An Asymptotic Orthogonality Relation for
Abstract.
Orthogonality is a fundamental theme in representation theory and Fourier analysis. An orthogonality relation for characters of finite abelian groups (now recognized as an orthogonality relation on GL(1)) was used by Dirichlet to prove infinitely many primes in arithmetic progressions. Asymptotic orthogonality relations for GL, with , and applications to number theory, have been considered by various researchers over the last 45 years. Recently, the authors of the present work have derived an explicit asymptotic orthogonality relation, with a power savings error term, for GL. Here we we extend those results to GL .
For , our results are contingent on the Ramanujan conjecture at the infinite place, but otherwise are unconditional. In particular, the case represents a new result. The key new ingredient for the proof of the case is the theorem of Kim-Shahidi that functorial products of cusp forms on GL(2)GL(3) are automorphic on GL(6). For (assuming again the Ramanujan conjecture holds at the infinite place), our results are conditional on two conjectures, both of which have been verified in various special cases. The first of these conjectures regards lower bounds for Rankin-Selberg L-functions, and the second concerns recurrence relations for Mellin transforms of GL Whittaker functions.
Central to our proof is an application of the Kuznetsov Trace formula, and a detailed analysis, utilizing a number of novel techniques, of the various entities—Hecke-Maass cusp forms, Langlands Eisenstein series, spherical principal series Whittaker functions and their Mellin transforms, and so on—that arise in this application.
1. Introduction
1.1. Brief description of the main result of this paper
Let be a rational integer, , and denote the ring of adeles over where denotes the finite adeles. The family of unitary cuspidal automorphic representations of and their standard L-functions
were first introduced by Godement and Jacquet [GJ72] and have played a major role in modern number theory. In the special case of the Euler products are just Dirichlet L-functions.
In this paper we focus on the unitary cuspidal automorphic representations of with trivial central character which are globally unramified. For these can be studied classically in terms of Hecke-Maass cusp forms on
where
is a generalization of the classical upper half-plane. In fact is isomorphic to the classical upper half-plane. For , Hecke-Maass cusp forms are smooth functions which are automorphic for with moderate growth and which are joint eigenfunctions of the full ring of invariant differential operators on and are also joint eigenfunctions of the Hecke operators. Such globally unramified Hecke-Maass forms can be classified in terms of Langlands parameters which (assuming the cusp form is tempered) are pure imaginary numbers that sum to zero. Further, the Hecke-Maass cusp forms with Langlands parameters can be ordered in terms of their Laplace eigenvalues given by
as proved by Stephen Miller [Mil02].
Let be a Hecke-Maass cusp form for for and set
to denote the Petersson norm of The Hecke-Maass cusp forms form a Hilbert space over with respect to the Petersson inner product.
Definition 1.1 (L-function of a Hecke-Maass cusp form).
Let be a Hecke-Maass cusp form for . Then for with sufficiently large we define the L-function where is the Hecke eigenvalue of
Definition 1.2 (Asymptotic orthogonality relation for ).
Let (with associated Langlands parameters ) denote an orthogonal basis of Hecke-Maass cusp forms for with L-function given by Fix positive integers . Then, for , we have
where and is a smooth function of the variables (for ) with support on the Laplace eigenvalues where
Remark 1.3 (Power savings error term).
The asymptotic orthogonality relation has a power savings error term if can be replaced with for some fixed The error terms will generally depend on . This type of asymptotic orthogonality relation was first conjectured by Fan Zhou [Zho14].
Remark 1.4 (Normalization of Hecke-Maass cusp forms).
The approach we take in proving asymptotic orthogonality relations for is the Kuznetsov trace formula presented in §4 where (which are independent of the way the are normalized) appears naturally on the spectral side of the trace formula leading to an asymptotic orthogonality relation of the form
(1.5) |
If we normalize so that its first Fourier coefficient is equal to one then it is shown in Proposition 4.4 that
This allows us (with a modification of the test function ) to replace the inner product appearing in (1.5) with the adjoint L-function as in Definition 1.2. The main reason for doing this is that there are much better techniques developed for bounding special values of L-functions as opposed to bounding inner products of cusp forms. So having in the asymptotic orthogonality relation instead of will allow us to obtain better error terms in applications.
Orthogonality relations as in Definition 1.2 have a long history going back to Dirichlet (for the case of GL(1)) who introduced the orthogonality relation for Dirichlet characters to prove infinitely many primes in arithmetic progressions. Bruggeman [Bru78] was the first to obtain an asymptotic orthogonality relation for GL(2) which he presented in the form
where goes over an orthogonal basis of Hecke-Maass cusp forms for . This is not quite in the form of Definition 1.2 but it can be put into that form with some work. Other versions of GL(2) type orthogonality relations with important applications were obtained by Sarnak [Sar87], and, for holomorphic Hecke modular forms, by Conrey-Duke-Farmer [CDF97] and J.P. Serre [Ser97].
The first asymptotic orthogonality relations for GL(3) with power savings error term were proved independently by Blomer [Blo13] and Goldfeld-Kontorovich [GK13] in 2013. In 2021 Goldfeld-Stade-Woodbury [GSW21] were the first to obtain a power-saving asymptotic orthogonality relation as in Definition 1.2 for GL(4).
A major breakthrough was obtained by Matz-Templier [MT21] who unconditionally proved an asymptotic orthogonality relation for , as in (1.5), for a wide class of test functions for all (with power savings) but without the harmonic weights given by the inverse of the adjoint L-function at 1. Their results were further strengthened in Finis-Matz [FM21]. The principal tool used to prove the asymptotic orthogonality relation in [MT21] was the Arthur-Selberg trace formula, whereas our approach is the natural generalization of the earlier results [Blo13], [GK13], [GSW21], which were based on the Kuznetsov trace formula. Blomer [Blo21] presented a very nice exposition comparing the Arthur-Selberg and Kuznetsov trace formulae which we now briefly summarize for the application to asymptotic orthogonality relations. The first key difference between these trace formulae is that the spectral side of the Kuznetsov trace formula has harmonic weights while the Arthur-Selberg trace formula does not have these harmonic weights. For with it is not currently known how to remove these weights (see [BZ20] for how to remove the weights on ). In [Blo21] Blomer remarks that “for applications to L-functions involving period formulae it is often desirable to have an additional factor in the cuspidal spectrum, but in other situations one may prefer a summation formula without an extra L-value.” The second major difference between these trace formulae is that the spectral side of the Kuznetsov trace formula does not contain residual spectrum while the Arthur-Selberg trace formula does. As pointed out by a referee the bulk of the work in Matz-Templier [MT21] consists in bounding the unipotent contribution on the geometric side of the Arthur trace formula so that it stays in line with the error term coming from the residual Eisenstein contribution on the spectral side given by Lapid-Mueller [LM09]. These residual Eisenstein series do not appear in the Kuznetsov trace formula which leads to a very strong conjectural error term in Theorem 1.1. In fact, the largest error term on the spectral side of the Kuznetsov trace formula arises from the tempered Eisenstein series coming from the maximal parabolic having Levi block decomposition. For explicit comparisons between our main theorem and the results of [MT21] see Remark 1.4. There are certain applications of our results using the Kuznetsov trace formula approach that go beyond the results in [MT21], [FM21]. Recall that denotes the Hecke eigenvalue of the Maass form . Fan’s thesis concerns the so-called vertical Sato-Tate problem which is a conjecture about the distribution of where is fixed and varies. This problem was studied by Bruggeman [Bru78] and Sarnak [Sar87] (for Maass forms), and Serre [Ser97] and Conrey-Duke-Farmer [CDF97] (for holomorphic forms), who showed by fixing and varying , that is an equidistributed sequence with respect to the Plancherel measure which depends on . Strikingly, as observed by Fan Zhou ([Zho14]), if we give each Hecke eigenvalue the weight , then the distribution involves the Sato-Tate measure which is independent of . Jana, in [Jan21], generalized the results of Zhou, but he only obtained an asymptotic formula without a power savings error term. A problem for the future would be to combine Jana’s approach with the methods of this paper. Jana also obtains bounds toward Sarnak’s density hypothesis using this strategy that are stronger than anything known using the Arthur-Selberg trace formula. The main aim of this paper is to explicitly work out an asymptotic orthogonality relation for via the Kuznetsov trace formula for a special choice of test function whose form is that of a Gaussian times a fixed polynomial. We do not address applications in this paper and leave that to future research. See [Blo21] for various applications of the Arthur-Selberg and Kuznetsov trace formulae and how they compare. We also point out that the Kuznetsov trace formula was generalized by Jacquet and Lai [JL85] who developed the relative trace formula which has had a wide following with new types of applications. See §1.1 for the statement of our main theorem. The proof we give assumes the Ramanujan conjecture at but it is possible to prove a weaker result by dropping this assumption. Otherwise the proof is unconditional for . In particular, the case represents a complete, new result. For , our result is conditional on two conjectures.
1.2. Ishii-Stade Conjecture
The Ishii-Stade Conjecture (see §8.2) concerns the normalized Mellin transform of the Whittaker function defined in Definition 2.3. Here, , and satisfies .
Suppose integers and , with and , are given. The Ishii-Stade Conjecture expresses as a finite linear combination, with coefficients that are rational functions of the ’s and ’s, of shifted Mellin transforms
where and the th coordinate of is . In other words, for such and , the conjecture expresses the Mellin transform in terms of shifts of this Mellin transform by at least units to the right in the variable .
Much as recurrence relations of the form
for Euler’s Gamma function imply concrete results concerning analytic continuation, poles, and residues of that function, so will the Ishii-Stade conjecture allow us to obtain explicit information about the behavior of beyond its original, a priori domain of definition. This explicit information will be crucial to the analysis of our test function , and consequently, to our derivation of an an asymptotic orthonality relation as in Definition 1.2.
We have been able to prove the Ishii-Stade Conjecture for with . See §8.2 below.
1.3. Lower bound conjecture for Rankin-Selberg L-functions
Fix . Let be a partition of with . The second conjecture we require for the proof of the asymptotic orthogonality relation for is a conjecture on the lower bound for Rankin-Selberg L-functions on the line where (for ) are Hecke-Maass cusp forms for , , respectively. For a Hecke-Maass cusp form with Langlands parameters , let
(1.1) |
denote the analytic conductor of as defined by Iwaniec and Sarnak [IS00].
