Tag0BTC]. morphism of locally noetherian schemes. According to [Sta21, Tag0BVG], the relative dualizing sheaf is the unique coherent -module such that for every pair of affine opens and with , we have a canonical isomorphism
If we further assume is flat, then by [Sta21, Tag0BVJ], there exists a global section such that whenever is finite, is identified with under the isomorphism.
Definition 4.8.
Let be a flat, proper morphism of noetherian schemes. The different is the annihilator of the cokernel
which is a coherent ideal sheaf .
By [Sta21, Tag0BW5], we have where is the quasi-coherent sheaf induced by the -module .
Lemma 4.9.
Let be a proper morphism of noetherian schemes. Let and be affine open subschemes such that . Let and suppose the Dedekind different is defined for . Then .
Proof.
Let be the prime of corresponding to and be the prime of corresponding to . We have
Let denote the closed subscheme associated to and let denote the Weil divisor associated to . The following is [Sta21, Tag0BWA].
Proposition 4.10.
Let be a proper morphism of noetherian schemes. If is invertible and is étale at the associated points of , then is an effective Cartier divisor and .
Explicitly, the Weil divisor associated to is
By definition, and . We will now show that the usual projection morphism satisfies the necessary properties to use the above results on the different. By [KM85, 5.5.1(1)], the map is finite and flat so the different is defined.
Suppose is a codimension 1 point. Let
denote the induced map on stalks. Since is of codimension 1 and is flat, we have is also of codimension 1. Consequently, both and are DVRs.
Proposition 4.11.
Let be a codimension 1 point. The induced map on stalks is a finite ring map. Furthermore, the induced map on fraction fields
is finite separable and is the integral closure of in .
Proof.
Let be an affine open containing and be an affine open containing such that . Since is finite, it is also integral so the induced map is integral. By [Sta21, Tag034K], the induced map on localizations remains integral so is integral. Since is an integral scheme, we have inclusions
Therefore . Similarly, we conclude Both and are normal schemes so (resp. ) is integrally closed in (resp. ). Having established
is integral, the integral closure of in is precisely .
Since is a finite map between two integral schemes, the extension of function fields
is a finite extension of characteristic zero fields, and hence separable. Note that and so is a finite separable extension. By [Ser79, I, §IV, Prop. 8], we can conclude is finite. ∎
Since is integral, it’s only associated point is its unique generic point. In Proposition 4.11, we deduced the map on function fields is finite separable so is étale at the generic point of . By Proposition 4.10, we have
(17) |
Since is a DVR in this case, is equal to the valuation of in . By Corollary 4.9 and Proposition 4.11, . Also, by Proposition 4.6(b), we can identify with
Let denote the valuation of the different in the DVR . Recall by [Liu02, 8.3.4] the codimension 1 points of are precisely the closed points of the generic fiber, and the generic points of the special fiber. We will compute when is a closed point of the generic fiber. If is a generic point of the special fiber, then we show all contributions are the same which taken together contribute nothing to the different of . However, one can compute the explicitly using strict Henselizations.
4.3 Computing for the closed points of the generic fiber
The generic fiber of is open so it suffices to compute over i.e. for the map
First we show the value of does not change after base changing . In particular, we can compute the value of over and use the classical theory of modular curves as compact Riemann surfaces.
Lemma 4.12.
Let be a finite type morphism of normal curves over a field and let be a field extension of . Let denote the usual projection morphism from base change and let . Then we have .
Proof.
We can equate with the valuation of the annihilator ideal of in . Since Kahler differentials are compatible with base change, we have
Since is flat, the map is flat. Furthermore, is finite over since is finite type. Thus, by Lemma 2.5, we have
Hence . ∎
Thus we can compute in the situation that our modular curves are over . Let denote the modular curve over of level . In this situation, is tamely ramified over so by Proposition 4.6(a), we have where is the ramification index of . We will investigate the ramification index of all points in under the usual projection map .
Let be a congruence subgroup. We also let denote the (complex) upper half plane and . As we will see shortly, investigating the ramification of a point amounts to understanding the stabilizer group
The points of can be classified depending on which elements of fix . Note that
for matrices not equal to . Hence the fixed points of a given are either conjugate complex numbers, a single real number, or two distinct real numbers. Based on this fact, the following definition from [Miy89, §1.3, p.27].
Definition 4.13.
We say is a/an
-
•
elliptic point of if there exists such that , or equivalently has two distinct fixed points and .
-
•
cusp of if there exists such that , or equivalently has a unique real fixed point .
-
•
hyperbolic point of if there exists such that , or equivalently has two distinct real fixed points.
-
•
ordinary point of if does not fix , excluding .
Since the matrices we consider are in , we will not have any hyperbolic points appearing. We denote . The following is [Miy89, Theorem 1.5.4] which describes the stabilizer groups .
