Exceptional theta functions and arithmeticity of modular forms on
Abstract.
Quaternionic modular forms on the split exceptional group were defined by Gan-Gross-Savin. A remarkable property of these automorphic functions is that they have a robust notion of Fourier expansion and Fourier coefficients, similar to the classical holomorphic modular forms on Shimura varieties. In this paper we prove that in even weight at least , there is a basis of the space of cuspidal modular forms of weight such that all the Fourier coefficients of elements of this basis are in the cyclotomic extension of .
Our main tool for proving this is to develop a notion of “exceptional theta functions” on . We also develop a notion of exceptional theta functions on . In the case of , these are level one, holomorphic vector-valued Siegel modular forms, with explicit Fourier expansions, that are the theta lifts from algebraic modular forms on anisotropic for the dual pair . In the case of split , our exceptional theta functions are level one quaternionic modular forms, with explicit Fourier expansions, that are the theta lifts from algebraic modular forms on anistropic for the dual pair .
As further consequences of this theory of exceptional theta functions, we also obtain the following corollaries: 1) there is an algorithm to determine if any cuspidal, level one Siegel modular form on of most weights is a lift from ; 2) the level one theta lifts from possess an integral structure, in the sense of Fourier coefficients; 3) in every even weight , there is a nonzero, level one Hecke eigen quaternionic cusp form on split with nonzero Fourier coefficient. Finally, we obtain evidence for a conjecture of Gross relating Fourier coefficients of certain quaternionic modular forms to -values.
1. Introduction
Holomorphic modular forms on Shimura varieties have a good notion of Fourier coefficients. It is theorem, going back to Shimura, that the Fourier coefficients of holomorphic modular forms on Hermitian tube domains give these modular forms an algebraic structure: there is a basis of the space of holomorphic modular forms so that all the Fourier coefficients of elements of this basis are algebraic.
Outside the realm of holomorphic modular forms on Shimura varieties, there is little one can say111One exception might be the case of globally generic cohomological automorphic forms, for which the Whittaker coefficients can be directly related to Satake paremeters. about the arithmeticity of Fourier coefficients of automorphic functions. In fact, at present, it is not clear in general how one might define a good notion of Fourier coefficients for spaces of non-holomorphic automorphic forms.
Nevertheless, the quaternionic exceptional groups possess a special class of automorphic functions called the quaternionic modular forms, which do have a good notion of Fourier expansion and Fourier coefficients. These automorphic forms go back to Gross-Wallach [GW94, GW96] and Gan-Gross-Savin [GGS02]. The precise shape of their Fourier expansion was determined in [Pol20a], extending and refining earlier work of Wallach [Wal03]. In particular, it is possible to define what it means for a quaternionic modular form to have Fourier coefficients in some ring . In [Pol22], we conjectured that the space of quaternionic modular forms of some fixed weight on a quaternionic exceptional group has a basis consisting of elements all of whose Fourier coefficients are algebraic numbers.222It seems useful to point out that there has been other recent work, specifically [PV21], that conjectures the existence of surprising algebraic structures on spaces of non-holomorphic automorphic forms.
In this paper, we provide substantial evidence toward this conjecture in the case of the split exceptional group . We setup the result now.
Suppose is a cuspidal quaternionic modular form on of weight . Let denote the unipotent radical of the Heisenberg parabolic of and its center. Then
is the Fourier expansion of . Here the range over non-degenerate characters of and the are the generalized Whittaker functions of [Pol20a]. The functions are locally constant and called the Fourier coefficients of . We say that has Fourier coefficients in a ring if for all characters of and all . We write for the space cuspidal quaternionic modular forms on of weight with Fourier coefficients in .
Let be the cyclotomic extension of .
Theorem 1.0.1.
Suppose is even. Then there is a basis of the cuspidal quaternionic modular forms of weight with all Fourier coefficients in . In other words, .
Our main tool for proving Theorem 1.0.1 is a notion of “exceptional” theta functions, that mirrors the classical theory of Siegel modular theta functions associated to pluriharmonic polynomials. Recall that these classical pluriharmonic theta functions are (often cuspidal) Siegel modular forms, with completely explicit Fourier expansions, that can be considered as arising from the Weil representation restricted to where is a rational quadratic space whose quadratic form is positive definite. In other words, they are the theta lifts from algebraic modular forms on with nontrivial archimedean weight to holomorphic Siegel modular forms. From the perspective of theta lifts and algebraic modular forms, the remarkable fact is that one can give the Fourier expansions of these lifts completely explicitly.
To prove Theorem 1.0.1, we develop a notion of exceptional theta functions on . These are quaternionic modular forms on whose Fourier coefficients we can tightly control. They arise as theta lifts from an anisotropic group of type . Using the Siegel-Weil theorem of [Pol23], we can prove that every cuspidal quaternionic modular form of even weight at least on is one of our exceptional theta functions. This establishes the theorem.
Our theory of exceptional theta functions on has a parallel–but easier–development on . As this theory is a bit easier, and might be more familiar to the reader, we begin with the -case.
1.1. Holomorphic theta functions
Let denote the the algebraic group of type that is split at every finite place and compact at the archimedean place. Let denote the simply connected group of type that is split at every finite place and the group at the archimedean place. There is a dual pair , and a corresponding theta lift from to using the automorphic minimal representation on [Kim93] that was studied by Gross-Savin [GS98]. This lift produces (in general) vector-valued holomorphic Siegel modular forms on . It makes sense to ask if, given an algebraic modular form on , one can give an explicit Fourier expansion of its theta lift to . This is easy if the weight of the algebraic modular form is trivial, but is not immediate (at least to us) if the weight is nontrivial. Theorem 1.1.1 computes this Fourier expansion in the level one case.
We now setup this theorem. Let denote the octonions over with positive definite norm. Denote by the -dimensional exceptional cubic norm structure consisting of the Hermitian matrices with coefficients in . Let denote the -dimensional space of trace octonions, let denote the standard representation of and set . There is a projection given by taking the trace projections of the off-diagonal entries of an element of . Consider the natural map
and denote by the image of under this map.
Let denote the highest weight of the representation of , and let denote the highest weight of the -dimensional adjoint representation . For non-negative integers , let denote the representation of with highest weight , embedded in . This representation is the one generated by where and are highest weight vectors for and for the same Borel subgroup of . Given , and , we can form the pairing
where we have shifted the in the exponent of .
Denote by Coxeter’s order of integral octonions, and set the elements whose diagonal entries are in and off-diagonal entries are in . Recall that Kim’s modular form on has Fourier expansion
where and if is rank one then where is the largest integer with .
Finally, set .
Theorem 1.1.1.
Suppose and is the level one algebraic modular form on with . Then the theta lift of is a vector-valued Siegel modular form of weight with Fourier expansion
When [GS98] proves that is a cusp form.
A simple restatement of Theorem 1.1.1 is as follows. Consider the projection map given by sending
Then if (Hermitian matrices with coefficients) is half-integral, the Fourier coefficient of is
(1) |
(The sum is finite and explicitly determinable.) We verify that the live in the highest weight submodule
so that really is a vector-valued Siegel modular form for the representation .
An important computational aspect of Theorem 1.1.1 is that one can use in the pairing instead of . This enables one to compute theta lifts much more quickly than if one had to use the algebraic modular form . Moreover, even if one does not know a priori that , the theorem still holds as stated. In particular, verifying that the right hand side of equation (1) is nonzero for a single shows that . The reader can, of course, easily check this claim directly.
The Fourier expansion in Theorem 1.1.1 can be seen as completely analogous to the Fourier expansion of classical pluriharmonic theta functions. Indeed, the becomes the pluriharmonic polynomial, and the rank one ’s become the lattice vectors over which one sums.
When the Siegel modular form is a Hecke eigenform, it has Satake parameters , one for each prime number , which are semisimple elements in . (Because we work with level one forms, we blur the distinction between and .) It is proved by Gross-Savin [GS98], Maagard-Savin [MS97], and Gan-Savin [GS21] that the theta lift is functorial for spherical representations; in fact, it is functorial for all representations, see Gan-Savin [GS22]. In particular, these conjugacy classes are in . Thus Theorem 1.1.1 can produce numerous explicit examples of level one vector-valued Siegel modular forms all of whose Satake parameters are in . We have taken the liberty of providing some small illustration of this, as follows. Let and . Then it is known from Chenevier-Taibi [CT20] that the space of vector-valued, level one cuspidal Siegel modular forms of these weights are one dimensional.
