<<

arXiv:1604.08551v6 [math.NT] 23 Feb 2019 field conductor 1.1. 00MteaisSbetCasfiain[S]cds 11R42 codes: subconvexity, [MSC] Keywords: Classification Subject Mathematics 2010 ..NnAcieenPae o L for Places Part Archimedean Exceptional Non for Places Archimedean 3.2. Non Estimations 3.1. Series Local Spherical in Vectors 3. Integral Classical Zeta Global of 2.4. Extension Formula Product Triple 2.3. Integral Regularized Regularized Zagier’s of 2.2. Extension Preliminaries 2.1. Miscellaneous Conventions and 2. Notations Method on 1.3. Discussion Result Main of 1.2. Statement 1.1. Introduction 1. ..Etmto o euaie L Regularized for Estimation Part Exceptional for 5.2. Estimation Estimations Global of 5.1. Complements Part Each for 5. Inversion Bounds Fourier series Generalized Eisenstein of 4.5. on Regroupment Truncation of Failure Integral 4.4. Truncated Interlude: of Bound to 4.3. Bound Reduction Period Global to 4.2. Reduction Result Main of 4.1. Proof 4. eerhprilyspotdb N-rn 001159 a 200021-125291 SNF-grant by supported partially Research .Apni:Dsaeiaini pca Case Special References a in Dis-adelization Acknowledgement Appendix: 6. ttmn fMi Result. Main of Statement F ih(sa)conductor (usual) with u oZge eeoe napeiu ae ooti h rele the obtain to paper previous a in series. developed Zagier to due Abstract. F .. h bound the i.e., ersnain oEsnti eis n eueaBurgess a deduce and series, Eisenstein to representations ihaayi conductor analytic with C ( π = ) egnrlz u rvosmto nsbovxt rbe f problem subconvexity on method previous our generalize We C | L L ( (1 π fnto vrnme ed ugs-yehbi bound. hybrid Burgess-type field, number over -function ∞ / UGS-IESBOVXT FOR SUBCONVEXITY BURGESS-LIKE 2 ) χ , C ) ( ≪ | π fin C F C ( ,teaslt ovrec for convergence absolute the ), ,ǫ χ ( .A anto,w pl h xeddter fregularized of theory extended the apply we tool, main a As ). C π If fin ( χ 4 π -Norms ) ep rhmda nltcconductor analytic archimedean resp. ) 4 1 sa(updl uoopi ersnaino GL of representation automorphic (cuspidal) a is -Norms / 4 − 1. (1 − Introduction 2 Contents θ A WU HAN ) / 16+ 1 ǫ o ayn ek characters Hecke varying for dDGSFgat00021L DFG-SNF-grant nd lk ucne on o ek characters, Hecke for bound subconvex -like attil rdc omlso Eisenstein of formulas product triple vant ℜ > s rGL or fteassociated the of 1 GL 2 χ 153647 × 1 vranme field number a over GL C 1 ( π ihcuspidal with ∞ ep analytic resp. ) d integral vranumber a over L -function 18 18 15 12 25 25 24 23 22 21 20 20 19 28 37 37 32 7 5 5 4 3 1 1 2 HAN WU

L(s, π) and the functional equation implies for any ǫ > 0, via the Phragm´en-Lindel¨of principle together with Iwaniec’s method [7, (19.16)] 1 to establish the necessary bound on the vertical line with ℜs =1+ ǫ with small ǫ> 0, the estimation at the central point

1/4+ǫ |L(1/2, π)| ≪F,ǫ C(π) , called the convex bound or the convexity. If the holds for L(s, π), then we have the optimal bound ǫ |L(1/2, π)| ≪F,ǫ C(π) , called the Lindel¨of Hypothesis. Reducing the exponent of C(π) from 1/4+ ǫ to 1/4 − δ + ǫ for some positive constant 0 <δ< 1/4 is called the subconvexity problem. More generally, for Q = C(π) resp. C(πfin) resp. C(π∞), an estimation

1/4−δ+ǫ |L(1/2, π)| ≪F,ǫ,C(π)/Q Q is called a (hybrid) subconvex bound resp. subconvex bound in the level aspect resp. subconvex bound in the archimedean aspect. In the simplest case, the first and most famous subconvex bound was obtained for the Riemann zeta- function by Weyl [21] (see for example [16, §6.6])

1/4−1/12+ǫ ζ(1/2+ it) ≪ǫ |t| , t ∈ R, which can be considered as (a special case of) a subconvex bound in the archimedean aspect for the Dirichlet L-functions. If χ is a Dirichlet of modulus q = C(χfin) ∈ N, Burgess [4] established his famous subconvex bound

it 1 − 1 +ǫ |L(1/2,χ|·|A )| = |L(1/2+ it,χ)| ≪t,ǫ q 4 16 . Later, Heath-Brown [9, 10] generalized Burgess’ result to include the t-aspect as the following hybrid bound 1 − 1 +ǫ |L(1/2+ it,χ)| ≪ǫ (q(|t| + 2)) 4 16 . Ever since, the subconvexity problem has become a venerable problem in analytic , in which both the optimal subconvex saving δ and the largest class of L-function mark the limit of techniques of analytic number theory. The saving δ =1/12 resp. 1/16 seem to be two natural barriers in the literature. They are called Weyl-type resp. Burgess-type subconvex bound for this historic reason. Moreover, it was discovered that for d > 1 the subconvexity problem of L(s, π) is intimately related with various equidistribution problems [6, 19]. More such relations can be found in [13, Lecture 5], as well as an application of the subconvexity problem in the level aspect for d = 1 and F imaginary quadratic. In this paper, we restrict to the case d = 1, i.e., when π = χ is a . In the case F = Q, many strong results are known besides the above bounds by Weyl, Burgess and Heath-Brown (for example [11]), especially in some special cases. For example, in [11] the case of q prime and of hybrid type is considered; in [15] with very strong result of sub-Weyl type, the case of q = pn a prime power and for the q-aspect is treated. Another interesting special case is when we restrict to χ = χq the quadratic character (and for q special, say square-free). Bounds of better quality than Burgess’ are known to hold for Weyl-type. For example, among many other good results Conrey and Iwaniec [5, Corollary 1.5] obtained 1 − 1 +ǫ |L(1/2+ it,χq)| ≪t,ǫ q 4 12 , which was recently generalized by Young [26, (1.5)] as

1 − 1 +ǫ |L(1/2+ it,χq)| ≪ǫ (q(|t| + 2)) 4 12 .

1The cited argument only treats the case for F = Q but it works for general number fields by replacing the relevant divisor function by the one for ideals of the ring of algebraic integers. BURGESS-LIKE SUBCONVEXITY FOR GL1 3

The above bound was further generalized for cube-free q by Petrow and Young [17]. Over a general number field, the best known result is the main theorem of Soehne [20, p.227], which follows the method 3 of Heath-Brown [9, 10] (it attains the Weyl-type bound if the usual conductor f = f0 is a cube): it 1/6+ǫ 1/2+ǫ 1/4+ǫ L(1/2+ it,χ) ≪ǫ,F C∞(χ|·|A )Nr(f) + Nr(f0) + Nr(f/f0) , where f0 is any ideal dividing f, the usual conductor of χ. In the work of Michel & Venkatesh [14, Theorem 5.1 & Section 5.1.7], a subconvex bound for Hecke characters χ was obtained with the subconvex exponent unspecified. We shall modify their approach and obtain a hybrid subconvex bound of Burgess-type for L-functions associated with Hecke characters over general number fields. Theorem 1.1. Let χ be a Hecke character of F with analytic conductor C(χ). We have

1 1 − 1−2θ +ǫ L( ,χ) ≪F C(χ) 4 16 , 2 ,ǫ

where θ is any constant towards the Ramanujan-Petersson conjecture.

Combining with Soehne’s bound, we deduce the following result. Corollary 1.2. Let χ be a Hecke character of F and t ∈ R. Denote by it T := C∞(χ · |·|A ), q := Cfin(χ). Then we have for any ǫ> 0 (T q)1/6 if T ≥ q1/2 1 ǫ 1/4 (1−2θ)/(3+2θ) 1/2 L( + it,χ) ≪ F (T q) · q if q ≤ T < q , 2 ǫ,  (T q)(3+2θ)/16 if T ≤ q(1−2θ)/(3+2θ)  where θ is any constant towards the Ramanujan-Petersson conjecture.

Proof. We apply Soehne’s bound with f0 = o the ring of integers of F, compare it with Theorem 1.1 and distinguish cases according to the relative size of T and q.  1.2. Discussion on Method. The proof is inspired by the method of our earlier work [22] where we established a Burgess-type subconvext bound for GL1 twists of a GL2 cuspidal representation π. 1 − 1−2θ +ǫ (1.1) |L(1/2, π × χ)| ≪π,F,ǫ C(χ) 2 8 . In this paper we show that it is possible to replace the cuspidal representation π by the Eisenstein series representation π(1, 1) and obtain the same bound. Theorem 1.1 then follows from the identity (1.2) L(s, π(1, 1) × χ)= L(s,χ)2. The main hurdle is to address the non square-integrability of Eisenstein series. For this we use a regular- ization process which we show does not harm the quality of the final out-come. By contrast, the original approach with truncation on Eisenstein series [14, §5.1.7] does destroy the Burgess-like quality (see §4.3 below for more details). It is worthwhile to give some comments on our method, which is quite different from the methods applied in the case F = Q by Burgess or Conrey-Iwaniec. Burgess’ method is based on the study of character sums of the shape χ(n), m1≤Xn≤m2 which makes use of Weil’s bound hence makes extensive use of the periodicity of the summand function n 7→ χ(n). Its direct generalization (1.3) χ(A), m1≤Nr(XA)≤m2 where A runs over integral ideals, loses the periodicity for the summand function. Our method can be viewed as a variant of Conrey-Iwaniec’s method (see [23, §1.1]). The main common feature is to bring 4 HAN WU a problem for GL1 into the setting for GL2, and to use the available knowledge on the spectral theory of automorphic representations for GL2. According to the comparison between [2] and [22], this method virtually consists of taking (1.3) into the fourth power and studying the cancellation of the resulted character sums by means of the “spectral decomposition of the shifted convolution sums” [1] instead of Weil’s bounds on character sums, hence it is also a variant of the original method of Burgess, “disguised” into the language of periods, which treats the archimedean aspect equally well. It would also be enlightening to point out the following explanation of the identity (1.2) in terms of an identity of periods. Any function in the induced model of π(1, 1) can be constructed from a Schwartz function Φ ∈ S(A2) as 1 2 +s 1+2s × fΦ(g) := |det g|A Φ((0,t)g)|t|A d t |s=0, × ZA whose Whittaker function is equal to

1 2 y × WΦ(a(y)) = |y|A F2(Φ)(t, )d t, × t ZA where F2(Φ) is the Fourier transform of Φ with respect to the second variable. Hence the period repre- senting the left hand side of (1.2) can be re-written, with the change of variables y 7→ yt, as

1 s− 2 × s s × × WΦ(a(y))χ(y)|y|A d y = F2(Φ)(t,y)χ(ty)|t|A|y|Ad td y. × × × ZA ZA ×A The right hand side of the above equation is exactly the integral representation of that of (1.2) if F2(Φ) is decomposable as the tensor product of two functions in S(A). In other words, if we read the above discussion inversely, we see that our method makes use of the “two dimensional Tate’s integral”, which brings in the structure of GL2 not present in the usual integral representation `ala Tate of L(s,χ). Remark 1.3. We also remark that the (global) factorizability χ(ty)= χ(t)χ(y) is responsable for such a link. For example, even though we have a similar identity of L-functions L(s, π(1, 1) × π)= L(s, π)2 where π is a cuspidal representation of GL2, an identity of the integral representations of the two sides does not seem to exist: the LHS is represented by the Rankin-Selberg integral for GL2 × GL2; the RHS is represented by twice/square the integral representation for the standard L-function for GL2. 1.3. Notations and Conventions. N is the set of natural numbers containing 0. All characters includ- ing Hecke characters are unitary. Non unitary ones will be called quasi-characters. If f is a meromorphic function around s = s0, we introduce the coefficients into its Laurent expansion (k) (k) f (s0) k f (s0) k f(s)= (s − s0) + (s − s0) . −∞

k hol k+k0 ∂ f k! ∂ k0 k (s0)= k+k (s − s0) f(s) . ∂s (k + k0)! ∂s 0 s=s0  hol Remark 1.4. The value f (s0) with notations as given above is intimately related with the finite part functional, denoted by f.p. in [8, Theorem (6.33)]. In addition to the notations given above, we import [22, Section 2.1], in which most of the notations are in fact standard. (For example, our normalization of measures is just the Tamagawa measure with the standard convergence factors.) We simply address the following points/differences. BURGESS-LIKE SUBCONVEXITY FOR GL1 5

(1) The number field is written in bold character F, with ring of algebraic integers o and ring of adeles A. v denotes a place of F. If v < ∞ is finite, we usually write v = p, which is identified with a prime ideal p of o. (2) We write the algebraic groups defined over F in bold characters such as G, N, B, Z etc, where G = GL2, B is the upper triangular subgroup of G, N ⊳ B is the unipotent upper triangular subgroup, and Z is the center of G. (3) K = Kv is the standard maximal compact subgroup of GL2(A), i.e. v Y SO2(R) if Fv = R K = SU (C) if F = C . v  2 v GL2(op) if v = p < ∞ × × (4) In GL2, for local or global variables x ∈ Fv or A, y ∈ Fv or A , we write 1 x y n(x)= , a(y)= . 1 1     (5) We use the abbreviation

[GL2]=GL2(F)Z(A)\GL2(A) = [PGL2].

(6) If f0 ∈ π(1, 1) the global principal series representation induced from trivial characters, which defines s −s a flat section fs ∈ π(|·|A, |·|A ), we normalize the usual Eisenstein series E(s,f0)=E(fs) by ∗ E (s,f0) := ΛF(1+2s)E(s,f0).

(7) In the above equation, ΛF(s) is the complete of ζF(s). More generally, L(·) ∗ denotes L-functions without factors at infinity. Λ(·) denotes the complete L-functions. We write ζF for the residue of ζF(s) at s = 1. We also introduce (−1) ΛF(−2s) λF (0) λF(s)= = + O(1). ΛF(2+2s) s Additional notations will be given in the course of proofs.

