• Keine Ergebnisse gefunden

Finally, we present an analog of Proposition 3.6.6 for the Maass-Selberg PoincarΓ© series.

Proposition 3.6.13. Let 𝛽 ∈ 𝐿′/𝐿 and π‘š ∈ Z+π‘ž(𝛽) with π‘š < 0. Then the PoincarΓ© series π‘€π‘˜,𝛽,π‘šπΏ (𝜏, 𝑠) has a Fourier expansion of the form

𝑣𝑠/2β„³π‘˜,𝑠/2(4πœ‹π‘šπ‘£)𝑒(π‘šπ‘’)(e𝛽+ (βˆ’1)(𝑏+βˆ’π‘βˆ’βˆ’2π‘˜)/2eβˆ’π›½) + βˆ‘οΈ

π›ΎβˆˆπΏβ€²/𝐿

βˆ‘οΈ

π‘›βˆˆZ+π‘ž(𝛾)

𝑏(𝑗)(𝛾, 𝑛, 𝑣, 𝑠)𝑒(𝑛𝑒)e𝛾, where the Fourier coefficients are given by

𝑏(𝛾, 𝑛, 𝑣, 𝑠) = 𝑣𝑠/2 βˆ‘οΈ

π‘βˆˆZβˆ–{0}

|𝑐|1βˆ’π‘˜βˆ’π‘ π»π‘,π‘˜πΏ (𝛽, π‘š, 𝛾, 𝑛)π½π‘˜,π‘š(𝑛, 𝑣, 𝑠, 𝑐).

Here 𝐻𝑐,π‘˜πΏ (𝛽, π‘š, 𝛾, 𝑛) is the generalised Kloosterman sum from Definition 3.3.1, and the integral function

π½π‘˜,π‘š(𝑛, 𝑣, 𝑠, 𝑐) :=π‘–π‘˜

∫︁ ∞

βˆ’βˆž

πœβˆ’π‘˜|𝜏|βˆ’π‘ β„³π‘˜,𝑠/2

(οΈ‚4πœ‹π‘šπ‘£ 𝑐2|𝜏|2

)οΈ‚

𝑒 (οΈ‚

βˆ’ π‘šπ‘’

𝑐2|𝜏|2 βˆ’π‘›π‘’ )οΈ‚

𝑑𝑒 is analytic in 𝑣 for 𝑣 >0 and holomorphic in 𝑠 for Re(𝑠)>1/2.

Proof. The proof runs completely analogous to the the one of Proposition 3.6.6, where we only have to consider the expression

∫︁ ∞

βˆ’βˆž

|𝜏|βˆ’π‘˜βˆ’π‘ 

βƒ’

βƒ’

βƒ’

βƒ’

β„³π‘˜,𝑠/2

(οΈ‚4πœ‹π‘šπ‘£ 𝑐2|𝜏|2

)οΈ‚βƒ’

βƒ’

βƒ’

βƒ’ 𝑑𝑒 (3.6.20)

instead of (3.6.7) to prove the existence of the present integral function π½π‘˜,π‘š(𝑛, 𝑣, 𝑠, 𝑐).

We estimate (3.6.20) by sup

π‘’βˆˆR

βƒ’

βƒ’

βƒ’

βƒ’

πœπ‘ βˆ’π‘˜β„³π‘˜,𝑠/2

(οΈ‚4πœ‹π‘šπ‘£ 𝑐2|𝜏|2

)οΈ‚βƒ’

βƒ’

βƒ’

βƒ’

Β·

∫︁ ∞

βˆ’βˆž

(𝑒2+𝑣2)βˆ’Re(𝑠)𝑑𝑒.

(3.6.21)

Using (3.6.18) we find 𝐢 >0 such that

βƒ’

βƒ’

βƒ’

βƒ’ β„³π‘˜,𝑠/2

(οΈ‚4πœ‹π‘šπ‘£ 𝑐2|𝜏|2

)οΈ‚βƒ’

βƒ’

βƒ’

βƒ’

≀𝐢|𝜏|βˆ’Re(𝑠)+π‘˜

for allπ‘’βˆˆR. Thus, by (3.6.8) the expression in (3.6.21) is bounded forRe(𝑠)>1/2.

for 𝑀, 𝑀′ ∈SL2(Z)and any 𝜏 ∈H. Since(𝑀, πœ‘π‘€)∘(𝑀′, πœ‘π‘€β€²) = (𝑀 𝑀′,Β±πœ‘π‘€ 𝑀′(𝜏)) the quantity 𝜎(𝑀, 𝑀′) is simply a sign, i.e., 𝜎(𝑀, 𝑀′) =Β±1, and this sign does not depend on the choice of𝜏 ∈Has the right-hand side of (3.7.1) is continuous in𝜏. We further set

πœŽπ‘˜(𝑀, 𝑀′) = 𝜎(𝑀, 𝑀′)2π‘˜ (3.7.2)

for 𝑀, 𝑀′ ∈ SL2(Z). Then 𝜌𝐿(𝐸2, 𝜎(𝑀, 𝑀′)) = πœŽπ‘˜(𝑀, 𝑀′) idC[𝐿′/𝐿] as we assume that 2π‘˜β‰‘π‘+βˆ’π‘βˆ’ mod 2 and thus

𝜌𝐿(^𝑀 𝑀′) =πœŽπ‘˜(𝑀, 𝑀′)𝜌𝐿(̃︁𝑀)𝜌𝐿(𝑀̃︁′) (3.7.3)

for 𝑀, 𝑀′ ∈SL2(Z), i.e., the Weil representation 𝜌𝐿 satisfies equation (1.14) in [Roe66].

