On the Resolution of Reductive Monoids and Multiplicativity of -Factors
Abstract.
In this article, we give a proof of multiplicativity for -factors, an equality of parabolically induced and inducing factors, in the context of the Braverman-Kazhdan/Ngo program, under the assumption of commutativity of the corresponding Fourier transforms and a certain generalized Harish-Chandra transform. Within our proof, we define a suitable space of Schwartz functions which we prove to contain the basic function. We also discuss the resolution of singularities and their rationality for reductive monoids, which are among the basic objects in the program.
2010 Mathematics Subject Classification:
11F70, 11F66, and 14E15Introduction
Every theory of -functions must satisfy the axiom of multiplicativity/inductivity, which simply requires that -factors for induced representations are equal to those of the inducing representations. This axiom is a theorem for Artin -functions and those obtained from the Langlands-Shahidi method [Sha10], and is a main tool in computing -factors, root numbers, and -functions. On the other hand, its proof in the cases obtained from Rankin-Selberg methods are quite involved and complicated. It is also central in proving equality of these factors when they are defined by different methods and in establishing the local Langlands correspondence (LLC) [Sha12, Sha17, HT01, Hen00, GT11, CST17] . Its importance as a technical tool in proving certain cases of functoriality [CKPSS04, Kim03, KS02] is now well established.
In this paper we will provide a proof of multiplicativity for -factors defined by the method of Braverman-Kazhdan/Ngo [BK02, BK10, BNS16, Ngô20] and L. Lafforgue [Laf14] in general under the assumption that the -Fourier transforms on the group and the inducing Levi subgroup commute with the -Harish-Chandra transform, a generalized Satake transform sending , where is a finite dimensional representation of the -group of by means of which the -factors are defined.
Within our proof, we define a space of -Schwartz functions for every as
(0.1) |
This definition is crucial since the -Schwartz functions defined in this way will be uniformly bi--finite (see equation (5.16) and Lemma ), making the descent to the inducing level possible, an important step in the proof of multiplicativity. While the -factor can be defined as the kernel of the Fourier transform, it is the full functional equation that allows our descent to the inducing level in a transparent fashion, using our definition of -Schwartz functions.
In [BK10], Braverman and Kazhdan defined their Schwartz space as a ”saturation” of ours. But our Schwartz space, which is denoted by in [BK10], covers a significant part of theirs and in particular, contains the -basic function as we prove in Proposition This is done using the extended Satake transform to almost compact functions [Li17] and the fact that it commutes with the Fourier transform induced from tori which is now defined in general, cf. Section and in partiular diagram .
The commutativity assumption allows us to extend the -Harish-Chandra transform to , commuting with and , respectively, where is the restriction of to the -group of . This construction of agrees with that of Braverman-Kazhdan in the case of doubling method [BK02, GPSR87, Li18, LR05, PSR86, Sha18, JLZ20, GL20], since being the interior of the defining monoid embeds as a unique open orbit into the Braverman-Kazhdan space (cf. [Li18]). Our proof is a generalization of Godement-Jacquet for , Theorem 3.4 of [GJ72].
One expects , where the latter is defined as the space of functions of compact support on , the monoid attached to whose restriction to are smooth (locally constant) and sending to itself inside . The group being smooth as a variety, the singularities of the monoid are outside of . Renner’s construction (Section 2) realizes as a closed subvariety for any Levi subgroup that contains , where is a fixed maximal torus of , used in the construction of .
The -Harish-Chandra transform cannot be defined on since the modulus character may vanish outside , but it can be defined on the image of inside , that lands in the image of in by the above discussion.
Our commutativity axiom, which implies multiplicativity and multiplicativity itself give rise to an inductive scheme that allows for a definition of Fourier transform by building from the case of conjugacy classes of Levi subgroups of . In fact, Theorem 5.3 gives the factors , an irreducible constituent of , equal to the inducing -factor , which in turn is defined through convolution by . For example, for , the Levi subgroups consist of split tori for which a canonical Fourier transform exists (c.f. [Ngô20]; see (6.2) here) and itself, which is equivalent to understanding supercuspidal -factors. We refer to section 5.4 for a more detailed discussion of this inductive construction.
In the case of , Laurent Lafforgue [Laf14] has defined a candidate distrubution which is shown formally to commute with the Harish-Chandra transform and evidence exists that it may give the correct supercuspial factors as observed by Jacquet, but it is still unknown if this is the right distribution. Work in this direction for tamely ramified representations is being pursued by the second author.
Although our definition of the space depends on the knowledge of how acts on , this seems to be the most efficient way of defining at present and sufficient for our purposes as a working definition, allowing us to begin making some initial steps toward understanding the general theory, and as observed earlier after equation (0.1), essential in proving the uniform -finiteness of -Schwartz functions.
One hopes that the geometry of will provide some insight into what this Fourier transform ought to be. In fact, the geometric techniques used to study the basic functions on reductive monoids via arc spaces in the function field setting [BNS16] tells us that the nature of the singularities of the monoid very much controls the asymptotics of the basic function. Taking cue from this, it is natural to consider the geometry of the singularities in the -adic case as well. As a first step, we may classify the singularities of our monoids via the theory of spherical varieties and we find that there is a good and explicit choice of -equivariant resolution of singularities [Bri89, Per14]. The resolution is moreover rational and so we may pass without trouble between differential forms on the monoid and its resolution. The geometric aspects of this theory are discussed in part in Section 3 of the present paper. Since our Schwartz spaces are, at least tentatively, linked by the definition of the Fourier transform via , we are able, at least speculatively, to unite the themes of this paper. Here is the outline of the paper.
