Computation of Hecke eigenvalues (mod ) via quaternions
Abstract.
In a 1987 letter [SL96], Serre proves that the systems of Hecke eigenvalues arising from mod modular forms (of fixed level coprime to , and any weight ) are the same as those arising from functions , where is some double quotient of and is the unique quaternion algebra over ramified at . We present an algorithm which then computes these Hecke eigenvalues on the quaternion side in a combinatorial manner.
1. Introduction
The study of Hecke eigenvalues originated from Ramanujan’s -function
a weight 12 and level 1 Hecke eigenform about which Ramanujan made a number of conjectures that have motivated much of the theory of modular forms over the past century. More recently, it is known [Del69] that a normalised Hecke eigenform in gives rise to a mod Galois representation . This representation is unramified away from and is characterised by the following trace and determinant of Frobenius data (mapped appropriately into ):
These representations play a central role in modern number theory, most notably in the conjectures of Serre. As such, the mod Hecke eigenvalues are objects of great interest, and so one may wish to enumerate these systems of Hecke eigenvalues as a source of examples.
At present, there already exist algorithms for computing these Hecke eigenvalues, for example using modular symbols in connection with the Eichler-Shimura theorem [Wie06]. Our approach will instead make use of the following theorem of Serre (where we have restricted his result to a fixed level ) [SL96], that tells us that one could instead compute these Hecke eigenvalues by working with a particular quaternion algebra. The computation of the Hecke eigenvalues then becomes a combinatorial one, which is perhaps a more elementary approach.
Theorem 1.1.
Let be the unique quaternion algebra over ramified at . The systems of Hecke eigenvalues () (with , and fixed coprime to ) coming from the modular forms (mod ) of level , are the same as those coming from the functions
We remark that in this theorem, we do not necessarily realise all the weight systems of eigenvalues from the modular form side as weight systems of eigenvalues on the quaternion side, only those arising from weight eigenforms not divisible by the Hasse invariant. They will still appear on the quaternion side when ranging over all weights.
Notation 1.2.
We need to explain some of this notation:
-
•
Let be any maximal order of .
-
•
Let and a maximal order, for any prime . So for .
-
•
Let denote the finite part of the adelic points of , in other words the restricted product with respect to the subgroups .
-
•
Let be a uniformiser of .
-
•
Let be the kernel of reduction mod , .
-
•
For , let be the subgroup of consisting of elements congruent to 1 mod , where is the highest power of dividing .
-
•
Let , an open subgroup of .
-
•
For , we will denote by the image in .
-
•
By a weight function we mean a function which satisfies , where acts on by multiplication in the -place.
Note that elements of correspond to isomorphism classes of invertible left -ideals with -structure, meaning a basis for as an -module. Explicitly, an adelic point corresponds to an ideal with for all . The -structure is then given by the reduction modulo of any element satisfying the congruences
for all (this congruence is vacuous for any , so such exists by the Chinese remainder theorem). We then quotient by and exactly to get the desired bijection.
For a prime , the Hecke operator on this space of functions is given by
for . Here means we pick a representative of , multiply this in the -place by the matrix (under an identification of and hence of ), and then take the image in . Each individual is not well defined in , but is well defined, provided we pick the same representative for each multiplication. This Hecke module structure has been studied before, as in the likes of [Koh01].
The algorithm we present computes a matrix for the Hecke operator on the space of (weight ) functions , where and are pairwise coprime. Note that we allow the cases as the definitions on the quaternion side still make sense here. We begin by computing an illustrative example for the case , and then generalise this to weight and level for any . Note that in this case the matrices we compute are exactly Brandt matrices. This is then extended to higher weight and level, essentially by keeping track of the -structure.