Conjecture 1.2 (Lower bounds for Rankin-Selberg L-functions).
Let be fixed. Then we have the lower bound
Remark 1.3.
Conjecture 1.2 follows from Langlands’ conjecture that is automorphic for . This can be proved via the method of de la Valée Poussin as in Sarnak [Sar04]. Interestingly, Sarnak’s approach can be extended to prove Conjecture 1.2 if is the dual of (see [GL18], [HB19]). Stronger bounds can also be obtained if one assumes the Lindelöf or Riemann hypothesis for Rankin-Selberg L-functions.
1.4. Constructing the test functions
Fix an integer . We now construct two complex-valued test functions on the space of Langlands parameters
that will be used in our proof of the orthogonality relation for
We begin by introducing an auxiliary polynomial that is used in constructing the test functions.
Definition 1.1 (The polynomial ).
Let and let be a Langlands parameter. Then we define
Note that if then is the square root of a polynomial in of degree , where
(1.2) |
By abuse of notation, we refer to as a polynomial although this is not strictly the case unless is even. For with bounded real and imaginary parts, say, and we have
(1.3) |
with an implicit constant depending on .
Definition 1.4 (The test functions and ).
Let and . Then for a Langlands parameter , we define
We observe that, by Stirling’s formula for the Gamma function and by (1.2) and (1.3), we have
(1.5) |
whenever is bounded and for . The implied constant in (1.5) depends on and .
Remark 1.6 (Positivity of ).
Writing with and for each , the function is positive. This is the case because for .
Remark 1.7 (Whittaker transform of the test function).
The symbol in the test function means this function is the Whittaker transform of . See § 8.
1.5. The Main Theorem
Theorem 1.1.
Fix Let denote an orthogonal basis of Hecke-Maass cusp forms for (assumed to be tempered at ) with associated Langlands parameter
and L-function . Fix positive integers . Then assuming the Ishii-Stade conjecture 8.3 and the lower bound conjecture for Rankin-Selberg L-functions 1.2, we prove that for ,
where is the Kronecker symbol, and are absolute constants which depend at most on and .
Remark 1.2.
Qiao Zhang [Zha22] recently proved the lower bound
(1.3) |
with . This improves on Brumley’s bound who obtained nearly the same result but with the term replaced by (see [Bru06] and the appendix of [Lap13]). Assuming (1.3) we can replace the error term in Theorem 1.1 with
So if one could prove (1.3) with this would give a power savings error term in our main theorem and would remove the assumption of the lower bound conjecture 1.2. In fact, the proof establishes a black box by which improvements to bounds on Rankin-Selberg L-functions result in better power savings error terms for the continuous spectrum contribution to the asymptotic orthogonality relation.
Remark 1.4.
A variant of Theorem 1.1 is obtained unconditionally in [MT21], [FM21], without the arithmetic weights and with different test functions, which are indicator functions of , where is a Weyl group invariant bounded open subset of , where is the Lie algebra of the subgroup of diagonal matrices with positive entries. Additionally, the results of [MT21], [FM21] do not entail the polynomial weights of size coming from (cf. (1.5)).
The error term obtained in [FM21], in the present setting of , is as . Here, is the dimension of the generalized upper half-plane , and the error term obtained by Finis-Matz has exponent equal to that dimension minus 1. By comparison, if one removes the polynomial weights from the error term in Theorem 1.1 above, then one obtains an error term that is . Also note that our main term is of a stronger form than that of [MT21], [FM21], in that ours entails a sum of different high order asymptotics.
More recently, Jana [Jan21] obtained a proof of the asymptotic orthogonality relation defined in 1.2, using the Kuznetsov trace formula and not the Selberg trace formula, with applications to the equidistribution of Satake parameters with respect to the Sato-Tate measure, second moment estimates of central values of L-functions as strong as Lindelöf on average, and distribution of low lying zeros of automorphic L-functions in the analytic conductor aspect. The paper of Jana does not contain a power saving error term.
Remark 1.5.
It is possible to remove the assumption of Ramanujan at the infinite place with more work which results in a weaker power savings error term in Theorem 1.1. For a Maass form with Langlands parameter , note that the test function is positive. This is true because, even if is a Langlands parameter of an element in the complementary spectrum, is a permutation of . A weaker version of Theorem 1.1 can be proved if one assumes that almost all (except for a set of zero density) are tempered. Such results have been obtained in [MT21], [FM21].
Proof of Theorem 1.1.
Computing the inner product of certain Poincaré series in two ways (see the outline in §1.6 below), we obtain a Kuznetsov trace formula relating the so-called geometric and spectral sides. The geometric side consists of a main term and a Kloosterman contribution . The spectral side also consists of two components: a cuspidal (i.e., discrete) contribution and an Eisenstein (i.e., continuous) contribution .
The left hand side of the theorem is precisely . The first set of terms on the right hand side comes from the asymptotic formula for given in Proposition 5.1. The power of in the error term comes from the bound for given in Theorem 7.1 (which also gives a factor of ). A bound for , which is a (finite) sum of terms , with the same power of but with the given power of follows as a consequence of Proposition 6.1. ∎
1.6. Outline of the key ideas in the proofs
Fix The orthogonality relation appears directly in the spectral side of the Kuznetsov trace formula for which we now discuss. The Kuznetsov trace formula is obtained by computing the inner product of two Poincaré series on in two different ways. The Poincaré series are constructed in a similar manner to Borel Eisenstein series by taking all translates of a certain test function which we choose to be the test function in Definition 1.4 multiplied by a character and a power function (see Definition 2.7).
The first way of computing the inner product of two Poincaré series is to replace one of the Poincaré series with its spectral expansion into cusp forms and Eisenstein series and then unravel the other Poincaré series with the Rankin-Selberg method. This gives the spectral contribution which has two parts: the cuspidal contribution and the Eisenstein contribution. The second way of computing the inner product is to replace one of the Poincaré series with its Fourier Whittaker expansion and then unravel the other Poincaré series with the Rankin-Selberg method. This is called the geometric contribution to the trace formula, which also consists of two parts: a main term, and the so-called Kloosterman contribution. The precise results of these computations are given in Theorems 4.1 and 4.1, respectively.
Bounding the Eisenstein contribution
The key component of the Eisenstein contribution to the Kuznetsov trace formula is the inner product of an Eisenstein series and the Poincaré series given in Definition 2.7. By unraveling the Poincaré series in the inner product (see Proposition 4.2) we essentially obtain the Fourier coefficient of the Eisenstein series multiplied by the Whittaker transform of . The explicit formula for the Fourier coefficient of the most general Langlands Eisenstein series given in Proposition 4.5 allows us to effectively bound all the terms in the integrals appearing in the Eisenstein contribution except for the product of adjoint L-functions
(1.1) |
appearing in that proposition. When considering the Eisenstein contribution to the Kuznetsov trace formula for all the adjoint L-functions in the above product are for cusp forms of lower rank . Now in the special case that , our Main Theorem 1.1 for gives a sharp bound for the sum of reciprocals of all adjoint L-functions of lower rank. This allows us to inductively prove a power savings bound for the product (1.1).
Asymptotic formula for the geometric contribution
We prove that the geometric contribution is a sum of expressions over elements in the Weyl group of . The are complicated multiple sums of multiple integrals weighted by Kloosterman sums (see (4.2)). If is the trivial element of the Weyl group then we obtain an asymptotic formula for (see Proposition 5.1) while for all other Weyl group elements , with , we obtain error terms with strong bounds for (see Proposition 6.1) which are bounded by the final error term on the right side of our main theorem.
The key terms in (4.2), the formula for , are the Kloosterman sums and two appearances of the test function : one that is twisted by the Weyl group element and one that is not. For the Kloosterman sums, we rely on bounds given by [DR98]. The task of giving strong bounds for occupies Sections 8, 9 and 10. We deal with the combinatorics of the twisted -function, and we combine the bounds for it, the other -function and the Kloosterman sums in Section 6.
The function is the inverse Whittaker transform of the test function given in Definition 1.4 above. Thanks to a formula of Goldfeld-Kontorovich [GK12], we can realize this as an integral of the product of , the Whittaker function (see Definition 2.3), and certain additional gamma factors. We then write the Whittaker function as the inverse Mellin transform of its Mellin transform: . This leads to the formula (valid for any ):
To estimate the growth of uniformly in and as , we shift the line of integration in the -integrals to with where for . We remark that this is precisely where the Ishii-Stade Conjecture is required. It is well known that
and hence understanding the values of for is straightforward by applying the functional equation for the gamma function or, equivalently, using an integral representation of the gamma function valid for . A similar strategy can be used when . However, for , the analogous method seems intractable because the Mellin transform is not just a ratio of gamma functions, but an integral of such. To overcome this difficulty, we apply the Ishii-Stade conjecture to describe the values of in terms of sums of the Mellin transform of shifts of the -variables. See also Remark 8.11 below.
The Cauchy residue formula allows us to express as a sum of the shifted -integral (termed the shifted term and denoted ) and many residue terms. The description of the shifted and residue terms is given in Section 8.3. In order to bound it is convenient to introduce the function .
The next step is to use a result of Ishii-Stade (see Theorem 8.5) which allows us to write the Mellin transform as an integral transformation of against certain additional gamma factors. It is important to note that can be expressed in terms of . By carefullly teasing apart the portion of which determines and that which doesn’t, we are able to separate out the gamma factors that don’t depend on and bound by the product of a power of and for a certain . This gives an inductive procedure, therefore, for bounding the shifted term.
In Section 10.2 we set notation for describing the -fold shifted residue terms. This requires generalizing a result of Stade (see Theorem 10.1) on the first set of residues of (i.e., those that occur at ) to, first, higher order residues (i.e., taking the residue with respect to multiple values ), and second, to residues which occur along the lines for . This result, together with a teasing out of the variables similar to that described above, allows us to bound an -fold residue term as the product of certain powers of and the variables times
Applying the bounds on that we inductively established for bounding the shifted term, and keeping careful track of all of the exponents and terms , we eventually show that the bound for the shifted main term is in fact valid for every residue term as well.
Remark 1.2.
In comparison to the results of [GK13] and [GSW21], we are using a slightly different normalization of the gamma functions and the auxiliary polynomial in the definition of the test functions and (see Definition 1.4). Adjusting for this difference the results obtained here when applied to and recover the previously proven asymptotic formulas.