Theorem 4.14.
-
a.
If is an elliptic point of , then is a finite cyclic group.
-
b.
If is a cusp of , then .
The following proposition, which is [Shi94, Prop. 1.37], relates the ramification index with the index of stabilizer groups.
Proposition 4.15.
Let be a finite index subgroup and consider the projection . The ramification index of a point under is equal to
We are now ready to compute the ramification index under our map .
Proposition 4.16.
Let denote the usual projection map and let . We have
Proof.
By [Shi94, Prop. 1.39], the congruence subgroup has no elliptic elements for any . Consequently, the ordinary points are precisely the points of . We will now apply Proposition 4.15 in the case and .
Note that the image of a cusp under is again a cusp and similarly for ordinary points. If is ordinary, then its stabilizer group is trivial so . If is a cusp, then there exists such that . Therefore
so it suffices to compute .
Let According to [Shi94, bottom p. 22], we have for that Furthermore, if and only if modulo , or equivalently . Since in our situation, we always have . Note that for any , we have . Therefore the order of in is equal to so
4.4 Relating the modular sheaf with
Let be a generic point of the special fiber. Recall the value of is equal to the valuation of the different ideal corresponding to the induced map on stalks
We will use the discussion in Section A.1 to provide an explicit description of . Recall in Theorem A.21 we have a commutative diagram
where and are the usual projection maps. By Theorem A.25, the restriction of to any irreducible component of is the map . We get a commutative diagram:
(23) |
In particular, the map is the same for each generic point of the special fiber. Hence, the value of is independent of in this situation.
Lemma 4.17.
We have
Proof.
Recall from equation (17) and the paragraphs proceeding it, we have
where the first sum is over closed points of the generic fiber and the second sum is over the generic points of the irreducible components of the special fiber. By Proposition 4.16, we have
where the first sum is over all the cusps in the generic fiber. As discussed above, the values appearing in the second sum are independent of . Using the fact that the special fiber is reduced, we have
Note that , when viewed as a divisor, is principal (see Proposition 2.13). Thus
as desired. ∎
Recall the map from (15):
Again, for convenience, we let denote the base change . To compute , our strategy is to first prove that using the isomorphism in (16):
Our next step is to provide a better description of .
Lemma 4.18.
We have an isomorphism of sheaves on
Proof.
Theorem 4.19.
We have
Corollary 4.20.
Let be an irreducible component of the special fiber of . We have
The following will allow us to calculate .
Lemma 4.21.
is equal to the number of cusps which intersect .
Proof.
Viewed as a divisor, is equal to the closure of all the cusps of the generic fiber of by Proposition A.15. By Theorem 2.11(d),
where the sum is indexed over i.e. over the cusps of the generic fiber.
All the cusps are rational, so by Corollary 2.15, intersects at a single irreducible component of the special fiber. Therefore
so is equal to the number of cusps which intersect as desired. ∎
For convenience, we let , the set of all cusps of the generic fiber. Using Lemma 4.21 and Proposition A.31, we obtain the following:
Proposition 4.22.
We have
To compute , we will use the following result in [Liu02, Theorem 9.1.37].
Theorem 4.23.
Let be a regular fibered surface, a closed point, and such that . Then we have
Corollary 4.24.
We have
Proof.
Recall that the space of cusp forms, by definition, are the global sections of .
Theorem 4.25.
Let be an integer. We have
and
Proof.
In general, taking tensor powers commutes with pullback of sheaves (see [Sta21, Tag01CD]). Hence and consequently
which gives our desired result. ∎
4.5 An Upper Bound
Recall at the end of Remark 2.20 we arrived at the following expression for the exponent
At the end of Section 2.4 we established
where the sum is over all irreducible components of the special fiber excluding . We will now provide an upper bound for . First we need to compute the sums , where is the entry of corresponding to row label and column label (see Section 3.2).
Lemma 4.26.
For , we have
Proof.
Proposition 4.27.
We have
Proof.
We split into two cases, depending on .
Theorem 4.28.
Let , , and be integers and be a prime such that . Let be the exponent of in the annihilator of . Then
Proof.
For any , we always have . Thus
By Corollary 3.21, . Also note that is an effective horizontal divisor since has no poles while is an effective vertical divisor. Since and do not have any common components, the intersection number is positive. Thus will always be negative. Continuing, we have
Using Theorem 4.25, the above expression becomes
Using Proposition 4.27, we have
which is maximized whenever is coprime to and attains a value of . ∎
By replacing with in the proof of Theorem 4.28, we obtain an upper bound for the exponent in the situation of cusp forms.
Corollary 4.29.
Let , and be integers and a prime such that . The exponent of in the annihilator of is bounded above by
Remark 4.30.