Corollary 1.1.2.
For both and , the cuspidal Siegel modular forms of these weights are lifts from . In particular, their Satake parameters all lie in .
Indeed, the proof of Corollary 1.1.2 is to produce a single and , and a single so that the right hand side of (1) is nonzero for these . It then follows that the are nonzero level one Siegel modular forms, and thus by the dimension computation of [CT20] are the unique cuspidal level one eigenforms of weights and . Producing many more such explicit examples would be possible. We remark that one can find conjectures here [BCFvdG17] of which small weight level one Siegel modular forms have their Satake parameters in .
More than just a couple of examples, however, Theorem 1.1.1 produces an algorithm to determine if any fixed level one Siegel modular cusp form of most weights is a lift from . To setup the result, note that for a weight , it is known [Ibu02] that there exists explicitly determinable finite sets of half-integral symmetric matrices so that if is a cuspidal level one Siegel modular form of weight , and if for all , then .
Corollary 1.1.3.
Suppose with , and is a level one Siegel modular cusp form of weight on , whose Fourier coefficients are given for all . Then there is an algorithm to determine if for some algebraic modular form on .
Let us indicate now some of the ingredients that go into the proof of Corollary 1.1.3. First, let us clarify that the theta lifts of level one algebraic modular forms that appear in Theorem 1.1.1 are defined as integrals
for a certain specific vector-valued element of in the minimal representation on . Here, in the integral, is a -equivariant pairing valued in . Let be a spanning set of , and . One can use Theorem 1.1.1 to explicitly compute the Fourier coefficients of associated to the various in , where . Using linear algebra, one can check algorithmically if is some linear combination of the ’s. Thus Theorem 1.1.1 gives an algorithm to determine if the Siegel modular form is a for some algebraic modular form on . Thus Corollary 1.1.3 will be proved if one knew the following claim:
Claim 1.1.4.
Suppose is a level one cuspidal Siegel modular form of weight with . Suppose is in the image of the theta correspondence for . Then for some algebraic modular form .
We prove this claim. The proof uses some powerful ingredients: The Howe Duality theorem of Gan-Savin [GS21]; an analysis of the minimal representation, as provided by Gross-Savin [GS98], Magaard-Savin [MS97] and Gan-Savin [GS21]; the existence of enough nonvanishing Fourier coefficients of of a certain form, as proved by Böcherer-Das [BD21]; and another argument from Gross-Savin [GS98].
Our proof of Theorem 1.1.1 is very simple. Let be the complexification of the eigenspace for the Cartan involution on the Lie algebra of . Write , the natural decomposition, so that annihilates the automorphic form on associated to any holomorphic modular form for this group. Let be a basis of and be the dual basis of . For an automorphic form on , set For an integer , let , so that
If (abusing notation) denotes the automorphic form on associated to Kim’s holomorphic modular form, we set .
With this definition, using Kim’s expansion of , we compute the Fourier expansion of . This is the main step in the proof of Theorem 1.1.1. We then use this to compute the Fourier expansion of . One obtains (the automorphic form associated to) a holomorphic function with the Fourier expansion as given in Theorem 1.1.1. We remark that the use of differential operators as we do has some overlap with the works [Ibu99, Cle22] of Ibukiyama and Clerc.
1.2. Quaternionic theta functions
Denote by the group of type which is split at every finite place and at the archimedean place and by the split exceptional group of this Dynkin type. Analogous to the dual pair is the dual pair , where is a specific form of that is split at all finite places and compact at the archimedean place, defined as the stabilizer of the identity matrix . We compute the theta lifts of certain algebraic modular forms on to , and obtain cuspidal quaternionic modular forms on together with their exact Fourier expansions.
More precisely, let denote the trace elements of . In other words, consists of the with , where is the symmetric non-degenerate pairing on determined by . There is a surjective -equivariant map from to the Lie algebra of . Denote by the kernel of this map. It is an irreducible representation of of dimension . For an integer , let denote the irreducible representation of with highest weight , generated by the tensor product of a highest weight vector of . It follows from the archimedean theta correspondence calculated in [HPS96] that algebraic modular forms on for the representation should lift to quaternionic modular forms on of integer weight . We explicitly compute the Fourier expansion of this lift, and as a result, obtain exceptional “pluriharmonic” cuspidal quaternionic theta functions on .
We setup the statement of the result. To do so, recall that the group has not one but two integral structures [Gro96], [EG96]. More precisely, if denotes one of these integral structures, then the double coset space has size two. Because of this, algebraic modular forms for can be described as follows. Denote by the subgroup of fixing the cubic norm. Set to be the subgroup of preserving the lattice and fixing the element . Recall the element of norm one from [EG96] or [Gro96]. Let denote the subgroup of preserving the lattice and fixing the element . Fix an element with . If is a representation of , let act on via , where the conjugation takes place in . Then a level one algebraic modular form for can be considered as a pair
In the case of , we rephrase this as follows. Let be the symmetric non-degenerate pairing on determined by ; one has
where is the symmetric trilinear form on satisfying , times the cubic norm on . Set to be the perpendicular space to under the pairing . One easily verifies that . Thus
and is invariant for the natural action of . It will be convenient for us to consider the algebraic modular form to be the pair .
Now, for and , write
We use the same notation for the pairing that extends this one via . Thus if , and , we can compute the quantity . We similarly define , by replacing the pairing with .
Set . For of rank one, set where is the largest integer with . In [Pol20b], we proved that the minimal modular form on has Fourier expansion
Here is the unipotent radical of the Heisenberg parabolic of , , denote constant terms, and is the generalized Whittaker function of [Pol20a].
Now, if and , set if . Moreover, with this notation, let and be the binary cubic forms given as
We prove the following.
Theorem 1.2.1.
Suppose is a level one algebraic modular form on for the representation with . Represent as a pair with being invariant and being invariant. Let be in , respectively so that and Then the theta lift is a cuspidal, level one, quaternionic modular form on of weight with Fourier expansion
A simple restatement of Theorem 1.2.1 is as follows. If is an integral binary cubic form, then the Fourier coefficient of is
These sums are finite.
The reason we do not state Theorem 1.2.1 in the case is because then the theta lifts will be non-cuspidal. In fact, the theta lifts obtained for are exactly the automorphic forms obtained in [GGS02, Section 10]. Thus Theorem 1.2.1 may be considered a generalization of [GGS02, Section 10].
Again, just like Theorem 1.1.1, the Fourier expansion given by Theorem 1.2.1 is completely parallel to the classical pluriharmonic theta functions: The in are the pluriharmonic polynomial, and the sum over of rank one is the sum over lattice vectors.
We now state a few corollaries of Theorem 1.2.1. For the first corollary, we can partially refine Theorem 1.0.1 in the case of level one.
Corollary 1.2.2.
There is a lattice and a lattice so that the level one theta lifts of elements of these lattices to have Fourier coefficients that are integers when evaluated at .
For the second corollary, recall that it was proved in [cDD+22] that if is a cuspidal automorphic representation on that corresponds to a level one quaternionic modular form of even weight , then the completed standard -function satisfies the exact functional equation , so long as the Fourier coefficient of is nonzero. At the time of the writing of [cDD+22], it was not known whether such exist. Using Theorem 1.2.1, one easily obtains the following.
Corollary 1.2.3.
Suppose is even. Then there is a cuspidal automorphic representation on that corresponds to a level one quaternionic modular form of weight with nonzero Fourier coefficient.
This corollary is, in fact, an ingredient in the proof of Theorem 1.0.1.
The third corollary we state has to do with the Fourier expansion of a particular cuspidal quaternionic modular form. To setup this corollary, recall that Dalal [Dal21] has recently given an explicit formula for the dimension of the level one quaternionic cuspidal modular forms of weights at least . From his dimension formula, one has that the first such nonzero cusp form appears in weight , and the space of weight cuspidal quaternionic modular forms is one-dimensional, spanned by a cusp form . Combining the (proof of) Corollary 1.2.3 with Corollary 1.2.2, one obtains that can be normalized to have integer Fourier coefficients, and that the Fourier coefficient associated to the cubic ring is nonzero.