However, 𝜌𝐿 does not satisfy equation (1.15) in [Roe66] as the negative identity matrix does not act as a scalar multiplication on C[𝐿′/𝐿] but as

e𝛾⃒

βƒ’π‘˜,πΏβˆ’1 = (βˆ’1)̃︁ 𝑏+/2βˆ’π‘βˆ’/2βˆ’π‘˜eβˆ’π›Ύ

(3.7.4)

for𝛾 βˆˆπΏβ€²/𝐿, as was already noted in (3.4.6). Therefore we need to work with a subspace of C[𝐿′/𝐿]on which 𝜌𝐿(βˆ’1)̃︁ does act as a simple scalar multiplication.

To motivate the definition of this subspace we first note that if 𝐹: H β†’ C[𝐿′/𝐿] is modular of weight π‘˜ with respect to𝜌𝐿 then 𝐹|π‘˜,πΏβˆ’1 =̃︁ 𝐹 and thus

𝑓𝛾 =πœ‰ π‘“βˆ’π›Ύ

for all 𝛾 βˆˆπΏβ€²/𝐿 where 𝐹 =βˆ‘οΈ€

π›ΎβˆˆπΏβ€²/𝐿𝑓𝛾e𝛾 and

πœ‰:= (βˆ’1)𝑏+/2βˆ’π‘βˆ’/2βˆ’π‘˜

is the fixed sign, determined by the weight π‘˜ action of βˆ’1̃︁ as in (3.7.4). So the image of 𝐹 does always lie in the subspace

𝒰 :=

{οΈƒ

βˆ‘οΈ

π›ΎβˆˆπΏβ€²/𝐿

πœ†π›Ύe𝛾 ∈C[𝐿′/𝐿] : πœ†π›Ύ =πœ‰ πœ†βˆ’π›Ύ for all 𝛾 βˆˆπΏβ€²/𝐿 }οΈƒ

of C[𝐿′/𝐿]. We fix the following notation: Let

𝐿′/𝐿={𝛾1,βˆ’π›Ύ1, . . . , π›Ύπ‘ž,βˆ’π›Ύπ‘ž, π›Ύπ‘ž+1, . . . , π›Ύπ‘Ÿ}

with 𝛾1,βˆ’π›Ύ1, . . . , π›Ύπ‘ž,βˆ’π›Ύπ‘ž, π›Ύπ‘ž+1, . . . , π›Ύπ‘Ÿ pairwise distinct and 2𝛾𝑗 = 0 for 𝑗 = π‘ž+ 1, . . . , π‘Ÿ, and set

^e𝑗 :=

{οΈƒ1

2(e𝛾𝑗 +πœ‰eβˆ’π›Ύπ‘—), if 1≀𝑗 β‰€π‘ž, e𝛾𝑗, if π‘ž+ 1 ≀𝑗 β‰€π‘Ÿ.

Then (^e1, . . . ,^eπ‘š) with π‘š = π‘Ÿ if πœ‰ = 1 and π‘š = π‘ž otherwise defines an orthonormal basis of the subspace𝒰 where we equip 𝒰 with the scalar product coming from 𝐢[𝐿′/𝐿].

Moreover, we find

𝜌𝐿( Λœπ‘‡)^e𝑗 =𝑒(π‘ž(𝛾𝑗))^e𝑗

and

𝜌𝐿( Λœπ‘†)^e𝑗 =

⎧

βŽͺβŽͺ

βŽͺβŽͺ

⎨

βŽͺβŽͺ

βŽͺβŽͺ

⎩

𝑒((π‘βˆ’βˆ’π‘+)/8)

βˆšοΈ€|𝐿′/𝐿|

π‘Ÿ

βˆ‘οΈ

β„“=1

(𝑒(βˆ’(𝛾𝑗, 𝛾ℓ)) +πœ‰ 𝑒((𝛾𝑗, 𝛾ℓ))) ^eβ„“, if 1≀𝑗 β‰€π‘ž, 𝑒((π‘βˆ’βˆ’π‘+)/8)

βˆšοΈ€|𝐿′/𝐿|

π‘Ÿ

βˆ‘οΈ

β„“=1

𝑒(βˆ’(𝛾𝑗, 𝛾ℓ))^eβ„“, if π‘ž+ 1≀𝑗 β‰€π‘š, for 𝑗 = 1, . . . , π‘š. Thus, the Weil representation 𝜌𝐿 fixes the subspace 𝒰, and we can therefore define the map

𝜈: SL2(Z)β†’Aut(𝒰), 𝑀 β†¦β†’πœŒπΏ( Λœπ‘€), (3.7.5)

which now satisfies

𝜈(βˆ’1)^e𝑗 =𝜌𝐿(βˆ’1)^̃︁ e𝑗 =𝑒(βˆ’π‘˜/2)^e𝑗 (3.7.6)

for 𝑗 = 1, . . . , π‘š. So by (3.7.3) and (3.7.6) the mapping 𝜈 is indeed a unitary multiplier system of weight π‘˜ for the group SL2(Z) and for the space 𝒰 in the sense of Roelcke (compare [Roe66], Section 1.6).