We remark that a similar computation has been performed by Pizer in [Piz80]. This has since been applied in, for example, [SW05] and [CS01]. Pizer was interested in computing the subspace of cusp forms on generated by theta series, and the Hecke operators on this subspace. The algorithm involves computing certain Brandt matrix series. The main difference between our approach and Pizer’s is that we incorporate level structure through coset representatives of . See also [Dem05] and [Dem07] for a similar approach in the context of Hilbert modular forms (the corresponding description of Hilbert modular forms on the quaternion side is more involved than for modular forms). Pizer instead works with orders of level , which allows them to work purely with the quaternion algebras, without having to write down explicit isomorphisms of the form , something that we must do (see Section 3). On the other hand, the use of matrices perhaps makes the contribution from the level structure more visible. The tradeoff is then between computing these isomorphisms with the Hensel-like argument involved, and computing orders of level . A related point of difference is that in Pizer’s argument one must compute the left ideal classes of an order of level , whereas we only need to compute these for our maximal order , because our level structure is already captured in , which is very concrete. Because a maximal order has minimal discriminant, the corresponding Minkowski bound is lower, and so the computation of these ideal classes in our case may be (somewhat) faster.
The author would like to thank his M.Phil. supervisors, Alex Ghitza and Chenyan Wu, for their careful reading of the script, and their valuable corrections and suggestions. Writing of this paper was partially supported by the University of Melbourne Robert George Williams Scholarship, as well as the Australian Government Research Training Program Scholarship.
2. An Example
Example 2.1.
Take to be 11. The quaternion algebra is then . In this example we will work with level as doing so greatly reduces the size of . We will then compute the Hecke eigenvalues on the space of functions on , where is not quite as we replace with . Note that this is the same as computing the Hecke eigenvalues on of weight . Indeed, the modularity condition tells us that these are functions on that are invariant under the action of , and so can be identified with functions on . The computation in this section is based on notes of Buzzard [Buz].
Firstly, let’s understand the set . By the same argument that corresponds to (isomorphism classes of) invertible left -ideals with -structure, we see that corresponds just to the invertible left -ideals . Running the following in MAGMA [BCP97], we compute a maximal order of , and then find that its left ideal class set has order 2, whose elements we call and . So has two elements. We can also compute integer bases for , and .
>>ΨD := QuaternionAlgebra< RationalField() | -1, -11>; >>ΨO := MaximalOrder(D); >>ΨBasis(O); >>ΨClasses := LeftIdealClasses(O); >>Ψ#Classes; >>ΨI1 := Classes[1]; >>ΨBasis(I1); >>ΨI2 := Classes[2]; >>ΨBasis(I2);
which outputs:
[ 1, i, 1/2*i + 1/2*k, 1/2 + 1/2*j ] 2 [ 1, -i, -1/2*i - 1/2*k, 1/2 - 1/2*j ] [ 2, -2*i, 1 - 3/2*i - 1/2*k, 1/2 - i - 1/2*j ]
We now want to rewrite and as adelic points in . For any invertible left -ideal , is a principal -ideal, generated by any nonzero element whose reduced norm has minimal -adic valuation. To see this, we refer to Corollary 16.6.12 in Voight [Voi21] that any invertible semi-order (lattice that contains 1 and has reduced norm equal to the ring ) is an order, so that is an invertible semi-order, which must then by . So if we are given a -basis for , then is generated by a basis element whose reduced norm has minimal -adic valuation. Computing the reduced norms of the given basis elements of and we get and respectively. So we see that (which was obvious anyway) and for all . Moreover, as 6 has minimal 2-adic valuation. Thus corresponds to and corresponds to (we use superscripts here to avoid overloading the subscripts, which we want to use for the places). An obvious choice of basis for the vector space of valued functions on are the characteristic functions and .
To compute the Hecke operator , we will need to work with matrices at the -place. Let’s compute and with respect to the basis . We will need isomorphisms and . In general, we have an isomorphism
for such that , when . For , note that has -basis , by using the basis for computed above and noticing that 2 is invertible in . So for any with , the above isomorphism restricts to an injection , which must in fact be an isomorphism by maximality. We could take for a root of in , choosing for example the root which is by Hensel’s lemma.
For we need to be a little more careful because we need to all map to elements in , rather than just . If we take for a root of in (taking for example the root congruent to by Hensel’s lemma), then we see that maps to , which is in . This gives us our desired isomorphism . Note that if we took instead , we would then map to , which is not in .
We now begin our computation of , starting with the value of . By definition,
so we reduce to checking, for example, whether is the same as or as an element of .