2. Preliminaries
2.1. Notational conventions
Definition 2.1 (Hat notation for summation).
Suppose that and . For any , define
Note that empty sums are assumed to be zero.
Definition 2.2 (Integration notation).
Let . We will often be working with - and -tuples of real or complex numbers. We will denote such tuples without a subscript and use subscripts to refer to the components. For example, we set , and such that
In such cases, we denote integration over all such variables subject to a condition(s) via
For example, given with , we denote integration over all such with for each via
We extend this notation liberally to integrals over , and and apply it also to integrals over the imaginary parts in the sequel.
Definition 2.3 (Polynomial notation).
Our analysis will often require us to bound certain polynomials in a trivial way. Namely, for complex variables with , if for each and is a polynomial, then . So, the relevant information about is its degree. This being the case, we will use the notation (with ) to represent an unspecified polynomial of degree less than or equal to in the variable(s) . Note that this notation agrees with the commonly employed practice (also used throughout these notes) of using to represent an unspecified positive real number whose precise value is not specified and may differ from one usage to another.
Definition 2.4 (Vector or matrix notation depending on context).
Given a vector , we shall define the diagonal matrix
2.2. Structure of
Suppose is a positive integer. Let denote the set of upper triangular unipotent matrices.
Definition 2.1 (Character of ).
Let . For an element of the form
(2.2) |
we define the character
(2.3) |
Definition 2.4 (Generalized upper half plane).
We denote the set of (real) orthogonal matrices , and we set
Every element (via the Iwasawa decomposition of [Gol06]) of has a coset representative of the form , with as above and
(2.5) |
where for each . The group acts as a group of transformations on by left multiplication.
Definition 2.6 (Weyl group and relevant elements).
Let denote the Weyl group of We consider it as the subgroup of consisting of permutation matrices, i.e., matrices that have exactly one in each row/column and all zeros otherwise. An element is called relevant if
where is the identity matrix of size and is a composition (a way of writing as a sum of positive integers; see Section 8.3). The long element of is .
Definition 2.7 (Other subgroups of ).
We define
and
where denotes the transpose of , i.e., the set of lower triangular unipotent matrices.
2.3. Basic functions on the generalized upper half plane
Definition 2.1 (Power function).
Let with . Let , where for . We define a power function on by
(2.2) |
where is the -th diagonal entry of the matrix as above.
Definition 2.3 (Jacquet’s Whittaker function).
Let with . Let with . We define the completed Whittaker function by the integral
which converges absolutely if for (cf. [GMW21]), and has meromorphic continuation to all satisfying .
Remark 2.4.
With the additional Gamma factors included in this definition (which can be considered as a “completed” Whittaker function) there are functional equations which is equivalent to the fact that the Whittaker function is invariant under all permutations of . Moreover, even though the integral (without the normalizing factor) often vanishes identically as a function of , this normalization never does.
If is a diagonal matrix in then the value of is independent of sign, so we drop the . We also drop the if the sign is .
Definition 2.5 (Whittaker transform and its inverse).
Assume . Let with . Set and as in Definition 2.4. Let be an integrable function. Then we define the Whittaker transform (where ) by
(2.6) |
provided the above integral converges absolutely and uniformly on compact subsets of . The inverse Whittaker transform [GK12, Theorem 1.6] is
provided the above integral converges absolutely and uniformly on compact subsets of .
Definition 2.7 (Normalized Poincaré series).
Let with for each . As with , we may think of as a matrix. Let . Then we define
(2.8) |
where is the (nonzero) constant given in Proposition 4.4. We extend the definition of and to all of by setting and .
Remark 2.9.
This definition, up to the normalizing factor , of the Poincaré series agrees with that used in [GSW21] with the minor caveat that takes on a slightly different normalization in terms of the polynomial and in the Gamma factors appearing in Definition 1.4. The normalizing factor is inserted so that in the Kuznetsov trace formula the cuspidal term is precisely the orthogonality relation in Theorem 1.1.
2.4. Fourier expansion of the Poincaré series
Definition 2.1 (Twisted Character).
Let
Let and consider the additive character (see (2.3)) of . Then for we define the twisted character by
Definition 2.2 (Kloosterman Sum).
Fix Let be characters of Let where is the Weyl group of Let
with Then the Kloosterman sum is defined as
with notation as in Definition 11.2.2 of [Gol06]. The Kloosterman sum is well defined (i.e. independent of the choice of Bruhat decomposition for ) if and only if it satisfies the compatibility condition It is defined to be zero otherwise. (See [Fri87].)
Proposition 2.3 ( Fourier coefficient of the Poincaré series ).
Let and satisfy and . If is sufficiently large for each , then
where
and denotes the transpose of a matrix .
Proof.
See Theorem 11.5.4 of [Gol06]. ∎
3. Spectral decomposition of
3.1. Hecke-Maass cusp forms for
Definition 3.1 (Langlands parameters).
Let . A vector is termed a Langlands parameter if .
Definition 3.2 (Hecke-Maass cusp forms).
Fix A Hecke-Maass cusp form with Langlands parameter for is a smooth function which satisfies for all , . In addition is square integrable, is an eigenfunction of the algebra of Hecke operators on , and is an eigenfunction of the algebra of invariant differential operators on , with the same eigenvalues under this action as the power function . The Laplace eigenvalue of is given by
See Section 6 in [Mil02]. The Hecke-Maass cusp form is said to be tempered at if the Langlands parameters are all pure imaginary.
Proposition 3.3 (Fourier expansion of Hecke-Maass cusp forms).
Assume Let be a Hecke-Maass cusp form for with Langlands parameters . Then for , we have the following Fourier-Whittaker expansion:
where , is the toric matrix in Definition 2.4 and is the Fourier coefficient of .
Proof.
See Section 9.1 of [Gol15]. ∎
Definition 3.4 (L-function associated to a Hecke-Maass form ).
Let with sufficiently large. Then the L-function associated to a Hecke-Maass cusp form is defined as
and has holomorphic continuation to all and satisfies a functional equation If is a simultaneous eigenfunction of all the Hecke operators then has the following Euler product:
3.2. Langlands Eisenstein series for
Definition 3.1 (Parabolic Subgroup).
For and consider a partition of given by with positive integers We define the standard parabolic subgroup
Letting denote the identity matrix, the subgroup
is the unipotent radical of . The subgroup
is the standard choice of Levi subgroup of .
Definition 3.2 (Hecke-Maass form associated to a parabolic ).
Let . Consider a partition with . Let For , let be either the constant function 1 (if ) or a Hecke-Maass cusp form for (if ). The form is defined on (where ) by the formula
where has the form , with In fact, this construction works equally well if some or all of the are Eisenstein series.
Definition 3.3 (Character of a parabolic subgroup).
Let Fix a partition with associated parabolic subgroup Define
(3.4) |
Let satisfy Consider the function (see Definition 2.1)
on , where
The conditions and guarantee that ’s entries sum to zero. When , with diagonal block entries , one has
so that restricts to a character of which is trivial on .
Definition 3.5 (Langlands Eisenstein series twisted by Hecke-Maass forms of lower rank).
For let denote the Langlands parameters of We adopt the convention that if then Then the Langlands parameters of (denoted ) are
(3.7) |
Definition 3.8.
(The Fourier coefficient of ) Let where Consider with associated Langlands parameters as defined in (3.7). Let . Then the term in the Fourier-Whittaker expansion of is
3.3. Langlands spectral decomposition for
Definition 3.1 (Petersson inner product).
Let . For we define the Petersson inner product to be
For , with
the measure is given by , with
The Langlands spectral decomposition for states that
We shall be applying the Langlands spectral decomposition to Poincaré series which are orthogonal to the residual spectrum.
Theorem 3.2 (Langlands spectral decomposition for ).
Let denote an orthogonal basis of Hecke-Maass forms for . Assume that are orthogonal to the residual spectrum. Then for we have
where the sum over ranges over parabolics associated to partitions , while the sum over (see Definition 3.2) ranges over an orthonormal basis of Hecke-Maass forms associated to . Furthermore, is a fixed non-zero constant.
4. Kuznetsov trace formula
The Kuznetsov trace formula is derived by computing the inner product of two Poincaré series in two different ways. More precisely, let with and , and consider the Petersson inner product .
In particular since (see [Fri87]), the inner product can be computed with the spectral expansion of the Poincaré series. The geometric approach utilizes the Fourier Whittaker expansion of the Poincaré series which involve Kloosterman sums.
The trace formula takes the following form.
(4.1) |
Here is the cuspidal contribution and is the Eisenstein contribution. See Theorem 4.1 for their precise definitions. The geometric side consists of terms corresponding to elements of the Weyl group. The identity element gives the main term , and the Kloosterman contribution is the sum of the remaining terms. See Theorem 4.1 for their precise definitions. The Kloosterman term and the Eisenstein contribution will be small with the special choice of the test function , and they constitute the error term in the main theorem.
4.1. Spectral side of the Kuznetsov trace formula
The first way to compute the inner product of the Poincaré series uses the spectral decomposition of the Poincaré series.
Recall also the definition of the adjoint L-function: where is the Rankin-Selberg convolution L-function as in §12.1 of [Gol15].
Theorem 4.1 (Spectral decomposition for the inner product of Poincaré series).
Fix and . Then For we have
With the notation of the Spectral Decomposition Theorem 3.2, the cuspidal contribution to the Kuznetsov trace formula is
and the Eisenstein contribution to the Kuznetsov trace formula is
for constants .
Proof.
The proof follows from the Langlands Spectral Decomposition Theorem 3.2 with the choices and . We have
We then insert the inner products given in Proposition 4.2 below. Doing so, we see that the cuspidal spectrum is
From Proposition 4.4, we see that
The cuspidal part is now immediate. The contributions from the Eisenstein series are computed in like manner using Proposition 4.5. ∎
Proposition 4.2 (The inner product of with an Eisenstein series or Hecke-Maass form).
Let . Consider the Eisenstein series , with associated Langlands parameters . Let denote a Hecke-Maass cusp form for with Langlands parameter and Fourier coefficient Then for ,
where the inner products on the left are defined by analytic continuation and is the nonzero constant (depending only on ) from Proposition 4.4.