As in equation (1.16) of [Roe66] we call a vector valued function 𝐺: Hβ†’ 𝒰 modular of weight π‘˜ in the sense of Roelcke if 𝐺|Rπ‘˜,𝐿𝑀 =𝐺for all 𝑀 ∈SL2(Z) where

(︀𝐺⃒

βƒ’

R π‘˜,𝐿𝑀)οΈ€

(𝜏) :=

(οΈ‚ πœ‘π‘€(𝜏)

|πœ‘π‘€(𝜏)|

)οΈ‚βˆ’2π‘˜

𝜈(𝑀)βˆ’1𝐺(𝑀 𝜏) (3.7.7)

for 𝑀 ∈SL2(Z) and 𝜏 ∈H. We then have the following identification:

Proposition 3.7.1. A function 𝐹: H β†’C[𝐿′/𝐿] is modular of weight π‘˜ with respect to the Weil representation 𝜌𝐿 if and only if the function 𝐺(𝜏) := Im(𝜏)π‘˜/2𝐹(𝜏) is modular of weight π‘˜ in the sense of Roelcke.

Proof. Since the elements of Mp2(Z)can be written as(𝑀,Β±πœ‘π‘€)with𝑀 ∈SL2(Z), and since (1,βˆ’1)∈ Mp2(Z) acts trivially on any vector valued function H β†’ C[𝐿′/𝐿] as we assume that 2π‘˜ ≑ 𝑏+βˆ’π‘βˆ’ mod 2, a function 𝐹: H β†’ C[𝐿′/𝐿] is modular of weight π‘˜ with respect to 𝜌𝐿 if and only if𝐹 |π‘˜,𝐿(𝑀, πœ‘π‘€) =𝐹 for all 𝑀 ∈SL2(Z). Moreover, it is easy to check that𝐹 |π‘˜,𝐿(𝑀, πœ‘π‘€) = 𝐹 if and only if

Im(𝜏)π‘˜/2𝐹(𝜏) =

(οΈ‚ πœ‘π‘€(𝜏)

|πœ‘π‘€(𝜏)|

)οΈ‚βˆ’2π‘˜

𝜌𝐿(𝑀)βˆ’1(οΈ€

Im(𝑀 𝜏)π‘˜/2𝐹(𝑀 𝜏))οΈ€

for 𝑀 ∈SL2(Z). This proves the claimed statement.

We denote the corresponding identification map by Ξ π‘˜, i.e., for 𝐹: H β†’ C[𝐿′/𝐿] we define

Ξ π‘˜(𝐹)(𝜏) := Im(𝜏)π‘˜/2𝐹(𝜏) (3.7.8)

66

for 𝜏 ∈H. Clearly,Ξ βˆ’1π‘˜ (𝐺) = Im(𝜏)βˆ’π‘˜/2𝐺(𝜏) for 𝐺:H β†’ 𝒰, and by the above consider-ations Ξ π‘˜(𝐹) is modular of weight π‘˜ in the sense of Roelcke if and only if 𝐹 is modular of weight π‘˜ with respect to 𝜌𝐿.

In order to state the spectral theorem given as a combination of Satz 7.2 and Satz 12.3 in [Roe67], which we will then transfer to the setting of (non-holomorphic) vector valued modular forms for the Weil representation, we need to introduce some more of Roelckes notation. For 𝐺, 𝐺′:Hβ†’ 𝒰 modular of weightπ‘˜ in the sense of Roelcke and measurable with respect to πœ‡we define the scalar product

(𝐺, 𝐺′)𝑅 :=

∫︁

SL2(Z)βˆ–H

⟨𝐺(𝜏), 𝐺′(𝜏)βŸ©π‘‘πœ‡(𝜏)

whenever the integral on the right-hand side exists. We then denote the Hilbert space of πœ‡-measurable functions𝐺: Hβ†’ 𝒰 that are modular of weight π‘˜ in the sense of Roelcke and square-integrable with respect to the above scalar product, i.e., (𝐺, 𝐺)𝑅 < ∞, by β„‹π‘˜(SL2(Z),𝒰, 𝜈). Further, we writeπ’žπ‘˜2(SL2(Z),𝒰, 𝜈)for the space of functions𝐺: Hβ†’ 𝒰 modular of weight π‘˜ in the sense of Roelcke which are two times continuously partially differentiable.

On π’žπ‘˜2(SL2(Z),𝒰, 𝜈) we consider Roelcke’s hyperbolic Laplace operator Ξ”π‘…π‘˜ :=βˆ’π‘£2

(οΈ‚ πœ•2

πœ•π‘’2 + πœ•2

πœ•π‘£2 )οΈ‚

βˆ’π‘–π‘˜π‘£ πœ•

πœ•π‘’, (3.7.9)

which is invariant under the corresponding weight π‘˜ action of SL2(Z), i.e., we have Ξ”π‘…π‘˜(︁

𝐺⃒

βƒ’

R π‘˜,𝐿𝑀)︁

=(οΈ€

Ξ”π‘…π‘˜πΊ)οΈ€ βƒ’

βƒ’