The condition that in is equivalent to saying that the -ideal is principal, where is defined by and for all . If this was the case, then we see that must be generated by an element of (since for all ) of reduced norm . We have previously computed a basis for . Writing , for , we compute
For this to be equal to 2, we must have and . So must be . This satisfies for , so we only need to check whether this works at 2. In other words, we need to check if
or
where our isomorphism sends to , and to . Equivalently, we need to check if
We see that neither is true. So we have computed one of the terms in , namely
As a sanity check, we verify that , so that
as we would expect. The isomorphism sends to . Let be the -ideal generated by the adelic point , and the ideal generated by . Then the condition that they correspond to the same element of is equivalent to the existence of some such that . Checking this locally, this means for , and at we have
(1) |
We can rewrite this as
so we see that , by inverting the matrix and clearing the denominator. It follows that if exists, we must have , and , by checking locally. There are finitely many possibilities for , namely . We need to check if any of these satisfies (1) when we replace with the corresponding matrix in . A computation shows that one can take , where we need that actually for our choice of square root.
Returning to our computation of , we need to compute the remaining terms and . By the same argument as above, it suffices to check whether
for the first term, and whether
for the second. The first two are true and the last two are not. So we deduce that
Therefore
We can also deduce that .
One similarly shows that and so . The required computation is to show that the -ideals generated by , for and , are principal -ideals. The generators can be taken to be and respectively.
It follows that and , so has matrix with respect to the basis . The eigenvalues are .
The Hecke operator can be computed in a similar way. We give one example evaluating the summand
of . One could do this by checking whether the -ideal generated by the adelic point in the square brackets is principal. If it is, then this value of is . This only works because has two elements. In general, we need to check if is in the same left ideal class as the -ideal generated by the adelic point . So we need to check if there exists such that . Checking this locally, we need (for the two isomorphisms and specified)
(2) |
and
(3) |
and for . The first two conditions imply that and . So , and . There are finitely many possibilities for .
Before we go through all the possibilities for and check whether they satisfy the two equations, we remark that for equation (2) we only really need to work with and the matrix modulo some power of 2 (and for equation (3) we only need modulo some power of 3). Indeed, using the fact that our choice of is congruent to , we have
Suppose has matrix . Then equation (2) tells us (after inverting the matrix) that we need
Note that the norm condition on will then guarantee that we land in not just . Working modulo 4, we need
Expanding this out, this is equivalent to
(4) |
Similarly, if has matrix , then equation (3) is equivalent to
(5) |
So now we look through of reduced norm 12. Write
for . We also compute the images of the basis elements in and , then reduce them modulo 4 and 3 respectively:
It follows that
Hence equations (4) and (5) tell us that there exists of reduced norm 12 satisfying equations (2) and (3), if and only if we can find such that
where the last equation tells us the reduced norm is 12. We can check that there are no such solutions. For example, we see that must be even, and then from the norm condition we see that . Trying , the norm condition tells us that we must have or , none of which satisfy the congruence conditions. For we similarly must have or which also does not work. And finally for we need . Since , in one case we have , and then in there are no solutions with and . Otherwise we have and with , and , which has no solutions, checking .
We deduce that
Performing the remaining calculations, using the relevant mod 3 and 4 matrices we have already written down above, we find that has matrix with respect to the basis . We compute the eigenvalues to be .
Comparing the matrices for and , they are simultaneously diagonalisable (as expected) with eigenvectors and . The corresponding eigenvalues are and , viewing as elements of . In view of Serre’s Theorem 1.1, one might wonder what mod 11 modular forms give rise to these eigenvalues. The first comes from , which has minimal weight filtration 120 (meaning that 120 is the lowest weight at which a mod 11 modular form has this -expansion), and the second from the Hasse invariant .
3. Explicit isomorphisms
We now work with any prime . Let be the unique quaternion algebra over ramified exactly at . It turns out that one can take and when respectively, (or ), and . In the remaining case , we can take for any prime with . This can be verified by computations with the Hilbert symbol as in [Voi21] Chapter 12. We will write for an appropriate .
In the example of the previous section, we wrote down explicit isomorphisms . In general this is tricky to do, complicated by the fact that might properly contain the order with -basis - we can have nontrivial denominators. See also [Voi13] for an algorithm for computing such isomorphisms, which arose as a byproduct of other more involved algorithms. Let have -basis . What we need to do is find matrices (corresponding to ) such that
and such that the matrices corresponding to , which a priori have entries in , actually have entries in . This latter condition just imposes some congruences on the entries of and . For example, in the case using the given basis for , we also require that
which only imposes congruence conditions when . If we can do this in the general case, then this gives us an isomorphism , which restricts to an injection . But is a maximal order in , so this injection must actually be our desired isomorphism.