Section 1 is a quick review of the method for as developed in [GJ72]. Renner’s construction of reductive monoids is briefly discussed in Section 2 which concludes with a treatment of the cases of symmetric powers for , describing all the objects involved in those cases. Section covers the geometric aspects studied in the paper. This includes the resolution of the singularities of reductive monoids, leading to a proof of rationality of these singularities. This allows a transfer of measures from the monoid to its resolution as discussed in Section and can be applied to the integration of basic functions on corresponding toric varieties in Example . Multiplicativity is stated and proved in Section , concluding with the example of in and a discussion of the inductive nature of Fourier transforms in . In proving multiplicativity, we have found it easier to work with the full functional equation rather than the definition given by convolutions. The cases of a tori and unramified data are addressed in Section . The paper is concluded with a brief discussion of the doubling construction of Piatetski-Shapiro and Rallis with relevant references cited.
Acknowledgements
The authors would like to thank J. Getz, D. Jiang, and B.C. Ngo for helpful conversations. A part of this paper was presented by the first author during the month long program “On the Langlands Program: Endoscopy and Beyond” at the Institute for Mathematical Sciences, National University of Singapore, December 17, 2018-January 18, 2019. He would like to thank the Institute and the organizers: W. Casselman, P.-H. Chaudouard, W.T. Gan, D. Jiang, L. Zhang, and C. Zhu, for their invitation and hospitality. Finally, we would like to thank Jayce Getz, Chun-Hsien Hsu, and Michel Brion for their comments after the paper was posted on arXiv. Both authors were partially supported by NSF grants DMS 1801273 and DMS 2135021
1. The case of standard representation for
We recall that the Godeement–Jacquet [GJ72] theory for standard -functions of , which this method aims to generalize, can be presented briefly through the definition of corresponding -factors.
Let be a -adic field and . Let be an irreducible admissible representation of . Given a Schwartz function on , i.e., , a smooth function of compact support on , one can define a zeta-function
where and , is a matrix coefficient and . Here is the contragredient of . Let
be the Fourier transform of with respect to the (additive) character of .
If , , then we can consider . The Godement–Jacquet theory defines a -factor which depends only on and and is a rational function of , satisfying
(1.1) |
for all and .
It is not hard to see that if we introduce the Int-invariant kernel, ,
of the Fourier transform, then
(1.2) |
by virtu of irreducibility of and the Schur’s lemma.
This formulation for the -factor is a quick and convenient way of introducing them which is amenable to generalization. We can therefore write
pointing to the significance of the kernel in defining the -factors.
2. The general case; monoids and Renner’s construction
To treat the general case we need to generalize . Let be an algebraically closed field of characteristic zero. A monoid is an affine algebraic variety over with an associative multiplication and an identity 1. For our purposes, we also need to be normal, i.e., is integrally closed in . We can always find a normalization in case is not normal, i.e., an epimorphism such that integral closure of in equals as we realize .
We thus let be a normal monoid and let , be the units of . We say is reductive if is. We now like to attach a monoid to a finite dimensional representation of , -group of , , where is a split reductive group. Let be a maximal torus and write
where is the set of weights of . Let be the set of cocharacters of or characters of and set . Next, denote by the convex span in of weights of and let be the cone in generated by rays through .
Let , be the “rational” dual cone to and the group algebra of . One can then identify as a subset of by going to , defind by
, and , where the sum is the one on the semigroup . We note that this is valid for any semigroup and
Now, assume has a character
such that
sends to . This means that for any weight of . In fact, for ,
and thus . Then and its existence implies that is strictly convex, i.e., has no lines in it. In fact, the cone is contained in the open half–space of vectors , , , satisfying . It is therefore strictly convex. (cf. [Ngô20],Proposition 5.1).
By the theory of toric varieties [CLS11], the strictly convex cone determines (uniquely) a normal toric variety, i.e., a normal affine torus embedding . Here is the monoid for attached to . More precisely, by Theorem 1.3.8, pg. 39 of [CLS11]. By definition 3.19 of [Ren05], is generated by and thus , the semigroup defining . The embedding , defines , a morphism of semigroups, into the character group of .
The dominant characters in all lie in and are those that extend to semigroup morphism (Proposition 3.20 of [Ren05]).
Finally we observe that is integral and dominant and thus .
To proceed, we remark that the Weyl group acts on and in the usual manner. Thus the dual rational cone may be identified with , both semigroups, since its group algebra generated by elements of or , is as we discussed earlier.
Let be a dominant (and integral) character. Then defines an irreducible finite dimensional (rational) representation of , , of highest weight . Since
we can extend to an irreducible rational representation of
Definition 2.1.
is called the irreducible representation of of highest weight , where is a dominant rational character of .
This in particular is valid for dominant elements in . We note that is one such.
Now choose so that generates . Let be the representation attached to . Set and . The character will be among these . We may assume . Define . We let be a normalization of .
We note that we may take to be among the fundamental weights of , with extending the trivial representation of since is a representation (character) of .
2.1. The case of symmetric powers of
As an example in this section we consider the symmetric power representations of and describe these objects in this case.