R π‘˜,𝐿𝑀

for 𝐺: Hβ†’ 𝒰 two times continuously partially differentiable and 𝑀 ∈SL2(Z). An easy calculation shows that the relation between Roelcke’s Laplace operator and the hyperbolic Laplace operator from (2.5.2) is given by

Ξ”π‘…π‘˜ (Ξ π‘˜(𝐹)(𝜏)) = Ξ π‘˜ (οΈ‚(οΈ‚

Ξ”π‘˜βˆ’π‘˜2βˆ’2π‘˜ 4

)οΈ‚

𝐹(𝜏) )οΈ‚

(3.7.10)

for 𝐹: H β†’ C[𝐿′/𝐿] two times continuously partially differentiable. As in [Roe66], equation (2.21), we now define the space

π’Ÿ2π‘˜(SL2(Z),𝒰, 𝜈) :={οΈ€

𝐺∈ π’žπ‘˜2(SL2(Z),𝒰, 𝜈) : 𝐺,Ξ”π‘…π‘˜πΊβˆˆ β„‹π‘˜(SL2(Z),𝒰, 𝜈)}οΈ€

.

In [Roe66], Satz 3.2, the author proves that the restriction of the operatorΞ”π‘…π‘˜ to the space π’Ÿ2π‘˜(SL2(Z),𝒰, 𝜈)has a self-adjoint continuationΞ”Λœπ‘…π‘˜, and this continuation is then used to show thatΞ”π‘…π‘˜ itself has a countable orthonormal system of eigenfunctions(πœ“π‘—)π‘—βˆˆN0 and a countable system of pairwise orthogonal eigenpackages (πœˆπ‘—,πœ†)π‘—βˆˆπ‘0 in the sense of [Roe66], Definition 5.1, such that the system of all these eigenfunctions πœ“π‘— and eigenpackages πœˆπ‘—,πœ† is complete in the Hilbert space β„‹π‘˜(SL2(Z),𝒰, 𝜈) (see [Roe66], Satz 5.7). In particular, the eigenfunctions πœ“π‘— are real analytic elements of the space π’Ÿ2π‘˜(SL2(Z),𝒰, 𝜈).

In [Roe67] Roelcke uses Eisenstein series to calculate a nicer expression for the terms coming from the eigenpackagesπœˆπ‘—,πœ† in the spectral expansion of a square-integrable func-tion (compare Satz 7.2 and 12.3 in [Roe67]). Let β„’ be the eigenraum for the eigenvalue 1 of the automorphism𝜈(𝑇) : 𝒰 β†’ 𝒰 with 𝑇 =(οΈ€1 1

0 1

)οΈ€, i.e., β„’:={π‘₯∈ 𝒰: 𝜈(𝑇)π‘₯=π‘₯}=

{οΈƒ

βˆ‘οΈ

π›ΎβˆˆπΏβ€²/𝐿 π‘ž(𝛾)=0

πœ†π›Ύe𝛾 ∈ 𝒰 }οΈƒ

.

Evidently, an orthonormal basis of β„’ is given by the set {^e𝑗: 1≀𝑗 β‰€π‘š and π‘ž(𝛾𝑗) = 0}.

Note thatβ„’might indeed be the zero-space ifπœ‰=βˆ’1and the group𝐿′/𝐿does not contain any norm-zero element of order larger than two. In this case there are no Eisenstein series to consider.

Now, given an element h ∈ β„’ Roelcke defines the non-holomorphic Eisenstein series associated to h via

πΈπ‘˜,h𝐿,𝑅(𝜏, 𝑠) := 1 2

βˆ‘οΈ

π‘€βˆˆβŸ¨π‘‡βŸ©βˆ–SL2(Z)

Im(𝜏)𝑠h

βƒ’

βƒ’

βƒ’

R π‘˜,𝐿𝑀 (3.7.11)

for 𝜏 ∈ H and 𝑠 ∈ C with Re(𝑠) > 1 (see [Roe67], eq. (10.1)). Note that we can omit some of the notation in [Roe67] as we only work with the full modular group SL2(Z) which has exactly one cusp, namely ∞. One can check that there are exactly dim(β„’) linearly independent Eisenstein series of the above type. Thus, it suffices to consider the Eisenstein series πΈπ‘˜,^𝐿,𝑅e

𝑗(𝜏, 𝑠)for 𝑗 = 1, . . . , π‘šwith π‘ž(𝛾𝑗) = 0.

We can now finally state [Roe67], Satz 7.2, where we replace the second part of the spectral expansion by the expression given in [Roe67], Satz 12.3:

Theorem 3.7.2 ([Roe67], Satz 7.2 and 12.3). Let 𝐺∈ β„‹π‘˜(SL2(Z),𝒰, 𝜈). Then 𝐺(𝜏) =

∞

βˆ‘οΈ

𝑗=0

(𝐺, πœ“π‘—)π‘…πœ“π‘—(𝜏) + 1 4πœ‹

βˆ‘οΈ

𝑗=1,...,π‘š π‘ž(𝛾𝑗)=0

∞

∫︁

βˆ’βˆž

(οΈ‚

𝐺, πΈπ‘˜,^𝐿,𝑅e

𝑗

(οΈ‚

Β·,1 2+π‘–π‘Ÿ

)οΈ‚)︂𝑅

πΈπ‘˜,^𝐿,𝑅e

𝑗

(οΈ‚

𝜏,1 2 +π‘–π‘Ÿ

)οΈ‚

π‘‘π‘Ÿ.