2. Miscellaneous Preliminaries 2.1. Extension of Zagier’s Regularized Integral. In this subsection we recall and summarize our extension of the theory of regularized integrals in [25, §5 & §6] without proofs. This extension fits well in the context of the Rankin-Selberg trace formula. It could not be well understood in the framework of the subconvexity problem. Hence we encourage the interested reader to read [25, §5 & §6] for a better understanding. We begin with the recall on the following space of functions on the automorphic quotient of GL2 over a general number field F with the ring of adeles A. Definition 2.1. ([25, Definition 5.14]) Let ω be a unitary character of F×\A×. Let ϕ be a smooth function on GL2(F)\GL2(A) with central character ω. We call ϕ finitely regularizable if there exist × × (1) K −1 unitary characters χi : F \A → C , αi ∈ C,ni ∈ N and smooth functions fi ∈ IndB(A)∩K(χi,ωχi ) for 1 ≤ i ≤ l, such that (1) for any M ≫ 1 ∗ −M ϕ(n(x)a(y)k)= ϕN(n(x)a(y)k)+ O(|y|A ), as |y|A → ∞, (2) we can differentiate the above equality with respect to the universal enveloping algebra of the lie algebra of GL2(A∞). Here we have written/defined the essential constant term

l 1 ∗ ∗ 2 +αi ni ϕN(n(x)a(y)k)= ϕN(a(y)k)= χi(y)|y|A log |y|A · fi(k). i=1 X 6 HAN WU

1 +αi In this case, we call Ex(ϕ)= {χi|·| 2 :1 ≤ i ≤ l} the exponent set of ϕ, and define 1 1 + +αi − +αi Ex (ϕ)= {χi|·| 2 ∈Ex(ϕ): ℜαi ≥ 0}; Ex (ϕ)= {χi|·| 2 ∈Ex(ϕ): ℜαi ≤ 0}. fr The space of finitely regularizable functions with central character ω is denoted by A (GL2,ω). fr Obviously A (GL2,ω) is stable under the right regular translation of GL2(A) and contains the Schwartz space with central character ω, hence the space of smooth cusp forms. It also contains any finite product fr of Eisenstein series ([25, Remark 5.19]). In the case ω = 1 and for any ϕ ∈ A (GL2, 1), the integral

∗ s−1/2 × R(s, ϕ) := (ϕN − ϕN)(a(y)κ)|y|A d ydκ × K ZA × is convergent for any ℜs >> 1 and admits meromorphic continuation. We use it to define the regularized integral as fr A (GL2, 1) → C,

reg 1 ϕ 7→ ϕ(g)dg := Ress=1/2R(s, ϕ)+ αi=−1 fi(κ)dκ . [PGL2] Vol([PGL2]) ni=0 K Z A(1) Z  Xχ1( )=1  GL2(A) −1 iµ  If f ∈ IndB(A) (χ1,χ2) such that χ1χ2 = |·|A for some µ ∈ R, we introduce the regularizing Eisenstein series as ([25, Definition 5.16])

F iµj reg Λ (1 − 2s − iµj) −1 2 (2.1) E (s,f)(g)=E(s,f)(g) − f(κ)dκ · χ1 (det g)|det g|A . ΛF(1+2s + iµ ) K j Z fr For any ϕ ∈ A (GL2,ω) with auxiliary data given in Definition 2.1 we define ([25, (5.3)])

nj nj ∂ ∂ reg (2.2) E(ϕ)= ℜα >0 E(αj ,fj)+ ℜα >0 E (αj ,fj), j nj j nj 1 ∂s 1 ∂s αj 6= +iµj αj = +iµj X 2 X 2 −1 2 −2iµj where µj ∈ R is defined only if ω χj (y)= |y|A . This defines a linear map fr fr A (GL2,ω) → A (GL2,ω), ϕ 7→ E(ϕ), 1 such that ϕ −E(ϕ) ∈ L (GL2,ω), which is GL2(A)-intertwining when Ex(ϕ) does not contain |·|A. We + − denote the image by E(GL2,ω). Moreover if Ex (ϕ) ∩Ex (ϕ) = ∅ then E(ϕ) is the unique element in 2 2 E(GL2,ω) such that ϕ −E(ϕ) ∈ L (GL2,ω) ([25, Proposition 5.25]), and we call it the L -residue of ϕ fr ([25, Definition 5.26]). In the case ω = 1, A (GL2, 1) is in the range of applicability of the regularized integral and ([25, Proposition 5.27]) reg ϕ(g)dg = (ϕ(g) −E(ϕ)(g))dg. Z[PGL2] Z[PGL2] In particular the above equation proves the GL2(A)-invariance of the regularized integral as a functional fr on A (GL2, 1), when Ex(ϕ) does not contain |·|A. In this case the above equality was originally due to Zagier [27]. We carefully generalized in [25, Theorem 5.12 & Definition 5.13] this theory into the adelic setting and proved the above equality without constraint on Ex(ϕ). In view of the inclusion ([25, Remark 5.19]) fr fr fr A (GL2,ω1) · A (GL2,ω2) ⊂ A (GL2,ω1ω2), we can consider the following bilinear form. Let πj , j =1, 2 be two principal series representations with central character ωj satisfying ω1ω2 = 1. Let Vj be the vector space of πj realized in the induced model ∞ from B(A) with subspace of smooth vectors Vj . We then get a GL2(A)-invariant bilinear form reg ∞ ∞ V1 × V2 → C, (f1,f2) 7→ E(f1)(g)E(f2)(g)dg, Z[PGL2] where E(fj) should be suitably regularized if πj is at a position which creates a pole/zero for the relevant Eisenstein series. We succeeded in [25, Theorem 6.5] to identify this bilinear form in the induced model. BURGESS-LIKE SUBCONVEXITY FOR GL1 7

In order to present the result, we need to introduce some extra notations. Precisely, if we identify for any s ∈ C the space of functions πs with H, where

GL2(A) s −s K πs := IndB(A) (|·|A, |·|A ), H := IndB(A)∩K1, then we can regard the intertwining operator Ms : πs → π−s as a map from H to itself. Using the flat section map H → πs,f 7→ fs, we mean

1 2 −s (Msfs)(a(y)κ) =: |y|A (Msf)(κ), i.e., Msfs = (Msf)−s .

Let e0 ∈ H be the constant function taking value 1. Define

PK : H → C, f 7→ f(κ)dκ, K Z where dκ is the probability Haar measure on K. We obtain a map from H to itself

Ms := Ms ◦ (I − PKe0), where I is the identity map. Since M is “diagonalizable”, we obtain the Taylor expansion as operators s f ∞ sn ∞ sn M f = M(n)f, resp. M f = M(n) f. s n! 0 1/2+s n! 1/2 n=0 n=0 X X Theorem 2.2. ([25, Theorem 6.5]) The regularized integralf of the productf of two unitary Eisenstein series is computed as: −1 −1 −1 −1 (1) If π1 = π(ξ1, ξ2), π2 = π(ξ1 , ξ2 ) resp. π2 = π(ξ2 , ξ1 ) and ξ1 6= ξ2, then reg (0) 2λF (0) (1) K K E(0,f1)E(0,f2)= (−1) P (f1f2) − P (M0 f1 ·M0f2), resp. Z[PGL2] λF (0) (0) λF (0) (1) K K K (−1) (P (f1M0f2)+P (f2M0f1)) − P (M0 f1 · f2). λF (0) −1 −1 (2) If π1 = π(ξ, ξ), π2 = π(ξ , ξ ), then

reg (2) (2) (1) (1) 4λF (0) 4λF (0) (1) E (0,f1)E (0,f2)= PK(f1f2)+ PK(f1 ·M f2) λ−1(0) λ−1(0) 0 Z[PGL2] F F (0) λF (0) (1) (1) 1 (3) (2) (1) K K K + −1 P (M0 f1 ·M0 f2) − P (M0 f1 · f2) − P (M0 f1 ·M0 f2). λF (0) 3 Here we have written ([25, (5.2)])

(−1) ∞ n ΛF(−2s) λF (0) s (n) λF(s) := = + λF (0). ΛF(2+2s) s n! n=0 X

2.2. Regularized Triple Product Formula. Let’s first complete the analysis for products of two Eisenstein series. We recall a lemma.

GL2(A) Lemma 2.3. ([25, Lemma 6.4]) Let f,f1,f2 ∈ ResK π(1, 1). For 0 6= s ∈ C small, we have for any n,n1,n2 ∈ N reg (n) reg,(n) 1 λF (s) E ( + s,f)= − PK(f); 2 (−1) Z[PGL2] λF (0) reg 1 1 Ereg,(n1)( + s,f )Ereg,(n2)( ,f )=0. 2 1 2 2 Z[PGL2] 8 HAN WU

We also recall the technique of deformation ([25, (6.1)]), inspired by the work of Michel & Venkatesh. fr 1 In general, if ϕ ∈ A (PGL2), E ∈E(PGL2) are given, so that ϕ − E ∈ L ([PGL2]), and if we can find fr continuous families ϕs ∈ A (PGL2), Es ∈E(PGL2) which coincide with ϕ, E at s = 0, then we have

reg reg reg (2.3) ϕ = ϕ −E = lim ϕs −Es = lim ϕs − Es . s→0 s→0 Z[PGL2] Z[PGL2] Z[PGL2] Z[PGL2] Z[PGL2] ! We turn to the study of regularized integrals of the form reg 1 1 E( ,f )E( ,f ). 2 1 2 2 Z[PGL2] Denote e = e1. We can write

reg 1 E (s,f)= f + (M f) + λF(s − )PK(f) e − e ; N s s −s 2 −s −1/2  f 1 n n Ereg,(n)( ,f)= f (n) + (−1)k(M(n−k)f)(k) N 2 1/2 k 1/2 −1/2 Xk=0   n+1 −1)f n (−1) λF (0) (n+1) n k (n−k) k K +P (f) · e−1/2 + (−1) λF (0)e−1/2 , ( n +1 k ) kX=1   reg reg,(n2) from which one easily deduce EN (1/2+ s,f1)EN (1/2,f2). We tentatively define

n2 3 n2 1 (n −k) Ereg(s) := E(n2)( + s,f f )+ (−1)kEreg,(k)( + s,f M 2 f ) 2 1 2 k 2 1 1/2 2 Xk=0   n2+1 (−1) f (−1) λF (0) reg,(n2+1) 1 +PK(f2) · E ( + s,f1) ( n2 +1 2

n2 n2 k (n2−k) reg,(k) 1 + (−1) λF (0)E ( + s,f1) k 2 ) Xk=1   1 +Ereg,(n2)( − s,f M f ) 2 2 1/2+s 1 reg,(n ) 1 reg,(n ) 1 + λF(s)PK(f ) · E f 2 ( − s,f ) − E 2 ( ,f ) . 1 2 2 2 2   Applying Lemma 2.3 with n1 = 0 together with (2.3), we get reg reg 1 reg,(n2) 1 reg 1 reg,(n2) 1 reg E ( ,f1)E ( ,f2) = lim E ( + s,f1)E ( ,f2) −E (s) 2 2 s→0 2 2 Z[PGL2] Z[PGL2] n2 k (n2) n2 (−1) (k) (n2−k) λF (−s) = lim λF (s)PK(f1M f2)+ PK(f2M1/2+sf1) s→0 (−1) 1/2 (−1) k λF (0) λF (0) Xk=0   n n2+1 (n2+1)f 2 (n2−k) f (k) (−1) λF (s) n2 k λF (0)λF (s) (n2) K K F +P (f1)P (f2) · + (−1) (−1) + λ (s)λF (−s) . ( n2 +1 k λF (0) ) kX=1   Taking Laurent expansions, we verify that the function in s in the range of the above limit is regular at s = 0, unlike its appearance. The symmetry

(k) (k) K K P (f1M1/2f2)=P (f2M1/2f1), ∀k ∈ N must be used. Moreover, it can be differentiatedf n1 timesf to deduce (3) of the following: BURGESS-LIKE SUBCONVEXITY FOR GL1 9

′ ′ ′ ′ ′ ′ Theorem 2.4. (1) If π1 6≃ π2 and ξ1 = ξ1, ξ2 6= ξ2 resp. ξ1 6= ξ1, ξ2 = ξ2 resp. ξ1 = ξ1, ξ2 = ξ2, 2 2 ξ1ξ2 6=1 and ξ1 ξ2 =1, then for any n1,n2 ∈ N e reg 1 1 reg 1 1 Ereg,(n1)( ,f ) · E(n2)( ,f )=0 resp. E(n1)( ,f ) · Ereg,(n2)( ,f )=0 2 1 2 2 2 1 2 2 Z[PGL2] Z[PGL2]

reg 1 1 resp. Ereg,(n1)( ,f ) · Ereg,(n2)( ,f )=0. 2 1 2 2 Z[PGL2] −1 −1 −1 −1 (2) If π1 = π(ξ1, ξ2), π2 = π(ξ1 , ξ2 ) resp. π2 = π(ξ2 , ξ1 ) with ξ1 6= ξ2, then for any n1,n2 ∈ N reg 1 1 E(n1)( ,f ) · E(n2)( ,f )=0, resp. 2 1 2 2 Z[PGL2]

is a linear combination with coefficients depending only on n1,n2 and λF(s) of

(n +n +1) (l) (l) K 1 2 K K P (M1/2 f1 · f2); P (M1/2f1 · f2)=P (f1 ·M1/2f2), 0 ≤ l ≤ max(n1,n2).

−1 −1 (3) If π1 = π(ξ, ξ), π2 = π(ξ , ξ ), then for any n1,n2 ∈ N reg 1 1 Ereg,(n1)( ,f ) · Ereg,(n2)( ,f ) 2 1 2 2 Z[PGL2]

is a linear combination with coefficients depending only on n1,n2 and λF(s) of

(l) (l) K K P (M1/2f1 · f2)=P (f1 · M1/2f2), 0 ≤ l ≤ max(n1,n2);

f (n +n +1) f K 1 2 K K P (M1/2 f1 · f2); P (f1)P (f2).

f Next, we give some complement of the main theorem of regularized integral [25, Theorem 5.12]. Let ∞ ξ1, ξ2,ω be Hecke characters with ξ1ξ2ω = 1. Let f ∈ π(ξ1, ξ2) and ϕ ∈ C (GL2,ω), i.e., a smooth function on GL2(F)\GL2(A) with central character ω. Suppose ϕ is finitely regularizable defined in Definition 2.1.

Proposition 2.5. For ℜs ≫ 1 sufficiently large,

1 ∗ s− 2 × R(s, ϕ; f) := (ϕN − ϕN)(a(y)κ)f(κ)ξ1(y)|y|A dκd y F× × K Z \A Z is absolutely convergent. It has a meromorphic continuation to s ∈ C. If in addition

Θ := max{ℜαj } < 0, j then we have, with the right hand side absolutely converging

R(s, ϕ; f)= ϕ · E(s,f), Θ < ℜs< −Θ. Z[PGL2] In the above region, the possible poles of R(s, ϕ; f) are −1 −1 (1) • 1/2+ iµ(ξ1ξ2 ) if ξ1ξ2 is trivial on A ; −1 • (ρ − 1)/2 where ρ runs over the non-trivial zeros of L(s, ξ1ξ2 ). In particular R(s, ϕ; f) is holomorphic for 0 ≤ℜs< min(−Θ, 1/2). 10 HAN WU

Proof. The proof is quite similar to that of [25, Theorem 5.12 (3)], except that Mfs is no longer explicitly computable. In fact, we have for T > 1, ℜs ≫ 1, using the standard Rankin-Selberg unfolding

ϕ · ΛT E(s,f)= R(s, ϕ; f) Z[PGL2] ∗ s−1/2 × − (ϕN − ϕN)(a(y)κ)f(κ)dκ ξ1(y)|y|A 1|y|A>T d y F× × K Z \A Z  ∗ −s−1/2 × − (ϕN − ϕN)(a(y)κ)Mfs(κ)dκ ξ2(y)|y|A 1|y|A>T d y F× × K Z \A Z 

l nj s+αj +iµj × (1) 1 ∂ T F A (1) + Vol( \ ) fj(κ)f(κ)dκ · 1χj ξ1(A )=1 · n  K n ! ∂s j s + α + iµ j=1 j j j X Z   l  ′ (−1)nj ∂nj T −s+αj +iµj (1) − fj(κ)Mfs(κ)dκ · 1χj ξ2(A )=1 · n ′ , K n ! ∂s j −s + α + iµ  j=1 j j j ! X Z ′  where µj resp. µj is such that

′ iµj iµj χj ξ1(y)= |y|A , resp. χj ξ2(y)= |y|A . We conclude by first shifting s to the desired region, then letting T → ∞. The possible poles are encoded −1 −1 in the possible poles of Mfs, which are included in those of L(1+2s, ξ1ξ2 ) in the above region (c.f. for example [24, Corollary 3.7, 3.10 & Lemma 3.18]). 