We compare our result in Corollary 4.29 with Edixhoven’s method in [Edi06]. He considers the situation of weight and level cusp forms where , and bounds via the inequality
We will show a similar inequality holds our situation. Let denote the irreducible components of where contains the cusp . Let be a non-zero cusp form such that and, without loss of generality, let be the minimum among the values in . By scaling, we can assume , , and for so that has non-negative valuation along every irreducible component i.e. . We can write the divisor associated to as
where is the horizontal part of the divisor . By Theorem 4.19, the sheaf of cusp forms is isomorphic to the relative dualizing sheaf . Since is an arithmetic surface, we can use intersection theory (see Theorem 2.11), along with Proposition 2.17, to compute:
Since has no poles, is effective so . By assumption, for ; we also have since and have no common components. Lastly, and intersect precisely at the supersingular points, so . Therefore we conclude
5 A Lower Bound
In this section, we will use Klein forms to build an explicit modular form in . By viewing these modular forms at level with coefficients in , we will obtain a lower bound for the exponent.
Definition 5.1.
A nearly holomorphic modular form is a modular form which is allowed to be meromorphic at the cusps.
Fix . Let , , and where . Define the Klein form
(25) |
The following result, which is [EKS11, Theorem. 2.6], builds upon results in [KL81, §2.1, §3.4] which establish a criterion for when a product of Klein forms is a nearly holomorphic modular form. For , we let denote the fractional part of . Note precisely when .
Theorem 5.2.
For an integer , let be a family of integers. Then the product
is a nearly holomorphic modular form for of weight if
Furthermore, for we have
Using this result when , we will choose a family of integers such that and for all . This guarantees that is a weight 2 (holomorphic) modular form of level . As , this also gives us a modular form of level as originally desired.
It may be natural to choose to be zero for most values of in order to simplify the quadratic condition in Theorem 5.2 and the expression for the order. Consider the situation when and are the only non-zero values for some distinct . By Theorem 5.2, we seek and such that
(26) |
(27) |
and
(28) |
for all .
We will choose the satisfying these conditions in three separate cases depending on the level: , , and with . It is not clear how one constructs similar modular forms of the remaining levels and . We first consider the case of level . To satisfy equation (26), we can take and . The next equation (27) becomes
which can be satisfied if we take to be a Pythagorean triple. As we now show, taking and suffices if . For convenience, we let denote for .
Proposition 5.3.
The product of Klein forms
is a weight 2 modular form of level for .
Proof.
Define by
By Theorem 5.2, the product defined by is a weight 2 nearly holomorphic modular form of level . It remains to show is holomorphic at the cusps. The order of is given by
where . We will show this expression is always non-negative.
Since , we can ignore this factor. Dividing through by 2, our goal is to show
or equivalently
is non-negative for all .
Let
noting that is the expression we are showing is non-negative. Since is periodic with period , it suffices to show for all . Since
we can conclude . Therefore it suffices to show for all .
We accomplish this by finding an explicit piecewise defined expression for and then computing its derivative. Note that
Similar expressions can be obtained for and . Putting these together, we have
Thus we can compute the derivative directly:
This shows that the pieces of are either strictly increasing, strictly decreasing, or constant in their appropriate interval. For each interval, the following table shows if is increasing, decreasing, or constant based on . The value of at the rightmost endpoint is also calculated for each interval.
Inc, Dec, Con | increasing | decreasing | constant | increasing | decreasing |
Thus we conclude for all . Consequently, is holomorphic at each cusp so is indeed a modular form. ∎
We compute the valuation of its -expansion at each cusp. The -expansion at for was given earlier in equation (5.2). In particular, the -expansion of is
For convenience, let and note . The -expansion of is
which has all integral coefficients.
To compute the -expansion of at the other cusps, we use the following identity for Klein forms, which can be found in [EKS11, Prop. 2.1]. For any , we have
We let which is a uniformizer of .
Proposition 5.4.
The -adic valuation of at its -expansions around the cusp is i.e. .
Proof.
Let ; note . The -expansion of at the cusp is given by
Observe that
Thus
Using equation (25), the -expansion of is
where . Therefore
for some . Note that is a uniformizer for for coprime to . Furthermore,
Thus the minimal valuation among the coefficients of the -expansion of is
Next we will handle the case of .
Proposition 5.5.
The product of Klein forms
is a weight 2 modular form of level . Furthermore, .
Proof.
Define by
Note that
and
By Theorem 5.2, the product defined by is a weight 2 nearly holomorphic modular form of level . It remains to show is holomorphic at the cusps. Similar to the proof in Proposition 5.3, it suffices to show
for all . Let
As before, it suffices to show for . Since we consider of the form , it suffices to show when or . We compute directly:
Consequently, is holomorphic at each cusp so is indeed a modular form.