Proof.
We outline the case of the Hecke-Maass forms. The series definition of the Poincaré series converges absolutely for sufficiently large (). It follows that for such we may unravel the Poincaré series in the inner product with the Rankin-Selberg Method. The inner product picks out the Fourier coefficient of multiplied by a certain Whittaker transform of . This Whittaker transform has analytic continuation in to a region including . For sufficiently large , we have from (2.8) that
(4.3) |
Note that . The integral in (4.3) converges (as a function of ) to a region which includes . It follows that the analytic continuation in to of the inner product satisfies
The proof for is the same. ∎
For , consider a Hecke-Maass cusp form for with Fourier Whittaker expansion given by Proposition 3.3. Assume is a Hecke eigenform. Let denote the first Fourier-Whittaker coefficient of Then we have
where is the Hecke eigenvalue (see Section 9.3 in [Gol15]), and .
Proposition 4.4 (First Fourier-Whittaker coefficient of a Hecke-Maass cusp form).
Assume Let be a Hecke-Maass cusp form for with Langlands parameters . Then the first coefficient is given by
where is a constant depending on only.
Proof.
See [GMW21]. ∎
Proposition 4.5 (The Fourier coefficient of ).
Let where Consider with associated Langlands parameters as defined in (3.7). Assume that each Hecke-Maass form (with ) occurring in has Langlands parameters with the convention that if then We also assume that each is normalized to have Petersson norm
Let denote the completed Rankin-Selberg L-function if ; otherwise define
where is the completed Riemann -function. Also define
with the convention that .
Let . Per our convention (Definition 2.4), we may think of as a vector or a diagonal matrix. Then the term in the Fourier-Whittaker expansion of is
where
(4.6) |
is the (or more informally the ) Hecke eigenvalue of , and
for some constant depending only on .
Proof.
See [GSW24].∎
4.2. Geometric side of the Kuznetsov trace formula
In this section, we obtain explicit descriptions of the terms and appearing on the geometric side of the Kuznetsov trace formula. In order to do this, we introduce Kloosterman sums for , which appear in the Fourier expansion of the Poincaré series. In the inner product , we replace with its Fourier expansion and unravel following the Rankin-Selberg method.
Theorem 4.1 (Geometric side of the trace formula).
Fix and ( is a nonzero constant; see Proposition 4.4). It follows that for ,
For the trivial element in the Weyl group , we define
where
(4.2) |
Proof.
We compute the inner product
Note that, as , the function (for any ) and . It follows from this and Proposition 2.3 above that
as claimed. ∎
5. Asymptotic formula for the main term
Proposition 5.1 (Main term in the trace formula).
Let satisfy and . There exist fixed constants (depending only on and ) such that the main term in the Kuznetsov trace formula (4.1) is given by
where and is the Kronecker symbol (i.e., if and ).
Proof.
It follows from the definition , making the change of variables , and noting that and , that
where the representation in terms of the norm of follows from the Plancherel formula in Corollary 1.9 of [GK12] and is a nonzero constant depending only on . Hence the main term for is thus
Let with . It then follows from Stirling’s asymptotic formula that
If we now make the change of variables for each , and we use the fact that the degree of is (see Definition 1.1) it follows that, if , as we have , where
and otherwise, the main term is zero. This gives the term in the statement of the proposition. The method of proof can be extended by using additional terms in Stirling’s asymptotic expansion for the Gamma function to obtain the additional terms. ∎
Remark 5.2.
Note that this doesn’t agree with [GSW21] in the case of because we have used a different normalization. Namely, the linear factors of agree with those defined previously, but we take a different power of each. Also, the gamma factors which appear in have a different : namely, what was in each gamma factor previously has been replaced by here.
6. Bounding the geometric side
The goal of this section is to use the bound given in Theorem 10.1 to prove the following, i.e., to bound the geometric side of the Kuznetsov trace formula.
Proposition 6.1.
Let be as above. Let . Let . Let as in (1.2). Then for sufficiently large and any , we have
where if with ,
Remark 6.2.
Assuming the lower bound conjecture for Rankin-Selberg L-functions, the resulting bound for the Eisenstein series contribution to the Kuznetsov trace formula (see Theorem 7.1) is of the magnitude to the power . Therefore, given Proposition 6.1 and Lemma A.13 (which says that ), in order for the bound from the geometric side of the trace formula to be less than the Eisenstein series contribution, it suffices that
which simplifies to give
Since we require that , we find that it suffices to take universally, meaning that the exponent of each term can be taken to be . In particular, for the case of , we see that this exponent is which is an improvement on the bound of obtained in [GSW21].
As remarked above, the main result that we will need is Theorem 10.1 or, more specifically, Remark 10.5, which states that for any , and for satisfying for each , that
(6.3) |
(The terms and are defined in Section 6.1 below. The function is defined in Theorem 9.2.)
This bound for is obtained via an integral representation denoted (see (8.4)) over variables valid for any with for each . The integral is taken over the lines . Essentially, the bound is then obtained by moving the lines of integration to for some .
The strategy for proving Proposition 6.1 will be to, first, introduce notation to rewrite in a simplified form. We do this in Section 6.1. Then, in Section 6.2 we give bounds for obtained by applying (6.3) to (with a parameter ) and to (with a parameter ) for general . In particular, we establish (6.2), bounding in terms of the product of three independent quantities , and . In Section 6.3, we will show that will converge provided that satisfies certain conditions (independent of ), and that for this choice of , also converges. We then determine (dependent on by and ) for which is also convergent. Finally, in Section 6.4, we complete the proof of Proposition 6.1 by simplifying the expression for the given choices of and .
6.1. Rewriting
Let and be the subgroups of consisting of diagonal matrices (with positive terms) and upper triangular unipotent matrices, respectively. Recall that if and , the modular character is defined to satisfy . Explicitly, it is given by
More generally, if , for
with we define
One checks that in the special case of for ,
(6.1) |
Similarly, if is the subgroup of consisting of lower triangular unipotent matrices and
then we can consider the character on which satisfies upon restricting the measure on to . It can be checked that
(6.2) |
6.2. Bounds for in terms of and
Since is determined by the Iwasawa decomposition of , we first make the change of variables . Then (6.3) implies that
(6.1) |
For the purposes of our analysis, we break up the integral in the -variables. To this end, let
For , define
Hence,
and (6.1) becomes
where
(6.2) |
Our strategy is now to, for each choice of , replace the terms with with the bound from (6.3) (in the first instance using a choice of , and in the second instance using ). Then we need to find choices of and for which the corresponding integrals converge and give good bounds.
Recall that if is the Iwasawa decomposition of an element , then . With this in mind, consider the Iwasawa decomposition , where , and . Then
is the Iwasawa form of , hence . Recall that the Iwasawa form of is assumed to be , meaning where , and . It can be shown [Jac67] that
(6.3) |
where for any . For example, in the case and , we find that and, for
that
In general, the values are always of the form plus a sum of squares of functions consisting of the entries of .
From (6.3), replacing with , we see that is bounded by
To similarly bound , we first remark that since
setting (and as usual), we see that
Therefore, it follows that
Recall that if is as in (6.3), if we define, for ,
and for a given choice of
then the bound on given in (6.2) can be replaced by
(6.4) |
We remark that in simplifying/finding , we have used (6.2). The basic strategy to prove Proposition 6.1 is now clear: we first find such that both and converge; then given this choice of , we determine a particular value of for which converges as well; finally, we work out the corresponding bounds on , and .
6.3. Restrictions on the parameters and
The trivial bound (see [DR98]) for the Kloosterman sum is given by
Hence is convergent whenever is chosen such that
From (6.1), if we set , then . More generally, converges in the following case:
(6.1) |
That this choice of makes converge is a consequence of the easily verifiable fact that
We assume henceforth that satisfies (6.1).
We next consider the convergence of . Recall that the Iwasawa form of is assumed to be , meaning where , and . Indeed, is given by (6.3). Then
The fact that the right hand side converges is a consequence of Jacquet [Jac67].
We now turn to the convergence of . Applying Lemma A.1 (which describes ), we see that
where
Hence, in order to bound (and thereby show that converges), we must choose such that converges. Clearly
(6.2) |
suffices, since making this choice implies that, for each ,
which converges (and gives the same value ) in either case.
6.4. Proof of Proposition 6.1
We have now shown that if and we choose as in (6.1) and via (6.2) accordingly, the right hand side of (6.4) converges, hence gives a bound for . Therefore, in order to complete the proof of Proposition 6.1, we need to first show that
and second that the given choice of and gives the claimed bound for the power of appearing in (6.4).
To complete the first of these tasks we note that, by (6.1) and the fact that is maximized (in ) when , we have
(6.1) |
for and . Similarly, using (6.1) and (6.2) we compute that, for and ,
(6.2) |
for sufficiently small. Note that the right hand side of (6.2)is a concave up parabola in , and therefore, on the interval , can attain its maximum only at or . So, if we can show that and both satisfy a suitable upper bound, then the same bound will hold for all .
We consider first the endpoint . Using (6.2) and the fact that , we find that
Again, is maximized when , so we conclude that
(6.3) |
for sufficiently small.
Next we consider the endpoint . From (6.2) we find that
(6.4) | ||||
the last step because . We find using calculus that, as a function of , the right hand side of (6.4) is maximized when . So
(6.5) | ||||
for small enough. From (6.3) and (6.5) it follows, again, that
for all and . This and (6.1) yield the desired bound on .
The second task is accomplished using Lemma A.9. ∎
7. Bounding the Eisenstein spectrum
Recall that if , with , then, by Theorem 4.1, the Eisenstein contribution to the Kuznetsov trace formula is given by
where
In this section we give bounds for in the case that and with .
7.1. The Eisenstein contribution to the Kuznetsov trace formula
The main result of this section is the following.
Theorem 7.1 (Bounding the Eisenstein contribution ).
Fix and a sufficiently large integer . Let , with . Then, assuming the Lower bound conjecture for Rankin-Selberg L-functions (see (1.2)), for we have the bound
7.2. Proof of Theorem 7.1.1
Proof.
We proceed by induction on , beginning with the case . In this case, the only parabolic subgroup is the minimum parabolic, or Borel, subgroup , and the only function corresponding to (see Definition 3.2) is the constant function . The Eisenstein contribution in this case, then, is simply the quantity .