Proposition 2.6. Let notations be as in the previous proposition with Θ ≤−1/2. Recall

l 1 ∗ ∗ 2 +αj ni ϕN(n(x)a(y)k)= ϕN(a(y)k)= χj (y)|y|A log |y|Afj(k). j=1 X (1) If ξ1 6= ξ2, then

hol (n+n ) n reg ′ j ∂ R 1 (n) 1 λF (0) ( , ϕ; f)= ϕ · E ( ,f) − PK(fj f), ∂sn 2 2 j (−1) [PGL2] λF (0)   Z X (1) where the summation is over j such that ξ1χj (A )=1, αj + iµ(ξ1χj )= −1/2. (2) If ξ1 = ξ2 = ξ, then

hol (n+n ) n reg ′ j ∂ R 1 reg,(n) 1 λF (0) ( , ϕ; f)= ϕ · E ( ,f) − PK(fjf) ∂sn 2 2 j (−1) [PGL2] λF (0)   Z X (n) −1 + λF (0) · PK(f · (ξ ◦ det)) · ϕ · (ξ ◦ det), Z[PGL2] where the summation over j is as in the previous case.

Proof. The case (1) being simpler, we only give details for (2). By twisting, we may assume ξ = 1. Let s be small with ℜs< 0. The L2-residue of ϕ · E(1/2+ s,f) is given by

(nj ) E(s) := E (s +1+ αj ,fj f), j X where the summation is over j such that ℜαj > −1. Define ′ ∗ reg reg,(nj ) (nj ) E (s) := E (s +1+ αj ,fjf)+ E (s +1+ αj ,fjf) j j X X BURGESS-LIKE SUBCONVEXITY FOR GL1 11

′ ∗ where is the summation as in the statement and is the rest. By the previous proposition, we j j have X X 1 1 reg 1 R( + s, ϕ; f)= ϕ · E( + s,f)= ϕ · E ( + s,f)+ λF(s)PK(f) · ϕ 2 2 2 Z[PGL2] Z[PGL2] Z[PGL2] 1 = ϕ · Ereg( + s,f) −Ereg(s) − (E(s) −Ereg(s)) 2 Z[PGL2]   Z[PGL2] + λF(s)PK(f) · ϕ. Z[PGL2] 1 Since Ereg,(n)(s) is the L2-residue of ϕ · Ereg,(n)( + s,f), we can compare the finite parts of both sides 2 and conclude by ′ reg (nj ) E(s) −E (s)= λF (s)PK(fj f). j X  Finally, we state and prove a special case of regularized triple product formulas. The method used in the proof is applicable in any general case but on the one treated in the following theorem is used in the current paper.

Theorem 2.7. Let fj ∈ π(1, 1), j =1, 2, 3. Then for any n ∈ N reg 1 E∗(0,f ) · E∗(0,f ) · Ereg,(n)( ,f ) 1 2 2 3 Z[PGL2] is the sum of ∂nR hol 1 , E∗(0,f ) · E∗(0,f ); f ∂sn 2 1 2 3     and a weighted sum with coefficients depending only on λF(s) of (l) PK(M0 f1 · f2)PK(f3), 0 ≤ l ≤ 3; (l) K P (f1 · f2 · M1/2f3), 0 ≤ l ≤ max(2,n)& l = n + 3; (l) K P ((f1M0f2 + f2Mf0f1) · M1/2f3), 0 ≤ l ≤ max(1,n)& l = n + 2; (l) PK(M f ·M f · M f ), 0 ≤ l ≤ n & l = n +1. 0 1 0 2 f 1/2 3

f Proof. We shall only point out how the computation is effectuated, since the precise formulas are quite long, useless for the purpose of the current paper and would only obscure the idea.

∗ 2 reg,(2) 1 1 reg,(1) 1 (1) (1) E(f ,f ) := (ΛF) · E ( ,f f )+ E ( ,f ·M f + M f · f ) 1 2 2 1 2 2 2 1 0 2 0 1 2  1 1 + Ereg( , M(1)f ·M(1)f ) 4 2 0 1 0 2  2 ∗ ∗ ∗ ∗ is the L -residue of E (0,f1)·E (0,f2). Let ϕ := E (0,f1)·E (0,f2)−E(f1,f2), then we need to compute reg 1 reg 1 ϕ · Ereg,(n)( ,f )+ E(f ,f ) · Ereg,(n)( ,f ). 2 3 1 2 2 3 Z[PGL2] Z[PGL2] The first term is computed by Proposition 2.6 (2), involving reg (Λ∗ )2 reg ϕ = E∗(0,f ) · E∗(0,f )= F E(1)(0,f ) · E(1)(0,f ), 1 2 4 1 2 Z[PGL2] Z[PGL2] Z[PGL2] which is treated in Theorem 2.2 (2). The second term is treated in Theorem 2.4 (3).  12 HAN WU

2.3. Extension of Global Zeta Integral. Fixing a central Hecke character ω over a number field F, fr we extend the global part of the Hecke-Jacquet-Langlands’ theory to A (GL2,ω) Definition 2.1, as well fr as an analogue of “approximate functional equation”. Note that Eisenstein series are in A (GL2,ω).

fr Definition 2.8. Let ϕ ∈ A (GL2,ω) for some Hecke character ω. For a Hecke character χ and s ∈ C, ℜs ≫ 1, we define the zeta-functional by

1 ∗ s− 2 × ζ(s,χ,ϕ)= (ϕ − ϕN)(a(y))χ(y)|y|A d y, F× × Z \A where we recall the essential constant term l 1 ∗ ∗ 2 +αi ni ϕN(n(x)a(y)k)= ϕN(a(y)k)= χi(y)|y|A log |y|Afi(k). i=1 X

Proposition 2.9. ζ(s,χ,ϕ) has a meromorphic continuation to s ∈ C with functional equation ζ(s,χ,ϕ)= ζ(1 − s,ω−1χ−1, w.ϕ).

(1) −1 −1 It has possible poles at s = −αj − iµ(χj χ) with χj χ(A )=1 resp. s = 1+ αj + iµ(χj ω χ ) with −1 −1 (1) (1) iµ(χ) χj ω χ (A )=1, with pure order nj +1. Here µ(χ) ∈ R is defined for χ(A )=1 as χ(t)= |t|A .

Proof. By the invariance of ϕ at left by w, we can re-write the zeta-integral as

1 ∗ s− 2 × ζ(s,χ,ϕ)= (ϕ − ϕN)(a(y))χ(y)|y| d y y∈F×\A× A Z |y|A≥1 1 ∗ −1 s− 2 × + (w.ϕ − w.ϕN)(a(y ))ωχ(y)|y| d y y∈F×\A× A Z |y|A≤1 1 ∗ s− 2 × − ϕN(a(y))χ(y)|y| d y y∈F×\A× A Z |y|A≤1 1 ∗ −1 s− 2 × + w.ϕN(a(y ))ωχ(y)|y| d y. y∈F×\A× A Z |y|A≤1 ∗ We can calculate the integral concerning ϕN and get

1 ∗ s− 2 × ζ(s,χ,ϕ)= (ϕ − ϕN)(a(y))χ(y)|y| d y y∈F×\A× A Z |y|A≥1 1 ∗ −1 −1 2 −s × ∗ + (w.ϕ − w.ϕN)(a(y))ω χ (y)|y| d y + ζF(1)· y∈F×\A× A Z |y|A≥1

nj +1 nj +1 (−1) fj (1) (−1) fj (w) n +1 + −1 −1 n +1 ,  (s + αj + iµ(χj χ)) j (1 − s + αj + iµ(χj ω χ )) j  χ χ| =1 χ ω−1χ−1| =1 j XA(1) j X A(1)   from which we easily deduce all the assertions. 

∗ −1 We turn to the special case ϕ(g)=E (s0,ξ,ωξ ; f)(g) the usual completed Eisenstein series resp. reg −1 E (s0,ξ,ωξ ; f)(g) (2.1), for which the local computation is the same as for a cusp form. Note that ∗ ∗ in this case ϕN = ϕN, hence ϕ(a(y)) − ϕN(a(y))=Σα∈F× Wϕ(a(αy)).

∗ −1 reg −1 Proposition 2.10. Let ϕ(g)=E (s0,ξ,ωξ ; f)(g) resp. E (s0,ξ,ωξ ; f)(g) where ξ,ω are Hecke characters and f ∈ πξ,ωξ−1 . The zeta-functional has a decomposition as an Euler product in which only BURGESS-LIKE SUBCONVEXITY FOR GL1 13 a finite number of terms are not equal to 1: −1 ζ(s,χ,ϕ)=Λ(s + s0,ξχ)Λ(s − s0, ωξ χ)· −1 2 L (1+2s ,ω ξ ) s− 1 v 0 v v W (s0)(a(y ))|y | 2 d×y resp. −1 fv v v v v F× v Lv(s + s0, ξvχv)Lv(s − s0,ωvξv χv) v Y Z −1 Λ(s + s0,ξχ)Λ(s − s0, ωξ χ) ζ(s,χ,ϕ)= −1 2 · Λ(1+2s0,ω ξ ) −1 2 L (1+2s ,ω ξ ) s− 1 v 0 v v W (s0)(a(y ))|y | 2 d×y . −1 fv v v v v F× v Lv(s + s0, ξvχv)Lv(s − s0,ωvξv χv) v Y Z

The way given in the proof of Proposition 2.9 is not the only way of the analytic continuation of the global zeta functional. Another version of truncation on the integral is closely related to the classical approximate functional equation. Let h0 be a smooth function supported in the inteval [0, 2), being equal to 1 on [0, 1]. For any A> 0, we denote by h0,A the function t 7→ h0(t/A). We then have for ℜs ≫ 1

1 ∗ s × (2.4) ζ( + s,χ,ϕ)= (ϕ − ϕN)(a(y))χ(y)|y|A(1 − h0,A(|y|A))d y 2 F× × Z \A ∗ −1 −1 −s −1 × + (w.ϕ − w.ϕN)(a(y))ω χ (y)|y|A h0,A(|y|A )d y F× × Z \A ∗ s × − ϕN(a(y))χ(y)|y|Ah0,A(|y|A)d y F× × Z \A ∗ −1 −1 −s −1 × + w.ϕN(a(y))ω χ (y)|y|A h0,A(|y|A )d y. F× × Z \A ∗ For the last two lines, it is not hard to compute their analytic continuation using the form of ϕN and the ′ analytic continuation of the Mellin transform of h0 as (since h0 is of compact support contained in (1, 2)) ∞ dt 1 M(h′ )(s +1)+1 M(h )(s)= h (t)ts = − 0 0 0 t s s Z0 N−1 N −1 (N) = (−1) (s + j) M(h0 )(s + N), ∀N ∈ N,s ∈ C. j=0 Y Remark 2.11. We also have, first for ℜs ≪−1 then for s ∈ C ∞ dt M(h′ )(s + 1) M(1 − h )(s)= (1 − h )(t)ts = 0 = −M(h )(s). 0 0 t s 0 Z0 Then the last two lines of (2.4) are defined for s ∈ C as, writing sj =1/2+ αj + iµ(χjχ)

l nj ∗ ∂ s+sj − ζF(1) fj(1)δχ χ A M(h0)(s + sj ) j ∂snj j=1 X  l nj +1 l ∗ (−1) ∗ = ζF(1) fj (1)δχ χ − ζF(1) fj (1)δχ χ· j (s + s )nj +1 j j=1 j j=1 X X nj ∞ 1 nj nj −k+1 ′ s+sj k δ(s+sj ) nj −k (log A) h0(t)t log tdt · A δ dδ− k 0 0 Xk=0   Z Z l ∞ 1 ∗ ′ nj δ(s+sj ) nj +1 ζF(1) fj (1)δχj χ h0(t) δ t dδ log tdt, j=1 0 0 X Z Z  14 HAN WU

′ −1 −1 and writing sj =1/2+ αj + iµ(χj ω χ )

l n j ′ ∗ nj ∂ s−s ′ −1 −1 j ζF(1) fj (w)δχj ω χ (−1) · A M(h0)(s − sj ) ∂snj j=1 X   l nj +1 l ∗ (−1) ∗ nj = ζF(1) fj(w)δχ ω−1χ−1 + ζF(1) fj (w)δχ ω−1χ−1 (−1) · j (s′ − s)nj +1 j j=1 j j=1 X X nj ∞ 1 n ′ ′ j nj −k+1 ′ s−sj k δ(s−sj ) nj −k (log A) h0(t)t log tdt · A δ dδ+ k 0 0 Xk=0   Z Z l ∞ 1 ′ ∗ nj ′ nj δ(s−s ) nj +1 −1 −1 j ζF(1) fj (w)δχj ω χ (−1) h0(t) δ t dδ log tdt. j=1 0 0 X Z Z  l ′ ′ We separate the terms in the sum Σj=1 according as sj = 0 and sj 6= 0 resp. sj = 0 and sj 6= 0. For ′ sj = 0 resp. sj = 0, the finite part at s = 0 is bounded, with implied constants depending only on F,nj ,h0, by l l nj +1 nj +1 δ 1 O(|f (1) log A|) resp. δ −1 −1 1 ′ O(|f (w) log A|). χj χ sj =0 j χj ω χ sj =0 j j=1 j=1 X X ′ For sj 6= 0 resp. sj 6= 0, they are of size at s = 0, with implied constants depending only on F, αj ,nj ,h0 and an arbitrary N ∈ N,

′ l ℜsj nj l −ℜs nj A |fj (1) log A| A j |fj (w) log A| δχ χ1s 6=0O resp. δ −1 −1 1s′ 6=0O . j j |s |N χj ω χ j |s′ |N j=1 j j=1 j ! X   X −1 The first resp. second line of (2.4) is supported in |y|A ∈ [A, ∞) resp. [(2A) , ∞), hence is well-defined ∗ for all s ∈ C by the rapid decay of ϕ − ϕN. For the second line at s = 0, we can apply Mellin inversion to see

∗ −1 −1 −1 × (w.ϕ − w.ϕN)(a(y))ω χ (y)h0,A(|y|A )d y F× × Z \A ∗ s 1 −1 −1 ds1 = ζF(1) A 1 ζ( + s ,ω χ , w.ϕ)M(h )(s ) 2 1 0 1 2πi Zℜs1=c1≫1 which is bounded, with implied constant depending only on F,h0 and an arbitrary N ∈ N, as |ζ( 1 + s ,ω−1χ−1, w.ϕ)| c1 2 1 |ds1| A O N−1 , ℜs =c ≫1 |s + m| 2π Z 1 1 m=0 1 ! −1 −1 where c1 can be chosen as any real number suchQ that the integral defining ζ(1/2+ s1,ω χ , w.ϕ) is absolutely convergent for ℜs1 ≥ c1. Similarly, we have for any B > 0 that

∗ × (ϕ − ϕN)(a(y))χ(y)(1 − h0,B)(|y|A)d y F× × Z \A 1 ds2 = − B−s2 ζ( + s ,χ,ϕ)M(h )(−s ) 2 2 0 2 2πi Zℜs2=c2≫1 is bounded, with implied constant depending only on F,h0 and an arbitrary N ∈ N as |ζ( 1 + s ,χ,ϕ)| −c2 2 2 |ds2| B O N−1 , ℜs =c ≫1 |s + m| 2π Z 2 2 m=0 2 ! where c2 can be chosen as any real number such thatQ the integral defining ζ(1/2+ s2,χ,ϕ) is absolutely convergent for ℜs2 ≥ c2. BURGESS-LIKE SUBCONVEXITY FOR GL1 15

Definition 2.12. For any function h : R+ → C and any Hecke character χ, we define the h-truncated fr (zeta-)integral on A (GL2,ω) as

∗ × ζ(h,χ,ϕ)= h(|y|A)χ(y)(ϕ − ϕN)(a(y))d y. F× × Z \A As a summary, we have obtained:

Proposition 2.13. Take h0 as indicated in the beginning, some positive constants 0 0. For any ϕ ∈ A (GL2,ω) with χi, αi,ni given in Definition 2.8, we write ′ −1 −1 sj =1/2+ αj + iµ(χj χ) resp. sj =1/2+ αj + iµ(χj ω χ ) −1 −1 (1) if χj χ resp. χj ω χ is trivial on A , and µ defined in Proposition 2.9. Then the difference 1 ζhol( ,χ,ϕ) − ζ(h,χ,ϕ) 2 is bounded, with implied constants depending only on F, αj ,nj ,h0 and an arbitrary N ∈ N, as the sum of (1) Degenerate Polar Part: l l nj +1 nj +1 δ 1 O(|f (1) log A|)+ δ −1 −1 1 ′ O(|f (w) log A|). χj χ sj =0 j χj ω χ sj =0 j j=1 j=1 X X (2) Normal Polar Part:

′ l ℜsj nj l −ℜs nj A |fj(1) log A| A j |fj (w) log A| δχ χ1s 6=0O + δ −1 −1 1s′ 6=0O . j j |s |N χj ω χ j |s′ |N j=1 j j=1 j ! X   X (3) Lower Part: |ζ( 1 + s,ω−1χ−1, w.ϕ)| |ds| Ac1 O 2 . N−1 2π Zℜs=c1≫1 m=0|s + m| ! (4) Upper Part: Q |ζ( 1 + s,χ,ϕ)| |ds| B−c2 O 2 . N−1 2π Zℜs=c2≫1 m=0|s + m| ! In (3) resp. (4), c1 > 0 resp. c2 > 0 is any real number such that the integral defining ζ(1/2 + −1 −1 Q s,ω χ , w.ϕ) resp. ζ(1/2+ s,χ,ϕ) is absolutely convergent for ℜs ≥ c1 resp. ℜs ≥ c2.