Lastly we handle the case with .
Proposition 5.6.
Suppose and . The product of Klein forms
is a weight 2 modular form of level . Furthermore, .
Proof.
Define by
Since , the values and are all distinct. We have
noting that so . Furthermore, so is a weight 2 level nearly holomorphic modular form. It remains to show has non-negative order at each cusp.
The order of is given by
where . We will show this expression is always non-negative. Similar to the proof of Proposition 5.3, we let
Writing for some , it suffices to show for all . Note that
Since and , we have
Therefore
which is always non-negative. Thus is holomorphic at each cusp.
We compute in the same way as in Proposition 5.4. Let ; the -expansion of at the cusp is given by
Observe that
Similar to our computation in the proof of Proposition 5.4, we have
for some explicit . Let denote the normalized valuation in so that . Then we have
Note that and Hence
Continuing,
as desired. ∎
Theorem 5.7.
Let and be integers and be a prime such that and . The exponent of in the annihilator of is equal to .
Appendix A Appendix – The modular curve and modular forms
We will summarize the formulation of the regular integral model of the modular curve of full level, as presented in [KM85]. Let be a ring and let denote the category whose objects are elliptic curves where is an -scheme, and whose morphisms are Cartesian squares.
We will concern ourselves with the representability of moduli problems . The functor induces a functor defined by
where denotes the isomorphism class of the pair . If is representable by , then is representable by . Indeed, from the bijection
we can associate to a morphism
in . We define the map by sending to the map . This correspondence is bijective and functorial in by the properties of .
Definition A.1.
Let be two moduli problems for elliptic curves. A morphism between moduli problems (over ) is a natural transformation .
Let be a morphism between two representable moduli problems over . We have an induced natural transformation . The map induces a map which on -points coincides with the map in the following diagram:
Definition A.2.
Let and be two moduli problems for elliptic curves over . We define the product, or simultaneous, moduli problem by
Suppose is representable by an elliptic curve and is relatively representable. According to [KM85, 4.3.4], is representable and so we naturally have a map . Following the notation in [KM85], we will usually denote by .
Definition A.3.
Let be an elliptic curve over a scheme . A section corresponds to a morphism333In general, a section of a morphism will be a locally closed embedding . When is separated, as in the case of an elliptic curve, the map becomes a closed immersion. The associated divisor is in general defined to be the closure of the scheme-theoretic image of in . whose composition with the structural morphism is the identity on . Since is separated, is a closed immersion. We denote by the scheme-theoretic image of in which is an effective Cartier divisor on . Consider the multiplication-by- isogeny whose kernel we denote by , an -group scheme.
A -structure on is a group homomorphism such that we have an equality of effective Cartier divisors
In this case, we call and the corresponding Drinfeld basis of .
We define to be the moduli problem which assigns to each elliptic curve the set of all -structures on . Let be a morphism in the category . Then
is defined by sending a Drinfeld basis of to . Indeed, is a Drinfeld basis of (see the paragraph proceeding [KM85, 1.4.1.2]).
We moreover consider the full level moduli problem over the cyclotomic integers , where is the th cyclotomic polynomial, by constructing a new moduli problem from as follows (see also [KM85, 9.4.3.1]). For a -algebra , let denote the image of mod under the map , following the convention in [KM85, 9.1.5]. Section 2.8 of [KM85] associates to the isogeny a bilinear pairing Define the moduli problem which assigns to each elliptic curve the set of all Drinfeld bases such that .
Theorem A.4.
Suppose . The moduli problem is represented by a regular scheme which is flat over of dimension 2. Moreover, is smooth over .
Proof.
The moduli problem is relatively representable and finite, flat over by [KM85, 5.1.1]. Furthermore, is rigid whenever by [KM85, 2.7.2]. Therefore, by Proposition [KM85, 4.7.0] and [KM85, 5.1.1], is representable by a regular scheme flat over of dimension 2. By [KM85, 9.1.8, 9.1.9], the same things hold true for the associated canonical moduli problem . Using [KM85, §5.1.1] again, is étale over , which, along with [KM85, 4.7.1], implies smoothness. ∎
Remark A.5.
Theorem A.4 also holds for the moduli problem over , for any ring extension . For any moduli problem and ring , let denote the moduli problem obtained by composing with the forgetful functor . According to [KM85, 4.13], if is relatively representable, then is also relatively representable by the same morphism for any -scheme . Furthermore, if is representable by , then is representable by the base change . In other words, we have .
By a process called "normalizing near infinity", as described in [KM85, §8.6], extends to a scheme which is proper over .
Definition A.6.
Let denote the closed subscheme of endowed with the reduced scheme structure, called the cuspidal locus of . If the modular curve is clear from context, we denote .