The main observation is that, for our purposes, we really only need the matrices corresponding to the modulo some sufficiently large power of . Consequently, we only need to compute and modulo some . This was hinted at in the computation of in the example above. We will describe a formula for a sufficiently large later. So we need to search for and in for some , which satisfy
and the congruence conditions imposed by the basis of (we take sufficiently large so that these congruences can be viewed as congruences modulo ). We know that such and exist because we know that there exists an isomorphism . Since have reduced trace 0, we can furthermore assume that and are trace-free. Then such a solution can be found by a finite enumeration of all matrices in (a more efficient method would be to use the lemma below repeatedly). The claim then is that these solutions can be lifted to our desired and (although we do not need to write down and , just know that and lift). Note that and will automatically satisfy the congruence conditions from the basis of because we are lifting from and . This is the content of the following lemma (for ).
Lemma 3.1.
Suppose we have trace-free matrices for an odd prime and , and suppose they satisfy
Then we can find trace-free matrices satisfying
with and . It follows that we can lift to desired matrices by induction.
Proof.
Our proof involves writing out all the matrix entries and multiplying them together. The requirement that the matrices be trace-free simplify our calculations. Let
Lifting these to (trace-free) matrices modulo , we will write
for trace-free matrices
A useful calculation is the following: if and are trace-free matrices, then
It follows that if we lift and to any trace-free matrices modulo , also denoted and , we have that
where the are all scalar matrices modulo . The conditions on and coming from those on and are the following:
for some scalars . We can rewrite this as
for some arbitrary vector . This is always possible if the vectors and are linearly independent. Since , linear dependence is equivalent to the existence of a nonzero scalar such that , or that either or is 0 mod . The latter two cases are not possible because otherwise or . Since , this means that divides or , which does not happen (we had to use in case and ). And in the first case, tells us that , which is false since . Hence we can always lift to . ∎
3.1. The case
When , the argument of Lemma 3.1 does not work, analogous to the difficulty of using a naive Hensel’s lemma for finding square roots in . We will make use of a generalised Hensel’s lemma for multiple variables, which can be found as Theorem 3.3 in Conrad’s notes [Con], specialised to the case of with its usual absolute value .
Theorem 3.2.
Let and satisfy
where is the Jacobian of - the determinant of its derivative matrix - and the norm of a vector is defined to be the maximum of the absolute values of its entries. Then there is a unique such that and .
To see how to apply this, we are looking for matrices satisfying
which also satisfy certain congruence conditions mod 4 (the highest power of 2 possibly dividing denominators in - see Proposition 3.7 to follow). Because we know , we know that such and exist, so we can find such a solution modulo some power of 2 bigger than 4 and try to lift to . If and with entries in , then the conditions
are equivalent to
(6) |
Lemma 3.3.
Suppose we have satisfying
(7) |
Then these are congruent to some modulo 128 respectively, satisfying equation (6).
Proof.
We have 3 equations and 6 variables, so to apply Theorem 3.2 we need to fix 3 variables. For example, if we fixed and considered the polynomials
then we know that . Note . The derivative matrix is
In order to lift to a solution , we need to not divide
Whether this holds depends on the we were given. If this does not hold for our particular , then we could instead fix some of the other variables, replacing with some , and check whether 16 divides the new value of . This means that the proof reduces to an analysis of several cases depending on the parities of the .
Notation 3.4.
Let denote the length 3 vector of polynomials in 3 variables obtained by fixing the variables other than . So for example we considered the case above.
We compute the following Jacobians:
(8) |
(9) |
(10) |
(11) |
(12) |
Now we split into several cases. Firstly, assume are odd. We will frequently use from equation (7) that is divisible by mod 128, and in particular is even.
- (i)
-
(ii)
odd and even. We see that are odd and is even. Then use (10) and the fact that .
