• Keine Ergebnisse gefunden

Finally, we state the theta lift from Definition 4.2.6 in the situation of the current lattice 𝐿. Let π‘˜ be a non-negative integer. Identifying H with H1 via the map 𝑧 ↦→𝑧𝑒3, and using the identity from (4.3.5) and the fact that π‘ž(Im(𝑧𝑒3)) = 𝑁Im(𝑧)2 for 𝑧 ∈ H we can write Shintani’s theta function for the lattice 𝐿 as

Θ𝐿,π‘˜(𝜏, 𝑧) = π‘βˆ’π‘˜Im(𝜏)1/2Im(𝑧)βˆ’2π‘˜ βˆ‘οΈ

π›½βˆˆπΏβ€²/𝐿

βˆ‘οΈ

πœ†βˆˆπΏ+𝛽

π‘„πœ†(𝑧,1)π‘˜π‘’(𝜏 π‘ž(πœ†π‘§) +𝜏 π‘ž(πœ†π‘§βŠ₯))e𝛽 (4.3.6)

for 𝜏, 𝑧 ∈ H. Here the quantities π‘ž(πœ†π‘§) and π‘ž(πœ†π‘§βŠ₯) are given as in Lemma 4.3.1. By Corollary 4.2.3 the theta functionΘ𝐿,π‘˜(𝜏, 𝑧)is modular of weightπ‘˜+ 1/2with respect to 𝜌𝐿 in 𝜏, and in the variable 𝑧 it transforms as

Θ𝐿,π‘˜(𝜏, 𝛾𝑧) = 𝑗(𝛾, 𝑧)2π‘˜Ξ˜πΏ,π‘˜(𝜏, 𝑧)

for 𝜏, 𝑧 ∈ H and 𝛾 ∈ Ξ“0(𝑁). Accordingly, the (regularized) theta lift of a real analytic function 𝐹:Hβ†’C[𝐿′/𝐿] modular of weight π‘˜+ 1/2 with respect to𝜌𝐿 is given by

Ξ¦πΏπ‘˜(𝑧;𝐹) = (𝐹,Θ𝐿,π‘˜(Β·, 𝑧))reg

for 𝑧 ∈ H, and assuming that the (regularized) inner product exists the lift Ξ¦πΏπ‘˜(𝑧;𝐹) of 𝐹 is modular of weight 2π‘˜ and level 𝑁.

For an integer π‘˜ β‰₯ 2 the theta lift Ξ¦πΏπ‘˜(𝑧; Β·) is essentially the classical Shimura theta lift, mapping (vector valued) holomorphic cusp forms of weight π‘˜+ 1/2 to holomorphic cusp forms of weight 2π‘˜. It was introduced by Shimura in [Shi73] and later expressed as a theta lift by Niwa in [Niw75]. In particular, the Shimura lift of a holomorphic PoincarΓ© series is up to some constant the corresponding Zagier cusp form introduced in Section 2.4.2 (see for example [Oda77] or [KZ81]). More precisely, let𝛽 βˆˆπΏβ€²/𝐿and π‘šβˆˆZ+π‘ž(𝛽) with π‘š >0. Then an easy unfolding argument shows that

Ξ¦πΏπ‘˜(𝑧;π‘ƒπ‘˜+1/2,𝛽,π‘š) = 2 (π‘˜βˆ’1)!

πœ‹π‘˜ π‘“π‘˜,𝛽,4𝑁 π‘š(𝑧) (4.3.7)

for π‘˜ ∈ Z with π‘˜ β‰₯ 2 and 𝑧 ∈ H. Here π‘“π‘˜,𝛽,4𝑁 π‘š(𝑧) is the meromorphic modular form introduced in Definition 2.4.6. Note that we can ignore the regularization of the inner product defining the theta lift in this case since π‘ƒπ‘˜+1/2,𝛽,π‘š(𝜏) is a cusp form for π‘š >0.

for 𝑋 ∈ 𝑉. Then (𝑉, π‘ž) is a quadratic space of signature (2,2) with associated bilinear form given by (𝑋, π‘Œ) = βˆ’tr(π‘‹π‘Œ#) for 𝑋, π‘Œ ∈ 𝑉 where (οΈ€π‘Ž 𝑏

𝑐 𝑑

)οΈ€#

= (οΈ€ 𝑑 βˆ’π‘

βˆ’π‘ π‘Ž

)οΈ€. Given a positive integer 𝑁 we consider the lattice

𝐿:=

{οΈ‚(οΈ‚ π‘Ž 𝑏 𝑁 𝑐 𝑑

)οΈ‚

: π‘Ž, 𝑏, 𝑐, π‘‘βˆˆZ }οΈ‚

in𝑉. Then𝐿 is an even lattice of level 𝑁 with dual lattice given by 𝐿′ =

{οΈ‚(οΈ‚π‘Ž 𝑏/𝑁

𝑐 𝑑

)οΈ‚

: π‘Ž, 𝑏, 𝑐, π‘‘βˆˆZ }οΈ‚

,

and the discriminant group𝐿′/𝐿is isomorphic to the quotient(Z/𝑁Z)2. It is well-known that the action of the identity component SO+(𝑉) of the orthogonal group of 𝑉 can be realized by the action of the group SL2(R)Γ—SL2(R)on 𝑉(R)via

(𝜎, πœŽβ€²).𝑋 :=πœŽπ‘‹πœŽβ€²βˆ’1

for 𝜎, πœŽβ€² ∈ SL2(R) and 𝑋 ∈ 𝑉(R). Here the four elements (±𝜎,Β±πœŽβ€²) define the same element inSO+(𝑉). Further, one easily checks that the discrete subgroupΞ“0(𝑁)Γ—Ξ“0(𝑁) of SL2(R)Γ—SL2(R) acts on 𝐿′, and that its subgroup Ξ“(𝑁)Γ—Ξ“(𝑁) fixes the classes of the discriminant group 𝐿′/𝐿, i.e., we have

Ξ“(𝑁)Γ—Ξ“(𝑁)βŠ†Ξ“(𝐿) and Ξ“0(𝑁)Γ—Ξ“0(𝑁)βŠ†SO+(𝐿), (4.4.1)

under the above identification of SO+(𝑉)with the productSL2(R)Γ—SL2(R). HereΞ“(𝑁) is the usual principal congruence subgroup given by

Ξ“(𝑁) :=

{οΈ‚(οΈ‚π‘Ž 𝑏 𝑐 𝑑

)οΈ‚

∈SL2(Z) :

(οΈ‚π‘Ž 𝑏 𝑐 𝑑

)οΈ‚

≑ (οΈ‚1 *

0 1 )οΈ‚

mod 𝑁 }οΈ‚

, which clearly is a subgroup of Ξ“0(𝑁).