Let and , the -th symmestric power of the standard representation of . Write with the basis . Let denote the weights of . Then we can order them as
. We have
inside which equals
The dual cone to is
Thus the dual to is the of . It is a lattice in the shaded area, corresponding to
We use , and to denote and in the semigroup algebra as before, i.e., and so on, then
The corresponding toric variety is
the variety defined by the zeros of , and
The monoid for (Renner’s construction): The dual cone in
is generated by and . The vectors and are -conjugate and therefore we have as our dominant weights . They correspond, respectively, to std, the standard representation, and (to be explained) and thus
Then
The character for :
Recall that the fibered product of and giving the units of the monoid for is (cf. [Sha17])
We then have the commuting diagram
Thus
and the left vertical arrow, the Proj2, gives
for which
3. Some geometry of reductive monoids as spherical varieties
Renner’s classification of reductive monoids uses the “extension principle” [Ren05]. The extension principle follows in the spirit of many similar classification results for spherical varieties that rely on the existence of an open -orbit where is a Borel subgroup of , that is in Renner’s case adapted to account for the monoid structure. By Renner’s classification, the category of Reductive monoids is equivalent to the category of tuples , where is any maximal torus in and is a Weyl-group stable toric variety. A morphism of data are given by a pair where is a morphism of reductive groups and a morphism of toric varieties such that the restriction of each morphism to the maximal torus agree . In the following, we reframe these results in terms of the theory of spherical varieties, in order to state the existence of a -equivariant resolution of singularities [BK07], [Rit03].
Let be a split reductive group defined over a characteristic zero field . Let be a variety defined over with an rational action . In this case we say is a -variety. Let be the sheaf of regular on . If is affine, we will identify with the coordinate algebra . In this case induces as usual a co-action map by with , where the latter is a finite sum. Thus each determines a finite dimensional -module. Because is reductive and we are in characteristic , each finite dimensional -module decomposes as a finite sum of irreducible representations indexed by their highest weight vector with weight . As such we may decompose , indexed by the that appear in .
Definition 3.1.
A -variety is spherical if has an open -orbit for some (hence any) Borel in .
Suppose is spherical. Then as above, by highest weight theory, each dominant integral character of that appears in has a highest weight vector . The line is the unique line stabilized by on which acts through the character . In other words is a semi-invariant. Suppose and are semi-invariants that are -eigenfunctions appearing in . Then the rational function is -invariant. As the -orbit in is dense, this implies is constant. Hence for spherical varieties, each that appears can only appear with multiplicity one. For general reductive groups actions, even the “naive” (categorical) quotient is reasonably well behaved.
Theorem 3.2.
Let be an affine -variety for a reductive group . Then the ring of -invariants is finitely generated, say . Then defines a good surjective quotient which is moreoever a categorical quotient . Each fiber of contains a unique closed -orbit in , and is normal if is.
Definition 3.3.
A spherical variety is simple if it has a unique closed -orbit.
We are interested in reductive monoids, which have open -orbit and are spherical with respect to with an open dense borel orbit.
Proposition 3.4.
Suppose has an open -orbit. Then has a unique closed orbit.
Proof.
The reductive quotient is constant on orbits, in particular on the open orbit. Hence . The fiber contains a unique closed orbit by Theorem 3.2 ∎
Therefore such are simple. Once again let us consider an affine simple as a variety. Decomposing where is the highest weight module for . The eigenfunction is -invariant. Thus one may consider taking -invariants are therefore generated as a vector space by the . Using the following
Theorem 3.5.
Let be reductive with maximal unipotent subgroup , and let be a -variety. Then is finitely generated. Moreover is normal if is.
Proof.
One first establishes the theorem for i.e. is finitely generated and in fact is a geometric quotient (a so called horospherical variety). One has a map where the quotient is by the diagonal action. On coordinate rings: a -invariant defines a -invariant functions . Thus by Theorem 1.2 is finitely generated. ∎
We can conclude that the variety is a variety, on which acts on through the character . In other words, we have a ring graded by . By Theorem , this is a finitely generated monoid. Each summand has a diagonalizable action by the torus , giving an equivalent characterization of toric varieties: hence defines an affine embedding. If is a reductive monoid, this must therefore be equal to Renner’s cone. Moreover, if is normal the associated toric variety is normal, hence the cone of weights of defining the toric variety is saturated.
Remark 3.6.
Recall that a -variety is simple if it contains a unique closed orbit. When is affine, it is enough that embeds as an open subvariety to imply is simple. We state without proof the following:
Proposition 3.7.
Any -variety can be covered by simple -varieties.
To classify a general spherical variety one needs the following additional data.
Definition 3.8.
Let denote the -stable discrete valuations on .
Definition 3.9.
Let denote the set of -stable prime divisors of .
Definition 3.10.
Let be a -orbit in . Then is the set of -stable prime divisors containing .
Definition 3.11.
Let denote the set of irreducible -stable divisors.
Proposition 3.12.
For a divisor , either
-
(1)
The -orbit is open and dense in the open orbit of .
-
(2)
is -stable.
Briefly (although see [Kno91] for details) a simple spherical variety is determined by its weight monoid, plus the data of which -stable boundary divisors containing the unique closed orbit are -stable and which are not. More precisely, each divisor defines a valuation by first restricting which defines a valuation on the multiplicative group of rational functions on (the valuation is the order of vanishing of a rational function on ). This defines a so-called colored cone, defined by the span of the finite number of valuations as above in which the colors are the valuations that are -stable but not -stable (or dually, their corresponding divisors have a -open-orbit).
Thus the set gives the set of colors of the simple spherical variety . The cone generated by and the natural image of in the set of valuations is the colored cone determined by the data that determines up to isomorphism the spherical variety . For reductive monoids, this cone is equivalent to the one constructed in the earlier section via highest weight theory.
Example 3.13.