-
(iii)
even and odd. We see that are odd and is even. Then use (8) and the fact that
. -
(iv)
both odd. Then . From the way we defined , we have that (when neither is equal to 2). By symmetry, suppose . Then we see that we must have . By symmetry, suppose is odd and is even but not divisible by 4. Then since is even, we must have that is even. If we then have even, use (11), and . If instead is odd, then because we know . But and are odd, so and therefore . It follows that is divisible by 4. Then we use (10), where is divisible by 8 but not by 16. The other cases are symmetric.
Finally we consider the case when either or is 2 (we cannot have both). Firstly, consider , so .
-
(i)
both even. We see that are odd. Since and is even, exactly one of is even. But then is a contradiction.
-
(ii)
odd and even. We see that are all odd. We also have from that , and therefore . Then use (12) and the fact that .
- (iii)
- (iv)
The only property of that we use is that . If we took , then is also 1 mod 4. So this case is symmetric to the above. ∎
The upshot of all this work is the following:
Corollary 3.5.
Let be any prime, and let be any integer at least 2, where we also ask that . Suppose we can find trace-free matrices satisfying
Then these lift to matrices with and such that
So suppose we can find for , or (for example by exhaustion of finitely many matrices in , or working through the Hensel argument for ) such matrices and which satisfy the congruence conditions determined by the basis of . In other words, if we mapped and , then the induced map on sends them to well defined matrices modulo (the 2 accounts for denominators when writing the in terms of - see below). Then and lift to such that the map determined by and gives an isomorphism .
Remark 3.6.
The following Proposition in Pizer’s paper [Piz80] (Proposition 5.2) gives explicit -bases for a maximal order of our quaternion algebra . We see in particular that the only primes dividing a denominator are 2 and (when and ), with power at most and .
Proposition 3.7.
A maximal order of is given by the -basis:
where is some integer such that .
This means we can explicitly write down the congruences we want our matrices and in the above Corollary to satisfy. We know that these matrices exist for by the ramification properties of the quaternion algebra .
Condition 3.8.
We want to find matrices , for and satisfying
and also
where we remind ourselves that when , we need is a prime congruent to with , and is some integer such that , which we choose in writing down the basis in Proposition 3.7. Note that these last four congruences are vacuous unless or .
4. Weight 0 mod and level 1, for general
We now work with any prime . Let be the unique quaternion algebra over ramified exactly at , a maximal order, and let , where . We will compute the Hecke operator on , which will give us the weight eigenvalues on . The argument in the example above largely generalises to this case. Recall the definition of .
Definition 4.1.
For a prime , the Hecke operator on the space of functions is given by
for . Recall that means we pick a representative of , multiply this in the -place by the matrix (under an identification of and hence of ), and then take the image in .
We explain part of this using the matrices of the previous section. Suppose we had some and corresponding to left ideal classes of (so almost all can be taken to be 1), and we wanted to know if , for some appearing in . Denote by the left -ideal with local generators , where we view as an element of . Denote by the left -ideal with local generators . Then we need to check whether there exists such that (the question of determining when two -ideals are in the same ideal class, for an Eichler order, has been studied in [KV12], but in our case we have local generators for the ideals, meaning that a basis for an ideal is not obvious). This means that for all , we require
(13) |
and also
(14) |
We deduce the following conditions on : let be the set of primes at which at least one of and is not 1, throwing out and . For all , we temporarily define . Then for all , equation (13) is equivalent to
(15) |
For , because is ramified at , equation (13) is equivalent to
(16) |
where denotes the usual -adic valuation on . For , equation (13) tells us that
(17) |
Additionally, since , and because has -adic valuation at most by definition, we see that
(18) |
A similar argument using equation (14) shows that
(19) |
and
(20) |
Note that we can rewrite, for , the condition (being a unit is then guaranteed by the norm condition) as saying
where denotes the standard involution in . We can do the same for , using the adjugate matrix of , which gives an extra factor of . Since , to check this condition we only really need , and modulo . Similarly, we only really need , and modulo . This means we can make use of the matrices computed in the previous section to rephrase our conditions on . Let be an integer basis for , taking for instance the basis of Proposition 3.7. Since , we can write
for variables which are to take values in . Let be a prime in . We can identify with a matrix , where are -linear functions in which can be computed. We can also, for any , identify and with matrices .