We now choose vectors 𝑒1 :=

(οΈ‚0 1 0 0

)οΈ‚

, 𝑒2 :=

(οΈ‚0 0 1 0

)οΈ‚

, 𝑒3 :=

(οΈ‚1 0 0 0

)οΈ‚

, 𝑒4 :=

(οΈ‚0 0 0 βˆ’1

)οΈ‚

.

Then𝑒1 ∈𝐿is primitive and isotropic and𝑒2 βˆˆπΏβ€² is isotropic with(𝑒1, 𝑒2) = 1. Moreover, the hyperbolic plane spanned by 𝑒1 and 𝑒2 is orthogonal to the plane spanned by the isotropic vectors 𝑒3 and 𝑒4 where (𝑒3, 𝑒4) = 1. We define the sublattice 𝐾 =πΏβˆ©π‘’βŠ₯1 βˆ©π‘’βŠ₯2, the corresponding quadratic spaceπ‘ˆ =πΎβŠ—Qof signature(1,1)and the induced complex planeβ„‹={𝑋+π‘–π‘Œ βˆˆπ‘ˆ(C) :π‘ž(π‘Œ)>0}. Then𝐾 ={(οΈ€π‘Ž0

0 𝑑

)οΈ€:π‘Ž, 𝑑 ∈Z},π‘ˆ ={(οΈ€π‘Ž0

0𝑑

)οΈ€: π‘Ž, π‘‘βˆˆ Q} and thus

β„‹ =

{οΈ‚(︂𝑧 0 0 𝑧′

)οΈ‚

: 𝑧, 𝑧′ ∈C with Im(𝑧) Im(𝑧′)<0 }οΈ‚

.

Hence by choosing H2 as the connected component H𝑒3βŠ•H𝑒4 of β„‹ we can identify the generalized upper half-planeH2 with the productHΓ—H. Let𝒦be the complex manifold defined in (4.1.1) associated to the space 𝑉(C). We choose𝒦+ to be the image ofH2 in 𝒦 under the map 𝑍 ↦→[𝑍𝐿] where

𝑍𝐿 =𝑍 βˆ’(π‘ž(𝑍) +π‘ž(𝑒2))𝑒1+𝑒2 =

(︂𝑧 βˆ’π‘§π‘§β€² 1 βˆ’π‘§β€²

)οΈ‚

for 𝑍 =𝑧𝑒3+𝑧′𝑒4 ∈H2. Conversly, given 𝑧, 𝑧′ ∈H with 𝑧=π‘₯+𝑖𝑦, 𝑧′ =π‘₯β€² +𝑖𝑦′ we set 𝑋𝐿(𝑧, 𝑧′) :=

(οΈ‚π‘₯ π‘¦π‘¦β€²βˆ’π‘₯π‘₯β€² 1 βˆ’π‘₯β€²

)οΈ‚

and π‘ŒπΏ(𝑧, 𝑧′) :=

(︂𝑦 βˆ’π‘₯π‘¦β€²βˆ’π‘₯′𝑦 0 βˆ’π‘¦β€²

)οΈ‚

such that𝑋𝐿(𝑧, 𝑧′) = Re(𝑍𝐿),π‘ŒπΏ(𝑧, 𝑧′) = Im(𝑍𝐿)for𝑍 =𝑧𝑒3+𝑧′𝑒4. Thenπ‘ž(𝑋𝐿(𝑧, 𝑧′)) = π‘ž(π‘ŒπΏ(𝑧, 𝑧′)) =𝑦𝑦′, and the identification ofH2 with the Grassmannian Gr(𝑉) is given by the map 𝑧𝑒3+𝑧′𝑒4 ↦→R𝑋𝐿(𝑧, 𝑧′)βŠ•Rπ‘ŒπΏ(𝑧, 𝑧′).

In order to understand the action of the groupSL2(R)Γ—SL2(R) onH2 we compute 𝑗((𝜎, πœŽβ€²), 𝑍) = ((𝜎, πœŽβ€²).𝑍𝐿, 𝑒1) =𝑗(𝜎, 𝑧)𝑗(πœŽβ€², 𝑧′)

(4.4.2)

for 𝜎, πœŽβ€² ∈ SL2(R) and 𝑍 = 𝑧𝑒3 +𝑧′𝑒4 ∈ H2. Here the automorphy factors on the right are the classical ones defined at the beginning of Section 2.4. Hence we find

((𝜎, πœŽβ€²).𝑍)𝐿=𝑗((𝜎, πœŽβ€²), 𝑍)βˆ’1(𝜎, πœŽβ€²).𝑍𝐿=

(οΈ‚πœŽπ‘§ βˆ’πœŽπ‘§Β·πœŽβ€²π‘§β€² 1 βˆ’πœŽβ€²π‘§β€²

)οΈ‚

= (πœŽπ‘§ 𝑒3 +πœŽβ€²π‘§β€²π‘’4)𝐿 (4.4.3)

for 𝜎, πœŽβ€² ∈ SL2(R) and 𝑍 = 𝑧𝑒3 +𝑧′𝑒4 ∈ H2, i.e., (𝜎, πœŽβ€²).𝑍 = πœŽπ‘§ 𝑒3+πœŽβ€²π‘§β€²π‘’4. Therefore, the identification of H2 with the product HΓ—H is SL2(R)Γ—SL2(R)-equivariant where the latter group naturally acts on HΓ—H via(𝜎, πœŽβ€²).(𝑧, 𝑧′) = (πœŽπ‘§, πœŽβ€²π‘§β€²).