For a reductive monoid , there is a beautiful description of the boundary in terms of -stable boundary divisors in the form of an extended Bruhat decomposition: Let be the Zariski closure of the normalizer of a maximal torus in . Let be the set of idempotents in , and note that reductive monoids are regular (in Renner’s sense), that is, we can decompose . Then we can construct the Renner monoid (sometimes called the Rook monoid) . Because reductive monoids are regular, makes sense as a finite monoid whose unit group is the Weyl group , and having the property that
From this description, it may be deduced that the set , with the unique closed orbit in , is given by the codimension one orbits for the simple reflection in the Weyl group determined by the simple root .
Thus, for a monoid with reductive group embedded as its unit group, the colors of are all the stable irreducible divisors of , and thus the monoid is determined purely by the data or equivalently . We state for convenience this form of the classification.
Theorem 3.14.
Let be a reductive group. The irreducible, normal algebraic monoids with unit group are the strictly convex polyhedral cones in generated by and a finite set of elements in .
The theory of reductive monoids affords us an explicit description of purely in monoid-theoretic terms.
Definition 3.15.
A spherical variety is toroidal if is empty.
Proposition 3.16.
Suppose the spherical variety is toroidal and let . Let be the stabilizer of . Then is a parabolic and moreover satisfies the local structure theorem, i.e. there is a Levi , depending only on and a closed -variety such that
is an isomorphism. Moreover, is a toric variety under L/[L,L].
As a consequence of the above isomorphism, the orbits of correspond to orbits in . Note that when is a reductive monoid, this is precisely Renner’s extension theorem [Ren05] with and . The proposition implies that the singularities of are those determined by the cone of the toric variety .
Let us recall Renner’s extension theorem for normal reductive monoids. It states that a morphism of reductive monoids is given by the data and and a morphism is equivalent to and . Briefly, in one has an analogue of the open cell which is the image of an open embedding which has codimension in . Thus one gets an equivariant morphism . By normality of , the codimension condition extends the map uniquely to , and one verifies this is in fact a morphism of monoids.
Remark 3.17.
The above map in Renner uses an open Bruhat cell analog in the context of monoids, which yields a structure theorem parallel to Proposition . More generally, -equivariant dominant morphisms of spherical varieties are in bijection with linear maps , where and are the lattices of co-weights of the underlying group , such that the image of the colored fan is contained within .
In view of the above structure theory, the singularities of a spherical variety are determined by those of its associated toric variety . Smooth toric varieties are classified as follows.
Theorem 3.18.
An affine toric variety is smooth if and only if the extremal rays of the rational polyhedral cone generated by its weight lattice is a basis for for the character lattice .
As a general toric variety is glued from affine toric varities, it is given by a fan consisting of rational polyhedral cones. Thus a toric variety is smooth if and only if its fan consists of rational polyhedral cones whose extremal rays generate (as a -module) . Moreover any peicewise linear morphism of rational polyhedral cones defines a -equivariant morphism of toric varietie, thus, by Theorem , one obtains an algorithm giving a resolution of singularities of a toric variety .
Theorem 3.19.
Let the monoid of weights of . Starting from an extremal ray , successively subdividing the cone such that each resulting cone is smooth, defines a smooth fan consisting of smooth cones such that the inclusion map defines a canonical -equivariant resolution of singularities .
Finally, one may define a -equivariant resolution from Remark and Theorems and :
Theorem 3.20.
Let be a reductive monoid. Then there exists a smooth spherical -variety that is toroidal, and a proper -equivariant morphism
Proof.
Take the colored cone determed by . Deleting all colors, construct a fan generated by a subdivision of into smooth colorless cones (which always exists, see [CLS11]). Then is a resolution of toric varieties, and has an affine chart given by the open cell by Proposition , and therefore is smooth. The natural inclusion of into defines a dominant -equivariant morphism , giving us our resolution. ∎
Example 3.21.
The case of and monoids are determined by their respective toric varieties , which are realized finite quotients of by cyclic group of order . These are the well known finite quotient surface singularities of type and their resolution are give by blow-ups at the origin. Likewise, the resolution of is given by -blow-ups at of .
Next we discuss the rationality of singularities. Recall that has rational singularities if a (and hence any) resolution has vanishing higher cohomology: if . Moreover, recall that for a resolution the fiber over the singular locus is the exceptional divisor.
Theorem 3.22.
Let be a normal variety with rational singularities. Let denote the embedding of the smooth locus of into . If we define , by extending (uniquely, by normality) algebraic top-dimensional differentials form to . If is any resolution of singularities, then extends over to an algebraic differential form on .
Theorem 3.23.
Normal spherical varieties, and hence reductive monoids have rational singularities.
Proof.
is a flat deformation onto a toric variety, and toric varieties have rational singularities. Rationality “descends” under flat deformation (use flat base change, see [Elk78]), hence has rational singularities. ∎
4. Integration on singular varieties
Let be a top-differential form on a reductive monoid . Taking the canonical resolution as above we get a well defined differential form on . Restricing to the open set on over which is an isomorphism, we have upon passing to -points
where is the Jacobian of and is the measure constructed by Weil from the top-differential form on .
Example 4.1.
As a first basic example, we may consider the basic function on toric varieties. Recall that the basic function on a toric variety defined by cone with dual cone . Suppose the generators of the weight monoid of are . The are co-characters . For each co-character in the weight monoid, is an open neighbood on which the value of the basic function is . It is known [Stu95, Cas17] that the function
as an integer valued function of the weight monoid is quasi-polynomial, i.e., it is in the sub-algebra of functions on generated by polynomials and periodic functions. In particular, this means one may dominate the basic function with a polynomial function on the lattice.