Now, Benedict Gross has suggested that the Fourier coefficients of certain non-tempered cuspidal quaternionic modular forms on of weight should be related to square-roots of twists of -values of classical modular forms of weight by Artin motives associated to totally real cubic fields. As pointed out to the author by Mundy, the quaternionic modular form should be one of these non-tempered lifts, to which Gross’s conjecture applies. Now, for congruent to or modulo , denote by the quadratic ring of discriminant . Then on the one hand, in the case of , Gross’s conjecture implies that the square Fourier coefficients associated to the cubic ring should be related to the central333We here use the classical normalization of -functions, instead of the automorphic normalization. -value of the twist of Ramanujan’s function by the quadratic character associated to . Denote by , the Shimura lift of . On the other hand, following Waldspurger [Wal81], Kohnen-Zagier [KZ81] relate the squares of Fourier coefficients to the same -value, . It thus makes sense to ask, in light of Gross’s conjecture, if there is some relationship between and . It turns out, the numbers are equal:
Corollary 1.2.4.
Normalize so that , and normalize so that . Then for all .
1.3. Acknowledgements
It is pleasure to thank Wee Teck Gan, Nadya Gurevich, and Gordan Savin for engaging with the author in a “Research in Teams” project in Spring 2022 at the Erwin Schrodinger Institute, which helped to stimulate thinking about exceptional theta correspondences. We thank them for fruitful discussions. We thank Tomoyoshi Ibukiyama for sending us his note [Ibu02], and we also thank Gaetan Chenevier, Chao Li, Finley McGlade, Sam Mundy, Cris Poor and David Yuen for helpful discussions.
2. Groups and embeddings
In this section, we explain various group theoretic facts regarding the groups with which we work.
2.1. Some exceptional groups
We begin by defining the group . Thus let be our exceptional cubic norm structure, . A typical element of we write as an ordered four-tuple , so that , and . We put on Freudenthal’s symplectic form, and quartic form. The group is defined as the algebraic -group preserving these two forms. We let be the group that preserves the forms on up to similitude, and let be this similitude.
The Siegel parabolic of is defined as the stabilizer of the line . Write for the group of linear automorphisms of that preserve the norm on , up to scaling. Let be this scaling factor, and the kernel of . A Levi subgroup of can be defined as the subgroup that also stabilizes the line . This is isomorphic to the group of pairs with . Such a pair acts on as . We write for this group element of .
2.2. The first dual pair
We now explain how embeds in . To accomplish this, we describe a linear isomorphism between and . Here is the trace octonions, is the -dimensional defining representation of , and is the kernel of the contraction map . We let be the standard symplectic basis of .
-
•
If , we map to where and .
-
•
We map to and to .
-
•
Set and (with indices taken modulo ). We map to and to .
Via the natural action of on , we obtain an action of this group on . It is clear that the action is faithful. To obtain the embedding into , we must check that preserves the symplectic and quartic form on , up to scaling:
Proposition 2.2.1.
The defined action of on gives an embedding .
To prove the proposition, we first make a few lemmas. We work over a general field of characteristic .
Lemma 2.2.2.
Suppose is in and so that the element is in . Then with
-
(1)
-
(2)
-
(3)
.
Proof.
This is direct computation.∎
Let now and consider the element in the Lie algebra of . Recall the element , in the Lie algebra of , that acts on as .
Lemma 2.2.3.
Under the above identification , the operator acts as .
Proof.
We first compute how acts on . Writing out the action of on this element, we obtain
The coefficient of comes from terms:
-
•
:
-
•
:
-
•
:
-
•
:
The coefficient of again comes from terms:
-
•
:
-
•
:
-
•
:
-
•
: .
Putting the above computations together, one obtains where .
Finally, one immediately obtains that and . The lemma follows. ∎
The following lemma is immediate.
Lemma 2.2.4.
The element of acts on as . The element acts on as .
Lemma 2.2.5.
Suppose , so that . The action of this element on is in ; in particular, it preserves the symplectic and quartic form on .
Proof.
We compute the action of on when . First, one computes that acts on taking to , where is the cofactor matrix of . This gives that the part of maps to . The vector (i.e., ) part of moves as .
Completely similarly (it is the same calculation), the part of maps to . The vector part of moves as .
To finish the proof of the lemma, one must verify that the map given by
scales the norm on by . This can be done, for example, by repeatedly applying the Cayley-Dickson construction and using the formulas of [Pol18, Section 8.1]. ∎
We can now prove Proposition 2.2.1.
Proof of Proposition 2.2.1.
Exponentiating the action of , we see that the action of lands in . The group is generated by these elements, together with the and with . The proposition thus follows from the above lemmas. ∎
2.3. Action of the maximal compact
Let denote the standard maximal compact subgroup of , so that
Let denote the subgroup of that fixes the line spanned by . This is a maximal compact subgoup of . Let be the complexification of the part for the Cartan involution on the Lie algebra of for this choice of maximal compact. Write for its two factors. We now work out how acts on .
We begin by discussing the Cayley transform [Pol20a, section 5]. To do, we review some notation from [Pol20a]:
-
•
;
-
•
, with, for , ;
-
•
for , ;
-
•
is the Lie algebra of , which acts on as given in [Pol20a, section 3.4].
The Cayley transform is defined as , so that . It satisfies:
-
(1)
-
(2)
-
(3)
.
For , we set as
For and , the factor of automorphy and the action of on is defined as .
Recall that we let , for and such that , act on as
For set . Here is the positive definite square root of , and for , is the map defined as . For set . Then .
Lemma 2.3.1.
One has
-
(1)
-
(2)
.
Consequently, . Moreover,
and
Proof.
The first parts are direct verifications. For the second, recall that is the subgroup of that stabilizes the lines and , while is the subgroup of that stabilizes the lines spanned by and . The last part is again a direct verification. ∎
Here is the statement of the result.
Proposition 2.3.2.
Suppose , so that . Let , with perpendicular to , so that . Then with .
Proof.
We know from lemmas above that for some . To compute it, we apply both sides to . The right-hand side gives , which is . The left-hand side gives .
To further compute this left-hand side, note that we have is identified with . Applying , one gets the action of on the . ∎
2.4. More exceptional groups
2.5. The second and third dual pairs
We will now define two more dual pairs, and .
For the second dual pair, , we explicate an identification , as follows. Here is the standard representation of , with standard basis . Identify with . We claim that under this identification, and the natural action of on the left, this group action preserves the symplectic and quartic form on .
It is clear that preserves the symplectic and quartic form. For the action of , the proof is similar to (but much easier than) that given above for . One simply needs to observe that and .
For the third dual pair, , we proceed as follows. Observe that from the action of on and . Now notice that . Because is simply connected, we thus obtain a map from to the centralizer of the group in . Thus we have a unique map so that acts on naturally, and the differential of the map is the Lie algebra embedding .
We now relate these two dual pairs. Thus let be the Levi of the Heisenberg parabolic in . Observe that fixes the line spanned by in , because it does so in . Thus because is the Levi of the Heisenberg parabolic in , we obtain a map .
Proposition 2.5.1.
The above-described two maps are identical.
Proof.
Indeed, it is clear that the ’s act exactly the same way. As for , we have two algebraic representations of on . By the formulas for the Lie bracket on , it is easy to see that the differential of these representations of are identical. Thus, they agree on the level of algebraic groups, as desired.∎
3. Facts about
We set down some notations and results we will need concerning the representations and algebraic modular forms on .
3.1. Special elements of representations of
Let be the trace subspace. Let be the identification of the with its dual given by the trace pairing. Recall that if and then is defined as
If then one defines . This defines a map . (Here is the Lie algebra of the subgroup of that preserves the trace pairing, or equivalently, fixes the element .) As , this map factors through the projection .
Set
One can construct special elements of using the following two lemmas.
Lemma 3.1.1.
Suppose is rank one. Suppose and . Then .
Proof.
Observe that so that
Keeping this expression in mind, we now recall the identity
valid for all in . If , then symmetrizing this identity in gives
Consequently, taking , and we obtain
This gives the lemma. ∎
Lemma 3.1.2.
Suppose and are such that . Then .
Proof.
The point is that one has and (in fact, implies ) if and only if is rank at most one in . But so the lemma follows.
Of course, one can also give a more direct proof: One has
Thus if , then . ∎
We now write down some specific elements with the following property:
-
•
The four dimensional space
is isotropic and singular in the sense that symplectic form on restricted to is and every element of has rank at most one.
Note that is isotropic and singular is equivalent to:
-
•
-
•
are rank at most one, and
-
•
, , and .
If satisfy the above properties, then is a highest weight vector for some Borel. Indeed, one can show that the span of is “amber”, in the sense of Aschbacher [Asc87, 9.3-9.5], see also [MS97, Definition 7.2, Proposition 7.3, Proposition 7.4(1)]. Consequently, such will allow us to construct explicit elements of
Lemma 3.1.3.