Here(πœ“π‘—)π‘—βˆˆN0 βŠ† π’Ÿ2π‘˜(SL2(Z),𝒰, 𝜈)is an orthonormal system of real analytic eigenfunctions of Roelcke’s hyperbolic Laplace operator Ξ”π‘…π‘˜ as introduced above, and the elements^e𝑗 for 𝑗 = 1, . . . , π‘šwith π‘ž(𝛾𝑗) = 0 form an orthonormal basis of the 1-eigenraum β„’ of 𝜈((οΈ€1 1

0 1

)οΈ€).

Using the identification given in (3.7.8) we can reformulate the above result in the setting of this work, i.e., for vector valued functions modular with respect to the Weil representation. In oder to do so we quickly comment on the necessary translations:

(1) For𝐹, 𝐹′:Hβ†’C[𝐿′/𝐿]modular of weightπ‘˜with respect to𝜌𝐿 and measurable with respect to πœ‡we have

(Ξ π‘˜(𝐹),Ξ π‘˜(𝐹′))𝑅= (𝐹, 𝐹′).

68

In particular, 𝐹 is square-integrable with respect to the scalar product (Β·,Β·) if and only if Ξ π‘˜(𝐹) is square-integrable with respect to (Β·,Β·)𝑅. Defining

β„‹π‘˜,𝐿 :={𝐹 ∈ π’œπ‘˜,𝐿: 𝐹 is πœ‡-measurable and (𝐹, 𝐹)<∞}, (3.7.12)

where π’œπ‘˜,𝐿 is the space of functions 𝐹: H β†’ C[𝐿′/𝐿] modular of weight π‘˜ with respect to 𝜌𝐿, we thus find that Ξ π‘˜(β„‹π‘˜,𝐿) = β„‹π‘˜(SL2(Z),𝒰, 𝜈).

(2) As in [Roe66], Satz 3.2, the restriction of the hyperbolic Laplace operator Ξ”π‘˜ to the space

π’Ÿπ‘˜,𝐿 := Ξ βˆ’1π‘˜ (π’Ÿπ‘˜2(SL2(Z),𝒰, 𝜈)) (3.7.13)

has a self-adjoint continuation which we denote by Ξ”Λœπ‘˜. Let (πœ“π‘—)π‘—βˆˆN0 be an orthonor-mal system of eigenfunctions of Roelcke’s Laplace operator Ξ”π‘…π‘˜ as in the above the-orem. Then (Ξ βˆ’1π‘˜ (πœ“π‘—))π‘—βˆˆN0 is an orthonormal system of eigenfunctions of the hyper-bolic Laplace operator Ξ”π‘˜ which is complete in the Hilbert space β„‹π‘˜,𝐿 up to some eigenpackages corresponding to Eisenstein series.

(3) A direct computation shows that Ξ βˆ’1π‘˜ (οΈ€

πΈπ‘˜,^𝐿,𝑅e

𝑗(𝜏, 𝑠))οΈ€

= 1 2πΈπ‘˜,𝛾𝐿

𝑗(𝜏, π‘ βˆ’π‘˜/2) for 𝑗 = 1, . . . , π‘šwith π‘ž(𝛾𝑗) = 0.

Therefore, Theorem 3.7.2 translates to the following statement:

Theorem 3.7.3. Let 𝐹 ∈ β„‹π‘˜,𝐿. Then 𝐹(𝜏) =

∞

βˆ‘οΈ

𝑗=0

(𝐹, πœ“π‘—)πœ“π‘—(𝜏)

+ 1 16πœ‹

βˆ‘οΈ

𝑗=1,...,π‘š π‘ž(𝛾𝑗)=0

∞

∫︁

βˆ’βˆž

(οΈ‚

𝐹, πΈπ‘˜,𝛾𝐿

𝑗

(οΈ‚

Β·,1βˆ’π‘˜ 2 +π‘–π‘Ÿ

)οΈ‚)οΈ‚

πΈπ‘˜,𝛾𝐿

𝑗

(οΈ‚

𝜏,1βˆ’π‘˜ 2 +π‘–π‘Ÿ

)οΈ‚

π‘‘π‘Ÿ.

Here (πœ“π‘—)π‘—βˆˆN0 βŠ† π’Ÿπ‘˜,𝐿 is an orthonormal system of real analytic eigenfunctions of the hyperbolic Laplace operator Ξ”π‘˜, and

𝐿′/𝐿={𝛾1,βˆ’π›Ύ1, . . . , π›Ύπ‘ž,βˆ’π›Ύπ‘ž, π›Ύπ‘ž+1, . . . , π›Ύπ‘Ÿ}

with 𝛾1,βˆ’π›Ύ1, . . . , π›Ύπ‘ž,βˆ’π›Ύπ‘ž, π›Ύπ‘ž+1, . . . , π›Ύπ‘Ÿ pairwise distinct and 2𝛾𝑗 = 0 for 𝑗 = π‘ž+ 1, . . . , π‘Ÿ.

Further, we have π‘š=π‘Ÿ if (βˆ’1)𝑏+/2βˆ’π‘βˆ’/2βˆ’π‘˜= 1 and π‘š =π‘ž otherwise.