It is also classical that locally in some coordinate system on , the map is given by a monomial transformation, i.e., has the form . For example, the toric variety is the cone and is resolved affine-locally by the classical “cylinder” resolution . Therefore the is in local coordinates a product of the form . For the cylinder resolution it is . Thus as the Jacobian grows as the inverse of an exponential while the basic function is dominated by polynomial growth. Hence the pullback of the differential form makes sense as on .
5. Multiplicativity
Every theory of -functions is expected to satisfy “multiplicativity”, i.e., the equality of -factors for parabolically inducing and induced data. Our goal in the rest of the paper is to establish, under natural assumptions on related Fourier transforms, the multiplicativity in the context of this theory in general. Our proof is a generalization of the standard case of Godement–Jacquet [GJ72]. We first need to connect Renner’s construction to parabolic induction.
5.1. Renner’s construction and parabolic induction
We start by observing that Renner’s construction [Ren05] respects parabolic induction. More precisely, let be a parabolic subgroup of , with unipotent radical and a Levi subgroup which we fix by assuming . Now is a reductive group with a maximal torus to which Renner’s construction applies. Let , where is the connected component of the -group of .
Now gives the same weights as does and thus shares the same toric variety coming from . Let . Then each orbit breaks up to a disjoint union of orbits and thus
where
(5.1) Conclusion: The monoid attached to by Renner’s construction for , , is the same as the closure of as a subgroup of upon action on , i.e., .
We also need to remark that the character discussed earlier, when restricted to maybe considered as the corresponding character of , i.e., . In fact, and both take values in the centers of and , respectively. The natural embedding of by means of the root data of and which are dual to root data of and who share the maximal torus , identifies as a subgroup of . We therefore have the commutative diagram
and consequently as needed.
The shift in general. To get the precise -factor in general one needs to shift by or should change to , with notation as in [Ngô20], where is half the sum of positive roots in a Borel subgroup of and the hightest weight of .
In our setting, we need to deal with the representation of as well which is not necessarily irreducible. Let be the hightest weights of . We may assume . The shift will then be .
Let us define , and set
Finally, let be the modulus character of .
5.2. The –Harish–Chandra transform
We now recall the -Harish–Chandra transform, a generalization of Satake transform. Given , define its Harish–Chandra transform by
(5.1) |
Next define the -Harish–Chandra transform, –HC in short, by
(5.2) |
Fourier transforms
The conjectural Fourier transform (kernel) is supposed to give the -factor for every irreducible admissible representation of through the convolution
(5.3) |
where is an irreducible admissible representation of and is a matrix coefficient of . In the context of parabolic induction from we will also have defined by .
Fourier transforms and Schwartz spaces
In this section we define suitable spaces of Schwartz functions on and , assuming how Fourier transforms and act on and , respectively. Then
and
Assume and commute with the –Harish–Chandra transform , i.e.,
(5.4) |
or the following diagram commutes
(5.5) |
We now define the Schwartz spaces and as follows:
(5.6) |
and
(5.7) |
As we pointed out in the introduction, these are subspaces of the conjectural -Schwartz spaces and suitable to our purposes. Moreover, as we prove in Proposition 5.3, they contain the and -basic functions. We recall that the -basic function is the unique one for which
where is the normalized spherical matrix coefficient of with defined as in .
Note that sends into and thus (5.5) implies that it also sends
We thus have:
Proposition 5.1.
The –Harish–Chandra transform sends into . In particular, equation is valid for our spaces of -Swartz functions on and .
We remark that this definition will be needed in our proof of multiplicativity, Theorem , the discussion after equation (5.16).
Remark 5.2.
This definition of Schwartz spaces agrees with ideas of Braverman–Kazhdan [BK02, GL20] and with the case of standard representation of . To wit consider , i.e., the Tate’s setting, and check it for the , i.e., the corresponding “basic function”. Let . Now is just the standard Fourier transform
(5.8) |
It can be easily checked that
where is the maximal ideal of .
This simple calculation allows us to prove the following general result:
Proposition 5.3.
The space contains the -basic function.
Proof.
Note that when is a maximal torus, the -Harish-Chandra transform becomes (a twist of) the Satake transform, and in this case the above diagram can be extended to the class of almost compact (ac) spherical functions as defined by Wen-Wei Li in [Li17], and we note that the -basic function is amongst this class (see [Sak18]). The above computation for can be extended to show that the function is also a sum in , where is the standard embedding of a maximal torus of into affine space.
Let be this extended Satake transform. Given a decomposition , with , the commutativity of (5.5) implies that the standard basic function on , lies in as defined in .
Note that here Sat is an isomorphism of -spherical compactly supported functions on and the Weyl-invariant compactly supported functions on . The basic function on , and the functions in its decomposition as are invariant under permutations of the coordinates, and so the above maps are well-defined in the remarks above.
We can deduce the analogous case for a general from the standard case above as follows: Let be a maximal torus in with representation of the dual group of . One obtains a canonical map that extends to a map , the target of this map being the toric variety constructed in section (see Section ). The -Schwartz space on can be defined as the image of
the pushforward by . Then the torus basic function can be expressed as Moreover, this pushforward is compatible with the -Fourier transform on tori, as in diagram . That is,
which shows that . Finally, the commutativity of diagram allows us to lift this decomposition to a decomposition of the basic function as
∎
Multiplicativity. As we discussed earlier every theory of -functions must satisfy multiplicativity, an axiom that is a theorem for all the Artin -functions and is the main tool in computing -factors and -functions. To explain, let be a parabolic subgroup of with a Levi subgroup , uniquely fixed such that , the maximal torus of fixed in our construction throughout. Let be an irreducible admissible representation of and let be a finite dimensional complex representation of and as before. For each irreducible admissible representation of , we can define the -factors and . Multiplicativity states that:
(5.9) |
Here we suppress the dependence of the factors on the non-trivial additive character of .