Suppose is rank one, and that is such that and . Set . Then and is isotropic and singular.
Proof.
First suppose is rank one and is arbitrary. Then if , then and .
Moreover, observe that if is rank one then , so that . Thus , and if we arrange that is rank one, then .
To ensure that we use . Indeed,
To ensure has rank at most one, we use . Indeed,
Because and and , we get for free that . (Of course, one can also check this directly.)
To complete the proof, we must verify that . For this, we compute
∎
3.2. Algebraic modular forms
Suppose is a representation of . By an algebraic modular form for , we mean an automorphic form satisfying for all and . If has level one, then because the double coset has size two, such can be described by two elements of . In this subsection, we make this identification explicit.
Recall the elements of norm , see [EG96]. Define to be the stabilizer of in and to be the stabilizer of in . From the point of view of double cosets, the element arises as follows. Let be representatives for . Using strong approximation on , we can write with etc. We can choose and so that . Indeed, observe that
-
(1)
-
(2)
so that has finite part in .
We have . If is a representation of , we let act on via .
We make the following notations:
-
•
Let be the image of in . Thus
-
•
Let be the image of in . Thus
-
•
Let be the image in of . Thus
Lemma 3.2.1.
One has
Proof.
Observe
and
Using that , the lemma follows. ∎
Suppose is a representation of and
satisfies for all , and . Then is determined by its values at and . Moreover, because acts freely on , we have
Note that the measures of the two open sets are in the proportion . Consequently, if , one has
Set . Consider , and let . Observe the following lemma:
Lemma 3.2.2.
If then .
Proof.
One has
∎
If one writes to be the perpendicular space to under the pairing , then . This follows from the lemma with . Moreover, observe that, for the action of on , is stabilized by the action of . Thus, we can think of our algebraic modular form as being the pair
4. Theorems on Siegel modular forms
In this section, we give the proof of Theorem 1.1.1 and its corollaries, Corollary 1.1.2 and Corollary 1.1.3. We do this assuming Theorem 4.0.1 stated below, which is the main technical ingredient in the proof of Theorem 1.1.1. Theorem 4.0.1 will be proved in section 6.
To setup the statement of Theorem 4.0.1, suppose is a representation, and is a function satisfying . In this scenario, let be the -valued function defined as
where is a -basis of . It is easily checked that is again -equivariant. Recall also that is the factor of automorphy defined in section 2. Let be the representation of on .
Theorem 4.0.1.
Suppose and . Let be an integer. There is a nonzero constant , independent of and , so that
Moreover, lies in the highest weight submodule of .
Recall that if is a level one algebraic modular form for the representation , then we defined the theta lift of as
where we have normalized the measure so that has measure . By rescaling or this measure, we can (and will) ignore the term .
We first state the fact that is the automorphic form corresponding to a Siegel modular form of weight .
Proposition 4.0.2.
For and the Siegel upper half-space of degree three, define
for any with in . Then is well-defined, and is a level one Siegel modular form of weight . If , it is a cusp form.
Proof.
The fact that is well-defined comes from the action on , which was determined in section 2. To see that it’s a holomorphic modular form of the correct weight, we use [GS98, Theorem 3.5]. The cuspidality when is proved in [GS98, Corollary 4.9].
In more detail, if is a linear form on , then one can write as a sum of terms of the form , where
is a usual scalar-valued theta lift. By [GS98, Theorem 3.5], as functions of , these lifts all lie in the holomorphic discrete series representation with minimal -type . Moreover, by the -equivariance that proves that is well-defined, the vector-valued function exactly encompasses the minimal -type in . Here we use that this minimal -type appears in with multiplicity one. Consequently, is a holomorphic Siegel modular form. It is clearly level one. Finally, [GS98, Corollary 4.9] shows that all the are cusp forms if , thus so is .
This completes the proof. ∎
Proof of Theorem 1.1.1.
Proof of Corollary 1.1.2.
Let span a null, two-dimensional subspace of the trace zero elements . That they are null means that . We set . It is easy to see that . Indeed, we can choose a Borel subgroup of to be the one that stabilizes the flag , and then it is clear that is a highest weight vector in . A computer calculation shows that, if is defined as above with and , then the
Fourier coefficient of is nonzero. Similarly, if is defined as above with and , then the Fourier coefficient of is nonzero.
To actually do the computation on a computer, we proceed as follows. First, we set to be the quaternion algebra over ramified at and the archimedean place. Let be its usual basis. We obtain the octonion algebra via the Cayley-Dickson construction using . This means that the addition in is component-wise and the multiplication is
Set and . Then, the following are a basis of , Coxeter’s ring [Cox46]: . These are the simple roots of the root lattice, with the extended node, the branch vertex, and going along longways.
For we take elements inside of as
Finally, to compute the Fourier coefficient, where
we must explain how to enumerate the rank one with . The point is that
being rank one with implies , . Thus one only must search through a finite list of (namely, those with these norms) to find all such . ∎
Proof of Claim 1.1.4.
Recall that we assume , that is a level one Siegel modular form of weight and we wish to prove that if is in the image of the theta correspondence from , then for a level one algebraic modular form .
We first use the Howe Duality theorem of Gan-Savin [GS21] to reduce to the case of eigenforms. Thus write as am orthogonal sum of eigenforms with each nonzero. Then the Petersson inner product for each . Let be the automorphic cuspidal representation of generated by . Then, by changing the order of the theta integral, one sees that the big theta lift of to is nonzero. By Howe Duality [GS21] and the argument in [Gan22, Proposition 3.1], is an irreducible representation of .
We now argue that . Let the archimedean component of , which a priori could depend upon . For any fixed vector in the finite part of , and any fixed vector in the finite part of one can consider the map
given by the theta lift. By the archimedean correspondence proved in Gross-Savin [GS98], the image is a holomorphic discrete series representation with lowest -type . Because for some choice of and we must obtain a theta lift that is not orthogonal to , we must have and .
It now follows from [GS98], because , that consists of cusp forms. Thus again by Howe Duality [GS21] and the argument in [Gan22, Proposition 3.1], is irreducible, so . By the results of [GS98, MS97, GS21] that apply to spherical representations, it must be that is unramified at every finite place. Thus we have a level one algebraic modular form , and we must show that is a nonzero multiple of . We have now reduced ourselves to the case of eigenforms, so will drop the from , and .
Let be a basis of , the level one automorphic form on corresponding to , and let be a basis of . By changing variables in the integral defining the theta lift from to , one sees that there exists in so that
Set and . Then we have
Let
where the action is diagonal on the minimal representation and the vector space . Because is -equivariant and is -equivariant, one has
Because the -type appears in with multiplicity one [GS98], we can write
where is the basis of dual to the basis . We wish to show that we can set to be the spherical vector in for all in this equality.
Fixing the vectors and letting the vary, the theta integral gives an equivariant map
This map is nonzero by assumption, so by -adic Howe Duality again [GS21] the map is uniquely determined up to scalar multiple. Our goal is to show that the image of the spherical vector on the left-hand side is nonzero on the right-hand side. We will do this by a global argument.
It suffices to show that, for each prime , in the unique up to scalar map , the image of the spherical vector is nonzero. Because this map is unique up to scalar multiple, we must only find some which is spherical at , some which is spherical at , and so that . By [GS98, Proposition 4.5], it suffices to check that has an appropriate period. More precisely, let be an odd prime with , and let be the quaternion algebra over ramified at and infinity. Then it suffices to show that has a period. But finally, by Bocherer-Das [BD21], for some odd prime , has a nonzero Fourier coefficient associated to the maximal order in . Consequently, does have a period and we have shown that the unique map is nonzero on the spherical vector.
Let
where is the spherical vector in . We have shown
is nonzero, and thus a nonzero multiple of . Because
we obtain is nonzero, and thus is nonzero multiple of . The claim is proved. ∎
Proof of Corollary 1.1.3.
We have explained, given , how to compute individual Fourier coefficients of . It remains to explain how to enumerate a spanning set of . To do this, we define elements which are a basis of , as follows.
-
•
-
•
-
•
.
-
•
-
•
-
•
-
•
-
•
Now, in terms of these elements, a basis for root spaces of can be found in [Pol19]. In particular, for the standard Borel chosen in [Pol19], will be a highest weight vector for . By the Poincare-Birkoff-Witt theorem, the representation is spanned by where here
-
•
span the negative root spaces of the simple roots of ;
-
•
the element can be explicitly computed using the formulas of [Pol19];
-
•
one can come up with easy bounds for the integer .