In particular, for π‘˜ ∈ Z a function 𝐹: H2 β†’ C is modular of weight π‘˜ with respect to the group Ξ“(𝑁)Γ—Ξ“(𝑁) in the sense of Definition 4.1.2 if and only if the function 𝑓(𝑧, 𝑧′) = 𝐹(𝑧𝑒3 +𝑧′𝑒4) for 𝑧, 𝑧′ ∈ H is modular of weight π‘˜ for the group Ξ“(𝑁) in the classical sense in both variables, i.e., if

𝑓(𝛾𝑧, 𝛾′𝑧′) = 𝑗(𝛾, 𝑧)π‘˜π‘—(𝛾′, 𝑧′)π‘˜π‘“(𝑧, 𝑧′) for all 𝛾, 𝛾′ βˆˆΞ“(𝑁) and 𝑧, 𝑧′ ∈H.

Forπœ†βˆˆπ‘‰ and 𝑧, 𝑧′ ∈H we denote the orthogonal projection of πœ† onto the positive or negative definite subspace R𝑋𝐿(𝑧, 𝑧′)βŠ•Rπ‘ŒπΏ(𝑧, 𝑧′) or𝑋𝐿(𝑧, 𝑧′)βŠ₯βˆ©π‘ŒπΏ(𝑧, 𝑧′)βŠ₯ in 𝑉(R) by πœ†π‘§,𝑧′ orπœ†(𝑧,𝑧′)βŠ₯, respectively. As π‘ž(Im(𝑍)) = Im(𝑧) Im(𝑧′)we can restate equation (4.1.2) in the present setting as

π‘ž(πœ†π‘§,𝑧′) = |(πœ†, 𝑍𝐿)|2 4 Im(𝑧) Im(𝑧′) (4.4.4)

for πœ† ∈ 𝑉 and 𝑧, 𝑧′ ∈ H with 𝑍 = 𝑧𝑒3 +𝑧′𝑒4. Further, given 𝑧, 𝑧′ ∈ H we denote the majorant of π‘ž associated to the pair (𝑧, 𝑧′) as defined in (4.1.3) by

π‘žπ‘§,𝑧′(πœ†) =π‘ž(πœ†π‘§,𝑧′)βˆ’π‘ž(πœ†(𝑧,𝑧′)βŠ₯), and using (4.4.4) we thus find

π‘žπ‘§,𝑧′(πœ†) = |(πœ†, 𝑍𝐿)|2

2 Im(𝑧) Im(𝑧′)βˆ’π‘ž(πœ†) (4.4.5)

for πœ†βˆˆπ‘‰ and 𝑧, 𝑧′ ∈H with 𝑍 =𝑧𝑒3+𝑧′𝑒4. 86

Eventually, we present the theta liftΞ¦πΏπ‘˜(𝑍;𝐹)from Section 4.2 in the case of𝐿being our current lattice of signature(2,2). Recall thatπ‘ž(Im(𝑍)) = Im(𝑧) Im(𝑧′)for 𝑍 =𝑧𝑒3+𝑧′𝑒4 with 𝑧, 𝑧′ ∈ H. Let now π‘˜ be a non-negative integer. Identifying H2 with the product HΓ—H we can write Shintani’s theta function for the lattice𝐿 as

Θ𝐿,π‘˜(𝜏, 𝑧, 𝑧′) = 𝑣 Im(𝑧)βˆ’π‘˜Im(𝑧′)βˆ’π‘˜ βˆ‘οΈ

π›½βˆˆπΏβ€²/𝐿

βˆ‘οΈ

πœ†βˆˆπΏ+𝛽

(πœ†, 𝑍𝐿)π‘˜π‘’(︁

π‘’π‘ž(πœ†) +π‘–π‘£π‘žπ‘§,𝑧′(πœ†))︁

e𝛽 (4.4.6)

for𝜏, 𝑧, 𝑧′ ∈Hwith 𝜏 =𝑒+𝑖𝑣. The theta functionΘ𝐿,π‘˜(𝜏, 𝑧)is modular of weightπ‘˜ with respect to 𝜌𝐿 in𝜏, and in the variables 𝑧 and 𝑧′ it transforms as

Θ𝐿,π‘˜(𝜏, 𝛾𝑧, 𝛾′𝑧′) = 𝑗(𝛾, 𝑧)π‘˜π‘—(𝛾′, 𝑧′)π‘˜Ξ˜πΏ,π‘˜(𝜏, 𝑧, 𝑧′)

for 𝜏, 𝑧, 𝑧′ ∈H and 𝛾, 𝛾′ βˆˆΞ“(𝑁). Therefore, the (regularized) theta lift of a real analytic function 𝐹:Hβ†’C[𝐿′/𝐿] modular of weight π‘˜ with respect to𝜌𝐿 is given by

Ξ¦πΏπ‘˜(𝑧, 𝑧′;𝐹) = (𝐹,Θ𝐿,π‘˜(Β·, 𝑧, 𝑧′))reg

for 𝑧, 𝑧′ ∈H, and the lift Ξ¦πΏπ‘˜(𝑧, 𝑧′;𝐹) of 𝐹 is modular of weight π‘˜ for the group Ξ“(𝑁)in both variables 𝑧 and 𝑧′ whenever the (regularized) inner product exists. In this setting, the lift Ξ¦πΏπ‘˜(𝑧, 𝑧′;𝐹) is sometimes called the Doi-Naganuma lift of 𝐹, which classically is a lift from scalar valued modular forms to Hilbert modular forms (see [DN67, DN69]).