By Theorem 4.1 in the case , we have
where . Now note that, by (3.7), . Moreover, by Definition 1.1, we have , so by Definition 1.4, we have
Furthermore, we see from Proposition 4.5 that
Then
We may restrict our integration to the domain , since decays exponentially otherwise. On this domain, we use Stirling’s bound (9.1) for the Gamma function, as well as the Vinogradov bound
We get
from which it follows immediately that So our desired result holds in the case .
We now proceed to the general case. For , in order to establish bounds for , we need to know that our main theorem is true for all . The reason this is needed is because we have to bound Rankin-Selberg L-functions with . This will require knowing the Weyl law with harmonic weights (Theorem 7.3) for . We may assume by induction, however, that this is indeed the case, i.e., the Weyl law with harmonic weights holds for all .
Now recall that, for the parabolic associated to a partition , we have
where is given by (see (3.7))
Since for all we see that
Now, for any where we have
It follows that
By Proposition 4.5, the coefficient of is given by
up to a non-zero constant factor with absolute value depending only on . To bound the divisor sum above we will use the bound of Luo-Rudnick-Sarnak [LRS99] for the Hecke Fourier coefficient of a (for Hecke-Maass cusp form given by
(A slightly stronger result has been obtained by Kim and Sarnak [Kim03]. However, the stated result above is sufficient for our purposes.) We immediately obtain the following bound for the divisor sum
It follows that
Lemma 7.1.
Proof.
This follows immediately from Lemma A.27. ∎
Next
We obtain the bound
Next, we bound the -integral above. It follows from Langlands conjecture (see 1.2) that for we have the bound
This together with the bound
implies that
(7.2) |
Since each (for , we can apply our inductive procedure together with the following theorem to bound
Theorem 7.3 (Weyl law with harmonic weights for with ).
Suppose with . Let be an orthogonal basis of Hecke-Maass cusp forms for ordered by eigenvalue. If are the Langlands parameters of , then
(7.4) |
Proof.
In [GSW21], all that was needed to prove this statement for was to have it be true for and , which was already known. A similar induction argument works in general. ∎
It immediately follows from the bound (7.2) and (7.4) that
Recall that which implies that
Next, by Lemma A.22 and It follows that
To complete the proof, we need to sum over all parabolics . It suffices, therefore, to consider the “worst case scenario” among the possible partitions for which the expression
is minimized. It is easy to see that this occurs when and , giving the bound . It follows that
Using (1.2), this immediately implies the desired result. ∎
Remark 7.5.
8. An integral representation of
Recall (see (1.4)) that
Using the formula for the inverse Lebedev-Whittaker transform given in [GK12], it follows that
where .
The strategy in this section for giving a representation of follows the same general outline as was used to obtain the results for and given in the papers [GK13] and [GSW21], respectively. As in the prior works, we express the Whittaker function as the inverse Mellin transform of its Mellin transform. (See Section 8.1.) Plugging this into the above formula gives an integral representation of in terms of an additional variable .
8.1. Normalized Mellin transform of Whittaker function
We introduce (as in [IS07]) the following Mellin transform and its inverse.
Definition 8.1 (Normalized Mellin transform of Whittaker function).
Let and such that . Let be the Whittaker function of Definition 2.3. The Mellin transform is
(8.2) |
and the inverse Mellin transform is given by
(8.3) |
As a consequence of this definition, we have
(8.4) | ||||
where with each .
We use the following theorem to make (8.4) explicit and to begin setting up an inductive method to bound for all .
Theorem 8.5 (Ishii-Stade).
Let and . Fix a Langlands parameter . Let with . Then
(8.6) |
where
and
8.2. A shifted term and the Ishii-Stade Conjecture
Our goal is to insert (8.6) into (8.4) and then shift the lines of integration in to , to the left of some of the poles of , which (see Theorem 10.1) occur at for every and . By Cauchy’s residue formula, this allows us to describe in terms of a the sum of a shifted term and finitely many shifted residue terms.
Definition 8.1 (shifted term).
Let be an integer and . The shifted term is given by the same formula as (8.4) but with replaced by :
(8.2) | ||||
One might be tempted to insert (8.6) into (8.2), but this is invalid if , because Theorem 8.5 requires that for each . To overcome this difficulty, we use shift equations as given in the following conjecture. This allows us to evaluate for .
Conjecture 8.3 (Ishii-Stade).
Let with ; let . Let . Then there exists a positive integer and, for each with , a polynomial and an -tuple , such that
(8.4) |
where the th coordinate of each is at least . Moreover, for each , we have
Proof of conjecture for .
Note that the case of the conjecture is trivial. Moreover, for a given and with , it’s enough to prove the conjecture for . The case then follows by applying the case to itself iteratively.
For and or , the conjecture follows immediately from the explicit formulae
respectively, together with the functional equation The case and is a consequence of [ST21, equations (21), (29), and (31)].
We now consider the case and . Note that it suffices to derive the appropriate recurrence relations for and (that is, for the variables and ); the cases and then follow from the invariance of under the involution
We follow an approach developed by Taku Ishii (personal correspondence). First, consider the case : we wish to show that
(8.5) |
is equal to a finite sum of terms where the first coordinate of each is at least one, and for each . To this end, let
(8.6) |
note that . Since , we have
(8.7) |
But for indeterminates , we have
(8.8) |
where is the elementary symmetric polynomial of degree in . So by equation (8.7) above, we have
(8.9) | ||||
since .
Now let , for , be the four-tuple with a one in the th place and zeroes elsewhere. By [IO14, Proposition 3.6], we have
(as operators acting on functions in the variable ), where the “Capelli elements” annihilate , and
. |
So by (8.9),
(8.10) | ||||
Recalling that the ’s are the elementary symmetric polynomials of degree in their arguments, we see that
for indeterminates . So (8.10) gives
the last step by the definition (8.6) of the ’s. This is our desired shift equation in .
The shift equation in is derived analogously. A fundamental difference in this derivation is that, in place of (8.8), we use the following expression involving Schur polynomials (see [Mac79, §I.3], especially Exercise 10 of that section):
Here,
and is the determinant of the matrix
The Schur polynomials are symmetric polynomials in the ’s, and are therefore expressible in terms of the elementary symmetric polynomials in the ’s. Techniques like those employed above, in the case , therefore apply. We omit the details. ∎
Remark 8.11.
The above proof, in the case (that is, for the variable —and therefore also for the variable ), generalizes to the case of , for any . For , we do not yet have a proof that works for all , though we expect that the above ideas and techniques should prove relevant. Indeed, using the above methods, and applying Mathematica to help with the more arduous calculations, we have been able to verify Conjecture 8.3 in full generality for .
We further note that, alternatively, one might continue in the ’s by shifting or deforming the lines of integration in (8.6). Unfortunately such an approach has, thus far, failed to yield suitable results. In particular, the residues that one obtains in moving these lines of integration past poles of the integrand are quite complicated, and do not seem to lend themselves to bounds of the type required to estimate .
8.3. is a sum of a shifted term and residues
Besides the shifted term (because we cross poles of upon shifting the lines of integration) there are also many residue terms. The residue terms will be parameterized by compositions of . Recall that a composition of length of a positive integer is a way of writing as a sum of strictly positive integers. Two sums that differ in the order define different compositions. Compare this, on the other hand with partitions which are compositions of for which the order doesn’t matter.
Definition 8.1.
(-admissible compositions) Let . A composition is termed -admissible if
The set of -admissible compositions of length greater than one is
Remark 8.2.
At times we may also notate a composition as an ordered list .
Definition 8.3 (-fold residue term).
Suppose that and is given by . Let
with for each . If has length two, we write . We define the -fold residue term
(8.4) | ||||
Remark 8.5.
In the shifted integral (8.4), if for some , there will be no residues coming from the integral in because we are not shifting past any poles. For this reason, one only obtains residue terms in the case that is -admissible. That said, equation (8.4) makes perfect sense even if is not -admissible. In this case, is identically zero.
Proposition 8.6.
Suppose that . Then there exists constants such that
Before giving the proof, we make some preliminary remarks and observations.
Remark 8.7.
Notice that an element of the symmetric group (i.e., the group of permutations of a set of elements) acts on and, by extension, on via
We can consider the analog to (8.4) obtained by replacing each instance of with :
We make two observations:
-
•
As varies over all compositions of length and varies over all possible permutations and varies over all , one obtains all possible -fold residues coming from shifting the lines of integration in . This is a consequence of Theorem 10.1 below.
-
•
The action of on ordered subsets of given by permuting the indices is trivial on , i.e., , and on the function
This implies that relabeling the variables by everywhere (1) doesn’t change the value of the integral, and (2) recovers the original integral given in (8.4).
Remark 8.8.
The constant is the size of the (generic) orbit of the action of on the set
Hence, defining the stabilizer of to be
we see that
Since the exact value of is irrelevant to our application, we omit its proof below and leave it instead to the interested reader.
Proof of Proposition 8.6.
Beginning with (8.4), we see that for any with for each . In order to compare this with , we successively shift the lines of integration in the variables for each such that (in descending order). If then shifting the line of integration from to doesn’t change the value of the integral in . In other words, there is a residue term if and only if the composition is admissible.
Beginning with the fact that
we may shift the line of integration in to . In doing so, provided that , we pass poles at for each . Hence, taking into account Remark 8.7, and considering (denoted ), it follows that
(8.9) | ||||
where is a constant (which can be verified to agree with the description given in Remark 8.8.)
We now shift the line of integration in to . As before, provided that , the Cauchy residue theorem and Remark 8.7 give
(8.10) | ||||
for constants for each of as claimed.
We next repeat this process shifting the integrals in for each of the terms on the right of (8.10), and then again for and so forth (skipping those for which ) until all of the lines of integration have been moved to for every possible integral. The claimed formula is now evident. ∎
8.4. Example:
We now consider the special case of where
Fix . Recall that . If we now shift the lines of integration to where , then we get additional residue terms corresponding to each composition and each as follows.
In general the composition (by abuse of notation, we also think of this as a vector so that ) corresponds to taking an -fold residue in the variables . Here is a table of the residues corresponding to the different compositions.
In each case . Not included in the table are the triple residues in for each . These correspond to the composition and .