The corollary follows. ∎
5. Theorems on quaternionic modular forms
In this section we prove Theorem 1.2.1, assuming a crucial technical result, Theorem 5.0.1, which is proved in section 7. We also prove Corollaries 1.2.2, 1.2.3, and 1.2.4.
For that is positive semi-definite, and an integer, let be the generalized Whittaker function [Pol20a] on associated to and . Similarly, for a real binary cubic form that is positive semi-definite, let be the associated generalized Whittaker function on .
Let now denote the complexification of the part for the Cartan involution on for the Cartan involution defined in [Pol20a]. Recall that if is a smooth function, then is defined as , where is a basis of .
Theorem 5.0.1.
Suppose and with . Then there is a nonzero constant so that for all rank one and , one has
Here, the -equivariant pairing is defined as follows. By virtue of the exceptional Cayley transform of [Pol20a] and the explanations of subsection 2.5, one has , and . Thus
Thus if , and we obtain an element in .
We can use the theorem to compute the Fourier expansion of the theta lift of a level one algebraic modular form on . We begin with the statement that these lifts are quaternionic modular forms of weight on .
Proposition 5.0.2.
Suppose , and . Then
is a quaternionic modular form on of weight . If , it is a cusp form.
Proof.
First note that, if , then
Thus if is the level one algebraic modular form on with and , then
Now, for and , consider the theta lift
where is the level one automorphic function on associated to . This lift gives an equivariant pairing . By [HPS96], it thus gives a map . Finally, by the -equivariance of , we see that is the minimal -type of this copy of , so it is a quaternionic modular form of weight .
We now show the cuspidality of the theta lifts if . Let and , respectively, be the two standard maximal parabolic subgroups of , so that is the Heisenberg parabolic. Let be the Heisenberg parabolic of , and the standard maximal parabolic with . In terms of the root system underlying the group , with long roots and short roots , is the parabolic for which is in its unipotent radical. The Levi subgroup of is of absolute Dynkin type . Write , , , and for the Levi decompositions.
We must check that the constant terms and are , if is an automorphic form in a representation with with . We first observe the following claim:
Claim 5.0.3.
One has an equality of constant terms and .
Granting the claim for the moment, we obtain that and . The constant terms and , restricted to their Levi subgroups, were determined in [Gan00]. For the first one, see page 174 of [Gan00], it is a sum of terms from a one-dimensional representation of and the minimal representation of . Both of these have integral against , because by [HPS96], the representation with does not participate in the theta correspondence for the dual pair . For the second one, see page 176 of [Gan00], the constant term restricted to is the trivial representation. Thus this too has integral against .
It remains to explain the proof of Claim 5.0.3. For the equality , note that is a sum of terms of the form . But, using that if is rank one with then , we find that the only with and is . Thus . One makes a completely similar argument for the constant term . The proposition is proved. ∎
Proof of Theorem 1.2.1.
Suppose , and is the level one algebraic modular form on with and . Then it follows from Proposition 5.0.2 that the theta lift of to is a quaternionic modular form of weight , and cuspidal if . Up to the constant , its Fourier expansion is given exactly as in the statement of Theorem 1.2.1. The general case, where , is explained in subsection 5.1 below.∎
We explain the proof of Corollary 1.2.2.
Proof of Corollary 1.2.2.
Over , we have a decomposition , where is -stable. Because is a pure inner form of split , one can use the results of [BGW15, sections 2.1, 2.2] to give such a decomposition over : , where is a rational representation of the algebraic group whose complexification is . Now let be the intersection of with ; it is immediately seen to be an integral lattice in , so that . But now, from the explicit formula from Theorem 1.2.1, it is clear that if , then has integral Fourier coefficients. One makes a similar argument for . This proves the corollary. ∎
We now explain the proof of Corollary 1.2.3. To do so, we first construct a special . Thus let
-
•
be an imaginary quadratic field, so that is split, such as .
-
•
, with but
-
•
with
-
•
with and .
Then, as in section 3, we set
-
•
-
•
-
•
and .
We set . It is proved in section 3 that .
We require the following lemma.
Lemma 5.0.4.
Let the notation be as above, with even. Set . Then
Proof.
Proof of Corollary 1.2.3.
We now explain the proof of Corollary 1.2.4.
Proof of Corollary 1.2.4.
First note that the cubic ring is associated to the binary cubic form . Indeed, setting and in , this is a good basis, and computing its multiplication table gives rise to the binary cubic form . The ring is of this form with and .
To explicitly compute the Fourier coefficients of , we now make a specific choice of as in the proof of Corollary 1.2.3. Namely, we take , , , and . We obtain
with
-
•
-
•
-
•
-
•
-
•
-
•
.
Now, given the binary cubic , in order to compute the associated Fourier coefficient of , we must compute the set of so that , , and rank one. We assume . Then is rank one if and only if and , which implies that is rank one. As mentioned above, the set of rank one in with consists just of the three elements .
Suppose that . Then because . So, the entry of . It now follows easily, using that , that is of the form
with and . Substituting , , , , one finds that is an integer and . The contribution to the Fourier coefficient for such a term is computed to be .
One makes a completely similar calculation if or . One obtains
where the sum is over pairs in such that .
But, this is clearly the Fourier coefficient of a harmonic theta function in , associated to the lattice . This space has dimension , spanned by . By looking at the coefficient of , one obtains the corollary. ∎
5.1. Theta lifts
In this subsection, we complete the proof of Theorem 1.2.1. More specifically, we will now use our knowledge of the computation of for arbitrary to compute .
Let and denote by the -Fourier coefficient of this automorphic form. Suppose . To begin, observe that
Now, if , we prove
where is the binary cubic if . Observe that , where . Combining these two facts, one obtains
Here, if , then
and one extends to . We have used Lemma 3.2.2 to get
Thus, putting everything together,
This completes the proof of Theorem 1.2.1.
6. The exponential derivative
The purpose of this section is to prove Theorem 4.0.1, which we restate here:
Theorem 6.0.1.
Suppose and . Let be an integer. There is a nonzero constant , independent of and , so that
Moreover, lies in the highest weight submodule of .
Our proof of this theorem is to, essentially completely explicitly, calculate the derivatives
(2) |
To do this, we use the Iwasawa decomposition , write each as a sum in terms of this decomposition, and calculate the derivatives for each piece.
6.1. Preliminaries
Recall from above the Cayley transform , which satisfies
-
(1)
-
(2)
-
(3)
.
Given , let denote its action on [Pol20a, section 3.4].
Proposition 6.1.1.
One has the following identities:
-
(1)
-
(2)
-
(3)
If , then .
-
(4)
If , then .
6.2. Some computations
For , define . To warm up, we will compute . Let denote the function . Let denote any element of for which .
Lemma 6.2.1.
One has the following computations:
-
(1)
;
-
(2)
;
-
(3)
;
-
(4)
if , then .
Proof.
The first statement follows from the identity . The second statement follows from the fact that , so that . For the third statement, we have
Now so is the coefficient in of , which is . Thus as claimed. For the fourth statement, observe that , , so . As , the statement follows. ∎
We now explain the computation of (2). To setup the result, define inductively as follows: , and for ,
Here we write
and we interpret if is constant. Thus
-
(1)
.
-
(2)
Define now as
and extending to by linearity.
Proposition 6.2.2.
Let the notation be as above. Then .
Proof.
We proceed by induction. Observe that
-
(1)
.
-
(2)
.
-
(3)
.
-
(4)
If , then .
Now, if , then
Thus . Consequently,
As
one can now easily verify the proposition. ∎
We now prove:
Lemma 6.2.3.
Let be as above and let . Set . Then
(3) |
In other words, only the leading term contributes.
Proof.
Suppose . Observe
where we have used that is equivariant for the action of . (This follows, for example, from the recursive formula.)
Now, because of the above equivariance, we can assume . Indeed, the two sides of the equality to be proved are linear in and .
We now compute in the basis of . More precisely, we take a basis of that is a basis of union a basis of , where and is a basis of . We compute in this basis. Then the only terms that contribute to the left-hand side of (6.2) are those with for all . Then, still in order for these terms to contribute in a nonzero way to (6.2), we must have for and for . Consequently, all the that contribute to the sum in our basis satisfy , , and . It now follows from our recursive formula for that only the leading term contributes. ∎
It follows immediately from Lemma 6.2.3 that
Combining this with Proposition 6.2.2, we obtain
Proposition 6.2.4.