5 Realizing non-holomorphic

Eisenstein series as theta lifts

In this chapter we realize the non-holomorphic Eisenstein series introduced in Section 2.6 as regularized theta lifts in two different ways. On the one hand we obtain averaged versions of these Eisenstein series as the theta lift of Selberg’s PoincarΓ© series of the first kind using the lattice of signature (2,1)from Section 4.3. Remarkably, we indeed obtain averaged versions of hyperbolic, parabolic an elliptic Eisenstein series as the theta lift of a single type of PoincarΓ© series.

On the other hand we realize the hyperbolic kernel function defined in Section 2.6.4 as the theta lift of a more elementary PoincarΓ© series introduced in Section 3.6, using the lattice of signature (2,2) from Section 4.4. We can then use this representation of the hyperbolic kernel function to obtain individual hyperbolic, parabolic and elliptic Eisenstein series as variations of this theta lift.

5.1 Regularized theta lifts of non-holomorphic PoincarΓ© series

Let (𝑉, π‘ž) be a quadratic space of signature (2, 𝑛) and let 𝐿 be an even lattice in 𝑉. We further fix vectors 𝑒1 ∈ 𝐿 and 𝑒2 ∈ 𝐿′ such that 𝑒1 is primitive with π‘ž(𝑒1) = 0 and (𝑒1, 𝑒2) = 1 as in Section 4.1, and we let H𝑛 be the induced generalized upper half-plane.

For𝛽 βˆˆπΏβ€²/𝐿 and π‘šβˆˆZ+π‘ž(𝛽) we set

𝐿𝛽,π‘š:={πœ† ∈𝐿+𝛽: π‘ž(πœ†) = π‘š}, and we define the subset 𝐻𝛽,π‘šπΏ of H𝑛 by

𝐻𝛽,π‘šπΏ := ⋃︁

πœ†βˆˆπΏπ›½,π‘š πœ†ΜΈ=0

{𝑍 ∈H𝑛: (πœ†, 𝑍𝐿) = 0}.

(5.1.1)

Note that𝐻𝛽,π‘šπΏ =βˆ…ifπ‘šβ‰₯0sinceπœ†βˆˆπΏπ›½,π‘šcan only be orthogonal to some𝑍𝐿=𝑋𝐿+π‘–π‘ŒπΏ with 𝑍 ∈H𝑛 ifπœ†= 0 orπ‘ž(πœ†)<0as Rπ‘‹πΏβŠ•Rπ‘ŒπΏ defines a2-dimensional positive definite subspace of 𝑉(R).

Letπ‘˜ be a non-negative integer and setπœ…:= 1 +π‘˜βˆ’π‘›/2. Then πœ…satisfies the condition given in (3.4.1). In the present section we study the regularized theta lift of two of the non-holomorphic PoincarΓ© series given in Definition 3.4.3, namely Selberg’s PoincarΓ© series of the first kind π‘ˆπœ…,𝛽,π‘šπΏ (𝜏, 𝑠)and the simpler, non-standard PoincarΓ© series π‘„πΏπœ…,𝛽,π‘š(𝜏, 𝑠).

Definition 5.1.1. Letπ‘˜ ∈Z with π‘˜β‰₯0, πœ…= 1 +π‘˜βˆ’π‘›/2,𝛽 βˆˆπΏβ€²/𝐿 and π‘šβˆˆZ+π‘ž(𝛽).

(a) We define the regularized theta lift of Selberg’s PoincarΓ© series of the first kind π‘ˆπœ…,𝛽,π‘šπΏ (𝜏, 𝑠) by

Ξ¦Sel,πΏπ‘˜,𝛽,π‘š(𝑍, 𝑠) = Ξ¦πΏπ‘˜(𝑍;π‘ˆπœ…,𝛽,π‘šπΏ (Β·, 𝑠)) for 𝑍 ∈Hπ‘›βˆ–π»π›½,π‘šπΏ and 𝑠 ∈Cwith Re(𝑠)>1 +𝑛/2.

(b) We define the regularized theta lift of the PoincarΓ© series π‘„πΏπœ…,𝛽,π‘š(𝜏, 𝑠)by Ξ¦Q,πΏπ‘˜,𝛽,π‘š(𝑍, 𝑠) = Ξ¦πΏπ‘˜(𝑍;π‘„πΏπœ…,𝛽,π‘š(Β·, 𝑠))

for 𝑍 ∈H𝑛 and π‘ βˆˆC with Re(𝑠)>1 +𝑛/2.

Here Ξ¦πΏπ‘˜(𝑍; Β·) denotes the regularized theta lift from Definition 4.2.6. We will see in Theorem 5.1.3 that the above definition is indeed well-defined, i.e., that the corresponding (regularized) integrals exist in all the given cases. Thus,Ξ¦Sel,πΏπ‘˜,𝛽,π‘š(𝑍, 𝑠)and Ξ¦Q,πΏπ‘˜,𝛽,π‘š(𝑍, 𝑠) are modular of weight π‘˜ in the variable𝑍 with respect to the groupΞ“(𝐿) defined in (4.1.4).

Before we start studying the above lifts, we quickly recall some technicalities. For 𝑍 ∈H𝑛 the majorant of the present quadratic formπ‘ž associated to𝑍 is given byπ‘žπ‘(πœ†) = π‘ž(πœ†π‘)βˆ’π‘ž(πœ†π‘βŠ₯)forπœ† βˆˆπ‘‰ (see equation (4.1.3)). Hereπœ†π‘ and πœ†π‘βŠ₯ denote the orthogonal projections of πœ†onto the positive or negative definite subspaces of𝑉 corresponding to 𝑍 and 𝑍βŠ₯, respectively. The majorant π‘žπ‘ defines a positive definite quadratic form on 𝑉, and thus the Epstein zeta function associated to π‘žπ‘ and the lattice 𝐿′ defined by

𝑍(𝑠;π‘žπ‘, 𝐿′) := βˆ‘οΈ

πœ†βˆˆπΏβ€²βˆ–{0}

π‘žπ‘(πœ†)βˆ’π‘  (5.1.2)

converges absolutely and locally uniformly for 𝑠 ∈ C with Re(𝑠) > 1 +𝑛/2 (see for example [Eps03]).