The following definition captures many of the desirable properties of .
Definition A.7.
Let be a Dedekind domain. We call a regular, integral, projective, flat -scheme of dimension 2 an arithmetic surface.
Theorem A.8.
Let . The scheme is an arithmetic surface over . Moreover, is geometrically connected with reduced closed fibers and smooth over .
Proof.
We have already established that is regular and flat over of dimension 2. By [KM85, 10.9.1(2)], there exists an open neighborhood of which is smooth over . In particular, is regular and flat over so is regular and flat over . Since is connected and regular, is integral.
Let and let be a noetherian, regular, excellent -algebra. We will construct an invertible sheaf on whose global sections will be defined as the space of modular forms. Let denote the universal elliptic curve. We define to be the pushforward of the sheaf of Kahler differentials on . If is the identity section of our elliptic curve, then we have . Hence , being the pullback of an invertible sheaf, is an invertible sheaf on . According to [KM85, 10.13.2], there is a canonical way to extend to an invertible sheaf on , which we denote by or simply if the modular curve is clear from context.
Proposition A.9.
The invertible sheaf canonically extends to an invertible sheaf on .
Definition A.10.
We call the invertible sheaf on the modular sheaf (of weight 2k). The global sections are known as modular forms of weight and level . The global sections are known as cusp forms of weight 2k and level .
The formation of the modular sheaf behaves well under base change. The following proposition is [KM85, 10.13.6].
Proposition A.11.
Let be an extension of noetherian, regular, excellent rings and let
denote the induced base change map. Then we have an isomorphism of invertible sheaves
Let be an elliptic curve over a smooth -scheme . According to [KM85, 10.13.10] (see also [Kat73, A1.4]), we have map of -modules, known as the Kodaira-Spencer map, which becomes an isomorphism precisely when represents a moduli problem which is étale over . In particular, if is smooth over (e.g. if is a unit in ), we get an isomorphism .
Theorem A.12.
Let be a noetherian, regular, excellent -algebra containing . The Kodaira-Spencer isomorphism on extends to an isomorphism on
Proof.
Suppose are two representable moduli problems that are finite over and normal near infinity. Any morphism of moduli problems is compatible with the usual morphism . Thus the induced map fits in the commutative diagram
Hence extends to a map on the normalizations near infinity.
Proposition A.13.
Let be an excellent, noetherian, regular ring and let and be two representable moduli problems, both finite over and both normal near infinity. Let be a morphism of moduli problems over . Under the induced map we have
Proof.
This is [KM85, 10.13.5(2)]. ∎
A.1 Cusps and the Special Fiber
In this section we will investigate the cusps and special fiber of and recall the theory of -expansions for modular forms. First we establish that the formation of the cuspidal locus and the formation of behaves well under base change. The following is [KM85, 8.6.6, 8.6.7].
Proposition A.14.
Let and be excellent, noetherian, regular -algebras. For any extension of scalars , we have
For two groups and , let denote the set of surjective group homomorphisms from to . Any subgroup acts on the set
via for any . Let
where is the identity matrix in .
According to [KM85, 10.9.1], the cusps of is a disjoint union of many sections of . Furthermore, the formal completion of along is the disjoint union of copies of the formal spectrum for some integer dividing , dependent on the index . Refer to [KM85, 10.2.5] which provides a canonical bijection between a cuspidal section and the corresponding label in .
Proposition A.15.
Let be a Dedekind domain with fraction field . As Weil divisors, we have
where the sum is indexed over .
Proof.
Now we recall how the -expansion map
is defined for a modular form over a -algebra at a specified cusp. Let be a cusp of , which corresponds to the image of a section and is a connected component of where is a cusp of the generic fiber. According to [DR73, §VII, 2.3.2], the cusp corresponds to a Tate curve over (see [DR73, §VII, Définition 1.16]) and a closed immersion fitting in the diagram
By [DR73, §VII, Corollaire 2.4] (see also [KM85, 10.9.1]), this closed immersion induces an isomorphism of formal schemes
where denotes the formal completion of along the cusp . According to [KM85, 8.8(T.2)] (see also [DR73, §VII, 1.16.1]) we have an isomorphism
between formal Lie groups. Let , where is the standard invariant differential on . Pulling back by allows us to identify with .
Definition A.16.
Let and let be a cusp of . Viewing in the completed stalk , we write for some . We call the -expansion of at . This gives a map
More generally, for a -algebra , we obtain the -expansion map
by tensoring with . We will always use as a local generator for to obtain the -expansion at .
The following result, which is [DR73, §VII, Théoréme 3.9] and known as the -expansion principle, essentially says -expansions can detect the “ring of definition”.
Proposition A.17.