Suppose , and . Then
for a nonzero constant
We can now prove Theorem 4.0.1.
Proof of Theorem 4.0.1.
Observe that by appropriate -equivariance, and by equivariance for the unipotent radical of the Siegel parabolic, it suffices to check the equality of the statement of Theorem 4.0.1 when , the Levi of the Siegel parabolic. But this follows from Proposition 6.2.4 and the definition of ; see the proof of Lemma 2.2.5 for the action of on . The proof of the theorem now follows from Lemma 6.2.5 below.∎
Recall that is the kernel of the contraction
Lemma 6.2.5.
If , then .
Proof.
By equivariance and linearity, it suffices to verify the claim of the lemma for . Now suppose . Then
and
Thus contracting yields the term
This is . The lemma follows. ∎
7. The quaternionic Whittaker derivative
The goal of this section is to prove Theorem 5.0.1, which we restate here:
Theorem 7.0.1.
Suppose and with . Then there is a nonzero constant so that for all rank one and , one has
We begin with the following proposition.
Proposition 7.0.2.
Let . Suppose and is rank one. Then the function defined as is quaternionic.
Proof.
Fix a rank one . Let be the character of given as . Here is the image of in , the abelianization of . Let be the unique (up to scalar multiple) moderate growth linear functional satisfying for all and . (Such an exists by a global argument: The global minimal representation has nonzero Fourier coefficients, so there is an for some . Now for an appropriate is the desired functional. The uniqueness of follows from [Pol20a].)
Now let be a basis of the minimal -type of and in the dual basis. Note that . Then is in . One obtains that . This latter space maps -equivariantly to
Finally, mapping to , we obtain a invariant element in . By Huang-Pandzic-Savin [HPS96], this is either or the minimal type of . Contracting now against some , we obtain some (possibly ) multiple of . Applying , it follows that is quaternionic. ∎
7.1. General strategy
As before, let be the generalized Whittaker function of weight associated to , which is positive semi-definite. Set be the component multiplying . (See [Pol20a] for the definition of .) Here is an -equivariant pairing on .
Consider the quantity . Then
Thus if we can compute for where is assumed singular and isotropic, then we can compute this quantity for general .
What we actually do is compute for where is assumed singular and isotropic. To setup the computation, we fix a basis of , fix a basis of that is the union of bases of and of the tensor product basis of . Then we fix our basis of to be the tensor product basis of with the above fixed basis of .
Now, it is clear that only contains the terms in where the equal one of
Note also that because is isotropic, the above Lie algebra elements all commute. We obtain that
We now observe the following fact: if , then
Thus, if we can compute the right hand side, then we can compute the left hand side.
To compute the quantity , we will represent as an integral, and differentiate under the integral sign. This is inspired by the work of McGlade-Pollack [MP22]. More exactly, set
Here is the projection onto the Lie algebra of the long root and we write for short.
We prove the following theorem. To setup the theorem, recall from [Pol20a] that if with , then
Moreover, let be the similitude character on the Levi of the Heisenberg parabolic of . There is an identification (see [Pol20a, Lemma 4.3.1]), and is the similitude character of via this identification. I.e., for , satisfies for all and Freudenthal’s symplectic form on .
Theorem 7.1.1.
Let the notation be as above. Let consist of the with . Suppose , the Levi of the Heisenberg parabolic of and is positive semidefinite. Set . There is a nonzero constant , independent of and independent of so that the integral
is equal to
Note that the case of Theorem 7.1.1 represents the function as a integral, and the cases of this theorem compute (by exchaning the order of integration and differentiation) the derivatives . We justify this exchange of integration and differentiation in subsection 7.6.
Corollary 7.1.2.
Suppose is in the Levi of the Heisenberg parabolic of , and . Then
for a nonzero constant , independent of and .
Proof.
Suppose . First consider the case . We must simplify the quantity
If with and , then
Consequently, if , the Heisenberg Levi on , and , then , so
In general, if , with and , one finds that
is equal to
if .
The cancels the from Theorem 7.1.1, giving the corollary. ∎
Proof of Theorem 5.0.1.
Both sides of the desired equality transform on the left under in the same way, and on the right under in the same way. Moreover, for , they are both known to be quaternionic functions. Thus to prove their equality, it suffices to pair against , and evaluate on in the Levi of the Heisenberg parabolic. But this is precisely what is done in Corollary 7.1.2, so the theorem is proved. ∎
7.2. Some derivatives
We now focus on proving Theorem 7.1.1. We must consider some derivatives of the function
Suppose . Then , and thus . Consequently, the function can be considered constant for the purposes of differentiating with respect to . We therefore must just differentiate the function
We obtain
We write and and Thus,
Now suppose , and is such that is isotropic and singular. Because it is isotropic, . Because it is singular, consists of rank one elements of . Recall that is rank one means for all , where is the bilinear form proportional to the Killing form defined in [Pol20a, section 4.2.2]. Thus, by symmetrizing and using that commute, one arrives at . Thus, . Using this, we differentiate again to obtain
For , define to be the symmetric sum of terms of the form
Then we have the following proposition.
Proposition 7.2.1.
Suppose are such that
-
(1)
for all with
-
(2)
singular and isotropic.
Then
Proof.
We proceed by induction, noting that the proposition is true for and as checked above. Note that
Thus, making the induction assumption, is equal to
plus
But now note that
The proposition follows. ∎
The next step is to calculate the and , then integrate.
7.3. Some Lie algebra calculations
We set and .
Then if , , so where and .
Now recall the long root triple of given by
-
•
,
-
•
and
-
•
.
For we have . Note that the exceptional Cayley transform [Pol20a, section 5] takes to . We obtain
where444This is slightly different than how is defined in [Pol19]. .
We now compute . We have
-
(1)
, which under goes to .
-
(2)
, which goes to
-
(3)
, which goes to
-
(4)
, which goes to .
Denote by the projection from . Let’s suppose . Write also with . Now
Here and .
We now compute
Similarly we compute
We therefore obtain
We now compute . We assume , . Then for , we have
Thus
We summarize what we’ve proved in a proposition.
Proposition 7.3.1.
Suppose , . Then one has
-
(1)
;
-
(2)
;
-
(3)
;
-
(4)
.
Here , , and .
Rewriting the above, we have
-
(1)
-
(2)
-
(3)
-
(4)
-
(5)
-
(6)
-
(7)
7.4. More calculations
Putting together the work we have done above, we obtain the following lemma. Define , , and to be the quantity one obtains by replacing every with a and every with a .
Lemma 7.4.1.
Suppose . Consider
This quantity is
(Interpret this formula to mean that it is if .)
Proof.
We evaluate . The point is that every nonzero term that goes into the definition of this sum is the same. One obtains that any such term is equal to
There are
such nonzero terms. Thus
and the lemma follows. ∎
Putting everything together, we obtain the following proposition. Set
Proposition 7.4.2.
One has
For a positive real number set
where
Set . We have proved:
Proposition 7.4.3.
Set . One has
The following proposition now implies Theorem 7.1.1.
Proposition 7.4.4.
One has
7.5. Evaluation of an integral
In this subsection, we prove Proposition 7.4.4.
For , from equation (5) of [Wei], one obtains
Set, for , ,
Changing variables, one obtains
where . Differentiating under the integral sign, and using that
we get
Finally, making the variable change , one gets
Set
Lemma 7.5.1.
Suppose . One has
Proof.
The proof is by induction, using the recurrence . See also [oIS]. ∎
7.6. Technical justification
We still must justify our differentiation under the integral. In other words, we must justify the identity
Here the are in and differentiate with respect to the variable.
Write if under the identification . To do this justification, it suffices to show (for all non-negative integers ) that there is a small neighborhood of , and a positive function on , so that for all ,
The function can depend upon and , but we drop them from the notation.
It is easy to see (e.g., by induction) that the derivative is of the form
where consists of sums of products of terms of the form , etc. The key point is that, if , then when written as a polynomial in , has degree at most . The coefficients of this polynomial depend on and , but are easily seen to be bounded for in a small compact set around .
To finish the proof, we now must bound
Here is the -equivariant norm on , where is the Cartan involution. Write in the Iwasawa decomposition. We take small open neighborhoods around and , and let be the product of these neighborhoods. Then, if , is bounded away from for small, independent of , and is bounded below by for large, with independent of . The existence of with the desired properties follows.