Lemma 5.1.2. Let 𝛽 βˆˆπΏβ€²/𝐿, π‘šβˆˆZ+π‘ž(𝛽) and 𝑍 ∈Hπ‘›βˆ–π»π›½,π‘šπΏ . Then the sum

βˆ‘οΈ

πœ†βˆˆπΏπ›½,π‘šβˆ–{0}

π‘ž(πœ†π‘)βˆ’π‘ 

converges absolutely and locally uniformly for π‘ βˆˆC with Re(𝑠)>1 +𝑛/2.

Proof. Firstly, we note that for πœ† ∈ 𝐿𝛽,π‘š we have π‘ž(πœ†π‘) = 0 if and only if (πœ†, 𝑍𝐿) = 0 by (4.1.2), which is the case if and only if πœ† = 0 or 𝑍 ∈ 𝐻𝛽,π‘šπΏ . Thus the given sum is well-defined. Next, we can choose a finite subset 𝐾 βŠ‚ 𝐿′ such that π‘žπ‘(πœ†)β‰₯ 2|π‘š| for all πœ†βˆˆπΏβ€²βˆ–πΎ since π‘žπ‘ is positive definite. Hence we obtain

π‘ž(πœ†π‘) = π‘žπ‘(πœ†) +π‘š

2 = π‘žπ‘(πœ†)

4 + π‘žπ‘(πœ†) + 2π‘š

4 β‰₯ π‘žπ‘(πœ†) 4 for πœ†βˆˆπΏπ›½,π‘šβˆ–πΎ, giving

βˆ‘οΈ

πœ†βˆˆπΏπ›½,π‘š πœ†ΜΈ=0

βƒ’βƒ’π‘ž(πœ†π‘)βˆ’π‘ βƒ’

⃒≀ βˆ‘οΈ

πœ†βˆˆπΏπ›½,π‘šβˆ©πΎ πœ†ΜΈ=0

π‘ž(πœ†π‘)βˆ’Re(𝑠)+ 4Re(𝑠) βˆ‘οΈ

πœ†βˆˆπΏπ›½,π‘šβˆ–πΎ πœ†ΜΈ=0

π‘žπ‘(πœ†)βˆ’Re(𝑠).

Here the first sum is finite and the second sum is dominated by the Epstein zeta function associated toπ‘žπ‘ and 𝐿′ given in (5.1.2), which converges absolutely and locally uniformly for π‘ βˆˆC with Re(𝑠)>1 +𝑛/2.

90

We can now prove that the theta lifts given in Definition 5.1.1 are indeed well-defined.

Moreover, if π‘˜ >0 we already obtain a meromorphic continuation of the two lifts to the half-plane defined by Re(𝑠)>1 +𝑛/2βˆ’π‘˜/2

Theorem 5.1.3. Let π‘˜ ∈Z with π‘˜ β‰₯0, πœ…= 1 +π‘˜βˆ’π‘›/2, 𝛽 βˆˆπΏβ€²/𝐿 and π‘šβˆˆZ+π‘ž(𝛽).

(a) The regularized theta lift of Selberg’s PoincarΓ© seriesπ‘ˆπœ…,𝛽,π‘šπΏ (𝜏, 𝑠)defines a real analytic function in 𝑍 ∈ Hπ‘›βˆ–π»π›½,π‘šπΏ and a holomorphic function in 𝑠 for Re(𝑠) > 1 +𝑛/2, which is given by

Ξ¦Sel,πΏπ‘˜,𝛽,π‘š(𝑍, 𝑠) = 2 Ξ“(𝑠+π‘˜) 4π‘ πœ‹π‘ +π‘˜

βˆ‘οΈ

πœ†βˆˆπΏπ›½,π‘š πœ†ΜΈ=0

π‘ž(πœ†π‘)βˆ’π‘ (πœ†, 𝑍𝐿)βˆ’π‘˜.

For π‘˜ >0 and 𝑍 ∈ Hπ‘›βˆ–π»π›½,π‘šπΏ the sum on the right-hand side yields a holomorphic continuation of Ξ¦Sel,πΏπ‘˜,𝛽,π‘š(𝑍, 𝑠) in 𝑠 to the half-plane Re(𝑠)>1 +𝑛/2βˆ’π‘˜/2.

(b) The regularized theta lift of the PoincarΓ© series π‘„πΏπœ…,𝛽,π‘š(𝜏, 𝑠) defines a real analytic function in 𝑍 ∈ H𝑛 and a holomorphic function in 𝑠 for Re(𝑠)> 1 +𝑛/2, which is given by

Ξ¦Q,πΏπ‘˜,𝛽,π‘š(𝑍, 𝑠) = 2 Ξ“(𝑠+π‘˜)

(2πœ‹)𝑠+π‘˜ π‘ž(Im(𝑍))βˆ’π‘˜ βˆ‘οΈ

πœ†βˆˆπΏπ›½,π‘š

πœ†ΜΈ=0

π‘žπ‘(πœ†)βˆ’π‘ βˆ’π‘˜(πœ†, 𝑍𝐿)π‘˜.

For π‘˜ >0 and 𝑍 ∈H𝑛 the sum on the right-hand side yields a holomorphic continu-ation of Ξ¦Q,πΏπ‘˜,𝛽,π‘š(𝑍, 𝑠) in 𝑠 to the half-plane Re(𝑠)>1 +𝑛/2βˆ’π‘˜/2.

Proof. Since (a) and (b) are poven analogously, we only give a detailed proof for part (a), and afterwards comment on the necessary adaptions of the given proof in the case of (b).