8.5. The integral in terms of an explicit recursive formula for
At first glance, the following definition appears to be relevant only for the shifted -term, as it is essentially equal to , and not for the shifted residue terms. However, it will turn out to be pivotal to bounding the residue terms as well.
Definition 8.1 (The integral ).
Let be an integer and . Then we define
(8.2) |
As alluded to above, inserting the result of Theorem 8.5 into (8.2), we find that
Hence, giving a bound for requires only that we bound in the case of . However, much more is true: we will show that if is the composition , then can be bounded by the same product of ’s as above times a certain power of and a product of the form
for certain values which depend on the value of .
The significance of this fact should not be understated. Without it, we would be required to treat nearly every possible composition (hence each possible residue term) individually. Indeed, returning to the case of , as noted in Section 8.4 above, there were seven residue terms. The only symmetries that we were able to exploit in [GSW21] to help were that the and residues were equivalent, and the and residues were equivalent as well. This left five individual distinct cases, each of which required several pages of work to bound. So, although the method of this paper does require dealing with some tricky notation and combinatorics, it eliminates the need to treat each residue on its own terms.
9. Bounding
Recall that for satisfying and ,
(9.1) |
Theorem 9.2.
Let be as above and set . Then for any ,
where
The implicit constant depends on , and .
Theorem 8.5 allows us to write in terms of an integral of the product of several gamma functions and the lower rank Mellin transform where
Using this, we are able siphon off the contribution to the integrand of 9.1 that is independent of the variable . This in turn allows us to relate to and prove the result inductively.
9.1. Symmetry of integration in
Since the integrand of (9.1) is invariant under the action of acting on , we may restrict the integration to a fundamental domain. A choice of such a fundamental domain is
(9.1) |
Hence, (9.1) is equal, up to a constant, to the same integral but restricted to satisfying (9.1). In the sequel we will always assume that (9.1) holds.
9.2. Extended exponential zero set
Recall that Stirling’s asymptotic formula (for fixed and with ) is given by
(9.1) |
Definition 9.2 (Exponential and Polynomial Factors of a Ratio of Gamma Functions).
We call the polynomial factor of , and is called the exponential factor. For a ratio of gamma functions, the polynomial (respectively, exponential) factor is composed of the polynomial (respectively, exponential) factors of each individual Gamma function.
Lemma 9.3 (Extended Exponential Zero Set).
Assume that is a Langlands parameter satisfying
Then the integrand of (as a function of ) has exponential decay outside of the set , where
Remark 9.4.
Proof.
We first prove Lemma 9.3 in the case that . In the formula (9.1) for , replace with . Then assuming (9.1), the exponential factor is where
We see, therefore, that the exponential factor is negative unless
as claimed.
Let us suppose that and with (). In order to prove Lemma 9.3 using induction on , we make use of the change of variables
Observe that . By Lemma A.19 in the case that ,
Then in the integrand for we may substitute the formula for given in Theorem 8.5. We also use the fact (see Lemma A.26) that
and, via Stirling,
Note that (9.1) implies that , hence,
By the induction hypothesis, the second row of this expression has exponential decay outside of the set
(9.5) |
for each . (Recall that .)
The assumption and the definition of above imply that
Thus, the exponential factor coming from the final line in the expression above is where
We know that the integral defining is convergent. Therefore, it must be the case that . In order to find where , i.e., where there is not exponential decay, we seek for values for which
(9.6) |
In order for the -variables to cancel it is clear that for each it need be true that . With this assumption, equation 9.6 simplifies:
In order for this to hold true, it is necessary that for all , since otherwise, the coefficients of on each side of the inequality wouldn’t match. On the other hand, for all is sufficient as well since
This unique solution to (9.6) implies, therefore, that there is exponential decay in the integrand of above unless The inductive assumption (9.5) implies that
and
thus yielding the desired bounds on .
To complete the proof, we remark that if , in order to use the result of Theorem 8.5, we need to first apply the shift equations given in Corollary 9.8 below. This will allow us to rewrite as a sum over terms all of which have the same basic form as that for with . Each of these terms has precisely the same exponential factor since this depends only on the imaginary parts of the arguments of the Gamma functions, hence the same exponential zero set is determined in general. ∎
For each , we define
(9.7) |
Using this, the following corollary is easily deduced. (See [GSW21] for the case of .)
Corollary 9.8.
Let . There exists a sequence of shifts and polynomials such that
where
9.3. Proof of Theorem 9.2 in the case
Proof.
As in the proof of Lemma 9.3, we can replace with and estimate using Stirling’s bound. We may, moreover, restrict to the exponential zero set to see that
Due to the presence of the term , we may assume moreover that . Thus, we have the bound
In the statement of Theorem 9.2, the claimed bound is , where is as defined in Theorem 9.2. We have in fact proved that , where
If, , then we may shift the integral over to be as close to as desired; indeed, we may make the shift to the point that the error can be absorbed into the term in the power of . Therefore, since for all , the Theorem follows. ∎
9.4. Proof of Theorem 9.2 for general
Proof.
Let and assume that Theorem 9.2 has been shown to be true for all integers . It follows from Corollary 9.8 with that
By Theorem 8.5,
Next, we use the functional equation for the gamma function to rewrite
Additionally, we use the fact that the integrand has exponential decay unless , and by Lemma 9.3, each of the variables are bounded in terms of . This means that we may replace the polynomials with the bound . Note that in doing so, the dependence on is removed:
Notice that the conclusion of Proposition 9.2 follows from the last several steps by simply replacing by in the integrand (or, equivalently, replace by in the domain of integration), and then at the step where the functional equation of Gamma is used to remove from the gamma functions, we remove in the exact same fashion.
We deduce that
Note that we have also made the change of variable for each , and we are using the notation . (Using the terminology of Lemma A.19 in the case of , we have .) As in the case of , due to the presence of the exponential terms, we see that the integral has exponential decay unless .
Lemma 9.1.
Let and be as above. In particular, they are purely imaginary with . Suppose, moreover, that is in -general position. Then
Proof.
Let denote the integral we are seeking to bound.
The polynomial part (see Definition 9.2) of the Gamma functions in is
and the exponential factor (when taking all in unison) is negative for any outside of the interval defined in Lemma 9.3. That lemma together with the presence of the other exponential terms in our integral allow us to take trivial bounds for the polynomial part, namely that . (Recall that .) Thus we see that
The desired result now follows easily from this and the statement of Lemma A.3. ∎
Combining Lemma 9.1 with the bound for given immediately before the statement of the lemma, and applying Lemma A.19, Lemma A.26 and Lemma A.27 (in the case that and ), we now have the bound
To be more explicit, the polynomial is the portion of which involves the terms .
At this point, we combine each of the terms in the final row with the corresponding term in . Strictly speaking, what is actually happening here is that this has the effect of reducing the power of each factor of by at most
Since each of the corresponding exponents remains positive, the net result is to reduce the overall power of by
Using this, and accounting for the integration in (which may be assumed to take place only for ), we now may write
Obviously, at this point we want to apply the inductive hypothesis. Since at this point we only need to do so in the case that (i.e., ) for all , the reduction in the powers of the exponents of any one of the factors of , as occurred above, leaves the overall power positive. Therefore, there is no issue, and we can assert (additionally applying Lemma A.5) the bound
Taking gives the claimed bound. Since , it follows that
as claimed. ∎
In the course of proving Theorem 9.2 we also established the following result that we record here since it will be useful in its own right.
Proposition 9.2.
Suppose that . Then
As a shorthand for this result, we write .
10. Bounding
In this section we prove the following.
Theorem 10.1.
Let and . Suppose that satisfies for each . Let be the set of compositions with . Then, for
and as defined in Theorem 9.2, we have
(10.2) |
where
(10.3) |
and
(10.4) |
The implicit constant depends on both and .
Remark 10.5.
10.1. Explicit single residue formula
In order to bound the terms we need an explicit formula for the residues of the Mellin transform of the Whittaker function. The following result establishes this for the case of single residues (i.e., when the composition has length ) as a corollary of Conjecture 8.3 combined with a theorem of Stade [Sta01] for the “first” residues, i.e., for those residues corresponding, in the notation of the theorem, to .
Theorem 10.1.
Let be the Mellin transform of the Whittaker function on with purely imaginary parameters in general position. Let act on via
The poles of occur, for each , at
The residue at is equal to a sum over shifts of terms of the form
where
(10.2) |
with and being the portion of corresponding to and respectively. It is the case that . Note that we take as definition that . The same formula holds for the residue at by replacing each instance of with .
Remark 10.3.
Another way of writing the above expression for the residue would be to take the product over all with and replace with . The two versions are equivalent because if , then and
Sketch of proof.
In the case that , this result (for ) agrees with [Sta01, Theorem 3.1]. If , we need to first apply Conjecture 8.3 to rewrite the expression for around as a sum over shifts (with for each ) of terms . Of all of these terms, the only ones for which there is a pole at are those for which , in which case we can use the above referenced theorem of Stade to write down the residue. Doing so, we obtain the alternate expression referenced to in Remark 10.3. ∎
10.2. Explicit higher residue formulae
In order to generalize Theorem 10.1, we first establish notation related to the -fold residue of at
To this end, let where . By abuse of notation, we write
which agrees with the original but removes .
Similarly, if , we define
and
If then by we mean that for each .
With this notation in place, we can now state a generalization of Theorem 10.1.
Corollary 10.1.
Let (), and set as above. For each , let with
Let for . There exist positive shifts with such that the iterated residue of at
is equal to a sum over all such shifts of
where
Proof.
This follows easily by induction with the base case being Theorem 10.1. ∎
Remark 10.2.
Although it is possible to rewrite each of the terms appearing in the statement of Corollary 10.1 in terms of the variables and for various and , the exact description is unnecessary for our purposes.
10.3. Proof of Theorem 10.1
10.4. Bounds for single residue terms
In this section111Note that this section will be superseded by Section 10.5 which will prove the bound for any admissible with . This section treats the case . we bound in the case that of the composition . Since is a composition of length two, we may take (see Definition 8.3) .
Proof of (10.4) when .