8. Arithmeticity of modular forms on
The purpose of this section is to prove Theorem 1.0.1, which we recall here. Suppose is a cuspidal quaternionic modular form on of weight . Then
is its Fourier expansion. The locally constant functions are its Fourier coefficients. We say that has Fourier coefficients in a ring if for all characters of and all . We write for the space cuspidal quaternionic modular forms on of weight with Fourier coefficients in .
Let be the cyclotomic extension of .
Theorem 8.0.1.
Suppose is even. Then there is a basis of the cuspidal quaternionic modular forms of weight with all Fourier coefficients in . In other words, .
The proof of this theorem has the following steps:
-
(1)
Set the subspace of consisting of theta lifts from algebraic modular forms on . It is clear that it is a submodule. Moreover, as an application of Theorem 5.0.1, it is easy to see that is defined over , i.e., that it has a basis consisting of elements with Fourier coefficients in .
- (2)
-
(3)
For the previous step to go through, a certain archimedean Zeta integral must be shown to be non-vanishing. One shows the non-vanishing of this integral by a global method, using Corollary 1.2.3.
We will break the proof into various lemmas.
Lemma 8.0.2.
Suppose is with and the spherical vector. Let be the algebraic modular forms on of level and for the representation . Then there is a lattice so that if , then has Fourier coefficients in .
Proof.
Write
(4) |
Now
where . In other words,
Thus the integral over is a finite sum of terms where is rational. But now Theorem 5.0.1 implies that if the Fourier coefficients of are in some ring , and is in an appropriate lattice, then the Fourier coefficients of are in . Thus we obtain the fact that the Fourier coefficients of theta lifts can all be made in , as soon as we prove the same result for the Fourier coefficients of .
For the latter, simply observe the following identities: suppose . Then we can write with in the , , and . Then we can further write with and . This follows from strong approximation on the simply connected group . Thus, if denotes a Fourier coefficient of (level one), then
Additionally,
This last term is Thus since all the are integral, all . It thus follows that if with , then all the Fourier coefficients of are in . This completes the argument. ∎
Write for the space of all lifts as in equation (4).
Lemma 8.0.3.
The subspace is a -submodule.
Proof.
This is clear. ∎
Recall the projection Let be the set of all for which for all . We set , and let be the composite projection.
To prove that it suffices to show that if generates an irreducible representation , then
(5) |
for some . Indeed, in this case, the submodule has orthocomplement equal to . Moreover, we can assume that is a pure tensor in .
Suppose is a totally real cubic étale extension of , and let be the group of type defined in terms of that has compact. See [Pol23] for a precise definition. To prove (5), it then further suffices to show that
(6) |
for some such .
The integral in (6) can be evaluated using the main theorem of [Pol23] and the Rankin-Selberg integral studied in [GS15, Seg17] (see also [Pol19]). To set up the result, following [Pol23], write for a certain simply connected group of absolute Dynkin type , defined in terms of and split over .
We have
for elements . Thus by Corollary 9.4.8 of [Pol23],
(7) |
Here is the Siegel-Weil Eisenstein series on the group .
The integral of (7) can now be written as a partial -function times some local Zeta integrals at bad finite places (including the archimedean place). Specifically, we have the following proposition. Moreover, these local zeta integrals at the finite places can be trivialized with Siegel-Weil inducing data for the Eisenstein series on . Specifically, we have the following proposition.
Suppose is a unitary character. To setup the proposition, we define
This is the local archimedean Zeta integral that comes from the Rankin-Selberg integral (7). The notation is from [Pol19, Theorem 5.2], which is a restatement of a Theorem of [GS15, Seg17]. Here also is the Siegel-Weil inducing section from [Pol23]. It follows from Proposition 8.0.8 below that converges absolutely for .
Proposition 8.0.4.
Let be a unitary character of for which . Recall that from the theory of binary cubic forms one associates to is a rank three -module in a cubic étale algebra over . Suppose is a set of finite places of that satisfies the following conditions:
-
(1)
If , then is unramified and is spherical at ;
-
(2)
If , then is a ring, and in fact the maximal order of ;
-
(3)
.
Then the finite vector can be chosen so that is spherical outside , and the integral (7) is equal to . Moreover, if is unramified at all finite primes, and in , then may be chosen to be empty.
Observe that if is level one and has Fourier coefficient nonzero, then we may take to be empty.
Proof.
The fact that the global integral represents the partial -function is from [GS15, Seg17], for a slightly larger set . In [Pol19] the set is shrunk to that in the statement of the proposition, except that [Pol19] includes in all cases. Then, in [cDD+22], the case where is level one and is is handled.
That the bad local integrals may be trivialized with some data is in [Seg17, Section 7]. What we state and use is slightly stronger. Specifically, we must verify that the bad local integrals can be trivialized for Siegel-Weil inducing sections. This follows simply because the Siegel-Weil inducing sections make up all of the induced representation , where is the Heisenberg parabolic of . To see this, recall that is generated by any vector which is not annihilated by the long intertwining operator, which turns out to be given by an absolutely convergent integral. The restriction of the spherical vector from is positively valued on , so it cannot be annihilated by the long intertwining operator.
It turns out the -value is always nonzero. This is a direct consequence of the main theorem of [Mui97].
Theorem 8.0.5.
Let be a cuspidal automorphic representation of over . Then the Euler product defining the partial standard -function converges absolutely for .
Proof.
The local factors are unitarizable. In [Mui97], the unitary dual of -adic is completely and explicitly determined. In particular, when is spherical, one has tight bounds on the Satake parameters of . These bounds imply the absolute convergence statement of the theorem. ∎
Finally, it turns out that if is even, the archimedean Zeta integral is nonzero for all non-degenerate .
Proposition 8.0.6.
Suppose is even. Then is nonzero for all non-degenerate .
Proof.
A change of variables in the integral shows that the non-vanishing of is equivalent for all . So, we take with . We will prove that this integral is nonvanishing by a global argument, using Proposition 8.0.4, Corollary 1.2.3, and Theorem 8.0.5. Fix now .
Let be as in the proof of Corollary 1.2.3. Set . Then one sees that has Fourier coefficient equal to , so does as well.
We can write as a finite sum of level one cuspidal eigenforms forms , with . Thus there is some such with Fourier coefficient nonzero; fix this .
We have . So
Let . Then
where is some finite group. Here we are using that .
But this latter integral is nonzero, because it is nonzero after pairing with . Consequently, we have deduced that is nonzero. ∎
Proof of Theorem 8.0.1.
We have proved that the space of lifts has a structure, from the Fourier coefficients. We have also prove that if is even, then . This proves the theorem.∎
We end with some of the technical details that were used in the proofs above.
Lemma 8.0.7.
Let denote the space of the representation , and suppose is an functional. Given , there is a Schwartz-Bruhat function on so that
is equal to , independent of .
Proof.
Write in the notation of [Pol19]. We have
The function is on , and we have . In this decomposition, we can write , where are a basis of , the trace elements of .
We take a pure tensor, . We make be the characteristic function of a set very close to . Let be the center of the Levi of the Heisenberg parabolic on . Then implies for some . Here is the elements of that are modulo in the matrix representation of . Here we are using that if , with and , with , then there is so that and . (Indeed, this latter fact can be proved by using and also unipotent elements in .)
Thus we must evaluate
We choose to be a pure tensor in our root basis of , so that . But . By choosing so that its Fourier transform is supported near , we see that we can trivialize the integral to a constant multiple of . This proves the lemma. ∎
We now prove the absolute convergence of the archimedean Zeta integral. The integral in question is
(8) |
Proposition 8.0.8.
The integral (8) converges absolutely for .
Proof.
Let be a Sschwartz function on . We obtain an inducing section from as
We will check that every inducing section in is of this form, and we will prove the proposition for these inducing sections.
For the first part, observe that if , then restricting to we obtain for all . These are the for which , and note that the negative sign does indeed occur. Now let be an arbitrary even smooth function on and a smooth compactly supported function on . We set ; this is a Scwhartz function.
We have now
We have used the evenness of . But because gives an arbitrary function on , we see that every inducing section in is an .
We thus now proceed to prove the absolute convergence of the double integral
when .
The bound we use below on is independent of , so it suffices to integrate over . Without loss of generality, we can assume that is a pure tensor of the appropriate sort. We thus must bound the integral
Here is a Schwartz function on the matrices , and the rest of the notation is as in Lemma 8.0.7.