For (a) fix 𝑍 ∈ H𝑛 βˆ–π»π›½,π‘šπΏ and 𝑠 ∈ C with Re(𝑠) > 1 + 𝑛/2. Evaluating the inner product βŸ¨οΈ€

π‘ˆπœ…,𝛽,π‘šπΏ (𝜏, 𝑠),Θ𝐿,π‘˜(𝜏, 𝑍)βŸ©οΈ€

and splitting the sum over matrices 𝑀 ∈ βŸ¨π‘‡βŸ©βˆ–SL2(Z) coming from the PoincarΓ© series into matrices 𝑀 with lower left entry 𝑐̸= 0 and 𝑐= 0 we can write the theta lift Ξ¦Sel,πΏπ‘˜,𝛽,π‘š(𝑍, 𝑠) as

CT𝑑=0 [οΈƒ

𝑇limβ†’βˆž

∫︁

ℱ𝑇

βˆ‘οΈ

π‘€βˆˆβŸ¨π‘‡βŸ©βˆ–SL2(Z) 𝑀̸=Β±1

Im(𝑀 𝜏)𝑠+πœ…π‘’(π‘šπ‘€ 𝜏)Θ𝐿,π‘˜,𝛽(𝑀 𝜏, 𝑍)π‘£βˆ’π‘‘π‘‘πœ‡(𝜏) ]οΈƒ

(5.1.3)

+ 2 CT𝑑=0 [οΈ‚

𝑇limβ†’βˆž

∫︁

ℱ𝑇

𝑣𝑠+πœ…βˆ’π‘‘π‘’(π‘šπœ)Θ𝐿,π‘˜,𝛽(𝜏, 𝑍)π‘‘πœ‡(𝜏) ]οΈ‚

. Here we use the notation𝜏 =𝑒+𝑖𝑣,πœ…= 1 +π‘˜βˆ’π‘›/2and Θ𝐿,π‘˜,𝛽(𝜏, 𝑍) = ⟨Θ𝐿,π‘˜(𝜏, 𝑍),eπ›½βŸ©.

Letβ„›:={𝜏 ∈H: |Re(𝜏)| ≀1/2}. Then β„› is a rectangular fundamental domain for the action of βŸ¨π‘‡βŸ© on H, and by Corollary 4.2.5 we have

∫︁

β„›βˆ–β„±

βƒ’

βƒ’

⃒𝑣𝑠+πœ…π‘’(π‘šπœ)Θ𝐿,π‘˜,𝛽(𝜏, 𝑍)

βƒ’

βƒ’

βƒ’π‘‘πœ‡(𝜏)≀𝐢

∫︁ 1 0

𝑣Re(𝑠)βˆ’π‘›/2βˆ’2

𝑑𝑣 <∞ (5.1.4)

for some constant𝐢 > 0asRe(𝑠)>1 +𝑛/2. Hence we can plug in𝑑= 0 in the first term in (5.1.3), take the limit 𝑇 β†’ ∞, interchange summation and integration and apply the usual unfolding trick to find

CT𝑑=0 [οΈƒ

𝑇limβ†’βˆž

∫︁

ℱ𝑇

βˆ‘οΈ

π‘€βˆˆβŸ¨π‘‡βŸ©βˆ–SL2(Z) 𝑀̸=Β±1

Im(𝑀 𝜏)𝑠+πœ…π‘’(π‘šπ‘€ 𝜏)Θ𝐿,π‘˜,𝛽(𝑀 𝜏, 𝑍)π‘£βˆ’π‘‘π‘‘πœ‡(𝜏) ]οΈƒ

= 2

∫︁

β„›βˆ–β„±

𝑣𝑠+πœ…π‘’(π‘šπœ)Θ𝐿,π‘˜,𝛽(𝜏, 𝑍)π‘‘πœ‡(𝜏).

In the second term in (5.1.3) we split the integral at the horizontal line 𝑣 = 1, i.e., CT𝑑=0

[οΈ‚

𝑇limβ†’βˆž

∫︁

ℱ𝑇

𝑣𝑠+πœ…βˆ’π‘‘π‘’(π‘šπœ)Θ𝐿,π‘˜,𝛽(𝜏, 𝑍)π‘‘πœ‡(𝜏) ]οΈ‚

=

∫︁

β„±1

𝑣𝑠+πœ…π‘’(π‘šπœ)Θ𝐿,π‘˜,𝛽(𝜏, 𝑍)π‘‘πœ‡(𝜏) + CT𝑑=0

[οΈƒ

𝑇limβ†’βˆž

∫︁

β„±π‘‡βˆ–β„±1

𝑣𝑠+πœ…βˆ’π‘‘π‘’(π‘šπœ)Θ𝐿,π‘˜,𝛽(𝜏, 𝑍)π‘‘πœ‡(𝜏) ]οΈƒ

. Here we can drop the regularization in the first term on the right-hand side as β„±1 is compact. Putting everything back together we obtain

Ξ¦Sel,πΏπ‘˜,𝛽,π‘š(𝑍, 𝑠) = 2

∫︁ 1 0

∫︁ 1/2

βˆ’1/2

𝑣𝑠+πœ…π‘’(π‘šπœ)Θ𝐿,π‘˜,𝛽(𝜏, 𝑍)𝑑𝑒𝑑𝑣 𝑣2 (5.1.5)

+ 2 CT𝑑=0 [οΈƒβˆ«οΈ ∞

1

∫︁ 1/2

βˆ’1/2

𝑣𝑠+πœ…βˆ’π‘‘π‘’(π‘šπœ)Θ𝐿,π‘˜,𝛽(𝜏, 𝑍)𝑑𝑒𝑑𝑣 𝑣2

]οΈƒ

. In both terms the integral with respect to 𝑒 is simply given by

∫︁ 1/2

βˆ’1/2

𝑒(π‘šπ‘’)Θ𝐿,π‘˜,𝛽(𝜏, 𝑍)𝑑𝑒.