Remark 2.14. We will use the bound for the normal polar part in the case sj is bounded away from 0. 2.4. Classical Vectors in Spherical Series. Let F be a non-archimedean local field with uniformizer ̟, absolute valuation |·|, valuation ring o & ideal p and cardinality of the residue class field q. Denote GL2(F) s −s by πs = IndB(F) (|·| , |·| ) the principal series representation of PGL2(F), where s ∈ C. For s ∈ iR, πs are unitary with the underlying Hilbert spaces identified with the same one

GL2(F) K ResK πs = IndB(F)∩K(1, 1) =: H, K = GL2(o).

We regard πs,s ∈ C as a family of representations of GL2(F) on H. By the Branching law, there is a canonical decomposition as K-representations

H = Hn, n≥0 where Hn is an irreducible K-subspace of H generatedM by a unitary vector en, such that {e0, ··· ,em} m form an orthonormal basis of the K0[p ]-invariant subspace of H for any m ∈ N. These vectors en, called “classical vectors” 2 in [22, Definition 5.4], are defined up to a factor of modulus 1. We determine/choose them as follows. First of all, we impose

e0(κ)=1, ∀κ ∈ K.

2They are called “paramodular” vectors by Brooks Roberts and Ralf Schmidt. 16 HAN WU

′ Lemma 2.15. Let en ∈ πs be defined by qs + q−s e′ = π (a(̟−1)).e − e , 1 s 0 q1/2 + q−1/2 0 ′ −n −1/2 s −s −n+1 −1 −n+2 en = πs(a(̟ )).e0 − q (q + q )πs(a(̟ )).e0 + q πs(a(̟ )).e0, ∀n ≥ 2. ′ ′ m Then if s ∈ iR, {e0,e1,...,em} is an orthogonal basis of the K0[p ]-invariant subspace of H for any m ∈ N. −1 −m m Proof. If s ∈ iR, {e0, πs(a(̟ )).e0,...,πs(a(̟ )).e0} is a basis of the K0[p ]-invariant subspace of ′ H for any m ∈ N. Then use the Macdonald formula [3, Theorem 4.6.6] to verify that en is orthogonal to −m πs(a(̟ )).e0 for 0 ≤ m ≤ n − 1.  Lemma 2.16. If we define q1/2 + q−1/2 q +1 q−ns e := e′ , e := e′ , ∀n ≥ 2, 1 qs+1/2 − q−s−1/2 1 n q − 1 1 − q−1−2s n r m then en is independent of s and {e0, ··· ,em} form an orthonormal basis of the K0[p ]-invariant subspace of H for any m ∈ N. Moreover, the dimension dn of Hn is given by n n−2 d0 =1, d1 = q, dn = q − q ,n ≥ 2.

n n+1 Proof. If we write on = ̟ o − ̟ o,n ≥ 1, then a b D = B(o)wN(o)= ∈ K : c ∈ o× , 0 c d    a b D = B(o)N (o )= ∈ K : c ∈ o , 1 ≤ n ≤ m, n − n c d n ′   ∞  Dm = K0[m]= ∪n=mDn are the double cosets of K w.r.t. B(o) and K0[m]. From the computation a̟−n b c−1d̟n 1 c−1(ad − bc) ∗ = , c̟−n d −1 0 0 c̟−n       a̟−n b 1 0 d−1̟−n(ad − bc) ∗ = , c̟−n d −d−1c̟−n 1 0 d       one easily deduces that qs+1/2 − q−s−1/2 qs+1/2 − q−s−1/2 e′ | = · (−q−1/2), e′ | = · q1/2, k ≥ 1; 1 D0 q1/2 + q−1/2 1 Dk q1/2 + q−1/2 ′ ′ −1−2s n(s+1/2) −1 en |Dk =0, 0 ≤ k ≤ n − 2, en |Dn−1 = (1 − q )q (−q ), ′ −1−2s n(s+1/2) −1 en |Dk = (1 − q )q (1 − q ), k ≥ n. The assertion follows since the mass wn of Dn, assuming the mass of K is 1, is given by q q−(n−1) w = , w = (1 − q−1),n ≥ 1. 0 q +1 n q +1 For the “moreover” part, it suffices to notice 1/2 (2.5) en(κ)= dn hκ.en,eni and to evaluate the above equation at κ = 1. 

Corollary 2.17. We record some special values of en: 1 n = 0; 1 n = 0; 1/2 −1/2 en(1) = q n = 1; en(w)= −q n = 1;  n n−2 1/2  (q − q ) n ≥ 2;  0 n ≥ 2.   BURGESS-LIKE SUBCONVEXITY FOR GL1 17

−1 Lemma 2.18. The two base {e0,e1,... } and {e0, πs(a(̟ )).e0,... } of the subspace of classical vectors in H are related as follows. (1) For n ≥ 2, we have q1/2 + q−1/2 qs + q−s e = π (a(̟−1)).e − e ; 1 qs+1/2 − q−s−1/2 s 0 qs+1/2 − q−s−1/2 0 q +1 q−ns e = π (a(̟−n)).e − q−1/2(qs + q−s)π (a(̟−n+1)).e + q−1π (a(̟−n+2)).e . n q − 1 1 − q−1−2s s 0 s 0 s 0 r (2) For n ≥ 2, we have  qs+1/2 − q−s−1/2 qs + q−s n π (a(̟−1)).e = e + e ; π (a(̟−n)).e = c(n,l; s)e , s 0 q1/2 + q−1/2 1 q1/2 + q−1/2 0 s 0 l Xl=0 where the coefficients c(n,l; s)= cp(n,l; s) are given by − n (n+1)s −(n+1)s (n−1)s −(n−1)s q 2 q − q q − q c(n, 0; s)= − q−1 ; 1+ q−1 qs − q−s qs − q−s   − n−1 −1−2s q 2 1 − q c(n, 1; s)= (qns − q−ns) ; 1+ q−1 1 − q−2s −1−2s − n−l ns (2l−2−n)s 1 − q q − 1 c(n,l; s)= q 2 (q − q ) , 2 ≤ l ≤ n. 1 − q−2s q +1 r (3) For n ≥ 2, we have n qs+1/2 − q−s−1/2 qs + q−s π (a(̟)).e = w.e + e ; π (a(̟n)).e = c(n,l; s)w.e , s 0 q1/2 + q−1/2 1 q1/2 + q−1/2 0 s 0 l Xl=0 where w ∈ K is the Weyl element and the coefficients c(n,l; s) are the same as in (2).

Proof. (1) is merely a re-statement of Lemma 2.15 and 2.16. For (2), we first use Lemma 2.15 to deduce a relation of two formal power series ∞ ∞ ′ n −n n −1/2 s −s −1 2 enX = πs(a(̟ )).e0X 1 − q (q + q )X + q X n=2 n=0 ! X X   −1 −1/2 s −s − e0 − πs(a(̟ )).e0X + q (q + q )e0X ∞ −n n −1/2 s −s −1 2 = πs(a(̟ )).e0X 1 − q (q + q )X + q X n=0 ! X   qs + q−s − e − e′ − q−1 e X. 0 1 q1/2 + q−1/2 0   Reverting it, we obtain ∞ ∞ q(n+1)s − q−(n+1)s π (a(̟−n)).e Xn = q−n/2 Xn s 0 qs − q−s n=0 n=0 ! X X ∞ qs + q−s · e + e′ − q−1 e X + e′ Xn 0 1 q1/2 + q−1/2 0 n n=2 !   X and conclude by inserting Lemma 2.16. (3) follows from (2) by noting ̟n π (a(̟n)).e = π (wa(̟−n)w−1)π ( ).e = w.π (a(̟−n)).e . s 0 s s ̟n 0 s 0    18 HAN WU

Corollary 2.19. If R(s): πs → π−s is the normalized intertwining operator, sending e0 to e0, then R(s) acts on Hn by multiplication by −(1−2s) −2ns 1 − q µ(n; s)= µp(n,s)= q . 1 − q−(1+2s)

Proof. This is a special case of the computation in [24, §3.4.3]. Here is another proof. R(s)en is equal to q +1 q−ns π (a(̟−n)).e − q−1/2(qs + q−s)π (a(̟−n+1)).e + q−1π (a(̟−n+2)).e q − 1 1 − q−1−2s −s 0 −s 0 −s 0 r   1 − q−(1−2s) = q−2ns · 1 − q−(1+2s) q +1 qns π (a(̟−n)).e − q−1/2(qs + q−s)π (a(̟−n+1)).e + q−1π (a(̟−n+2)).e , q − 1 1 − q−1+2s −s 0 −s 0 −s 0 r   the last line being equal to en since it is independent of s.  Remark 2.20. For F archimedean, we have similar computations already available in [3, Proposition 2.6.3] and [22, §2.7]. We recall them without proof.

• F = R. Hn is the subspace of vectors v such that cos α sin α .v = einαv − sin α cos α   and Hn 6= {0} only if 2 | n ∈ Z. We have |n|−2 k +1 − 2s µ(n; s)= µv(n,s)= . 2|k=0 k +1+2s Y • F = R. Hn is the subspace on which SU2(C) acts as the unitary irreducible representation of dimension n +1 and Hn 6= {0} only if 2 | n ∈ N. We have n/2 k − 2s µ(n; s)= µv(n,s)= . 2|k=1 k +2s Y We write e0 ∈ H0 for the spherical function taking value 1 on K = SO2(R) or SU2(C). Definition 2.21. For n,l ∈ Z, we write l  n to mean either 0 ≤ l ≤ n or n ≤ l ≤ 0. We extend the definition of en resp. µ(n; s) resp. c(n,l; s) for n,l ∈ N to n,l ∈ Z,l  n by requiring

e−n := w.en, µ(−n; s)= µ(n; s), c(−n, −l; s)= c(n,l; s).

3. Local Estimations 3.1. Non Archimedean Places for Exceptional Part. We work on a non archimedean place p and omit the subscript p for simplicity of notations. Recall en defined in Lemma 2.16 and Definition 2.21, but change s to s0. To emphasize the dependence on s0, we write en,s0 ∈ πs0 for the flat section associated with en, and Wn(s0, ·) the associated Kirillov function in the Kirillov model K(πs0 , ψ) of πs0 , with respect to an unramified additive character ψ of F. Recall the local zeta functional

s−1/2 × ζ(s, W ) := W (y)|y| d y, W ∈ K(πs0 , ψ). F× Z Lemma 3.1. The ratios of zeta-functions

ζ(1/2+ s, Wn(s0, ·)) ζp,n(s,s0)= ζn(s,s0) := , n ∈ Z ζ(1/2+ s, W0(s0, ·)) are determined by BURGESS-LIKE SUBCONVEXITY FOR GL1 19

• ζ0(s,s0)=1 and

1/2 −1/2 s0 −s0 q + q −s q + q ζ1(s,s0)= q − ; qs0+1/2 − q−(s0+1/2) qs0+1/2 − q−(s0+1/2) • if n ≥ 2, then

−ns0 q +1 q −ns −1/2 s0 −s0 −(n−1)s −1−(n−2)s ζn(s,s0)= q − q (q + q )q + q ; q − 1 1 − q−1−2s0 r   • if n< 0, we have ζn(s,s0)= ζ−n(−s,s0). Proof. From the relation of zeta functions ζ(s,a(̟n).W )= |̟|−nsζ(s, W )= qnsζ(s, W ), the desired formulas are simple consequences of those in Lemma 2.18 (1) and (3).  Corollary 3.2. Assume ℜs = ǫ> 0 small, |n| is bounded by a constant and 0 ≤ k ≤ 2, then we have k ∂ nǫ−|n|/2 k ζn(s, 1/2) ≪ǫ q (log q) . ∂sk 0

3.2. Non Archimedean Places for L4-Norms. Using the notations of the previous subsection, we define the Rankin-Selberg local zeta ratios for n1,n2,n ∈ Z and s1,s2,s ∈ C

Wn (s1,g)Wn (s2,g)en,s(g)dg n1 n2 n n1 n2 n N(F)\PGL2(F) 1 2 (3.1) ζp( )= ζ( ) := . s1 s2 s s1 s2 s W0(s1,g)W0(s2,g)e0,s(g)dg R N(F)\PGL2(F) Lemma 3.3. (1) We have R −n −n −n n n n ζ( 1 2 )= ζ( 1 2 ). s1 s2 s s1 s2 s

(2) Let n2 =0= s2. The ratio is non-vanishing only if |n1| = |n|. (3) Recall the dimension dn of Hn computed in Lemma 2.16. We have for n ≥ 2 −1/2 −(s+1/2) n 0 n q +1 dn 2 + (n − 1)(1 − q ) ζ( )= · q−ns 0 0 s q − 1 1 − q−1 1+ q−(s+1/2) r  2 + (n − 2)(1 − q−(s+1/2)) 2 + (n − 3)(1 − q−(s+1/2)) −2q−1/2−(n−1)s + q−1−(n−2)s , 1+ q−(s+1/2) 1+ q−(s+1/2)  while for n =1, 1 0 1 q1/2 + q−1/2 2q−s 2 ζ( )= − . 0 0 s q1/2 − q−1/2 1+ q−(s+1/2) q1/2 + q−1/2   Proof. (1) is a consequence of the w-invariance of the Rankin-Selberg local zeta functional. For (2), we may assume n ≥ 0 by (1). It suffices to notice that

en(κ)Wn1 (s1,gκ)dκ K Z −1/2 is non-vanishing only if |n1| = |n|, since by (2.5) en(κ)κdκ is dn times the orthogonal projection K Z onto the en-vector of Hn. In particular, we deduce for n ≥ 0

−1/2 en(κ)Wn(s1,gκ)dκ = dn Wn(s1,g). K Z Hence for (3), we are reduced to computing

s−1/2 × Wn(0,a(y))W0(0,a(y))|y| d y. F× Z 20 HAN WU

By Lemma 2.18 (1), we are again reduced to computing

−n s−1/2 × W0(0,a(y̟ ))W0(0,a(y))|y| d y F× Z∞ = q−m/2(m + 1) · q−(n+m)/2(n + m + 1) · q−(n+m)(s−1/2) m=0 X 2 + (n − 1)(1 − q−(s+1/2)) = q−ns · , (1 − q−(s+1/2))3 and conclude from it. 