We note that, since the induced representation may not be irreducible, the -factor is defined to be , where is any irreducible constituent of . The -factor will not depend on the choice of as the proof below establishes.
A proof of (5.9) is usually fairly hard for -factors defined by Rankin–Selberg method [JPSS83, Sou93]. On the contrary, (5.9) is a general result within the Langlands–Shahidi [Sha10] method with a very natural proof.
Our aim here is to give a general proof of (5.9) within the Braverman–Kazhdan/Ngo and Lafforgue programs using (5.4) and Int-invariance of . It follows the arguments given in [GJ72]. We now proceed to give a proof of (5.9) which we formally state as:
Theorem 5.4.
Let be an irreducible admissible representation of and let
. Let be an irreducible finite dimensional complex representation of and let . Assume the validity of (5.4) and thus (5.5) for Schwartz functions, and that , , a maximal compact subgroup of satisfying , which follows from the expected Int-invariance of the kernel . Then
Proof.
Let be the contragredient of . Choose and . Then a
matrix coefficient for can be written as
(5.10) | |||||
where and are values of the functions in and , respectively, and is a matrix coefficient for , and . Let be a -Schwartz function in . We absorb the complex number in by replacing by and then ignoring it throughout the proof.
Now we have the zeta function
(5.11) |
As explained earlier, the shift allows us to get the precise -function at , rather than a shift of , when is the basic function of for a spherical representation . Using (5.10), (5.11) equals
(5.12) |
Write , , , . Then
With notation as in [GJ72], define:
(5.13) |
Thus
(5.14) |
Recall the HC–transform :
Then (5.14) equals
(5.15) |
Since is compact and and are smooth functions, there exist matrix coefficients of and continuous functions on such that
Similarly, there are Schwartz functions in and continuous symmetric functions on such that
(5.16) |
This is clearly true if , since it will then be uniformly smooth. Otherwise, using (5.16) we have
where for simplicity. But Lemma 5.6, proved later, implies
for all and in and thus (5.16) holds for all . Consequently
with by Proposition 5.1. Let
Then we have
(5.17) |
For simplicity of notation, let for each , . We now calculate
One needs to be careful since the involution will now play an important role. We should also point out that needs to be inserted to take into account the appearance of in the left hand side of the functional equation as it is the case in equation (1.1) for . We note that and that will appear as , and thus included in as did in as . Note that in the case
of and , , and as reflected in (1.1).
We have
Changing to , (5.18) equals
(5.19) |
Write , , , . Then
(5.20) |
and (5.19) equals
(5.21) |
which finally equals
(5.22) |
Again, for simplicity for each , let denote its -Fourier transform
(5.23) |
To proceed, we need:
Lemma 5.5.
Let and be in . Then
Proof.
We have
since and are in whose modulus character is 1. Now using Int-invariance of
, (5.24) equals
completing the proof. ∎
Remark 5.6.
In terms of we have proved:
(5.25) |
We now apply Lemma 5.5 to equation (5.22) to get:
(5.26) |
But
and
and therefore (5.26) equals
(5.27) |
We can now apply the commutativity of –Harish–Chandra transform and Fourier transforms and , i.e., equation (5.4) to conclude that (5.27) equals
(5.28) |
But
by the functional equation for . On the other hand the functional equation for gives (5.28) as
which equals
by (5.17). The equality
is now immediate. ∎
5.3. The case of .
We now determine in the case and , i.e., that of Godement–Jacquet [GJ72] and show that it agrees with calculations in Lemma 3.4.0 of loc. cit., after a suitable normalization. We thus assume is the standard maximal parabolic subgroup of , containing the subgroup of upper triangular elements , , with . Recall that we need to determine . But for
Thus
(5.29) |
Moreover
(5.30) |
and thus
(5.31) |
We now verify that Lemma 3.4.0 of [GJ72] is equivalent to our commutative diagram (5.5).
Let be the standard Fourier transform for and its restriction to . With notation as in pages 37–38 of [GJ72],
(5.32) |
where and , is the analogue of our HC–transform. In fact, (5.32) can be written as
(5.37) | |||||
(5.38) |
by (5.31), where and .
In the notation of [GJ72], Lemma 3.4.0 of [GJ72] states that
(5.39) |
for with its standard Fourier transform.
Then by (5.33) the right hand side of (5.34) equals
(5.42) | |||||
by (5.31), where
as defined by (5.1).
Similarly from the left hand side of (5.34), using a change of variables as in (5.33), we have
(5.46) | |||||
Thus (5.34) is equivalent to (5.4) for and .
5.4. Inductive definition of
In the introduction we mentioned that multiplicativity plus a definition of Fourier transform that acts through the correct scalar factors equal to the gamma factors on supercuspidal representations/characters, is enough to characterize the full Fourier transform. Indeed, if we assume that is a good distribution in the sense of Braverman-Kazhan [BK10], then we can identify with a rational, scalar valued function , where is defined by .