We have that if and only if we can find some such that the following conditions hold:
(21) |
The first condition tells us that satisfy some quadratic equation. Because the reduced norm is a positive definite quadratic form in the (due to ramification at ), this means that there are only finitely many solutions to the quadratic equation. We can then enumerate them (for example, diagonalising the quadratic form, computing the finitely many solutions with the new basis, and then solving for ). Once we do so, it remains to check whether they satisfy the last two conditions of (21). By expanding them out, these conditions can be interpreted as congruence conditions on modulo for , or more precisely modulo and . This allows us to determine whether , and hence compute a matrix representative for the Hecke operator . We are now ready to present the algorithm. This should be read in conjunction with the following Table of Notation. The column for ’Corresponding Matrices’ refers to matrices generated using the methods of Section 3.
Table of Notation
Notation | Definition | Corresponding Matrices |
---|---|---|
Quaternion algebra ramified at . | ||
A maximal order of . | ||
Representatives of the left ideal classes of . | ||
-bases for . | ||
The set of primes for which in some all elements have reduced norm divisible by , excluding and including . These are the primes at which we need to compute matrices. | ||
Generators for . | for each . | |
A -basis for . | for each . | |
The elements of corresponding to the . The square brackets means the double coset represented by the adelic point . From the way we compute this, almost all will be 1. | corresponding to the for each . | |
, for , and | . | |
The matrices . | ||
The characteristic function of the point . | ||
for and . The multiplication means , with multiplication occuring only in the -th place. | ||
The set of primes such that at least one of and is not 1, excluding and . |
Algorithm 4.2.
Input: distinct primes and .
Output: a matrix representing the action of the Hecke operator on the space of all functions .
-
(i)
Define a quaternion algebra over , ramified exactly at . Define a maximal order with integer basis given as in Proposition 3.7, for which we denote the basis elements . Compute the left ideal classes of , and bases for them.
-
(ii)
Compute the points corresponding to for each as follows. We take to be any generator of (which we know is principal). To do this, for our basis , compute the reduced norm of each of the four elements and set to be any of these elements whose reduced norm has minimal -adic valuation. Note that for almost all this valuation is zero, so we can instead take , and do so when possible.
-
(iii)
Determine the set , defined in the Table of Notation. For each , compute and . For each , compute matrices satisfying Condition 3.8. Using these, compute matrices corresponding to the , which in any case are well defined modulo . Denote these by for each . By expressing each as a -linear combination of , we can compute matrices corresponding to the .
Remark 4.3.
So far we have only mentioned as that it needed to be added to . If one wanted to compute several Hecke operators at once, they could add all the primes for the operators into . Then in the above steps we computed all the relevant matrices, to save us repeating the calculations if we were to compute the Hecke operators one by one.
-
(iv)
Let be the characteristic functions of the points . This is a basis for the vector space of -valued functions on . Let be the matrices
. Define the quantity for and byThen by definition we have the formula
This is then the -th entry of the matrix for with respect to the basis . Hence it remains to compute this quantity .
-
(v)
Fix . Determine the set . Compute the quantities
and
-
(vi)
Check if there exist integers such that the following conditions hold:
where is the matrix . If such exist, set to be 1, and otherwise 0. Note that the only dependence on is in the last condition.
5. Introducing weight and level
We now discuss what modifications must be done when computing the Hecke eigenvalues mod of general weight and level on the quaternion side. In the previous section we checked whether two elements of are the same by checking if they determine isomorphic invertible left -ideals. The point is that if we now attach -structure to our ideals, we need to determine whether this isomorphism sends one -structure to the other. Our aim is to compute the Hecke operators , for , on the space of functions . Recall this was given by
for . If we wanted to isolate the Hecke eigenvalues arising from weight , then we need to restrict to the functions which satisfy , where acts on by multiplication in the -place. In our case, we can identify with , which is closed under multiplication, where and are the generators of ( can be viewed as a uniformiser of ).