Corollary 3.4. For any integer k ≥ 0, we have

∂k n 0 n ζ( ) ≪ q−|n|(log q)k. ∂sk 0 0 1/2 k

4. Proof of Main Result The main structure of proof is similar to our former work [22]. We shall only emphasize on the differences and the extra difficulties. We shall not recall the intuition of the method in terms of the equidistribution of certain lines approaching the low lying horocycles, but refer the reader to the first two pages of [22, §3].

4.1. Reduction to Global Period Bound. The fixed GL2 automorphic representation π = π(1, 1) is realized as completed Eisenstein series E∗(0, ·). We imitate the cuspidal case by choosing

∗ ϕ0 =E (0,f0) ϕ = n(T ).ϕ0, where f0 is the spherical function taking value 1 on K in the induced model of π(1, 1). Writing the normalized Whittaker functions as

∗ W0,v := ζv(1)W0,v = ζv(1)Wf0,v, we get by Proposition 2.10 an expression of the relevant L-function

−1 ∗ × × n(Tv).W (a(yv))d yv 1 2 ∗ × Fv 0,v 1 L( ,χ) = n(Tv).W0,v(a(yv))d yv 1 ζ( ,χ,ϕ), 2  F× 2  2 v v<∞ R Lv( 2 ,χv) vY|∞ Z Y   where the global zeta-integral is defined in Definition 2.8 and reduces in our case to

s−1/2 × ζ(s,χ,ϕ)= (ϕ(a(y)) − ϕN(a(y)))χ(y)|y|A d y, F× × Z \A whose value at s =1/2 must be interpreted via analytic continuation, unlike the cuspidal case.

Proposition 4.1. We can choose Tv with |Tv|v ∈ [C(χv), 2C(χv)] such that

∗ × × n(Tv).W (a(yv))d yv 1 ∗ × Fv 0,v − F 2 n(Tv).W0,v(a(yv))d yv 1 ≫ Q , F× 2 v v<∞ R Lv( 2 ,χv) vY|∞ Z Y where Q = C(χ) = ΠvC(χv) is the analytic conductor of χ (we keep this notation in what follows). Proof. This is a special case of [23, Proposition 2.4] for the “Option (B)”.  BURGESS-LIKE SUBCONVEXITY FOR GL1 21

4.2. Reduction to Bound of Truncated Integral. We are reduced to bounding ζ(1/2,χ,ϕ). It can be defined only via analytic continuation. However, we can still approximate it with truncated integral, just as what the classical approximate functional equation does. The outcome is that we essentially only need to estimate an integral of compact domain, which is equivalent to a finite sum in the classical setting. Recall Definition 2.12. We shall apply Proposition 2.13 with A = Q−κ−1,B = Qκ−1 for some κ ∈ (0, 1) to be chosen later, with c1 = c2 =1/2+ ǫ and with h0 and h specified there. Lemma 4.2. Assume Q is bounded away from 0. Then we have for any small ǫ> 0

1 − κ +ǫ ζ( ,χ,ϕ)= ζ(σ ∗ h,χ,ϕ)+ OF (Q 2 ) 2 ,h0,ǫ κ κ−1 × − 2 +ǫ 2 +ǫ = σ ∗ h(|y|A)ϕ(a(y))χ(y)d y + OF,h0,ǫ(Q + Q ), F× × Z \A where σ is the following average of Dirac measures 1 −1 σ = δ|̟ | |̟ ′ | 2 v v v v′ ME ′ v,vX∈IE with a parameter E > 0 to be chosen later and E I = {v< ∞ : q ∈ [E, 2E],T =0},M = |I | ≫ . E v v E E log E

Proof. We only need to consider the case for h since σ ∗ h gives bounded translations. Compared to [22, Lemma 3.2], the new situation is: • The normal polar part is non-vanishing; ∗ × • At v | ∞, W0,v is no longer of compact support in Fv , hence [22, Corollary 4.3] used in [22, Section 6.1] for local archimedean bounds on the vertical line ℜs = −1/2 − ǫ need to be re-considered; • There is a new passage from the first line to the second line, for which the estimation of an integral against the constant term ϕN need to be done. We proceed to bound each part appearing in Proposition 2.13 one by one. (0) The degenerate polar part is vanishing. iµ (1) The normal polar part and the integral against ϕN are non-vanishing only if χ = |·| for some µ ∈ R, [Fv :R] [F:Q] in which case C(χv) = 1 for all v< ∞ and C(χv) ≍ |µ| for all v | ∞. Hence Q ≍ |µ| . Note that by [25, Proposition 7.33], there are µ1,µ2 ∈ C depending only on F such that 1 1 2 2 × (4.1) ϕ0,N(zn(x)a(y)κ)= µ1|y|A + µ2|y|A log|y|A, ∀z ∈ Z(A), x ∈ A,y ∈ A ,κ ∈ K. If we write 1 1 2 2 ϕN(zn(x)a(y)κ)= |y|A f1(κ)+ |y|A log|y|Af2(κ), then we can easily calculate F 2 − [ v:R] f1(1) = µ1, f2(1) = µ2, f1(w)= µ1 (1 + |Tv| ) 2 , v|∞

F Y F 2 − [ v:R] 2 − [ v:R] f2(w)= µ2 (1 + |Tv| ) 2 log (1 + |Tv| ) 2 . v|∞ v|∞   We thus find that the normal polarY part, which is of the formY A1/2|f (1)| A1/2|f (1) log A| A−1/2|f (w)| A−1/2|f (w) log A| O 1 + 2 + O 1 + 2 |1/2+ iµ|N |1/2+ iµ|N |1/2 − iµ|N |1/2 − iµ|N     can be bounded as O(Q−N ) for any N ∈ N, due to the arbitrarily large denominators. For the integral −κ−1 κ−1 against ϕN, since h(t) has support contained in [Q ,Q ] with |h(t)| ≤ 1, we find that (note that n(T ).ϕN(a(y)) = ϕN(a(y)))

κ−1 × 2 +ǫ (4.2) h(|y|A)ϕN(a(y)χ(y)d y = OF,h0,ǫ(Q ). F× × Z \A 22 HAN WU

(2) We turn to the lower part. Recall the choice c1 = 1/2 + ǫ. The relevant zeta-function has a decomposition as a finite product

1 −1 1 −1 2 ∗ −1 s × ζ( + s,χ , w.ϕ)= L( + s,χ ) · wn(Tv).W0,v(a(yv))χv (yv)|yv|vd yv × 2 2 Fv vY|∞ Z ∗ −1 s × F× wn(Tv).W0,v(a(yv))χv (yv)|yv|vd yv · v . 1 −1 2 v<∞ R Lv( 2 + s,χv ) Y At an archimedean place v, say Fv = R, to the local integral

∗ −1 s × wn(Tv).W0,v(a(yv))χv (yv)|yv|vd yv × ZR 2 −1 2 s ∗ −1 s × = χv(1 + Tv ) (1 + Tv ) W0,v(a(y))ψv (yTv)χv (yv)|yv|vd yv × ZR 1/2+ǫ [23, Lemma 3.12 (2)] gives a bound as O(|Tv|v ). For Fv = C the argument is similar, using [23, Lemma 3.13 (2)]. At v< ∞, [22, Corollary 4.8] is still applicable. We thus deduce that, using the convex bound of L(1/2+ s,χ−1),

1 −1 1 ǫ 1 +ǫ ζ( + s,χ , w.ϕ) ≪F | + s| C(χ) 2 . 2 ,ǫ 2

− κ+1 1 +ǫ − κ +ǫ The desired bound is thus O(Q 2 Q 2 )= O(Q 2 ). − κ +ǫ (3) The treatment of the upper part is similar and simpler. It gives the desired bound O(Q 2 ). 

4.3. Interlude: Failure of Truncation on Eisenstein series. We have approximated ζ(1/2,χ,ϕ) by some smoothly truncated integral

h h ϕ(a(y))χ(y)d×y where := h(|·|) F× × F× × F× × Z \A Z \A Z \A ! as in the cuspidal case. We then would like to apply the Cauchy-Schwarz inequality 2 h h h × × 2 × ϕ(a(y))χ(y)d y ≤ d y · n(T ). |ϕ0(a(y))| d y, F×\A× F×\A× F×\A× Z Z Z 2 and apply Fourier inversion to |ϕ0| , interchange the order of summation and estimate each component 2 as in the cuspidal case. This is not possible because ϕ0 is no longer of rapid decay hence |ϕ0| is not square integrable any more. A first idea, which was already employed in [14, §5.1.7], is to (smoothly) X truncate the Eisenstein series ϕ0 up to some height X, denoted by Λ ϕ0. For example, if we naively choose X no less than the height of the truncation on the integral (namely, Qκ+1 in the notation of [22]), we find h h × X × ϕ(a(y))χ(y)d y = n(T ).Λ ϕ0(a(y))χ(y)d y. F× × F× × Z \A Z \A X 2 X 2 We could continue the argument by replacing ϕ0 with Λ ϕ0. But then some L -Sobolev norm of Λ ϕ0 come in as a multiplicative factor of the final estimation of the above integral. This causes no problem 4 in the cuspidal case since the relevant norm is bounded by some L -Sobolev norms of ϕ0, which depends X only on π. This is no longer the case for Λ ϕ0 since its norms all depend on X, hence some positive power of Q = C(χ). This means that in the final optimization just before [22, Remark 3.11], we would have to replace EQ−1/4+θ/2 by something like XEQ−1/4+θ/2, which completely destroys the Burgess-like quality. Indeed in the thesis version of [22] we have pursued this idea and were only able to obtain a saving (1 − 2θ)/12 instead of the Burgess-like saving (1 − 2θ)/8. A better way, which is also the main innovation of this paper, is to generalize the spectral decomposi- tion/Fourier inversion into a space of functions suitably larger than the square-integrable ones. As we BURGESS-LIKE SUBCONVEXITY FOR GL1 23

fr 2 have seen in §2.2, A (GL2, 1) is a good candidate: it contains |ϕ0| and differs from smooth vectors in the L2-space only by (non-unitary) Eisenstein series. Precisely, we shall decompose

h h h 2 × 2 × × n(T ). |ϕ0(a(y))| d y = n(T ). |ϕ0| −E (a(y))d y + n(T ).E(a(y))d y, F×\A× F×\A× F×\A× Z Z   Z 2 2 where we have written E = E(|ϕ0| ) for simplicity. Without amplification, the norms of |ϕ0| −E depend 3 only on ϕ0 hence π; with amplification the relevant norms have contributions as small as (log E) (see Theorem 5.4) where E denotes the length of the amplifiers, which is negligible. Hence we can treat the 2 term related with |ϕ0| −E in the same way as in the cuspidal case without harming the quality of the bound. Since E is determined explicitly by ϕ0, the term related with it is explicitly estimable. We will treat the estimation and see that its contribution does not harm the quality of the final bound, neither. Namely, the generalized spectral decomposition fits as well with the estimation of the integrals as the ordinary one in the cuspidal case. Note that the simpler ϕ0 is, the simpler E is.

4.4. Regroupment of Generalized Fourier Inversion. We make the strategy described in the above remark more precise. Introducing

′ 1 ̟p σ = χ δ −1 , χ 2 ̟p̟ M ̟p′ p′ E p p′ ,X∈IE   we can apply the Cauchy-Schwarz inequality

2 2 × ′ × σ ∗ h(|y|A)ϕ(a(y))χ(y)d y = h(|y|A)σχ ∗ ϕ(a(y))χ(y)d y F×\A× F×\A× Z Z

× ′ 2 × ≤ h(|y|A)d y · h(|y|A)|σ χ ∗ ϕ(a(y))| d y. F× × F× × Z \A Z \A

The first integral in the last line is of size OF(log Q), hence negligible. Opening the square, we get

′ 2 1 ̟p1 −1 ̟p2 ̟p1 ̟p2 |σχ ∗ ϕ(a(y))| = 4 χ χ a .ϕ · a .ϕ (a(y)) M ̟p′ ̟p′ ̟p′ ̟p′ E p ,p′ ,p ,p′ ∈I 1 2 1 2 ! 1 1 X2 2 E         ̟ ̟ ′ 1 ̟p2 p1 p2 = 4 χ~pa n(T ). a .ϕ0 · ϕ0 (a(y)), M ̟p′ ̟p′ ̟p E ~p∈I4  2    1 2   XE where we have abbreviated

̟p1 −1 ̟p2 (1) χ~p := χ χ ∈ C . ̟p′ ̟p′  1   2  Decomposing the non square-integrable function as

̟ ̟ ′ ̟ ′ p1 p2 p2 (4.3) a .ϕ0 · ϕ0 = a .ϕ0(~p)+ E0(~p), ̟p′ ̟p ̟p  1 2   2  where the L2-residual part (2.2) or [25, Definition 5.26] is given an abbreviated notation

̟ ̟ ′ p1 p2 (4.4) E0(~p) := E a .ϕ0 · ϕ0 , ̟p′ ̟p   1 2   24 HAN WU applying to ϕ0(~p) = ϕ0(~p)N + ϕ0(~p)cusp + ϕ0(~p)Eis the Fourier inversion decomposition in the sense of [22, Theorem 2.18] and regrouping the two constant terms, we can rewrite the second integral as

′ 2 × h(|y|A)|σχ ∗ ϕ(a(y))| d y F× × Z \A

1 ̟p1 ̟p2 × = 4 χ~p h(|y|A) a .ϕ0 · a .ϕ0 (a(y))d y M F×\A× ̟p′ ̟p′ E ~p∈I4  1   2  !N XE Z 1 1 + χ ζ(h, 1,n(T ).ϕ (~p) )+ χ ζ(h, 1,n(T ).ϕ (~p) ) M 4 ~p 0 cusp M 4 ~p 0 Eis E ~p∈I4 E ~p∈I4 XE XE 1 ~ + 4 χ~pζ(h , 1,n(T ).E0(p)), M ̟p′ /̟p2 E ~p∈I4 2 A XE where we have used the h-truncated zeta-integral in Definition 2.12. Note that we can drop ̟ ′ /̟ in p2 p2 the second integrand, since its adelic norm is contained in [1/2, 2]. 4.5. Bounds for Each Part. Lemma 4.3. We have for any ǫ> 0

1 ̟p1 ̟p2 × −2+ǫ ǫ κ−1+ǫ F 4 χ~p h(|y|A) a .ϕ0 · a .ϕ0 (a(y))d y ≪ ,ǫ E Q + Q . M F×\A× ̟p′ ̟p′ E ~p∈I4 Z  1   2  !N XE

Proof. We write and decompose

̟p1 ̟p2 × SN(~p; h) := h(|y|A) a .ϕ0 · a .ϕ0 (a(y))d y, F× A× ̟p′ ̟p′ Z \  1   2  !N W ∗ SN(~p; h)= SN (~p; h)+ SN(~p; h),

∗ ̟p1 ̟p2 × SN(~p; h)= h(|y|A)a .ϕN(a(y)) · a .ϕN(a(y))d y. F× A× ̟p′ ̟p′ Z \  1   2  The treatment of 1 W χ SN (~p; h) M 4 ~p E ~p∈I4 XE −2+ǫ ǫ is the same as [22, Lemma 3.4], which gives a term ≪F,ǫ E Q . Using (4.1), we find 1 1 2 2 ∗ 2 ̟p2 ̟p1 ̟p1 ̟p2 SN(~p; h)= |µ1| + µ1µ2 log + µ2µ1 log ̟p′ ̟p′ ̟p′ ̟p′ 2 A 1 A! 1 A 2 A