Our results on mulitplicativity allow in principle for us to construct in an inductive fashion a distribtion on by formally inducing from for each conjugacy class of Levi subgroup . In fact, our setup and definitions, culminating in Theorem , are normalized so as to make induction of representations adjoint to our -Harish-Chandra transform, that is, we have an equality
Here is a supercuspidal character of a representation on . The ajdunction allows us to identify the and actions on the Bernstein components of on and the Bernstein component of on , respectively. In , we started with an assumption of knowledge of and and we showed that this is equivalent to an equality of gamma factors. However, the gamma factors determine the distribution uniquely, and so one can in principle characterize completely a distribution by specifying its action on supercuspidal representations on , and postulating multiplicativity as an axiom. More concretely, if we inductively know for conjugacy classes of parabolic subgroups , we may formally induce to provide a definition of with a correct action, at least on functions whose spectral decomposition consists solely of induced data from :
The distribution can be a priori defined by the above in order to meet this adjunction, and in fact will then be represented by the conjugation-invariant function
where the are chosen representatives of -conjugacy classes of elements that are -conjugate to , and and are the respective discriminant functions on and . (Here we are identifying with the invariant function representing it).
That satisfies the first adjunction, and therefore multiplicativity, follows from the formula for the trace, and the expression of the distribution character in terms of , adapted to the -setting.
6. Example: The case of Tori and unramified data
We now consider the case of tori, which for present purposes we assume are split. Let be a split torus over . When is a maximal split torus in a reductive group , the upcoming discussion gives the first term of the inductive construction defining the Fourier transform for , with minimal parabolic which is a Borel subgroup. Let be a finite dimensional represenation of . Our notation is justified if we assume , where is a representation of . Let . Then
Write
(6.1) |
where the , are the weights of . We note that they are not necessarily distinct. If we realize these weights of as co-characters of , we get a map (defined over , as is split), which being dual to , is given by (c.f. [Ngô20])
We can extend this to a monoid homomorphism
where is the corresponding toric variety. As in [Ngô20], define the trace function by
and set
by , where is our fixed non-trivial character of .
Denote by the kernel
for as defined in Section 1, i.e., the standard Fourier transform on . We use again for its restrition to , the monoid for .
In [Ngô20] Ngo defines the kernel for the Fourier transform on by
(6.2) |
which equals to
(6.3) |
where is the kernel of . In Proposition 6 of [Ngô20], Ngo regularizes this integration into a principal value integral.
The space of Schwartz functions on are compactly supported functions in that are restrictions of standard Schwartz functions on to . Their further restriction to is in our notation.
Let be the push-forward of . We will verify that the diagram
(6.4) |
commutes, where is the image of under .
Let and define
(6.5) |
where . The commutativity of (6.4) is equivalent to
Lemma 6.1.
For , define by (6.5). Then
Proof.
By definition, for ,
(6.6) | ||||
using in (6.6), then the lemma follows. ∎
The push-forward can be restricted to
leading to
(6.7) |
where is the Weyl group , and denotes the corresponding Hecke algebra. Identifying, via the corresponding Satake isomorphisms
and
(6.8) |
in which defines the Fourier transform
restricted to . Consequently, at least on , the Fourier transform and commute with the Harish-Chandra transform.
7. The case of standard -functions for classical groups; the doubling method
We conclude by addressing multiplicativity in the case of standard -functions, twisted by a character, for classical groups as developed by Piatetski-Shapiro and Rallis, which has been addressed further by a number of other authors [BK02, JLZ20, Li18, Sha18] within our present context. We refer to the local theory developed by Lapid and Rallis. We will be brief and only mention the relevant statements.
The -Harish-Chandra transform is the one given in Proposition of [LR05] as with notation as in [LR05]. Our commutativity equation () in this case is equation () in Lemma of [LR05] in which , a normalized intertwining operator, while acts as the operator induced from with notation as in [LR05] in the context of doubling construction, or simply put .
References
- [BNS16] Alexis Bouthier, Bao Chau Ngo, and Yiannis Sakellaridis. On the formal arc space of a reductive monoid. American Journal of Mathematics, 138(1):81–108, 2016.
- [BK10] Alexander Braverman and David Kazhdan. -functions of representations and lifting. In Visions in Mathematics, pages 237–278. Springer, 2010.
- [BK02] Alexander Braverman and David A Kazhdan. Normalized intertwining operators and nilpotent elements in the Langlands dual group. Moscow Mathematical Journal, 2(3):533–553, 2002.
- [Bri89] Michel Brion. Spherical varieties; an introduction. In Topological methods in algebraic transformation groups, pages 11–26. Springer, 1989.
- [BK07] Michel Brion and Shrawan Kumar. Frobenius splitting methods in geometry and representation theory, volume 231. Springer Science & Business Media, 2007.
- [Cas17] William Casselman. Symmetric powers and the Satake transform. Bulletin of the Iranian Mathematical Society, 43(4 (Special Issue)):17–54, 2017.
- [CST17] J. W. Cogdell, F. Shahidi, and T.-L. Tsai. Local Langlands correspondence for and the exterior and symmetric square root numbers for GL(n). Duke Mathematical Journal, 166(11):2053 – 2132, 2017.
- [CKPSS04] James W. Cogdell, Henry H. Kim, Ilya I. Piatetski-Shapiro, and Freydoon Shahidi. Functoriality for the classical groups. Publications Mathématiques de l’IHÉS, 99:163–233, 2004.
- [CLS11] David A. Cox, John B. Little, and Henry K. Schenck. Toric Varieties, volume 124 of GSM. American Mathematical Soc., 2011.
- [Elk78] Renée Elkik. Singularités rationnelles et déformations. Inventiones Mathematicae, 47(2):139–147, 1978.
- [GT11] Wee Teck Gan and Shuichiro Takeda. The local Langlands conjecture for GSp (4). Annals of Mathematics, pages 1841–1882, 2011.