Firstly, we write down the elements of . For defined as previously, we have
Hence, for our points of , we need to multiply on the left by representatives of and at the appropriate places, in order to determine all the elements of . Ranging over each choice of and , the corresponding -double cosets cover , but need not be distinct. For example, for any triple
if we define as in the Updated Table of Notation below, then and define the same element of . This example comes from the fact that and (but and for ), so that
When interpreting elements of in terms of isomorphism classes of invertible left -ideals with -structure, this phenomenon is due to automorphisms of the left -ideals, which then shift around the -structure; in our example we always have the automorphism given by multiplication by , which sends , . It is possible to determine these automorphisms. Recall that the points give local generators for representatives of the left ideal classes of . The automorphisms of as a left -ideal are precisely given by right multiplication by the units of the right order of , where we define this as in [Voi21] to be:
Note that will always be a finite set, once again because the quaternion algebra is ramified at infinity.
Remark 5.1.
In Serre’s letter [SL96], the relationship between Hecke eigenvalues on the modular form and quaternion sides arises as a result of some generalisation of the Deuring correspondence. Classically, this establishes an equivalence of categories between supersingular elliptic curves mod under isogenies, and invertible left -ideals under nonzero left -module homomorphisms (for a maximal order of the quaternion algebra ramified at ). See Theorem 42.3.2 of [Voi21] for a reference. Then, if a supersingular elliptic curve corresponds to an invertible left -ideal , we see that . But the automorphism group of an elliptic curve is well understood. As in Theorem III.10.1 of [Sil09], the automorphism group of has order dividing 24, and if , then has order 2 and is just given by .
We have seen that in listing the elements of , we need to identify, for example, and when and , because . In general, if , we know
for and as defined in the Updated Table of Notation. To define this, we use the fact that because , we have for any , , and so . Let be the matrix congruent to mod for each .
From this, we are motivated to define the set
where we identify for any . Then we can enumerate the elements of as . Note that reduces to when we quotient by in . Once again, an obvious choice of basis for functions is given by the characteristic functions . One is then led to computing the following quantity:
for . Then we have the formula
If we index the rows and columns of the matrix for , with respect to this basis, by , then the -th entry is this value above. We observe that if , for as defined in step (iv) in Algorithm 4.2. This is because if , then they generate left -ideals with -structure that are isomorphic; in particular the ideals are isomorphic, and so . One can think of this as replacing in Algorithm 4.2 with a permutation matrix, depending on , keeping track of the -structure.
So now we see how to compute . It is 0 if . If , then from Algorithm 4.2 we compute some such that for all ,
and also
Pick representatives and for . What we need to check now is whether we also have
(22) |
and for
(23) |
for some and . In other words, we need to compute , which we know is in by definition of , and then check if this reduces to modulo the uniformiser , for some and . We also need to compute for each the terms , which we know are in , and then check whether they reduce to modulo , for the same and . To make sense of this, we need to write down matrices at the primes dividing as well, so we extend our set in the Updated Table of Notation (changes from the previous table in bold).
We are now ready to write out the algorithm for the general case.
Updated Table of Notation
Notation | Definition | Corresponding Matrices |
---|---|---|
Quaternion algebra ramified at . | ||
A maximal order of . | ||
Representatives of the left ideal classes of . | ||
-bases for . | ||
The units of the right orders of . | ||
The set of primes for which in some all elements have reduced norm divisible by , excluding , and including and all primes dividing . These are the primes at which we need to compute matrices. | ||
for each . | ||
for each . | ||
The elements of corresponding to the . The square brackets means the double coset represented by the adelic point . From the way we compute this, almost all will be 1. | corresponding to the for each . | |
The -adic valuation of . | ||
, for , and | . | |
The matrices . | ||
Identified with . |
Notation | Definition | Corresponding Matrices |
---|---|---|
For , let be the image of in . | ||
for ,
|
For , consider for all . This lives in , so we can compute the corresponding matrix in , and consider the reduction . Let be the matrix congruent to mod for each . | |
Here we identify for any . | ||
for | Here for and with reduction mod . To be precise, we should choose lifts of to and to . By construction, is well defined for any choice of representative of . | |
The characteristic function of the point . | ||
for . | ||
The set of primes such that at least one of and is not 1, excluding and , and including all primes dividing . |
Algorithm 5.2.
Input: a prime and level coprime to , together with a prime coprime to .
Output: a matrix representing the action of the Hecke operator on the space of all functions .
-
(i)
Define a quaternion algebra over , ramified exactly at . Define a maximal order with integer basis given as in Proposition 3.7, for which we denote the basis elements . Compute the left ideal classes of , and bases for them.