× · h(|y|A)|y| Ad y F× × Z \A 1 1 2 2 ̟p1 ̟p2 × + (µ1µ2 + µ2µ1) · h(|y|A)|y|A log|y|Ad y ̟p′ ̟p′ F× A× 1 A 2 A Z \ 1 1 2 2 2 ̟p1 ̟p2 2 × + |µ2| · h(|y|A)|y|A log |y|Ad y, ̟p′ ̟p′ F× A× 1 A 2 A Z \ from which we easily see, by the same consideration of (4.2),

∗ κ−1+ǫ 1 ∗ κ−1+ǫ SN(~p; h) ≪F Q ; χ SN(~p; h) ≪F Q . ,ǫ M 4 ~p ,ǫ E ~p∈I4 XE



BURGESS-LIKE SUBCONVEXITY FOR GL1 25

′ ′ Remark 4.4. There are nine different patterns of the positions of p1, p1, p2, p2, listed in [22, Proposition 3.5]. The estimation of the rest three terms depends on the pattern. For simplicity, we only treat the ′ ′ typical pattern in detail in what follows, i.e., when p1, p1, p2, p2 are distinct. The treatment of the other patterns is quite similar. Lemma 4.5. For any ǫ> 0 we have

1 ǫ κ−1 2 −2 χ ζ(h, 1,n(T ).E (~p)) ≪F (EQ) (Q E + E ). M 4 p 0 ,h0,ǫ E ~p∈I4 XE

Proof. This follows from Corollary 5.3 and the omitted calculation for other patterns. The situation is quite similar to that of [22, Section 6.2 - 6.4].  1 1 The estimation of χ ζ(h, 1,n(T ).ϕ (~p) ) and χ ζ(h, 1,n(T ).ϕ (~p) ) is essen- M 4 ~p 0 cusp M 4 ~p 0 Eis E ~p∈I4 E ~p∈I4 XE XE tially the same as the cuspidal case given in [22, Section 6.3 & 6.4]. In fact, the only difference appears

′ A ̟v1 ̟v2 4 in [22, (6.16)], where we could bound k∆∞ a .ϕ0 · a .ϕ0k easily by L -norm of ϕ0. For ̟v′ ̟v′  1   2  A′ ′ ′ the current case, we need to bound k∆∞ ϕ0(v1, v1, v2, v2)k defined in (4.3). Decomposing the relevant function into K∞-isotypic parts, we can apply Theorem 5.4. Thus unlike the cuspidal case, we get an extra (log E)3 into our estimation, which is harmless. Hence [22, Lemma 3.6 & 3.7] remain valid in the current case, giving Lemma 4.6. For any ǫ> 0 we have

1 ǫ 2 1/2−θ χ ζ(h, 1,n(T ).ϕ (~p) ) ≪F (EQ) E Q . M 4 ~p 0 cusp ,h0,ǫ E ~p∈I4 XE

Lemma 4.7. For any ǫ> 0 we have

1 ǫ (κ−1)/2 χ ζ(h, 1,n(T ).ϕ (~p) ) ≪F (EQ) EQ . M 4 ~p 0 Eis ,h0,ǫ E ~p∈I4 XE

We are finally lead to establishing (1.1) by

−1 −1/4+θ/2 −κ/2 (κ−1)/2 1/2 (κ−1)/4 (κ−1)/2 − 1−2θ min max(E ,EQ ,Q ,Q , E Q ,EQ )= Q 8 , κ,E with an optimal choice given by 1−2θ 1 θ E = Q 8 ,κ = + . 4 6 5. Complements of Global Estimations

5.1. Estimation for Exceptional Part. Recall E0(~p) defined in (4.4).

Definition 5.1. For ~p, we define ~n(~p) = (nv)v for v running over the set of places of F such that ′ ′ • nv =0 for v | ∞ and v∈{ / p1, p1, p2, p2}; ′ ′ • np =1 if p ∈{p1, p2}, np = −1 if p ∈{p1, p2}.

For ~n = (nv)v,~l = (lv)v with components in Z, we write ~l  ~n to mean lv  nv at each v, defined in Definition 2.21. We define for ~l  ~n ~ ~ ′ σ(l)= lv, klk = |lv|, e~n = ⊗venv , v v X X µ(~n; s)= µv(nv; s), c(~n,~l; s) := cp(np,lp; s) v p<∞ Y Y 26 HAN WU where the local components are defined in Lemma 2.18 (2), Corollary 2.19, Remark 2.20 and Definition 2.21. We write the Laurent expansion at s =1 of the complete zeta function ΛF(s) as

1 ∗ (5.1) ΛF(s)= ΛF + γF + O((s − 1)). s − 1 We recall Ereg(s,f) in (2.1) or [25, Definition 5.16] as well as the abbreviation ∂n Ereg,(n)(s,f) := Ereg(s,f). ∂sn

We compute E0(~p) explicitly as

∗ 2 reg,(2) 1 1 ∗ ′ reg 1 E (~p)= c(~n(~p),~l; 0) |ΛF| E ( ,e )+(2γF − ΛFµ (~l; 0))2γFE ( ,e ) 0 2 ~l 2 2 ~l ~lX~n(~p)  ∗ ∗ 1 ∗ ′ reg,(1) 1 + 2ΛFγF + Λ (2γF − ΛFµ (~l; 0)) E ( ,e ) , F 2 2 ~l    where the derivative µ′(~l;0) is taken with respect to s in µ(~l; s). Consequently, we obtain

∗ 2 reg,(2) 1 (5.2) ζ(h, 1,n(T ).E (~p)) = c(~n(~p),~l; 0) |ΛF| ζ(h, 1,n(T ).E ( ,e )) 0 2 ~l ~lX~n(~p)  1 ∗ ′ reg 1 + (2γF − ΛFµ (~l; 0))2γFζ(h, 1,n(T ).E ( ,e )) 2 2 ~l

∗ ∗ 1 ∗ ′ reg,(1) 1 + 2ΛFγF + Λ (2γF − ΛFµ (~l; 0)) ζ(h, 1,n(T ).E ( ,e )) . F 2 2 ~l    Thus we are reduced to bounding, for n =0, 1, 2 1 ζ(h, 1,n(T ).Ereg,(n)( ,e )). 2 ~l Lemma 5.2. For n =0, 1, 2 and any ǫ> 0 sufficiently small, we have

reg,(n) 1 ǫ κ−1 − 1 σ(~l) − 1 k~lk ζ(h, 1,n(T ).E ( ,e ) ≪F (EQ) (Q E 2 + E 2 ). 2 ~l ,h0,ǫ

Proof. Since the Mellin transform the integrand is explicitly related to L-functions, we depart from Mellin inversion as 1 1 ds ζ(h, 1,n(T ).Ereg,(n)( ,e )) = M(h)(−s)ζ(1/2+ s, 1,n(T ).Ereg,(n)( ,e )) , 2 ~l 2 ~l 2πi Zℜs≫1 and then shift the vertical line of integration to the left. There are poles of the integrand determined by Proposition 2.9. We calculate the constant terms in order to analyze the poles. We have

1 F µ(~l; + s) reg 1 1+s 2sΛ (−2s) 2 −s ~ EN ( + s,e~)(a(y)k)= |y|A e~(k)+ |y|A e~(k), l 6= 0; 2 l l ΛF(2+2s) 2s l

−s reg 1 1+s 2sΛF(−2s) |y|A − 1 EN ( + s)(a(y)k,e~0)= |y|A e~0(1) + e~0(1). 2 ΛF(2+2s) 2s

Hence we get for n =0, 1, 2, ~l 6= ~0 and with constants ck depending only on F n reg,(n) 1 n −k~lk k−k~lk+1 n−k n(T ).E ( ,e )(a(y)) = |y|A log |y|Ae (1) + c E (log E) log |y|Ae (1), N 2 ~l ~l k ~l k=Xk~lk−1 BURGESS-LIKE SUBCONVEXITY FOR GL1 27

reg,(n) 1 n n(T ).E ( ,e )(a(y)w) = (Ht(wn(T ))|y|A) log (Ht(wn(T ))|y|A)e (w)+ N 2 ~l ~l n −k~lk k−k~lk+1 n−k ckE (log E) log (|y|AHt(wn(T )))e~l(w); k=Xk~lk−1 while for ~l = ~0 n reg,(n) 1 n n−k+1 n(T ).E ( ,e )(a(y)) = |y|A log |y|Ae (1) + c log |y|Ae (1), N 2 ~0 ~0 k ~0 kX=0

reg,(n) 1 n n(T ).E ( ,e )(a(y)w) = (Ht(wn(T ))|y|A)(log Ht(wn(T ))|y|A) e (1)+ N 2 ~0 ~0 n n−k+1 ck(log|y|AHt(wn(T ))) e~0(1). Xk=0 1 1 By Proposition 2.9, ζ( + s, 1,n(T ).Ereg,(n)( ,e )) has 2 2 ~l • a pole at s = 1 with residue equal to

n Ht(wn(T )) log Ht(wn(T ))e~l(w), which is bounded as, using Corollary 2.17

−2 n − 1 σ(~l) Q log Q · E 2 .

• a pole at s = 0. • a pole at s = −1. We can thus write for 0 <ǫ< 1, using [22, (6.1) & (6.2)] and Proposition 2.9 1 1 1 ds ζ(h, 1,n(T ).Ereg,(n)( ,e )= ζ( + s, 1,n(T ).Ereg,(n)( ,e ))M(h)(−s) 2 ~l 2 2 ~l 2πi Zℜs=ǫ 1 ~ κ−1+ǫ − 2 σ(l) + OF,h0,ǫ(Q E ). To bound the integral on the vertical line ℜs = ǫ, we have by Proposition 2.10 1 1 ζ( + s, 1,n(T ).Ereg,(n)( ,e )) 2 2 ~l 1 1 n F F ∂ ζ ( 2 + s + s0)ζ ( 2 + s − s0) s × 1 = n |s0= · W0,v(s0,a(yv))ψv(yvTv)|yv|vd yv· 2 F × ∂s0  ζ (1+2s0) Fv  vY|∞ Z ζp(1+2s0) s × 1 1 W0,p(s0,a(yp))ψp(ypTp)|yp|pd yp· ζ ( + s + s )ζ ( + s − s ) F× p p 2 0 p 2 0 p <∞Y,Tp6=0 Z

C(ψ )s−s0 ζ (s,s ) , p p,lp 0  p∈{p ,p ,p′ ,p′ } 1 Y2 1 2  where ζp,lp (s,s0) is defined in Lemma 3.1. From Corollary 3.2 we deduce

k ∂ 1 ~ s−s0 − 2 klk+ǫ | 1 C(ψp) ζp,lp (s,s0) ≪F,ǫ E , k ≤ n ≤ 2. k s0= 2 ∂s0 ′ ′ p∈{p1,p2,p ,p } Y 1 2

28 HAN WU

−kp At p < ∞,Tp 6= 0, assuming Tp = ̟p we can calculate explicitly (using for example [22, Lemma 4.7])

ζp(1+2s0) s × 1 1 W0,p(s0,a(yp))ψp(ypTp)|yp|pd yp ζ ( + s + s )ζ ( + s − s ) F× p 2 0 p 2 0 Z p 1 1 − −s−s0 − −s+s0 1 2 1 2 −kp( +s−s ) 1 − qp −kp( +s+s )−2s 1 − qp = q 2 0 − q 2 0 0 p −2s0 p −2s0 1 − qp 1 − qp 1 1 − −s−s0 − −s+s0 −2kps0 1 2 2 −(kp−1)( +s−s ) (1 − qp )(1 − qp ) 1 − qp − q 2 0 . p −2s0 qp − 1 1 − qp Thus we obtain the following bound

k ∂ ζp(1+2s0) s × | 1 W (s ,a(y ))ψ (y T )|y | d y k s0= 1 1 0,p 0 p p p p p p p 2 F× ∂s0 ζp( + s + s0)ζp( + s − s0) p 2 2 Z −kpǫ k −k ≪ qp (kp log qp) ≪k ǫ .

At v | ∞, we can trivially bound

k k ∂ s × ∂ ǫ × | 1 W (s ,a(y ))ψ (y T )|y | d y ≤ | 1 W (s ,a(y )) |y | d y ≪ 1 k s0= 0,v 0 v v v v v v v k s0= 0,v 0 v v v v ǫ ∂s 2 F× F× ∂s 2 0 Z v Z v 0 using classical asymptotic estimation for Whittaker functions (or [12, Proposition 4.1]). Together with convex bounds for ζF we see for n =0, 1, 2

F 1 reg,(n) 1 [ :Q] +ǫ − 1 k~lk+ǫ ζ( + s, 1,n(T ).E ( ,e )) ≪F (1 + |s|) 2 E 2 . 2 2 ~l ,ǫ

We get the desired bound using [22, (6.1) & (6.2)] again. 

Corollary 5.3. For a typical pattern, we have for any ǫ> 0 ǫ κ−1 2 −2 |ζ(h, 1,n(T ).E0(~p))| ≪F,h0,ǫ (EQ) (Q E + E ).

Proof. This follows from (5.2) and the following bounds resulting from Lemma 2.18 (2), Corollary 2.19 and Remark 2.20: ′ ǫ − 1 (4−k~lk) −2+ 1 k~lk µ (~l; 0) ≪ǫ E , c(~n(~p),~l; 0) ≪ E 2 = E 2 . 

5.2. Estimation for Regularized L4-Norms. Recall E > 0 defined in Lemma 4.2. Choose a uni- formizer ̟p at every finite place p. Let ~n = (np)p,np ∈ N such that

• np = 0 unless E ≤ qp < 2E; • both the number of p such that np 6= 0 and |np| are bounded by some absolute constant. × −np ∗ Associated to ~n we define t = t(~n) ∈ A such that tv =1, v | ∞ and tp = ̟p . Take ϕj =E (0,fj ) for j =1, 2 where

• fj,v ∈ πv(1, 1) is a unitary vector, spherical at each v = p < ∞;

• at each v | ∞, fj,v lies in some Kv-isotypic Hnj,v , whose definition is recalled in Remark 2.20; • write ~nj = (nj,v)v|∞ or (nj,v)v with nj,p = 0 for p < ∞; • |nj,v|, v | ∞ are bounded by some absolute constant. Recall [25, Definition 5.26] and write for j =1, 2 2 Et = E(a(t)ϕ1 · ϕ2), ϕt = a(t)ϕ1 · ϕ2 −Et, Ej = E(|ϕj | ). 2 We are interested in the L -norm of ϕt in terms of E. BURGESS-LIKE SUBCONVEXITY FOR GL1 29

Theorem 5.4. With the above notations and conditions, we have

3 kϕtk ≪ (log E) .

Proof. Note that reg reg 2 2 2 2 (5.3) kϕtk = a(t)|ϕ1| − a(t)E1 (g) |ϕ2| −E2 (g)dg + |ϕ2| (g)a(t)E1(g)dg Z[PGL2] Z[PGL2] reg  reg  2 − a(t)E1(g)E2(g)dg + |Et(g)| dg Z[PGL2] Z[PGL2] reg reg 2 − 2ℜ a(t)ϕ1(g)ϕ2(g)Et(g)dg + a(t)|ϕ1| (g)E2(g)dg. Z[PGL2] Z[PGL2] We bound the right hand side term by term, which will occupy the rest of this subsection. Note that only the fourth and sixth terms have growing contribution as (log E)3 and (log E)6. 