Plugging in the definition of the 𝛽-th component of the theta function Θ𝐿,π‘˜(𝜏, 𝑍) and using that

βˆ‘οΈ

πœ†βˆˆπΏ+𝛽

∫︁ 1/2

βˆ’1/2

βƒ’

βƒ’

⃒𝑒(π‘šπ‘’)(πœ†, 𝑍𝐿)π‘˜π‘’(𝜏 π‘ž(πœ†π‘) +𝜏 π‘ž(πœ†π‘βŠ₯))

βƒ’

βƒ’

⃒𝑑𝑒= βˆ‘οΈ

πœ†βˆˆπΏ+𝛽

βƒ’βƒ’(πœ†, 𝑍𝐿)π‘˜π‘’βˆ’2πœ‹π‘£π‘žπ‘(πœ†)βƒ’

βƒ’<∞ as the sum defining Θ𝐿,π‘˜,𝛽(𝜏, 𝑍)is absolutely convergent, we obtain

∫︁ 1/2

βˆ’1/2

𝑒(π‘šπ‘’)Θ𝐿,π‘˜,𝛽(𝜏, 𝑍)𝑑𝑒 = 𝑣𝑛/2 π‘ž(Im(𝑍))π‘˜

βˆ‘οΈ

πœ†βˆˆπΏ+𝛽

(πœ†, 𝑍𝐿)π‘˜π‘’βˆ’2πœ‹π‘£π‘žπ‘(πœ†)

∫︁ 1/2

βˆ’1/2

𝑒(𝑒(π‘šβˆ’π‘ž(πœ†))𝑑𝑒.

Hereπ‘žπ‘(πœ†) = π‘ž(πœ†π‘)βˆ’π‘ž(πœ†π‘βŠ₯) as in (4.1.3), and the integral on the right-hand side is1 if π‘ž(πœ†) = π‘š and vanishes otherwise. Therefore, we can write (5.1.5) as

Ξ¦Sel,πΏπ‘˜,𝛽,π‘š(𝑍, 𝑠) = 2 π‘ž(Im(𝑍))π‘˜

(οΈƒβˆ«οΈ 1 0

𝑣𝑠+π‘˜βˆ’1 βˆ‘οΈ

πœ†βˆˆπΏπ›½,π‘š

(πœ†, 𝑍𝐿)π‘˜π‘’βˆ’4πœ‹π‘£π‘ž(πœ†π‘)𝑑𝑣

+ CT𝑑=0 [οΈƒβˆ«οΈ ∞

1

𝑣𝑠+π‘˜βˆ’π‘‘βˆ’1 βˆ‘οΈ

πœ†βˆˆπΏπ›½,π‘š

(πœ†, 𝑍𝐿)π‘˜π‘’βˆ’4πœ‹π‘£π‘ž(πœ†π‘)𝑑𝑣 ]οΈƒ)οΈƒ

.

92

Here the contribution of the element πœ†= 0 appearing only if π‘š =𝛽 = 0 is zero since

∫︁ 1 0

𝑣𝑠+π‘˜βˆ’1𝑑𝑣+ CT𝑑=0

∫︁ ∞ 1

𝑣𝑠+π‘˜βˆ’π‘‘βˆ’1𝑑𝑣= 1

𝑠+π‘˜ βˆ’CT𝑑=0

(οΈ‚ 1 𝑠+π‘˜βˆ’π‘‘

)οΈ‚

= 0.

Hence we can exclude the element πœ† = 0 in each of the two sums above. Moreover, we reunite the remaining terms, giving

Ξ¦Sel,πΏπ‘˜,𝛽,π‘š(𝑍, 𝑠) = 2

π‘ž(Im(𝑍))π‘˜CT𝑑=0 [οΈƒβˆ«οΈ ∞

0

𝑣𝑠+π‘˜βˆ’π‘‘βˆ’1 βˆ‘οΈ

πœ†βˆˆπΏπ›½,π‘š πœ†ΜΈ=0

(πœ†, 𝑍𝐿)π‘˜π‘’βˆ’4πœ‹π‘£π‘ž(πœ†π‘)𝑑𝑣 ]οΈƒ

. (5.1.6)

Next we note that π‘ž(πœ†π‘)>0 and (πœ†, 𝑍𝐿)ΜΈ= 0 for πœ†βˆˆπΏπ›½,π‘š with πœ†ΜΈ= 0 as𝑍 ∈Hπ‘›βˆ–π»π›½,π‘šπΏ . Thus, using equation (4.1.2) we find that

βˆ‘οΈ

πœ†βˆˆπΏπ›½,π‘š πœ†ΜΈ=0

∫︁ ∞ 0

βƒ’

βƒ’

⃒𝑣𝑠+π‘˜βˆ’1(πœ†, 𝑍𝐿)π‘˜π‘’βˆ’4πœ‹π‘£π‘ž(πœ†π‘)

βƒ’

βƒ’

⃒𝑑𝑣= Ξ“(Re(𝑠) +π‘˜)π‘ž(Im(𝑍))π‘˜/2 4Re(𝑠)+π‘˜/2πœ‹Re(𝑠)+π‘˜

βˆ‘οΈ

πœ†βˆˆπΏπ›½,π‘š πœ†ΜΈ=0

π‘ž(πœ†π‘)βˆ’Re(𝑠)βˆ’π‘˜/2 (5.1.7)

forRe(𝑠)+π‘˜ >0, and by Lemma 5.1.2 the sum on the right-hand side converges absolutely and locally uniformly for Re(𝑠)>1 +𝑛/2βˆ’π‘˜/2. Therefore we can simply plug in 𝑑= 0 on the right-hand side of (5.1.6), interchange summation and integration and use again (4.1.2) to proof the statement given in part (a) of the theorem.