-
(ii)
Compute the points corresponding to for each as follows. We take to be any generator of (which we know is principal). To do this, for our basis , compute the reduced norm of each of the four elements and set to be any of these elements whose reduced norm has minimal -adic valuation. Note that for almost all this valuation is zero, so we can instead take , and do so when possible.
-
(iii)
Determine the set , defined in the Updated Table of Notation. For each , compute and . For each , compute matrices satisfying Condition 3.8. Using these, compute matrices corresponding to the , which in any case are well defined modulo . Denote these by for each . By expressing each as a -linear combination of , we can compute matrices corresponding to the .
-
(iv)
Compute each , and for each , compute and for . Determine representatives for . Fix . We will compute for and among these representatives. Determine the set . Compute
and
-
(v)
We firstly compute as in Algorithm 4.2. Check if there exist integers such that the following conditions hold:
where is the matrix . If such exist, set to be 1, and otherwise 0. Note that the only dependence on is in the last condition.
-
(vi)
If , set for all , . Otherwise, taking our solution from the above step, compute the matrices
for , and
Writing in terms of the generators and of the quaternion algebra , let be the reduction modulo , where we view the elements of as , which is a group under multiplication.
Then, for representatives and of , define
(24) -
(vii)
If we index the rows and columns of the matrix for , with respect to the basis consisting of , by , then the -th entry is
This gives us the matrix for .
References
- [BCP97] Wieb Bosma, John Cannon, and Catherine Playoust, The Magma algebra system. I. The user language, J. Symbolic Comput. 24 (1997), no. 3-4, 235–265, Computational algebra and number theory (London, 1993). MR MR1484478
- [Buz] Kevin Buzzard, Computing modular forms on definite quaternion algebras, available at https://www.ma.imperial.ac.uk/ buzzard/maths/research/notes/.
- [Con] Keith Conrad, A multivariable Hensel’s lemma, available at https://kconrad.math.uconn.edu/blurbs/.
- [CS01] Caterina Consani and Jasper Scholten, Arithmetic on a quintic threefold, Internat. J. Math. 12 (2001), no. 8, 943–972. MR 1863287
- [Del69] Pierre Deligne, Formes modulaires et représentations -adiques, Séminaire Bourbaki 11 (1968-1969), 139–172 (fre).
- [Dem05] Lassina Dembélé, Explicit computations of Hilbert modular forms on , Experiment. Math. 14 (2005), no. 4, 457–466. MR 2193808
- [Dem07] by same author, Quaternionic Manin symbols, Brandt matrices, and Hilbert modular forms, Math. Comp. 76 (2007), no. 258, 1039–1057. MR 2291849
- [Koh01] David R. Kohel, Hecke module structure of quaternions, Class field theory—its centenary and prospect (Tokyo, 1998), Adv. Stud. Pure Math., vol. 30, Math. Soc. Japan, Tokyo, 2001, pp. 177–195. MR 1846458
- [KV12] Markus Kirschmer and John Voight, Corrigendum: Algorithmic enumeration of ideal classes for quaternion orders, SIAM J. Comput. 41 (2012), no. 3, 714. MR 2967473
- [Piz80] Arnold Pizer, An algorithm for computing modular forms on , Journal of Algebra 64 (1980), no. 2, 340–390.
- [Sil09] Joseph H Silverman, The Arithmetic of Elliptic Curves, Graduate texts in mathematics, Springer, Dordrecht, 2009.
- [SL96] J-P. Serre and Ron Livne, Two letters on quaternions and modular forms (mod ), Israel Journal of Mathematics 95 (1996), 281–299.
- [SW05] Jude Socrates and David Whitehouse, Unramified Hilbert modular forms, with examples relating to elliptic curves, Pacific J. Math. 219 (2005), no. 2, 333–364. MR 2175121
- [Voi13] John Voight, Identifying the matrix ring: algorithms for quaternion algebras and quadratic forms, Quadratic and higher degree forms, Dev. Math., vol. 31, Springer, New York, 2013, pp. 255–298. MR 3156561
- [Voi21] John Voight, Quaternion algebras, Graduate Texts in Mathematics, Springer Cham, 2021.
- [Wie06] Gabor Wiese, Mod modular forms, available at https://wstein.org/msri06/wiese/, 2006.