One has and is bounded above, so we must bound
Make the variable change , and set . Then we must bound
Now we claim is, for (assumed real now) bounded by . Indeed, is rapidly decreasing, so for an sufficiently large of our choosing. Thus
Dropping terms of the form (since we are fixing ), both integrals above are bounded by . Thus we must bound
We break this integral into two pieces, one where and the other with . The first integral has a compact domain, so can be ignored. To show the convergence of the second integral, it suffices to show that , where is a rapidly decreasing function. And since the -Bessel function is rapidly decreasing, it suffices to show that for all .
Both sides of the desired inequality scale linearly with the center of , so it suffices to check that
is bounded away from for . Furthermore, by a change of variables again (now or in the initial integral), we can assume . Then if
we wish to bound below the quantity
where .
Finally, to see that this rational function in is bounded below for , we work in polar coordinates. We have
If then so so and then , so the quantity is at least . If , then
because is increasing on . This proves the claim, and thus the proposition. ∎
References
- [Asc87] Michael Aschbacher, The -dimensional module for . I, Invent. Math. 89 (1987), no. 1, 159–195. MR 892190
- [BCFvdG17] Jonas Bergström, Fabien Cléry, Carel Faber, and Gerard van der Geer, Siegel modular forms of degree two and three, http://smf.compositio.nl, Retrieved October 6, 2022.
- [BD21] Siegfried Böcherer and Soumya Das, On fundamental fourier coefficients of siegel modular forms, Journal of the Institute of Mathematics of Jussieu (2021), 1–41.
- [BGW15] Manjul Bhargava, Benedict H. Gross, and Xiaoheng Wang, Arithmetic invariant theory II: Pure inner forms and obstructions to the existence of orbits, Representations of reductive groups, Progr. Math., vol. 312, Birkhäuser/Springer, Cham, 2015, pp. 139–171. MR 3495795
- [cDD+22] Fatma Çiçek, Giuliana Davidoff, Sarah Dijols, Trajan Hammonds, Aaron Pollack, and Manami Roy, The completed standard -function of modular forms on , Math. Z. 302 (2022), no. 1, 483–517. MR 4462682
- [Cle22] Jean-Louis Clerc, Construction à la Ibukiyama of symmetry breaking differential operators, J. Math. Sci. Univ. Tokyo 29 (2022), no. 1, 51–88. MR 4414247
- [Cox46] H. S. M. Coxeter, Integral Cayley numbers, Duke Math. J. 13 (1946), 561–578. MR 19111
- [CT20] Gaëtan Chenevier and Olivier Taïbi, Discrete series multiplicities for classical groups over and level 1 algebraic cusp forms, Publ. Math. Inst. Hautes Études Sci. 131 (2020), 261–323, https://otaibi.perso.math.cnrs.fr/levelone/siegel/siegel_genus3.txt. MR 4106796
- [Dal21] Rahul Dalal, Counting Discrete, Level-, Quaternionic Automorphic Representations on , arXiv e-prints (2021), arXiv:2106.09313.
- [EG96] Noam D. Elkies and Benedict H. Gross, The exceptional cone and the Leech lattice, Internat. Math. Res. Notices (1996), no. 14, 665–698. MR 1411589
- [Gan00] Wee Teck Gan, A Siegel-Weil formula for exceptional groups, J. Reine Angew. Math. 528 (2000), 149–181. MR 1801660
- [Gan22] by same author, Automorphic forms and the theta correspondence, Arizona Winter School notes (2022), https://swc-math.github.io/aws/2022/index.html.
- [GGS02] Wee Teck Gan, Benedict Gross, and Gordan Savin, Fourier coefficients of modular forms on , Duke Math. J. 115 (2002), no. 1, 105–169. MR 1932327
- [Gro96] Benedict H. Gross, Groups over , Invent. Math. 124 (1996), no. 1-3, 263–279. MR 1369418
- [GS98] Benedict H. Gross and Gordan Savin, Motives with Galois group of type : an exceptional theta-correspondence, Compositio Math. 114 (1998), no. 2, 153–217. MR 1661756
- [GS15] Nadya Gurevich and Avner Segal, The Rankin-Selberg integral with a non-unique model for the standard -function of , J. Inst. Math. Jussieu 14 (2015), no. 1, 149–184. MR 3284482
- [GS21] Wee Teck Gan and Gordan Savin, Howe duality and dichotomy for exceptional theta correspondences, 2021.
- [GS22] by same author, The local langlands conjecture for , 2022.
- [GW94] Benedict H. Gross and Nolan R. Wallach, A distinguished family of unitary representations for the exceptional groups of real rank , Lie theory and geometry, Progr. Math., vol. 123, Birkhäuser Boston, Boston, MA, 1994, pp. 289–304. MR 1327538
- [GW96] by same author, On quaternionic discrete series representations, and their continuations, J. Reine Angew. Math. 481 (1996), 73–123. MR 1421947
- [HPS96] Jing-Song Huang, Pavle Pandžić, and Gordan Savin, New dual pair correspondences, Duke Math. J. 82 (1996), no. 2, 447–471. MR 1387237
- [Ibu99] Tomoyoshi Ibukiyama, On differential operators on automorphic forms and invariant pluri-harmonic polynomials, Comment. Math. Univ. St. Paul. 48 (1999), no. 1, 103–118. MR 1684769
- [Ibu02] by same author, Vanishing of vector valued siegel modular forms, Unpublished note (2002).
- [Kim93] Henry H. Kim, Exceptional modular form of weight on an exceptional domain contained in , Rev. Mat. Iberoamericana 9 (1993), no. 1, 139–200. MR 1216126
- [KZ81] W. Kohnen and D. Zagier, Values of -series of modular forms at the center of the critical strip, Invent. Math. 64 (1981), no. 2, 175–198. MR 629468
- [MP22] Finley McGlade and Aaron Pollack, Quaternionic functions and applications, Work in progress (2022).
- [MS97] K. Magaard and G. Savin, Exceptional -correspondences. I, Compositio Math. 107 (1997), no. 1, 89–123. MR 1457344
- [Mui97] Goran Muić, The unitary dual of -adic , Duke Math. J. 90 (1997), no. 3, 465–493. MR 1480543
- [oIS] The On-Line Encyclopedia of Integer Sequences, A100861: Triangle of bessel numbers read by rows, https://oeis.org/A100861.
- [Pol18] Aaron Pollack, Lifting laws and arithmetic invariant theory, Camb. J. Math. 6 (2018), no. 4, 347–449. MR 3870360
- [Pol19] by same author, Modular forms on and their standard -function, Proceedings of the Simons Symposium ”Relative Trace Formulas” (accepted) (2019).
- [Pol20a] by same author, The Fourier expansion of modular forms on quaternionic exceptional groups, Duke Math. J. 169 (2020), no. 7, 1209–1280. MR 4094735
- [Pol20b] by same author, The minimal modular form on quaternionic , Jour. Inst. Math. Juss. (accepted) (2020).
- [Pol21] by same author, Exceptional algebraic structures and applications, Notes from a topics course at UCSD (2021), https://www.math.ucsd.edu/~apollack/course_notes.pdf.
- [Pol22] by same author, Modular forms on exceptional groups, Arizona Winter School notes (2022), https://swc-math.github.io/aws/2022/index.html.
- [Pol23] by same author, Exceptional siegel weil theorems for compact , Preprint (2023).
- [PV21] Kartik Prasanna and Akshay Venkatesh, Automorphic cohomology, motivic cohomology, and the adjoint -function, Astérisque (2021), no. 428, viii+132. MR 4372499
- [Seg17] Avner Segal, A family of new-way integrals for the standard -function of cuspidal representations of the exceptional group of type , Int. Math. Res. Not. IMRN (2017), no. 7, 2014–2099.
- [Wal81] J.-L. Waldspurger, Sur les coefficients de Fourier des formes modulaires de poids demi-entier, J. Math. Pures Appl. (9) 60 (1981), no. 4, 375–484. MR 646366
- [Wal03] Nolan R. Wallach, Generalized Whittaker vectors for holomorphic and quaternionic representations, Comment. Math. Helv. 78 (2003), no. 2, 266–307. MR 1988198
- [Wei] Eric W. Weisstein, Modified bessel function of the second kind, From MathWorld–A Wolfram Web Resource. https://mathworld.wolfram.com/ModifiedBesselFunctionoftheSecondKind.html.