Let be a -algebra, let be a subalgebra of , and let . If the -expansion of at every cusp of lies in , then is a modular form over i.e. .
Let and be integers and let be a prime such that . We will describe the special fiber of . For simplicity, we consider the modular curve over a -algebra which is a DVR of mixed characteristic in which is invertible and with fraction field and perfect residue field . Let denote the special fiber of .
We will recall the theory of Igusa curves, as developed in Section 12 of [KM85], to describe the irreducible components of . For , let denote the map induced by the th power of the Frobenius automorphism on . For any -scheme , we let denote the pullback under from which we also obtain a map called the absolute Frobenius. More generally, for any scheme , we let denote the pullback under from which we obtain a morphism of -schemes, called the th-fold relative Frobenius.
Let be a moduli problem on . We define the moduli problem on by extending scalars via so that
If is representable and finite over and normal near infinity, then the same holds for and we have
Definition A.18.
Let be an elliptic curve over an -scheme . An Igusa structure of level on is a point which generates the kernel of Verschiebung in the sense of [KM85, 1.4.1]. Let denote the moduli problem which assigns to each elliptic curve the set of all Igusa structures of level on .
According to [KM85, 12.7.1], if is a representable moduli problem finite over which is normal near infinity, the simultaneous moduli problem is representable over and normal near infinity. The following result is [KM85, 12.7.2] applied to the simultaneous moduli problem over for any . We denote
Proposition A.19.
-
a.
is a proper smooth curve over .
-
b.
The usual projection is finite and étale outside the supersingular points of for all .
Next we define another moduli problem closely related to Igusa structures, which will directly appear in the description of .
Definition A.20.
Let be an elliptic curve over an -scheme and fix . An exotic Igusa structure of level on is a point such that is a Drinfeld -basis of along with a point such that . Let
denote the moduli problem which assigns to each elliptic curve the set of all exotic Igusa structures of level on .
The following result, which is [KM85, 12.10.6] in the situation , relates Igusa structures with exotic Igusa structures and also establishes the representability of . We denote
Theorem A.21.
For , we have a canonical isomorphism sitting in the commutative diagram
where is the th-fold relative Frobenius .
We introduce the following definition from [KM85, §13.1] which will allow us to describe the special fiber of . Consider the general situation with a field, a smooth scheme over , and a finite flat morphism of schemes. Suppose there exists a nonempty finite set of -rational points of such that for each there exists a unique closed -rational point over such that for some . The points of are referred to as the supersingular points; indeed in the situation and are modular curves, will be taken to be the supersingular points of which correspond to supersingular elliptic curves. Furthermore, we assume there is a finite collection of -schemes with a morphism such that
-
•
for each and there exists a unique closed, -rational point over .
-
•
is finite flat over and is smooth over .
-
•
is a closed immersion and is an isomorphism over the complement of in .
By [KM85, 13.1.3], known as the “Crossings Theorem”, if is connected, then the are the irreducible components of . Furthermore, if each is reduced, then is also reduced.
Definition A.22.
In the situation just discussed, we say is the disjoint union of the ’s with crossings at the supersingular points.
Remark A.23.
Following [KM85, 13.1.7], we can relax the -rationality of and and instead require -rationality after extending scalars to a separable closure of . In this situation, we still say is the disjoint union of the ’s with crossings at the supersingular points.
Next we describe . We identify as the subgroup
of . According to [KM85, 13.7.1], the moduli problem on assigns to each elliptic curve the set of all Drinfeld -bases with . Let denote the homomorphism of -schemes corresponding to . Consider the diagram
where . By [KM85, 13.7.2(3)], a choice of -basis of defines an isomorphism , allowing us to view as a surjective homomorphism .
Definition A.24.
The component label of is the class of in
By [KM85, 13.7.4, 13.7.5], this establishes a canonical bijection between the irreducible components of the special fiber and the set of component labels. The following is [KM85, 13.7.6].
Theorem A.25.
The special fiber of is the disjoint union, with crossings at the supersingular points of , of the exotic Igusa curves over indexed by
Furthermore, is reduced.
Note that the claim is reduced comes from the fact that is reduced, together with the “Crossings Theorem” of [KM85, 13.1.3] and [KM85, 13.1.4].
Recall the closed subscheme of cusps of is a disjoint union of many sections of which, by Proposition A.15, can be viewed as the closure of points . By Corollary 2.15, intersects precisely one irreducible component of . The following, which is [KM85, 13.9.3], tells us precisely the component label of the irreducible component which intersects with, given the index of in (see the paragraph preceding Proposition A.15).
Theorem A.26.
The natural projection
assigns to a component of indexed by the irreducible component of that intersects with.
Corollary A.27.
Each irreducible component of intersects with the same number of cuspidal components of .