For the first term in (5.3), we use Cauchy-Schwarz inequality to get reg 2 2 a(t)|ϕ1| − a(t)E1 (g) |ϕ2| −E2 (g)dg Z[PGL2]   2 2 = a(t) |ϕ1| −E1 (g) |ϕ2| −E2 (g)dg Z[PGL2] 2  2  2 2 ≤ a(t) |ϕ1| −E1 · |ϕ2| −E2 = |ϕ1| −E1 · |ϕ2| −E2 , which is independent of t hence E. 

For the second & third term in (5.3), first notice that we can write

2 1 E = E(k)( ) 1 1,k 2 Xk=0 where E1,k(s) is a regularizing Eisenstein series (2.1) and the superscript (k) means taking derivative k times with respect to s. Moreover, E1,k(s) are spherical at finite places. We notice further that although the regularized integral is not GL2(A)-invariant, it is still K-invariant [25, Proposition 5.27 (2)]. Hence if we write the Hecke operator associated with ~n as

|np| T (~n) := R(κ1a(t)κ2)dκ1dκ2 = T (p ), p<∞ Kfin×Kfin Z Y where R(·) denotes the GL2(A)-translation, we have reg reg 2 2 |ϕ2| (g)a(t)E1(g)dg = |ϕ2| (g) (T (~n)E1(g)) dg, Z[PGL2] Z[PGL2] reg reg a(t)E1(g)E2(g)dg = (T (~n)E1(g)) E2(g)dg. Z[PGL2] Z[PGL2] E1,k(s) is a generalized eigenvector of T (~n) with eigenvalue [3, Theorem 4.6.6]

−n/2 (n+1)s −(n+1)s (n−1)s −(n−1)s qp qp − qp −1 qp − qp (5.4) λ(~n; s)= λp(|np|; s), λp(n; s)= − q p −1 s −s p s −s 1+ qp qp − qp qp − qp ! Y in the sense that (c.f. [25, Remark 5.17])

T (~n)E1,k(1/2+ s)= λ(~n;1/2+ s)E1,k(1/2+ s)+ ck(λ(~n;1/2+ s) − 1)λF(s), 30 HAN WU where λF(s) is a ratio of zeta functions of F (see for example Theorem 2.2) and ck ∈ C is some constant depending on ϕ1. Note that λ(~n;1/2) = 1. Consequently, we get

k k T (~n)E(k)(1/2) = λ(l)(~n;1/2)E(k−l)(1/2) 1,k l 1,k Xl=0   k−1 k λ(k+1)(~n;1/2)λ(−1)(0) + c λ(l+1)(~n;1/2)λ(k−l)(0) + F . k l +1 F k +1 Xl=0   Inserting the obvious bound λ(l)(~n;1/2) ≪ (log E)l, we see that the second and third term in (5.3) are bounded as ≪ (log E)3.

Remark 5.5. λp(n; s)= cp(n, 0; s) defined in Lemma 2.18 (2).

For the fourth term in (5.3), we first make Et explicit.

Definition 5.6. We extend the definition of ~l  ~n in Definition 5.1 into the case nv 6=0, v | ∞ by

• Fv = R: lv  nv means lv = nv; • Fv = C: lv  nv means 0 ≤ lv ≤ nv.

At v | ∞, we write env for some unitary vector in Hnv recalled in Remark 2.20 without specification. We can thus write ∗ ∗ (5.5) a(t).ϕ1 = c(~n,~l;0)E (0,e ~) ⇒ Et = c(~n,~l; 0)E(E (0,e ~)ϕ2). ~l~n ~n1+l ~l~n ~n1+l X X Recall the tensor product formulas for representations of Kv, v | ∞

• Fv = R: Hn ⊗ Hm ≃ Hn+m; • Fv = C: Hn ⊗ Hm ≃ Hn+m ⊕ Hn+m−2 ⊕···⊕ H|n−m|. Together with the explicit determination of constant terms

1 1 ∗ ∗ 2 1 ∗ ′ 2 EN(0,e ~)(a(y)κ)= ΛF(1)|y| log|y|A + 2γF − ΛF(1)µ (~n1 + ~l; 0) |y| e ~(κ), ~n1+l A 2 A ~n1+l     1 1 ∗ 2 1 ∗ ′ 2 ϕ N(a(y)κ)= ΛF(1)|y| log|y|A + 2γF − ΛF(1)µ (~n ; 0) |y| f (κ), 2, A 2 2 A 2     we deduce the existence of c ~m such that (5.6) ∗ reg,(2) 1 reg,(1) 1 reg 1 E(E (0,e ~)ϕ2)= c ~m c2E ( ,e ~)+ c1(~l)E ( ,e ~)+ c0(~l)E ( ,e ~) , ~n1+l 2 ~m+l 2 ~m+l 2 ~m+l ~mX~n1+~n2   where we have written

∗ 2 1 ∗ ′ 1 ∗ ′ c = c (~l)= |ΛF| , c (~l)= 2γF − ΛFµ (~n + ~l; 0) 2γF − Λ µ (~n ; 0) , 2 2 0 2 1 2 F 2    

1 ∗ ′ ∗ ∗ 1 ∗ ′ c (~l)= 2γF − ΛFµ (~n + ~l; 0) Λ +ΛF 2γF − Λ µ (~n ; 0) . 1 2 1 F 2 F 2     Lemma 5.7. For k1, k2 ≥ 0, the regularized integral reg 1 1 reg,(k1) reg,(k ) E ( ,e ~ )(g)E 2 ( ,e ~ )(g)dg 2 ~m1+l1 2 ~m2+l2 Z[PGL2] is non-vanishing only if ~m1 = ~m2,~l1 = ~l2. We also have reg 1 1 ~ reg,(k1) reg,(k2) −klk k1+k2+2 E ( ,e ~m+~l)(g)E ( ,e ~m+~l)(g)dg ≪ E (log E) . [PGL2] 2 2 Z

BURGESS-LIKE SUBCONVEXITY FOR GL1 31

Proof. This is a direct consequence of Theorem 2.4, together with k (k) d M e = λF(s − 1/2)µ(~m + ~l; s) e . 1/2 ~m+~l dsk ~m+~l s=1/2   f 

Applying the lemma and the estmations

~ ′ ′ ′ ′ − k~nk−klk µ (~n1 + ~l;0) = µ ( ~n1;0)+ µ (~l; 0), µ (~l; 0) ≪ log E, |c(~n,~l; 0)| ≪ E 2 we finally get

reg 2 2 ~ 2 ~ ~ |Et(g)| dg = |c(~n, l; 0)c ~m| ck1 (l)ck2 (l)· [PGL2] Z ~Xl~n ~mX~n1+~n2 k1X,k2=0 reg reg,(k ) 1 1 E 1 ( ,e )(g)Ereg,(k2)( ,e )(g)dg 2 ~m+~l 2 ~m+~l Z[PGL2] ≪ E−k~nk(log E)6.

For the fifth term in (5.3), note that we have computed/decomposed a(t).ϕ1 and Et in (5.5) and (5.6). Lemma 5.8. For any k ≥ 0, the regularized integral reg ∗ reg,(k) 1 E (0,e ~ )(g)ϕ2(g)E ( ,e ~ )(g)dg ~n1+l1 2 ~m+l2 Z[PGL2] is non-vanishing only if ~l1 = ~l2. We also have reg ∗ reg,(k) 1 −k~lk k+4 E (0,e )(g)ϕ2(g)E ( ,e )(g)dg ≪k E (log E) . ~n1+~l ~m+~l [PGL2] 2 Z

Proof. We apply Theorem 2.7. The “degenerate part”, i.e., the weighted sum involving PK is easily seen −k~lk k+4 to be non-vanishing only if ~l1 = ~l2, and in the case ~l1 = ~l2 = ~l, it is bounded by E (log E) since µ(n)(~l;1/2) ≪ E−klk(log E)n. For the “main part”, we note that

1 ∗ R( +¯s, E (0,e ~ )(g)ϕ2(g); e ~ ) 2 ~n1+l1 ~m+l2 is the product of ζF(s + 1)4ζF(2s + 2)−1, some local components at v | ∞ irrelevant for estimation, and l 0 l ζ ( 1,p 2,p ) p 0 0 1/2+ s defined in (3.1). Lemma 3.3 (2) implies the non-vanishing assertion. We need to estimate hol ∂kR 1 k! ∂k+4 1 , ··· = · s4R( + s, ··· ) . ∂sk 2 (k + 4)! ∂sk+4 2     s=0  

Corollary 3.4 concludes the bound. 

We deduce that reg 2 2 a(t)ϕ1(g)ϕ2(g)Et(g)dg = c ~m|c(~n,~l; 0)| · ck(~l) [PGL2] Z ~mX~n1·~n2 ~Xl~n kX=0 reg ∗ reg,(k) 1 E (0,e ~)(g)ϕ2(g)E ( ,e ~)(g)dg ~n1+l 2 ~m+l Z[PGL2] ≪ E−k~nk(log E)6. 32 HAN WU

For the last term in (5.3), first notice that we can write

2 1 E = Ereg,(k)( ,h ), 2 2 k Xk=0 where hk are spherical at p for which np 6= 0. We are reduced to bounding reg 1 |a(t).E∗(0,f )(g)|2Ereg,(k)( ,h )(g)dg, 1 2 k Z[PGL2] to which apply directly Theorem 2.7. The degenerate part contributes (log E)3. In fact, only (l) PK(M0 a(t)f1 · a(t)f1)PK(hk), 0 ≤ l ≤ 3 have non constant contribution. For the main part, using the GL2(Fp)-invariance of the local Rankin- Selberg zeta functions we easily deduce 1 2 4 F ∗ 1 ∗ 2 ζF(s + 1) 1 ζ (2s + 2)R( 2 + s, |E (0,f1)| ; hk) R( + s, |a(t).E (0,f1)| ; hk)= · λ(~n; + s) · , 2 ζF(2s + 2) 2 ζF(s + 1)4 1 where only the term λ(~n; + s) defined in (5.4) depends on E. We bound 2 k! ∂k+4 1 s4R( + s, |a(t).E∗(0,f )|2 ; h ) ≪ (log E)k+4 (k + 4)! ∂sk+4 2 1 k s=0

(l) l 6 from λ (~n;1/2) ≪ (log E) , and conclude that the last term in (5.3) is bounded as ≪ (log E) . 6. Appendix: Dis-adelization in a Special Case For the convenience of the readers not familiar with adelic language, and also for those who want to see the computation in the classical setting, we offer some essential part of the dis-adelized computation iµ in the case F = Q and χ = |·|A , i.e., the case for the Riemann-zeta function. On the adelic side, we shall work in the framework of PGL2, hence all previous groups such as N, B, K are considered as subgroups of PGL2 via the canonical projection. 1fin denotes the identity element in PGL2(Afin). ϕ is reserved for functions on [PGL2]. We restrict ourselves to SO2(R)-invariant functions. The finite places of Q correspond to prime numbers p = pZ. On the classical side, we shall confuse SO2(R)-invariant modular forms on Γ\PGL2(R) with functions on Γ\H, where Γ < PSL2(R) is a lattice, Γ is the subgroup generated by Γ and diag(1, −1). f is reserved for functions on Γ\PGL2(R). Hence e e y x f(z) := f , z = x + iy ∈ H. e 1   We also write e(x) := e2πix for x ∈ R. We assume the existence of a compact subgroup KΓ < PGL2(Afin) such that × × KΓ ∩ PGL2(Q)= Γ, det(KΓ)= Z , a(Z ) ⊂ KΓ. Then the strong approximation theorem, PGL (A ) = PGL (Q)K , yields e 2 fin b2 Γb PGL2(Q)\PGL2(A)/PSO2(R)KΓ ≃ Γ\PGL2(R)/PSO2(R) ≃ Γ\H. K -invariant functions ϕ are in bijective correspondence with functions f, determining each other by Γ e f(g)= ϕ(g, 1fin), ∀g ∈ PGL2(R). × Consequently, if p denotes the standard uniformizer at the place p, then since p ∈ Zq for any prime q 6= p and ϕ is invariant by a(Z×), we get (a(p−1).ϕ)(g, 1 )= ϕ(g, 1, ··· , 1,a(p−1), 1, ··· )) = ϕ(a(p)g, 1 )= f(a(p)g), b fin fin BURGESS-LIKE SUBCONVEXITY FOR GL1 33 or (a(p−1)f)(z) := f(pz). The constant term, defined by

ϕN(g) := ϕ(n(x)g)dx, ZQ\A is left-B(Q) right-KΓ invariant. From the strong approximation theorem, we deduce that

B(Q)\PGL2(Afin)/KΓ ≃ B(Q)\PGL2(Q)/Γ corresponds bijectively with the set of cusps of Γ. (Look at the right action of PGL (Q) on P1(Q)!) We e 2 take a system of representatives [Γ] ⊂ PGL2(Q) for the above double coset decomposition. Hence ϕN and the following finite collection determine each other

ϕN(g,γ), γ ∈ [Γ] ⊂ PGL2(Q) ⊂ PGL2(Afin). −1 For every γ ∈ [Γ], N(Afin) ∩ γKΓγ is a compact subgroup of N(Afin), hence equal to N(dγ Z) for some × dγ ∈ Q . The strong approximation theorem for Q implies b Afin = Q + dγ Z ⇒ Afin = α + dγ Z ⇒ Q\A = (dγ Z\R) × dγ Z, α G where α runs over a system ofb representatives for Q/dγZb. Hence b

−1 ϕN(g,γ)= ϕ(n(u)g,n(z)γ)dzdu = |dγ | ϕ(n(u)g,γ)du dγ Z\R dγ Z dγ Z\R Z Z b Z −1 −1 = |dγ | ϕ(γ n(u)g, 1fin)du, ∀g ∈ PGL2(R). Zdγ Z\R On the other hand, the function −1 −1 fγ (g) := f(γ g)= ϕ(γ g, 1fin) is a modular form for the lattice γΓγ−1. Since γΓγ−1 ∩ N(Q)= γK γ−1 ∩ N(Q)= N(d Z) ∩ N(Q)= N(d Z), e Γ γ γ the normalized constant term at the cusp ∞ of f is e γ b −1 fγ,N(g) := |dγ | fγ(n(x)g)dx = ϕN(g,γ). Zdγ Z\R Write fN = f1,N. Since both functions × A 7→ C, y 7→ ϕ(a(y)) and y 7→ ϕN(a(y)) are left-Q× right-Z× invariant, and since by class number 1 of Q Q×\A×/Z× ≃ {±1}\R× ≃ R , b >0 we can compute the zeta functional as b ∞ ∞ s−1/2 × s−1/2 × ζ(s, 1, ϕ)= (ϕ − ϕN)(a(t), 1fin)t d t = (f − fN)(a(t))t d t. Z0 Z0 K Write e0 ∈ H := IndB(A)∩K1 for the constant function taking value 1 on K. The normalized Eisenstein ∗ series E (s,e0) corresponds to the usual normalized real analytic Eisenstein series for the full modular group 1/2+s ∗ y E (s,z) := Λ(1 + 2s) (c,d)=1 1+2s . c,d∈Z |cz + d| X Introducing, first for ℜs ≫ 1 then by analytic continuation using the second expression, the function