In order to also proof part (b) we need to modify the above proof of (a) as follows: As in equation (5.1.6) we obtain

Ξ¦Q,πΏπ‘˜,𝛽,π‘š(𝑍, 𝑠) = 2

π‘ž(Im(𝑍))π‘˜CT𝑑=0 [οΈƒβˆ«οΈ ∞

0

𝑣𝑠+π‘˜βˆ’π‘‘βˆ’1 βˆ‘οΈ

πœ†βˆˆπΏπ›½,π‘š

πœ†ΜΈ=0

(πœ†, 𝑍𝐿)π‘˜π‘’βˆ’2πœ‹π‘£π‘žπ‘(πœ†)𝑑𝑣 ]οΈƒ

(5.1.8)

for 𝑍 ∈ H𝑛 (without the exclusion of 𝐻𝛽,π‘šπΏ ) and 𝑠 ∈ C with Re(𝑠) >1 +𝑛/2, where in (5.1.6) we find the term π‘’βˆ’4πœ‹π‘£π‘ž(πœ†π‘) = π‘’βˆ’2πœ‹π‘£π‘žπ‘(πœ†)π‘’βˆ’2πœ‹π‘£π‘š instead of π‘’βˆ’2πœ‹π‘£π‘žπ‘(πœ†) as in (5.1.8).

Here the missing factor, namely π‘’βˆ’2πœ‹π‘£π‘š, is exactly the term by which the two building blocks 𝑣𝑠𝑒(π‘šπœ)e𝛽 and 𝑣𝑠𝑒(π‘šπ‘’)e𝛽 defining the PoincarΓ© seriesπ‘ˆπœ…,𝛽,π‘šπΏ (𝜏, 𝑠)andπ‘„πΏπœ…,𝛽,π‘š(𝜏, 𝑠) differ. Now as in equation (5.1.7) we see that

βˆ‘οΈ

πœ†βˆˆπΏπ›½,π‘š πœ†ΜΈ=0

∫︁ ∞ 0

βƒ’

βƒ’

⃒𝑣𝑠+π‘˜βˆ’1(πœ†, 𝑍𝐿)π‘˜π‘’βˆ’2πœ‹π‘£π‘žπ‘(πœ†)

βƒ’

βƒ’

⃒𝑑𝑣 (5.1.9)

= Ξ“(Re(𝑠) +π‘˜)π‘ž(Im(𝑍))π‘˜/2 2Re(𝑠)πœ‹Re(𝑠)+π‘˜

βˆ‘οΈ

πœ†βˆˆπΏπ›½,π‘š πœ†ΜΈ=0

π‘ž(πœ†π‘)π‘˜/2π‘žπ‘(πœ†)βˆ’Re(𝑠)βˆ’π‘˜,

where we have used again the identity (4.1.2). As π‘ž(πœ†π‘)β‰€π‘žπ‘(πœ†) = π‘ž(πœ†π‘) +|π‘ž(πœ†π‘βŠ₯)|, and since the sumβˆ‘οΈ€

πœ†βˆˆπΏβ€²βˆ–{0}π‘žπ‘(πœ†)βˆ’π‘  converges forRe(𝑠)>1 +𝑛/2as remarked in (5.1.2), we find that the right-hand side of (5.1.9) converges for Re(𝑠)> 1 +𝑛/2βˆ’π‘˜/2. Therefore we can plug in 𝑑 = 0 on the right-hand side of (5.1.8), interchange summation and integration, and compute the remaining integral. This finishes the proof of part (b).

Remark 5.1.4.

(1) If𝛽 =βˆ’π›½in𝐿′/𝐿and2βˆ’π‘›βˆ’2πœ…β‰‘2mod4then the PoincarΓ© seriesπ‘ˆπœ…,𝛽,π‘šπΏ andπ‘„πΏπœ…,𝛽,π‘š vanish identically. This matches the two formulas given in the previous theorem since if 𝛽 = βˆ’π›½ then 𝐿𝛽,π‘š = βˆ’πΏπ›½,π‘š and thus the sums in (a) and (b) of Theorem 5.1.3 cancel completely if π‘˜ is odd as

π‘ž((βˆ’πœ†)𝑍) =π‘ž(πœ†π‘), π‘žπ‘(βˆ’πœ†) =π‘žπ‘(πœ†), (βˆ’πœ†, 𝑍𝐿) = βˆ’(πœ†, 𝑍𝐿)

for πœ†βˆˆπ‘‰ and 𝑍 ∈H𝑛. On the other hand, the congruence 2βˆ’π‘›βˆ’2πœ…β‰‘2 mod4 is satisfied if and only if the corresponding non-negative integerπ‘˜(withπœ…= 1+π‘˜βˆ’π‘›/2) is odd.

(2) Ifπ‘š = 0 thenπ‘ˆπœ…,𝛽,0𝐿 (𝜏, 𝑠) = π‘„πΏπœ…,𝛽,0(𝜏, 𝑠), and thus also the corresponding lifts need to agree, i.e., we have

Ξ¦Sel,πΏπ‘˜,𝛽,0(𝑍, 𝑠) = Ξ¦Q,πΏπ‘˜,𝛽,0(𝑍, 𝑠)

for 𝑍 ∈ H𝑛 and 𝑠 ∈ C with Re(𝑠) >1 +𝑛/2βˆ’π‘˜/2. This agrees with the formulas given in Theorem 5.1.3 since

π‘ž(πœ†π‘) = π‘žπ‘(πœ†)

2 and (πœ†, 𝑍𝐿) = 2π‘žπ‘(πœ†)π‘ž(Im(𝑍)) (πœ†, 𝑍𝐿)

for πœ†βˆˆπΏπ›½,0 with πœ†ΜΈ= 0 and 𝑍 ∈H𝑛. Here the first equality is clear asπ‘ž(πœ†) = 0, and the second equality follows from the identity in (4.1.2).

In the following two sections we now evaluate the above theta lifts for the special lattices of signature (2,1) and (2,2) from Section 4.3 and Section 4.4, respectively. Though we are mainly interested in the lift of Selberg’s PoincarΓ© series in the case of signature(2,1), and in the lift of the PoincarΓ© series π‘„πΏπœ…,𝛽,π‘š(𝜏, 𝑠) in the case of signature (2,2), we also present the opposite cases for the sake of completeness.

5.2 Averaged non-holomorphic Eisenstein series as