To the end of this section we further study eigenvalues of the operator Ξ”π‘˜ and their corresponding eigenfunctions, following [Roe66]. Recall that Ξ”Λœπ‘…π‘˜ denotes the self-adjoint continuation of the restriction of the operator Ξ”π‘…π‘˜ to the space π’Ÿπ‘˜2(SL2(Z),𝒰, 𝜈). Com-bining Satz 5.4, Satz 5.5 and the considerations on page 335 in [Roe66] we find

spec(οΈ€Ξ”Λœπ‘…π‘˜)οΈ€

βŠ†{οΈ€

πœ†π‘…0, πœ†π‘…1, . . . , πœ†π‘…π‘š

π‘˜

}οΈ€βˆͺ[πœ‡π‘…π‘˜,∞), (3.7.14)

where the left-hand side denotes the spectrum of the operator Ξ”Λœπ‘…π‘˜, and πœ†π‘…β„“ and πœ‡π‘…π‘˜ are given by

πœ†π‘…β„“ := 1 4 βˆ’

(οΈ‚|π‘˜| βˆ’1 2 βˆ’β„“

)οΈ‚2

, πœ‡π‘…π‘˜ := 1 4 βˆ’

(οΈ‚1βˆ’π‘˜0

2 )οΈ‚2

(3.7.15)

for β„“ = 0,1, . . . , π‘šπ‘˜. Here π‘šπ‘˜ := ⌊|π‘˜|βˆ’12 βŒ‹, and π‘˜0 ∈ [0,2) such that π‘˜0 ≑ π‘˜ mod 2. In particular, the finite set {πœ†π‘…0, . . . , πœ†π‘…π‘š

π‘˜}in (3.7.14) can be omitted if|π‘˜| ≀2as for |π‘˜|<1 we have π‘šπ‘˜<0, and for1≀ |π‘˜| ≀2we have π‘šπ‘˜ = 0 and πœ†π‘…0 =πœ‡π‘…π‘˜.

We can reformulate the above result as follows: If 𝐺 ∈ π’Ÿπ‘˜2(SL2(Z),𝒰, 𝜈) is an eigen-function of Ξ”π‘…π‘˜ we can write its eigenvalue πœ†π‘… as πœ†π‘… = 1/4 +π‘Ÿ2 with

π‘Ÿ ∈[0,∞), π‘Ÿβˆˆ [οΈ‚

𝑖1βˆ’π‘˜0 2 ,0

)οΈ‚

or π‘Ÿ =𝑖

(οΈ‚|π‘˜| βˆ’1 2 βˆ’β„“

)οΈ‚

(3.7.16)

for some β„“= 0,1, . . . , π‘šπ‘˜.

Next we again translate these results to our vector valued setting. Recall that we write Ξ”Λœπ‘˜ for the self-adjoint continuation of the restriction of the Laplace operator Ξ”π‘˜ to the spaceπ’Ÿπ‘˜,𝐿. Using the relation (3.7.10) it is easy to see that𝐺: Hβ†’ 𝒰 is an eigenfunction of Roelckes hyperbolic Laplace operator Ξ”π‘…π‘˜ with eigenvalue πœ†π‘… if and only ifΞ βˆ’1π‘˜ (𝐺) is an eigenfunction of the operator Ξ”π‘˜ with eigenvalue πœ† = πœ†π‘… + π‘˜2βˆ’2π‘˜4 . Therefore, the spectrum of Ξ”Λœπ‘˜ is simply shifted by π‘˜2βˆ’2π‘˜4 , i.e., (3.7.14) translates to

spec(οΈ€Ξ”Λœπ‘˜)οΈ€

βŠ† {πœ†0, πœ†1, . . . , πœ†π‘šπ‘˜} βˆͺ[πœ‡π‘˜,∞).

(3.7.17) with

πœ†β„“ :=

(οΈ‚1βˆ’π‘˜ 2

)οΈ‚2

βˆ’

(οΈ‚|π‘˜| βˆ’1 2 βˆ’β„“

)οΈ‚2

, πœ‡π‘˜ :=

(οΈ‚1βˆ’π‘˜ 2

)οΈ‚2

βˆ’

(οΈ‚1βˆ’π‘˜0 2

)οΈ‚2

(3.7.18) ,

and where π‘šπ‘˜ and π‘˜0 are given as above. Moreover, given an eigenfunction 𝐹 ∈ π’Ÿπ‘˜,𝐿 of Ξ”π‘˜ we can write its eigenvalue πœ† as

πœ†=

(οΈ‚1βˆ’π‘˜ 2

)οΈ‚2

+π‘Ÿ2 (3.7.19)

with π‘Ÿ as in (3.7.16).

In chapter 2 of [Roe66] the author also studies Fourier expansions of eigenfunctions.