Using Lemmas A.19, A.26 and A.28, we can rewrite
in terms of , and . Thus, together with Theorem 10.1, we see that Definition 8.3 in the case of a single residue term (i.e., ) satisfies the bound
In order to have the correct power of we need to shift the line of integration in to . Note that by Lemma A.14, no poles are crossed in doing so, and by Lemma A.15, taking and , we may replace the third to last line by
Let , and define similarly. Replacing the integral over by and factoring out the powers of , we see that
Note that by Proposition 9.2 we may remove the dependence on the shift . Hence
By (10.2),
Thus the integrals in and above are essentially the product of and . The only issue is that because, as seen in the fact that the variables and are shifted, we have
Therefore, we can rewrite the previous formula as
By Theorem 9.2, we have
Recall that . Hence, using the elementary identity
together with Lemma A.6,
This gives the desired bound provided that the exponent of the final is negative. Using the facts that is maximized when or and , we see that the final exponent is
(10.1) |
as claimed. ∎
10.5. Bounds for -fold residues
We consider a composition of of length given by . We may also write . Let as usual.
As a final piece of notation, let be defined via
Note that and more generally, defining , . Since (assuming that ) the Jacobians of the change of variables
and
are trivial, we see that (for )
(10.1) |
Proof of (10.4) when .
Using Remark A.29 and Corollary 10.1, we can bound by a sum over certain shifts each of the form
where
and
are the degrees coming from Remark A.29 and Corollary 10.1, respectively, and is as in Corollary 10.1. Note that, in addition to using the change of variables (10.1), we have used Lemma A.18 and Lemma A.20 to break up and rewrite the product of in terms of and .
The next step is to shift the lines of integration in the variables for (or, equivalently, for ) such that the real part of the exponent of each term is . In particular, this implies that we must shift the line of integration of to
(10.2) |
Provided that is sufficiently large, Lemma A.14 implies that this shift can be made without passing any poles. Moreover, Lemma A.15 implies that
(10.3) |
Note that the presence of the term implies that there is exponential decay for . As we will see momentarily, besides the polynomial terms , and (10.3), we just get a product of for some (to be determined) values . The upshot is that all of these polynomials can be bounded by to the degree of the polynomial plus . Hence, we can bound the expression above by
(10.4) |
where with and as above and
is the bound coming from the terms described in (10.3), simplified using Lemma A.21. Combining everything, we find that equals
Recall that the bound on is a sum of expressions of the form given in (10.4) for various shifts . However, using Proposition 9.2, we can remove the dependence on the shifts. Hence,
(10.5) |
Hence, setting where
we find that
Let . We now now apply Theorem 9.2 to each to obtain
Now we generalize the proof of Lemma A.6, keeping in mind that , to simplify the expression
Next, we write the sum over as
We plug this back in to get
from which it follows that the exponent of in the bound for above is
where
and
Hence,
Note that if and and , then this expression becomes
which agrees with (10.1).
Therefore, to complete the proof, we need only show that Indeed,
(with the final inequality being equality if and only if or ), as desired. ∎
Remark 10.6.
A critical step in the proof of (10.4) (either in the case of single residues, as is proved in Section 10.4 or higher order residues, as in Section 10.5) is to shift the lines of integration in the variables or . A feature of this work that is quite different from the case of as proved in [GSW21], is that no poles are crossed when making these shifts. This represents a major simplification. Recall from the discussion of Section 8.4 that in the case of there are two fundamentally different types of single residues, two different types of double residues and a triple residue. As it turned out, when making the additional shift for each of the single and double residues, one ends up with five additional residue terms. Taken all together, it was necessary to complete the analysis of writing down explicitly what the residues are in terms of gamma function s, finding the exponential zero set, applying Stirling’s formula and then obtaining a bound for ten(!) separate residues integrals. All of this was in addition to performing these steps for the shifted term.
Appendix A Auxiliary results
In an effort to avoid obstructing the flow of the argument in the main body of this paper, we will include here the many technical results that are used throughout. We remind the reader that the notational conventions that are used throughout the paper and this appendix are given in Section 2.1.
Lemma A.1.
Suppose that for some composition with . Then, if , is equal to
In particular,
Proof.
Let as above. In order to carefully analyze , we define This notation implies that . Now, let us think of the matrix as a block diagonal of the form where
Thus,
Let and and set
Then , the Iwasawa -variables of , satisfy . For , therefore, we see that
and for ,
from which the statement of the lemma follows directly. ∎
Definition A.2.
We say that is in -general position if the set
consists of distinct elements. We say that is in general position if it is in -general position for each .
Lemma A.3.
Suppose that there exists such that for each , the real part of is bounded by at least from any integer. Assume that is in -general position, for each , and . Assume that
and let . If , then
If there is an extra power of in the exponent (in which case the implicit constant will depend on ), and if , the integral is bounded by
Remark A.4.
The implicit –constant depends on , but in applications this will always be bounded.
Proof.
The bound in the case of is obvious, so we may assume henceforth that . Consider the set
For a fixed choice in -general position, let be the element of that has the greatest imaginary part, the next greatest imaginary part and so on. Hence .
Write . Note that . Upon applying Lemma A.3 from [GSW21], one obtains the bound
This is one of the possible summands on the right hand side of the statement of the lemma. Hence, regardless of the specific ordering which may arise for the given choice of , the claim follows. ∎
Lemma A.5.
Let . Then
Proof.
First, let us assume that . Then , hence
On the other hand, assuming that , we see that
as claimed. ∎
Lemma A.6.
Suppose that . Let
Then for any with ,
Proof.
We consider first the case of for some and all . For any , note that
(A.7) |
hence
Combining this with the other terms (which are easily shown to satisfy the analogous bound), the desired result is immediate. ∎
Remark A.8.
The function appears prominently in Theorem 10.1 and is critical in bounding the geometric side of the Kuznetsov trace formmula. Its graph is shown in Figure 1 in comparison to two other functions and .
In the case of , the function appears [GSW21] (see Theorem 4.0.1) as a bound for the function. Indeed, making necessary adjustments due to a different choice of normalization factors (see Remark 1.2), the result of [GSW21] is that
Theorem 10.1 establishes the same result but with replaced by . Although the improvement is slight, we remark that it is essential in Lemma A.6 and evidently allows the inductive method of the present paper to lead to the same asymptotic orthogonality relation as was established directly in [GSW21].
Lemma A.9.
Proof.
We first note that although the bound holds for any , for any , provided that is sufficiently close to a half integer. Lemma A.11 (as justified in Remark A.12) asserts that if is odd then elements from the set of all the possible values of and are indeed within of a half integer, and if is even then of values have this property. Hence,
(A.10) |
Since we see that
Therefore, summing over , we see (making use of the fact that ) that
Combining this with (A.10), the desired result is now immediate. ∎
Lemma A.11.
Let be a composition of with . Suppose that . Set , and for each we have and for each and we let Then
Remark A.12.
Note that the quantity given in Lemma A.11 in the case of odd is for any composition (with equality precisely when ). If is even then
Equality in this case occurs precisely when and are both odd and all other are even.
Proof.
For notational purposes, set
We first consider the case of odd, for which for all integers . Therefore, . As for , note that is equal to plus an integer as long as . Otherwise, . Hence .
In the case of even, exactly when is even. Hence . To the end of finding , we introduce the notation
for which it is clear that .
The cardinality of depends, obviously, on the integrality of . To determine this, we first assume that . Then
Therefore (since is even), if and only if . This implies that
or more concisely, . The determination of is similar: .
For , we see that
We see again that the integrality of depends on the parity of . If is odd,
and if is even,
One can check, arguing case by case as above, that in any event, the answer is . ∎
Lemma A.13.
Suppose that . The function
is invariant under permutations, i.e., for any , we have In particular, if is a partition of then is well defined. Moreover, among all partitions of (with ),
Proof.
Suppose that where
Then one can show by an elementary (albeit tedious) computation that In other words, is invariant under any transposition , hence invariant under all of .
Suppose that . If for some , then one shows via a straightforward computation that If with , it then follows, setting , that
Among all , the right hand side is minimized when . ∎
Recall that where , as defined at the beginning of Section 8.
Lemma A.14.
If and are fixed, then the function is holomorphic for all with .
Proof.
The fact that implies that is holomorphic is immediate, so the only question is what happens at the (simple) poles of . But these occur at for some integer which correspond to zeros of or . ∎
Lemma A.15.
For fixed and and , we have the bound
Proof.
This follows immediately from the Stirling bound . ∎
Definition A.16.
Let be Langlands parameters satisfying Let be a partition of with Then for each we define where
Remark A.17.
Note that for each In particular implies
Lemma A.18.
We have
Proof.
Computing directly, and using the fact that , we find that
as claimed. ∎
Lemma A.19.
Suppose that and satisfies Set for fixed , and define Then
Proof.
This is easily deduced as a special case of Lemma A.18 in the case that , , , and . ∎
Lemma A.20.
We continue the notation of Lemma A.18. Then
Proof.
Note that if , then for any and ,
and for any we have This immediately implies the desired formula. ∎
Lemma A.21.
Suppose that satisfies . Suppose that and set . Then
Proof.
We calculate
This final sum telescopes to give . Since , this implies the claimed result. ∎
The following result can be interpreted as a consequence—by counting (half) the number of gamma factors on each side of the equality—of Lemma A.20. Alternatively, proving it independent of Lemma A.20 gives further evidence that the product decomposition is correct.
Lemma A.22.
Let . We have
Proof.
We use induction on . If , the formula obviously holds. Let where . Then, by induction,
Since , it is evident that the desired formula holds. ∎
Lemma A.23.
Suppose . Then
Proof.
If the result is obviously true. Suppose that the result holds for . Write . Then
which can easily be shown now to simplify to , as claimed. ∎
Remark A.24.
Lemma A.26.
Let satisfy . Set , and let () and () be as in the previous lemma. We have
Proof.
This is easily deduced as a special case of Lemma A.20 when , , , and . ∎
We recall the definition of the polynomial given in Definition 1.1:
Also, we remind the reader of the polynomial notation given in Definition 2.3.
Lemma A.27.
Let , , and be as in Definition A.16. Set . Then
Proof.
This follows from the fact that if with then
Therefore, each constitutes a unique factor of for each . ∎
Lemma A.28.
Suppose that and . Then
where .
Proof.
Let . Then
From the definition of given in Definition 1.1, we see that for each such there are factors
of for which
This bound holds because the degree of the Pochhammer symbol in the denominator is , and by assumption, the degree of the numerator is . Combining all such terms with the remaining factors of , gives a polynomial of degree . ∎
Remark A.29.