Every group homomorphism is uniquely determined by its image on the basis vectors and . Let denote the map defined by
A complete list of representatives in is given by
With this labeling, the following is clear:
Corollary A.28.
The special fiber has many irreducible components.
We will record the number of cuspidal components of an irreducible component of intersects with. For convenience, we let , the set of all cusps of the generic fiber. The following is [Miy89, 4.2.10].
Lemma A.29.
For , we have
We will also need to know how to compute . The following is from [Miy89, 4.2.3, 4.2.4].
Lemma A.30.
-
a.
Let and be coprime integers. Then
-
b.
Let be prime and an integer. Then
Proposition A.31.
Each irreducible component of intersects precisely
many cuspidal components of .
Proof.
By Corollary A.27, this quantity is independent of irreducible component . Hence the number of cuspidal components intersects is equal to the total number of cusps of the generic fiber divided by the total number of irreducible components. By Corollary A.28, the number of irreducible components is .
Appendix B Appendix – Circulant Matrices
Definition B.1.
An circulant matrix over a field is any matrix of the form
where each column is equal to the previous column shifted downward by 1, looping around as appropriate.
The entries of an circulant matrix can be characterized by the equation
for all where the index is taken modulo among the residue classes in . Let denote a primitive th root of unity. The following provides an explicit description of the eigenvalues and corresponding eigenvectors of a circulant matrix.
Lemma B.2.
The eigenvalues of a circulant matrix are precisely
for . A corresponding eigenvector of is given by
Proof.
We will verify that for all . For , the entry of is
while the entry of is
which agrees with the entry of for all . ∎
Circulant matrices are always diagonalizable (see [KW01, §2]) so we can write where the columns of are the eigenvectors and is a diagonal matrix whose diagonal entries are the corresponding eigenvalues . Explicitly,
We see the -entry of is equal to for .
Lemma B.3.
The matrix is unitary i.e. the inverse of is equal to the conjugate transpose .
Proof.
Note that is symmetric and the conjugate of is . The -entry of is equal to
Hence so . ∎
Now that we know , we can describe the entry of , if it exists.
Proposition B.4.
Let be an invertible circulant matrix whose entries along the first column are in that order. Then the -entry of is equal to
where are the eigenvalues of for .
Proof.
We compute out the product . The matrix will be of the form
Thus the -entry of is equal to
and so the -entry of is equal to
Using our explicit description of the entries of , we can establish the following.
Corollary B.5.
The inverse of an invertible circulant matrix is circulant.
Proof.
Let be an circulant matrix whose entries along the first column are ordered so that for . By Lemma B.4, the -entry of is equal to
To show is circulant, we will establish the relationship where the index is taken modulo in the residue class . In the above expression for , the index only appears in the term . Observe that
Hence so is circulant. ∎
Appendix C Appendix – Inverse via the Woodbury Matrix Identity
Let be an invertible matrix, an invertible matrix where , an matrix, and a matrix. The Woodbury matrix identity states
A proof of this identity can be found in [HS81, §1.3]. This identity allows us to compute the inverse of provided we can easily compute the inverses of and .
In general, we can use the Woodbury Matrix Identity to compute the inverse of where is an invertible matrix and is an matrix of rank , provided is invertible. Let be the matrix whose columns are the linearly independent columns of . Let denote the vector in the th column of which can be expressed as
We define the th column of the matrix to consist of entries in that order. Consequently, where is the identity matrix.
Using the Woodbury Matrix Identity, we will provide a formula for the inverse in the following situation as encountered in Section 3.3. Let be a block diagonal matrix with two invertible blocks of sizes and . The inverse is also block diagonal with blocks of the same size. We write
We also consider the case
where the subscripts on the and entries are there to help keep track of their position.
Proposition C.1.
Let where and are the given matrices above. Let denote the -entry of . We have
where is the sum of all entries in the first block in and is the sum of all entries in the second block in .
Proof.
By the Woodbury Matrix Identity,
We first compute
Hence
Next we will compute
Lastly, we compute
We now describe the -entry of the above product in four different cases.
Case 1(): The -entry is equal to
Case 2 ( and ): The -entry is equal to
Case 3 ( and ): The -entry is equal to
Case 4 (): The -entry is equal to
Subtracting these quantities from corresponding -entry of gives our desired result. ∎
Acknowledgements
Most of the results in this paper originated from my PhD thesis. I would like to thank Bryden Cais for his guidance, support, and incredibly helpful feedback. I would also like to thank Brandon Levin, Doug Ulmer, and Hang Xue for their helpful comments. This research did not receive any specific grant from funding agencies in the public, commercial, or not-for-profit sectors.