More precisely, it is shown that if 𝐺: H β†’ 𝒰 is modular of weight π‘˜ in the sense of Roelcke and an eigenfunction of Ξ”π‘…π‘˜ with eigenvalue πœ†π‘… = 1/4 +π‘Ÿ2 for some π‘ŸβˆˆC, then 𝐺 has a Fourier expansion of the form

𝐺(𝜏) =𝑒(𝑣) +

π‘š

βˆ‘οΈ

𝑗=1

βˆ‘οΈ

π‘›βˆˆZ+π‘ž(𝛾𝑗) 𝑛̸=0

(︀𝑏𝐺(𝛾𝑗, 𝑛)π‘Šβˆ’sign(𝑛)π‘˜/2,π‘–π‘Ÿ(βˆ’4πœ‹|𝑛|𝑣) (3.7.20)

+ 𝑐𝐺(𝛾𝑗, 𝑛)π‘Šsign(𝑛)π‘˜/2,π‘–π‘Ÿ(4πœ‹|𝑛|𝑣))οΈ€

𝑒(𝑛𝑒)^e𝑗 70

for𝜏 =𝑒+𝑖𝑣 ∈Hand 𝑏𝐺(𝛿𝑗, 𝑛), 𝑐𝐺(𝛿𝑗, 𝑛)∈C(see [Roe66], equations (2.13) and (2.18)).

Here the function𝑒(𝑣) is given by 𝑒(𝑣) :=𝑣1/2βˆ’π‘–π‘Ÿ βˆ‘οΈ

𝑗=1,...,π‘š π‘ž(𝛾𝑗)=0

𝑏𝐺(𝛾𝑗,0)^e𝑗 +𝑣1/2+π‘–π‘Ÿ βˆ‘οΈ

𝑗=1,...,π‘š π‘ž(𝛾𝑗)=0

𝑐𝐺(𝛾𝑗,0)^e𝑗, (3.7.21)

where we need to replace the term𝑣1/2+π‘–π‘Ÿby𝑣1/2ln(𝑣)ifπ‘Ÿ= 0(compare [Roe66], equation (2.17)). If we further assume that 𝐺(𝜏) = 𝑂(𝑣𝛼) as 𝑣 β†’ ∞ uniformly in 𝑒 for some 𝛼 ∈ R as in [Roe66], Definition 1.1 (b), then we find 𝑏𝐺(𝛾𝑗, 𝑛) = 0 for all 𝑗 = 1, . . . , π‘š and 𝑛 ∈Z+π‘ž(𝛾𝑗), 𝑛̸= 0 (compare [Roe66], Lemma 2.1) , i.e.,

𝐺(𝜏) = 𝑒(𝑣) +

π‘š

βˆ‘οΈ

𝑗=1

βˆ‘οΈ

π‘›βˆˆZ+π‘ž(𝛾𝑗) 𝑛̸=0

𝑐𝐺(𝛾𝑗, 𝑛)π‘Šsign(𝑛)π‘˜/2,π‘–π‘Ÿ(4πœ‹|𝑛|𝑣)𝑒(𝑛𝑒)^e𝑗, (3.7.22)

where 𝑒(𝑣)is still given as in (3.7.21).

We now further assume that 𝐺 ∈ π’Ÿπ‘˜2(SL2(Z),𝒰, 𝜈). Then πœ†π‘… = 1/4 +π‘Ÿ2 with π‘Ÿ as in (3.7.16), i.e., we either have π‘Ÿ β‰₯0 orπ‘Ÿ ∈ 𝑖R. Moreover, 𝐺 is square-integrable and thus needs to behave as 𝑣1/2βˆ’πœ€ as 𝑣 β†’ ∞with πœ€ >0. Therefore 𝐺 has a Fourier expansion as in (3.7.22) where if π‘Ÿ is purely imaginary either the coefficients 𝑏𝐺(𝛾𝑗,0) or 𝑐𝐺(𝛾𝑗,0) of 𝑒(𝑣) all need to vanish, depending on whether Im(π‘Ÿ) is positive or negative, and if π‘Ÿ is real 𝑒(𝑣) needs to vanish identically. In other words, we have

𝐺(𝜏) = 𝑣1/2βˆ’|Im(π‘Ÿ)| βˆ‘οΈ

𝑗=1,...,π‘š π‘ž(𝛾𝑗)=0

𝑐𝐺(𝛾𝑗,0)^e𝑗 +

π‘š

βˆ‘οΈ

𝑗=1

βˆ‘οΈ

π‘›βˆˆZ+π‘ž(𝛾𝑗) 𝑛̸=0

𝑐𝐺(𝛾𝑗, 𝑛)π‘Šsign(𝑛)π‘˜/2,π‘–π‘Ÿ(4πœ‹|𝑛|𝑣)𝑒(𝑛𝑒)^e𝑗, (3.7.23)

with 𝑐𝐺(𝛾𝑗,0) = 0for 𝑗 = 1, . . . , π‘šif π‘Ÿ is real.

All of these considerations translate directly to the setting of vector valued eigenfunc-tions for the Weil representation. We quickly summarise them in the following lemma:

Lemma 3.7.4. Let 𝐹: H β†’ C[𝐿′/𝐿] be modular of weight π‘˜ with respect to the Weil representation and an eigenfunction of the hyperbolic Laplace operatorΞ”π‘˜ with eigenvalue

πœ†=

(οΈ‚1βˆ’π‘˜ 2

)οΈ‚2

+π‘Ÿ2 with π‘ŸβˆˆC.