Let and . The result of Lemma A.28 clearly generalizes to the case of taking multiple residues at for each (in reverse order). In this case, taking the product on the left hand side over all of the terms we obtain
where
Acknowledgments
Eric Stade would like to thank Taku Ishii for many helpful converations, and for the ideas constituting the proof of Conjecture 8.3 in the case . Michael Woodbury would like to thank the University of Colorado for wonderful accommodations while hosting him during the Spring 2022 semester. We would also like to thank the referees for many helpful comments.
References
- [Art79] James Arthur, Eisenstein series and the trace formula, Automorphic forms, representations and -functions (Proc. Sympos. Pure Math., Oregon State Univ., Corvallis, Ore., 1977), Part 1, Proc. Sympos. Pure Math., XXXIII, Amer. Math. Soc., Providence, R.I., 1979, pp. 253–274. MR 546601 (81b:10020)
- [Blo13] Valentin Blomer, Applications of the Kuznetsov formula on , Invent. Math. 194 (2013), no. 3, 673–729. MR 3127065
- [Blo21] by same author, The relative trace formula in analytic number theory, Relative trace formulas, Simons Symp., Springer, Cham, [2021] ©2021, pp. 51–73. MR 4611943
- [Bru78] R. W. Bruggeman, Fourier coefficients of cusp forms, Invent. Math. 45 (1978), no. 1, 1–18. MR 472701
- [Bru06] Farrell Brumley, Effective multiplicity one on and narrow zero-free regions for Rankin-Selberg -functions, Amer. J. Math. 128 (2006), no. 6, 1455–1474. MR 2275908
- [BZ20] Jack Buttcane and Fan Zhou, Plancherel distribution of Satake parameters of Maass cusp forms on , Int. Math. Res. Not. IMRN (2020), no. 5, 1417–1444. MR 4073945
- [CDF97] J. B. Conrey, W. Duke, and D. W. Farmer, The distribution of the eigenvalues of Hecke operators, Acta Arith. 78 (1997), no. 4, 405–409. MR 1438595
- [DR98] Romuald Dabrowski and Mark Reeder, Kloosterman sets in reductive groups, J. Number Theory 73 (1998), no. 2, 228–255. MR 1658031
- [FM21] Tobias Finis and Jasmin Matz, On the asymptotics of Hecke operators for reductive groups, Math. Ann. 380 (2021), no. 3-4, 1037–1104. MR 4297181
- [Fri87] Solomon Friedberg, Poincaré series for : Fourier expansion, Kloosterman sums, and algebreo-geometric estimates, Math. Z. 196 (1987), no. 2, 165–188. MR 910824
- [GJ72] Roger Godement and Hervé Jacquet, Zeta functions of simple algebras, Lecture Notes in Mathematics, Vol. 260, Springer-Verlag, Berlin-New York, 1972. MR 0342495
- [GK12] Dorian Goldfeld and Alex Kontorovich, On the determination of the Plancherel measure for Lebedev-Whittaker transforms on , Acta Arith. 155 (2012), no. 1, 15–26. MR 2982424
- [GK13] by same author, On the Kuznetsov formula with applications to symmetry types of families of -functions, Automorphic representations and -functions, Tata Inst. Fundam. Res. Stud. Math., vol. 22, Tata Inst. Fund. Res., Mumbai, 2013, pp. 263–310. MR 3156855
- [GL18] Dorian Goldfeld and Xiaoqing Li, A standard zero free region for Rankin-Selberg -functions, Int. Math. Res. Not. IMRN (2018), no. 22, 7067–7136. MR 3878594
- [GMW21] Dorian Goldfeld, Stephen D. Miller, and Michael Woodbury, A template method for Fourier coefficients of Langlands Eisenstein series, Riv. Math. Univ. Parma (N.S.) 12 (2021), no. 1, 63–117. MR 4284774
- [Gol06] Dorian Goldfeld, Automorphic forms and -functions for the group , Cambridge Studies in Advanced Mathematics, vol. 99, Cambridge University Press, Cambridge, 2006, With an appendix by Kevin A. Broughan. MR 2254662 (2008d:11046)
- [Gol15] by same author, Automorphic forms and L-functions for the group , Cambridge Studies in Advanced Mathematics, vol. 99, Cambridge University Press, Cambridge, 2015, With an appendix by Kevin A. Broughan, Paperback edition of the 2006 original [ MR2254662]. MR 3468028
- [GSW21] Dorian Goldfeld, Eric Stade, and Michael Woodbury, An orthogonality relation for (with an appendix by Bingrong Huang), Forum Math. Sigma 9 (2021), Paper No. e47, 83. MR 4271357
- [GSW24] by same author, The first coefficient of Langlands Eisenstein series for , Acta Arith. 214 (2024), 179–189.
- [HB19] Peter Humphries and Farrell Brumley, Standard zero-free regions for Rankin-Selberg -functions via sieve theory, Math. Z. 292 (2019), no. 3-4, 1105–1122. MR 3980284
- [IO14] Taku Ishii and Takayuki Oda, Calculus of principal series Whittaker functions on , J. Funct. Anal. 266 (2014), no. 3, 1286–1372. MR 3146819
- [IS00] H. Iwaniec and P. Sarnak, Perspectives on the analytic theory of -functions, no. Special Volume, Part II, 2000, GAFA 2000 (Tel Aviv, 1999), pp. 705–741. MR 1826269
- [IS07] Taku Ishii and Eric Stade, New formulas for Whittaker functions on , J. Funct. Anal. 244 (2007), no. 1, 289–314. MR 2294485
- [Jac67] Hervé Jacquet, Fonctions de Whittaker associées aux groupes de Chevalley, Bull. Soc. Math. France 95 (1967), 243–309. MR 0271275
- [Jan21] Subhajit Jana, Applications of analytic newvectors for , Math. Ann. 380 (2021), no. 3-4, 915–952. MR 4297178
- [JL85] H. Jacquet and K. F. Lai, A relative trace formula, Compositio Math. 54 (1985), no. 2, 243–310. MR 783512
- [JN19] Subhajit Jana and Paul D. Nelson, Analytic newvectors for , arXiv e-prints (2019), arXiv:1911.01880.
- [Kim03] Henry H. Kim, Functoriality for the exterior square of and the symmetric fourth of , J. Amer. Math. Soc. 16 (2003), no. 1, 139–183, With appendix 1 by Dinakar Ramakrishnan and appendix 2 by Kim and Peter Sarnak. MR 1937203
- [KS02] Henry H. Kim and Freydoon Shahidi, Functorial products for and the symmetric cube for , Ann. of Math. (2) 155 (2002), no. 3, 837–893, With an appendix by Colin J. Bushnell and Guy Henniart. MR 1923967
- [Lan76] Robert P. Langlands, On the functional equations satisfied by Eisenstein series, Lecture Notes in Mathematics, Vol. 544, Springer-Verlag, Berlin-New York, 1976. MR 0579181
- [Lap13] Erez Lapid, On the Harish-Chandra Schwartz space of , Automorphic representations and -functions, Tata Inst. Fundam. Res. Stud. Math., vol. 22, Tata Inst. Fund. Res., Mumbai, 2013, With an appendix by Farrell Brumley, pp. 335–377. MR 3156857
- [LM09] Erez Lapid and Werner Müller, Spectral asymptotics for arithmetic quotients of , Duke Math. J. 149 (2009), no. 1, 117–155. MR 2541128
- [LRS99] Wenzhi Luo, Zeév Rudnick, and Peter Sarnak, On the generalized Ramanujan conjecture for , Automorphic forms, automorphic representations, and arithmetic (Fort Worth, TX, 1996), Proc. Sympos. Pure Math., vol. 66, Amer. Math. Soc., Providence, RI, 1999, pp. 301–310. MR 1703764
- [Mac79] I. G. Macdonald, Symmetric functions and Hall polynomials, Oxford Mathematical Monographs, The Clarendon Press, Oxford University Press, New York, 1979. MR 553598
- [Mil02] Stephen D. Miller, The highest lowest zero and other applications of positivity, Duke Math. J. 112 (2002), no. 1, 83–116. MR 1890648
- [MT21] Jasmin Matz and Nicolas Templier, Sato-Tate equidistribution for families of Hecke-Maass forms on , Algebra Number Theory 15 (2021), no. 6, 1343–1428. MR 4324829
- [MW95] C. Mœglin and J.-L. Waldspurger, Spectral decomposition and Eisenstein series, Cambridge Tracts in Mathematics, vol. 113, Cambridge University Press, Cambridge, 1995, Une paraphrase de l’Écriture [A paraphrase of Scripture]. MR 1361168
- [Ram00] Dinakar Ramakrishnan, Modularity of the Rankin-Selberg -series, and multiplicity one for , Ann. of Math. (2) 152 (2000), no. 1, 45–111. MR 1792292
- [Sar87] Peter Sarnak, Statistical properties of eigenvalues of the Hecke operators, Analytic number theory and Diophantine problems (Stillwater, OK, 1984), Progr. Math., vol. 70, Birkhäuser Boston, Boston, MA, 1987, pp. 321–331. MR 1018385
- [Sar04] by same author, Nonvanishing of -functions on , Contributions to automorphic forms, geometry, and number theory, Johns Hopkins Univ. Press, Baltimore, MD, 2004, pp. 719–732. MR 2058625
- [Ser97] Jean-Pierre Serre, Répartition asymptotique des valeurs propres de l’opérateur de Hecke , J. Amer. Math. Soc. 10 (1997), no. 1, 75–102. MR 1396897
- [ST21] Eric Stade and Tien Trinh, Recurrence relations for Mellin transforms of Whittaker functions, J. Funct. Anal. 280 (2021), no. 2, Paper No. 108808, 25. MR 4166593
- [Sta01] Eric Stade, Mellin transforms of Whittaker functions, Amer. J. Math. 123 (2001), no. 1, 121–161. MR 1827280
- [Zha22] Qiao Zhang, Lower Bounds for Rankin-Selberg -functions on the Edge of the Critical Strip, arXiv e-prints (2022), arXiv:2211.05747.
- [Zho14] Fan Zhou, Weighted Sato-Tate vertical distribution of the Satake parameter of Maass forms on , Ramanujan J. 35 (2014), no. 3, 405–425. MR 3274875