- [GPSR87] Stephen Gelbart, Ilya Piatetski-Shapiro, and Stephen Rallis. Explicit constructions of automorphic L-functions, volume 1254 of SLN. Springer, 1987.
- [GL20] Jayce Robert Getz and Baiying Liu. A refined poisson summation formula for certain braverman-kazhdan spaces. Science China Mathematics, pages 1–30, 2020.
- [GJ72] Roger Godement and Hervé Jacquet. Zeta functions of simple algebras, volume 260. Springer, 1972.
- [HT01] Michael Harris and Richard Taylor. The Geometry and Cohomology of Some Simple Shimura Varieties.(AM-151), Volume 151, volume 151. Princeton university press, 2001.
- [Hen00] Guy Henniart. Une preuve simple des conjectures de Langlands pour GL(n) sur un corps p-adique. Inventiones mathematicae, 139(2):439–455, 2000.
- [JPSS83] Hervé Jacquet, Ilya I. Piatetski-Shapiro, and Joseph A. Shalika. Rankin-selberg convolutions. American Journal of Mathematics, 105(2):367–464, 1983.
- [JLZ20] Dihua Jiang, Zhilin Luo, and Lei Zhang. Harmonic analysis and gamma functions on symplectic groups. arXiv preprint arXiv:2006.08126, 2020.
- [Kim03] Henry Kim. Functoriality for the exterior square of and the symmetric fourth of . Journal of the American Mathematical Society, 16(1):139–183, 2003.
- [KS02] Henry H Kim and Freydoon Shahidi. Functorial products for and functorial symmetric cube for , with an appendix by C. Bushnell and G. Henniart. Annals of Math, 155(8):837–893, 2002.
- [Kno91] Friedrich Knop. The Luna-Vust theory of spherical embeddings. In Proceedings of the Hyderabad conference on Algebraic Groups, pages 225–249, 1991.
- [Laf14] Laurent Lafforgue. Noyaux du transfert automorphe de Langlands et formules de poisson non linéaires. Japanese Journal of Mathematics, 9(1):1–68, 2014.
- [LR05] E Lapid and Stephen Rallis. On the local factors of representations of classical groups. In Automorphic representations, L-functions and applications: progress and prospects, volume 11 of Ohio State Univ. Math. Res. Inst. Publ, pages 309–359. de Gruyter, Berlin, 2005.
- [Li17] Wen-Wei Li. Basic functions and unramified local l-factors for split groups. Science China Mathematics, 60(5):777–812, 2017.
- [Li18] Wen-Wei Li. Zeta integrals, Schwartz spaces and local functional equations, volume 2228 of SLN. Springer, 2018.
- [Los09] Ivan Losev. Proof of the Knop conjecture. In Annales de l’Institut Fourier, volume 59, pages 1105–1134, 2009.
- [Los10] Ivan Losev. Uniqueness properties for spherical varieties. Les cours du CIRM, 1(1):113–120, 2010.
- [Ngô20] Bao Châu Ngô. Hankel transform, Langlands functoriality and functional equation of automorphic L-functions. Japanese Journal of Mathematics, 15(1):121–167, 2020.
- [Per14] Nicolas Perrin. Introduction to spherical varieties. http://relaunch.hcm.uni-bonn.de/fileadmin/perrin/spherical.pdf, 2014.
- [PSR86] Ilya Piatetski-Shapiro and Stephen Rallis. factor of representations of classical groups. Proceedings of the National Academy of Sciences, 83(13):4589–4593, 1986.
- [Pop87] Vladimir L Popov. Contraction of the actions of reductive algebraic groups. Mathematics of the USSR-Sbornik, 58(2):311, 1987.
- [Ren05] Lex E Renner. Linear algebraic monoids, Encylopedia of Mathematical Sciences, volume 134. Springer, 2005.
- [Rit03] Alvaro Rittatore. Reductive embeddings are cohen-macaulay. Proceedings of the American Mathematical Society, 131(3):675–684, 2003.
- [Sak18] Yiannis Sakellaridis. Inverse satake transforms. In Werner Müller, Sug Woo Shin, and Nicolas Templier, editors, Geometric Aspects of the Trace Formula, pages 321–349, Cham, 2018. Springer International Publishing.
- [Sha10] Freydoon Shahidi. Eisenstein series and automorphic L-functions, volume 58, American Mathematical Society Colloquium Publications. American Mathematical Society, Providence, RI, 2010.
- [Sha12] Freydoon Shahidi. On equality of arithmetic and analytic factors through local Langlands correspondence. Pacific Journal of Mathematics, 260(2):695–715, 2012.
- [Sha17] Freydoon Shahidi. Local factors, reciprocity and Vinberg monoids. In “Prime Numbers and Representation Theory”, volume 2 of Lecture Series of Modern Number Theory, pages 200–256. Science Press, 2017.
- [Sha18] Freydoon Shahidi. On generalized Fourier transforms for standard L-functions. In “Geometric Aspects of the Trace Formula”, Simons Symposium on the Trace Formula, pages 351–404. Springer, 2018.
- [Sou93] David Soudry. Rankin-Selberg Convolutions for : Local Theory. Memoirs of AMS 105, no. 500. American Mathematical Soc., 1993.
- [Stu95] Bernd Sturmfels. On vector partition functions. Journal of Combinatorial theory, series A, 72(2):302–309, 1995.
- [Tat50] John Tate. Fourier analysis in number fields and hecke’s zeta functions. Thesis Princeton, 1950.
Purdue University
West Lafayette, IN 47907
[email protected]
[email protected]