∞ 2 e(−uy) y 1/2−s 1/2−s −πt+ t  s × K (s,y) := ΓR(1+2s)|y| du = |y| e t d t, y ∈ R, ∞ (1 + u2)1/2+s ZR Z0 34 HAN WU one computes the Fourier expansion at ∞ as y x (E∗ − E∗ )(s,z)=(E∗ − E∗ )(s,e ) , 1 N N 0 1 fin    (e+1)s −(e+1)s −1/2 p − p = e(nx)K∞(s,ny) · |n| 06=n∈Z pekn ps − p−s X s−1/2Y = e(nx)K∞(s,ny) · |n| σ−2s(|n|), 06=n∈Z

s X where σs(m) := d is the usual divisor sum function. Hence the zeta-functional is equal to d|m X ∞ ∗ s0−1/2 s−1/2 × ζ(s, 1, E (s0,e0)) = K∞(s0,ny) · |n| σ−2s (|n|)y d t 06=n∈Z 0 Z0 ∞X ∞ ∞ n2y2 −πt+ t  s0 s−s0 × × =2 σ−2s0 (n) e t y d td y n=1 0 0 Z Z ∞ ∞ X∞ s+s0 s−s0 s0−s −π(t+y) 2 2 × × = σ−2s0 (n)n e t y d td y n=1 0 0 X Z Z =ΓR(s − s0)ΓR(s + s0)ζ(s − s0)ζ(s + s0)=Λ(s − s0)Λ(s + s0). ∗ Proposition 4.1 studies a variant of the above equality by replacing s with 1/2+ iµ,E (s0,e0) with ∗ n(T ).E (0,e0) for some T ∈ R, |T | ≍ |µ|. Together with Lemma 4.2, we are reduced to bounding ∞ h(y) · (n(T ).E∗)(0,iy)d×y, Z0 where h is a positive function with support in [|µ|−κ−1, |µ|κ−1]. We would like to apply C-S ∞ 2 ∞ ∞ h(y) · (n(T ).E∗)(0,iy)d×y ≤ h(y)d×y · h(y) · (n(T ).|E∗|2)(0,iy)d×y Z0 Z0 Z0 and Fourier inverse |E∗(0,z)|2, which is not square integrable. Michel & Venkatesh’s idea is to Fourier inverse the restriction of |E∗(0,z)|2 on a large compact region containing the domain of integration, which fails to give the Burgess-like quality because the L2-norm of such restriction depends polynoimally on µ. Our idea is to find E, some linear combination of derivatives of Eisenstein series such that |E∗(0,z)|2 −E comes back to L2. The existence of E is due to Zagier [27] and our extension [25, §5]. Hence we can estimate ∞ h(y) · (n(T ).|E∗|2)(0,iy)d×y 0 Z ∞ ∞ = h(y) · (n(T ).(|E∗|2 −E))(0,iy)d×y + h(y) · (n(T ).E)(iy)d×y Z0 Z0 the two terms on the RHS seprately. To the first term we apply Fourier inversion to |E∗|2 −E. For the second, we compute directly. The advantage over Michel & Venkatesh’s approach is that the L2-norm of |E∗|2 −E is essentially constant, while the second term contributes no more than the constant part of the the first term. The truth is a little more complicated due to the large contribution of the one-dimensional part of |E∗|2 −E. Hence we need to use the method of amplification. We shall not be precise on the exact form of amplification since the goal is to give an idea how things look like. If p is a prime regarded as the × × uniformizer of Qp embedded in Afin, then ∞ −1 ∗ ∗ ∗ s−1/2 × ζ(s, 1,a(p ).E (s0,e0)) = (E − EN)(s0,ipy)y d y, Z0 from which we deduce that −1 ∗ −iµ ∗ ζ(1/2+ iµ, 1,a(p ).E (0,e0)) = p ζ(1/2+ iµ, 1, E (0,e0)). BURGESS-LIKE SUBCONVEXITY FOR GL1 35

∗ iµ −1 ∗ We can replace E (0,e0) with some balanced average of p a(p ).E (0,e0) without affecting the integral representation and Lemma 4.2 given above, since a(p−1) commutes with the n(T ),T ∈ R. After applying C-S, we are lead to Fourier inversing

−1 ∗ ∗ ∗ ∗ a(p ).E (0,e0) · E (0,e0) or E (0,pz) · E (0,z).

In the classical setting, we have regarded all forms from PSL2(Z)\H to Γ0(p)\H. Hence we need to find E(p) some linear combination of derivatives of Eisenstein series for Γ0(p), such that

∗ ∗ ϕ0(p)(z) := E (0,pz) · E (0,z) −E(p)(z) has slow increase at every cusp of Γ0(p). This is of course feasible in the classical setting, but with painful computations. The more convenient adelic computation goes as follows.

K GL2(A) s −s Definition 6.1. For e ∈ H := IndB(A)∩K1, we write es ∈ IndB(A) (|·|A, |·|A ) for the flat section associated with it. For any s0 ∈ C,n ∈ N we write ∂n e(n) := | e . s0 ∂sn s=s0 s It is easy to verify (n) (m) (n+m) e0,0 · e0 = e1/2 , ∀e ∈ H, ∀n,m ∈ N.

Let e1 ∈ H admitting the same local component as e0 at places ∞ and q 6= p. At p we take its local component as the e1 in Lemma 2.16. Precisely, −1/2 1/2 e1 |GL2(Zp)−K0(p)= −p , e1 |K0(p)= p .

E(s,e1) corresponds to an Eisenstein series for Γ0(p) given by y x y x E1(s,z)=E(s,e1) , 1fin = e1,s γ , 1fin 1 γ∈B(Q)\PGL2(Q) 1        yX x X = ^ e1,s γ , 1fin γ∈N(Z)\Γ0(p) 1     X 1 y x + ^ e1,s γ , 1fin γ∈N−(pZ)\Γ0(p) 1 1 X      ^ since we know the set {0, ∞} of cusps for Γ0(p) as 1 PGL (Q)= B(Q)Γ^(p) ⊔ B(Q) Γ^(p). 2 0 1 0   We insert the value of e1,s to obtain

1/2 1/2+s −1/2 1 1/2+s E1(s,z)= p ^ ℑ(γ.z) − p ^ ℑ( γ.z) . γ∈N(Z)\Γ0(p) γ∈N−(pZ)\Γ0(p) 1 X X   Note that it is regular at s = 1/2. We now turn to the determination of E(p). Using the Laurent expansion of Λ in (5.1), we can write the constant term as

∗ ∗ (1) EN(0,e0)=2γQe0,0 +ΛQe0,0. −1 ∗ −1 Computing the constant term of a(p ).E (0,e0) is convenient only if we can express a(p )e0,s as linear combination of flat sections. This is done in Lemma 2.18, yielding ps+1/2 − p−(s+1/2) ps + p−s a(p−1).e = e + e , 0,s p1/2 + p−1/2 1,s p1/2 + p−1/2 0,s

p1/2 − p−1/2 p1/2 − p−1/2 a(p−1).e(1) = e(1) + e(1) + (log p)e + (log p)e . 0,0 0,0 p1/2 + p−1/2 1,0 p1/2 + p−1/2 0,0 1,0 36 HAN WU

We thus obtain

−1 ∗ ∗ (a(p ).E )N(0,e0) · EN(0,e0) p1/2 − p−1/2 = |Λ∗ |2e(2) + |Λ∗ |2 e(2) Q 0,1/2 Q p1/2 + p−1/2 1,1/2 1/2 −1/2 ∗ 2 p − p ∗ ∗ (1) + |Λ | (log p)+2γQΛ +2γQΛ e Q p1/2 + p−1/2 Q Q 0,1/2   1/2 −1/2 1/2 −1/2 ∗ 2 ∗ p − p ∗ p − p (1) + |Λ | (log p)+2γQΛ +2γQΛ e Q Q p1/2 + p−1/2 Q p1/2 + p−1/2 1,1/2   1/2 −1/2 ∗ p − p 2 + 2γQΛ (log p)+4|γQ| e Q p1/2 + p−1/2 0,1/2   1/2 −1/2 ∗ 2 p − p + 2γQΛ (log p)+4|γQ| e . Q p1/2 1,1/2  

Writing for E(s,z) the usual (non-completed) real analytic Eisenstein series and for n ∈ N

∂n 1 ∂n Ereg,(n)(1/2,z) := E(s,z) − , E(n)(1/2,z)= E (s,z), ∂sn s − 1/2 1 ∂sn 1 s=1/2   s=1/2

we deduce via the adelic-classical correspondence of constant terms that

p1/2 − p−1/2 E(p)(z) := |Λ∗ |2Ereg,(2)(1/2,z)+ |Λ∗ |2 E(2)(1/2,z) Q Q p1/2 + p−1/2 1 1/2 −1/2 ∗ 2 p − p ∗ ∗ reg,(1) + |Λ | (log p)+2γQΛ +2γQΛ E (1/2,z) Q p1/2 + p−1/2 Q Q   1/2 −1/2 1/2 −1/2 ∗ 2 ∗ p − p ∗ p − p (1) + |Λ | (log p)+2γQΛ +2γQΛ E (1/2,z) Q Q p1/2 + p−1/2 Q p1/2 + p−1/2 1   1/2 −1/2 ∗ p − p 2 reg + 2γQΛ (log p)+4|γQ| E (1/2,z) Q p1/2 + p−1/2   1/2 −1/2 ∗ 2 p − p + 2γQΛ (log p)+4|γQ| E (1/2,z) Q p1/2 1   does the job. 2 With the above construction, we then showed that the (re-normalized) L -norm of ϕ0(p) on Γ0(p)\H 3 is ≪ (log p) , as well as its derivatives with respect to the Lie algebra of PGL2(R). Its Fourier inversion

hϕ0(p), 1i ϕ0(p)(z)= · 1+ Cp(f)f(z) Vol(Γ0(p)\H) f cusp form for Γ0(p) ∞ Xdτ ∞ dτ + C (0, iτ)E(iτ,z) + C (1, iτ)E (iτ,z) p 4π p 1 4π Z−∞ Z−∞ = ϕ0(p)N(z)+ Cp(f)f(z) f cusp form for Γ0(p) ∞ X dτ ∞ dτ + C (0, iτ)(E − EN)(iτ,z) + C (1, iτ)(E − E N)(iτ,z) p 4π p 1 1, 4π Z−∞ Z−∞ BURGESS-LIKE SUBCONVEXITY FOR GL1 37 with Fourier coefficients Cp(f), Cp(0, iτ), Cp(1, iτ) ∈ C, converges in uniformly in any compact (for the first version) resp. Siegel domain (for the second). Thus ∞ × ζ(h,n(T ).ϕ0(p)) := h(y) · ϕ0(p)(iy)d y 0 Z ∞ ∞ × × = h(y) · ϕ0(p)N(iy)d y + Cp(f) h(y) · f(iy)d y 0 f 0 Z ∞ ∞ X Z × dτ + Cp(0, iτ) h(y) · (E − EN)(iτ,iy)d y −∞ 0 4π Z ∞ Z ∞ × dτ + C (1, iτ) (E − E N)(iτ,iy)d y p 1 1, 4π Z−∞ Z0 and we estimate the RHS term by term.

Acknowledgement The preparation of the paper scattered during the stays of the author’s in FIM at ETHZ, YMSC at Tsinghua University, Alfr´ed Renyi Institute in Hungary supported by the MTA R´enyi Int´ezet Lend¨ulet Automorphic Research Group and in TAN at EPFL. The author would like to thank all four institutes for their hospitality. The author would like to thank Professor Paul Nelson for useful discussions. The author would also like to thank Professor Philippe Michel, his former thesis adviser, for his encouragement and enlightening strategic argument. The author would like to thank Dr. Bingrong Huang for bringing the paper of Soehne [20] into his attention. The author is grateful for the referees’ detailed check and suggestions.

References [1] Blomer, V., and Harcos, G. The spectral decomposition of shifted convolution sums. Duke Mathematical Journal 144, 2 (2008), 321–339. [2] Blomer, V., and Harcos, G. Twisted L-functions over number fields and Hilbert’s eleventh problem. Geometric and Functional Analysis 20 (2010), 1–52. [3] Bump, D. Automorphic Forms and Representations. No. 55 in Cambridge Studies in Advanced Mathematics. Cambridge University Press, 1998. [4] Burgess, D. On character sums and L-series II. Proceedings of the London Mathematical Society s3-13, 1 (1963), 524–536. [5] Conrey, J. B., and Iwaniec, H. The cubic moment of central values of automorphic L-functions. Annals of Mathematics 151, 3 (2000), 1175–1216. [6] Duke, W. Hyperbolic distribution problems and half-integral weight Maass forms. Inventiones mathematicae 92 (1988), 73–90. [7] Duke, W., Friedlander, J. B., and Iwaniec, H. The subconvexity problem for Artin L- functions. Inventiones mathematicae 149 (2002), 489–577. [8] Gelbart, S. S., and Jacquet, H. Forms of GL(2) from the analytic point of view. In Proceedings of Symposia in Pure Mathematics (1979), vol. 33, pp. 213–251. [9] Heath-Brown, D. Hybrid bounds for Dirichlet L-functions. Inventiones mathematicae 47 (1978), 149–170. [10] Heath-Brown, D. Hybrid bounds for Dirichlet L-functions II. The Quarterly Journal of Mathe- matics 2, 31 (1980), 157–167. [11] Huxley, M., and Watt, N. Hybrid bounds for Dirichlet’s L-function. Mathematical Proceedings of the Cambridge Philosophical Society 129, 3 (Nov. 2000), 385–415. [12] Jacquet, H. Integral representation of Whittaker functions. In Contributions to Automorphic Forms, Geometry & Number Theory, H. Hida, D. Ramakrishnan, and F. Shahidi, Eds. The Johns Hopkins University Press, 2004, ch. 15, pp. 373–419. 38 HAN WU

[13] Michel, P. Analytic number theory and families of automorphic L-functions. In Automorphic forms and applications (2007), vol. 12 of IAS/Park City Math. Ser., American Mathematical Society Providence, RI, pp. 181–295. [14] Michel, P., and Venkatesh, A. The subconvexity problem for GL2. Publications math´ematiques de l’IHES´ 111, 1 (2010), 171–271. [15] Milicevi´ c,´ D. Sub-Weyl subconvexity for Dirichlet L-functions to prime power moduli. Compositio Mathematica 152, 04 (Apr. 2016), 825–875. [16] Patterson, S. The Theory of The Riemann Zeta-Function, vol. 14 of Cambridge Studies in Ad- vanced Mathematics. Cambridge University Press, 1988. [17] Petrow, I., and Young, M. P. A generalized cubic moment and the Petersson formula for newforms. arXiv: 1608.06854v4 to appear in Math. Ann., August 2018. [18] Rudin, W. Real and Complex Analysis, 3rd ed. McGraw-Hill Book Company, 1986. [19] Sarnak, P. Estimates for Ranking-Selberg L-functions and quantum unique ergodicity. Journal of Functional Analysis 184 (2001), 419–453. [20] Soehne, P. An upper bound for Hecke zeta-functions with Groessencharacters. Journal of Number Theory 66, Article No. NT972167 (1997), 225–250. [21] Weyl, H. Zur Absch¨atzung von ζ(1 + ti). Mathematische Zeitschrift 10 (1921), 88–101. [22] Wu, H. Burgess-like subconvex bounds for GL2 × GL1. Geometric and Functional Analysis 24, 3 (2014), 968–1036. [23] Wu, H. Explicit Burgess-like subconvex bounds for GL2 × GL1. arXiv: 1712.04365 (2017). [24] Wu, H. A note on spectral analysis in automorphic representation theory for GL2: I. International Journal of Number Theory 13, 10 (June 2017), 2717–2750. [25] Wu, H. Rankin-Selberg trace formula for GL2: geometric side. arXiv: 1810.09437, October 2018. [26] Young, M. P. Weyl-type hybrid subconvexity bounds for twisted L-functions and Heegner points on shrinking sets. Journal of the European Mathematical Society 19 (2017), 1545–1576. [27] Zagier, D. The Rankin-Selberg method for automorphic functions which are not of rapid decay. Journal of the Faculty of Science 28 (1982), 415–438.

Han WU MA C3 604 EPFL SB MATHGEOM TAN CH-1015, Lausanne Switzerland [email protected]