References
- [BDP17] Massimo Bertolini, Henri Darmon and Kartik Prasanna “-adic -functions and the coniveau filtration on Chow groups” With an appendix by Brian Conrad In J. Reine Angew. Math. 731, 2017, pp. 21–86 DOI: 10.1515/crelle-2014-0150
- [Čes17] Kęstutis Česnavičius “A modular description of ” In Algebra Number Theory 11.9, 2017, pp. 2001–2089 DOI: 10.2140/ant.2017.11.2001
- [ČNS23] Kęstutis Česnavičius, Michael Neururer and Abhishek Saha “The Manin constant and the modular degree” In J. Eur. Math. Soc., 2023 DOI: DOI 10.4171/JEMS/1367
- [Con07] Brian Conrad “Arithmetic moduli of generalized elliptic curves” In J. Inst. Math. Jussieu 6.2, 2007, pp. 209–278 DOI: 10.1017/S1474748006000089
- [DR73] P. Deligne and M. Rapoport “Les schémas de modules de courbes elliptiques” In Modular functions of one variable, II (Proc. Internat. Summer School, Univ. Antwerp, Antwerp, 1972), 1973, pp. 143–316. Lecture Notes in Math.\bibrangessepVol. 349
- [Edi06] Bas Edixhoven “Comparison of integral structures on spaces of modular forms of weight two, and computation of spaces of forms mod 2 of weight one” With appendix A (in French) by Jean-François Mestre and appendix B by Gabor Wiese In J. Inst. Math. Jussieu 5.1, 2006, pp. 1–34 DOI: 10.1017/S1474748005000113
- [EKS11] Ick Sun Eum, Ja Kyung Koo and Dong Hwa Shin “A modularity criterion for Klein forms, with an application to modular forms of level 13” In J. Math. Anal. Appl. 375.1, 2011, pp. 28–41 DOI: 10.1016/j.jmaa.2010.08.035
- [HS81] H.. Henderson and S.. Searle “On deriving the inverse of a sum of matrices” In SIAM Rev. 23.1, 1981, pp. 53–60 DOI: 10.1137/1023004
- [JCP16] Estela De Lourdes Juárez-Ruiz, R. Cortés-Maldonado and Felipe Perez-Rodriguez “Relationship between the Inverses of a Matrix and a Submatrix” In Computacion y Sistemas 20, 2016, pp. 251–262 DOI: 10.13053/CyS-20-2-2083
- [Kat04] Kazuya Kato “-adic Hodge theory and values of zeta functions of modular forms” Cohomologies -adiques et applications arithmétiques. III In Astérisque, 2004, pp. ix\bibrangessep117–290
- [Kat73] Nicholas M. Katz “-adic properties of modular schemes and modular forms” In Modular functions of one variable, III (Proc. Internat. Summer School, Univ. Antwerp, Antwerp, 1972), 1973, pp. 69–190. Lecture Notes in Mathematics\bibrangessepVol. 350
- [KL81] Daniel S. Kubert and Serge Lang “Modular units” 244, Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences] Springer-Verlag, New York-Berlin, 1981, pp. xiii+358
- [KM85] Nicholas M. Katz and Barry Mazur “Arithmetic moduli of elliptic curves” 108, Annals of Mathematics Studies Princeton University Press, Princeton, NJ, 1985, pp. xiv+514 DOI: 10.1515/9781400881710
- [KW01] Dan Kalman and James E. White “Polynomial equations and circulant matrices” In Amer. Math. Monthly 108.9, 2001, pp. 821–840 DOI: 10.2307/2695555
- [Lan88] Serge Lang “Introduction to Arakelov theory” Springer-Verlag, New York, 1988, pp. x+187 DOI: 10.1007/978-1-4612-1031-3
- [Liu02] Qing Liu “Algebraic geometry and arithmetic curves” Translated from the French by Reinie Erné, Oxford Science Publications 6, Oxford Graduate Texts in Mathematics Oxford University Press, Oxford, 2002, pp. xvi+576
- [Miy89] Toshitsune Miyake “Modular forms” Translated from the Japanese by Yoshitaka Maeda Springer-Verlag, Berlin, 1989, pp. x+335 DOI: 10.1007/3-540-29593-3
- [Ser79] Jean-Pierre Serre “Local fields” Translated from the French by Marvin Jay Greenberg 67, Graduate Texts in Mathematics Springer-Verlag, New York-Berlin, 1979, pp. viii+241
- [Shi94] Goro Shimura “Introduction to the arithmetic theory of automorphic functions” Reprint of the 1971 original, Kanô Memorial Lectures, 1 11, Publications of the Mathematical Society of Japan Princeton University Press, Princeton, NJ, 1994, pp. xiv+271
- [Sta21] Stacks Project authors “The Stacks Project”, 2021 URL: https://stacks.math.columbia.edu/