(a) Then 𝐹 has a Fourier expansion of the form 𝐹(𝜏) = 𝑒(𝑣) +

π‘š

βˆ‘οΈ

𝑗=1

βˆ‘οΈ

π‘›βˆˆZ+π‘ž(𝛾𝑗) 𝑛̸=0

(οΈ‚

𝑏𝐹(𝛾𝑗, 𝑛)(βˆ’4πœ‹|𝑛|𝑣)βˆ’π‘˜/2π‘Šβˆ’sign(𝑛)π‘˜/2,π‘–π‘Ÿ(βˆ’4πœ‹|𝑛|𝑣) +𝑐𝐹(𝛾𝑗, 𝑛)(4πœ‹|𝑛|𝑣)βˆ’π‘˜/2π‘Šsign(𝑛)π‘˜/2,π‘–π‘Ÿ(4πœ‹|𝑛|𝑣)

)οΈ‚

𝑒(𝑛𝑒)^e𝑗

with

𝑒(𝑣) = 𝑣1/2βˆ’π‘˜/2βˆ’π‘–π‘Ÿ βˆ‘οΈ

𝑗=1,...,π‘š π‘ž(𝛾𝑗)=0

𝑏𝐹(𝛾𝑗,0)^e𝑗 +𝑣1/2βˆ’π‘˜/2+π‘–π‘Ÿ βˆ‘οΈ

𝑗=1,...,π‘š π‘ž(𝛾𝑗)=0

𝑐𝐹(𝛾𝑗,0)^e𝑗.

(b) If 𝐹(𝜏) = 𝑂(𝑣𝛼) as 𝑣 β†’ ∞ uniformly in 𝑒 for some 𝛼 ∈ R then 𝐹 has a Fourier expansion of the form

𝐹(𝜏) =𝑒(𝑣) +

π‘š

βˆ‘οΈ

𝑗=1

βˆ‘οΈ

π‘›βˆˆZ+π‘ž(𝛾𝑗) 𝑛̸=0

𝑐𝐹(𝛾𝑗, 𝑛)π’²π‘˜,1/2+π‘–π‘Ÿ(4πœ‹π‘›π‘£)𝑒(𝑛𝑒)^e𝑗,

where 𝑒(𝑣) is given as in part (a).

(c) If 𝐹 ∈ π’Ÿπ‘˜,𝐿 then πœ†= (1βˆ’π‘˜2 )2+π‘Ÿ2 with π‘Ÿ ∈[0,∞), π‘Ÿβˆˆ

[οΈ‚

𝑖1βˆ’π‘˜0 2 ,0

)οΈ‚

or π‘Ÿ=𝑖

(οΈ‚|π‘˜| βˆ’1 2 βˆ’β„“

)οΈ‚

,

where π‘˜0 ∈[0,2)with π‘˜0 β‰‘π‘˜ mod 2 andβ„“ ∈ {0,1, . . . , π‘šπ‘˜} with π‘šπ‘˜ =⌊|π‘˜|βˆ’12 βŒ‹. In this case 𝐹 has a Fourier expansion of the form

𝐹(𝜏) = 𝑣1/2βˆ’π‘˜/2βˆ’|Im(π‘Ÿ)| βˆ‘οΈ

𝑗=1,...,π‘š π‘ž(𝛾𝑗)=0

𝑐𝐹(𝛾𝑗,0)^e𝑗+

π‘š

βˆ‘οΈ

𝑗=1

βˆ‘οΈ

π‘›βˆˆZ+π‘ž(𝛾𝑗) 𝑛̸=0

𝑐𝐹(𝛾𝑗, 𝑛)π’²π‘˜,1/2+π‘–π‘Ÿ(4πœ‹π‘›π‘£)𝑒(𝑛𝑒)^e𝑗,

where 𝑐𝐹(𝛾𝑗,0) = 0 for 𝑗 = 1, . . . , π‘š if π‘Ÿ is real.

72

4 Borcherds’ generalized Shimura lift

In the following we give a quick introduction to the theory of regularized theta lifts in the sense of Borcherds, focussing on a generalized version of the classical Shimura lift from half-integral weight modular forms to forms of even weight. In the second half of this chapter, we specialize to certain lattices of signature (2,1) and (2,2).

Our main reference for this chapter is the original work of Borcherds, namely [Bor98].

Furthermore, we again use the notation from [Bru02].

4.1 Modular forms on orthogonal groups

Let (𝑉, π‘ž) be a quadratic space of signature (2, 𝑛) with 𝑛 β‰₯ 1. The Grassmannian of 𝑉 is given by the set of 2-dimensional positive definite subspaces of 𝑉(R) :=𝑉 βŠ—R, i.e., by

Gr(𝑉) :={𝑣 βŠ†π‘‰(R) : dim(𝑣) = 2 and π‘ž|𝑣 >0}.

For 𝑣 ∈ Gr(𝑉) we write 𝑣βŠ₯ for the orthogonal complement of 𝑣 in 𝑉(R) such that 𝑉(R) = π‘£βŠ•π‘£βŠ₯. Clearly, 𝑣βŠ₯ is a negative definite subspace of dimension 𝑛. Further, we denote the orthogonal projection of some vector 𝑋 βˆˆπ‘‰(R) onto 𝑣 or 𝑣βŠ₯ by 𝑋𝑣 or 𝑋𝑣βŠ₯, respectively. In particular, we have π‘ž(𝑋) = π‘ž(𝑋𝑣) +π‘ž(𝑋𝑣βŠ₯).

Next we equip the GrassmannianGr(𝑉) with a complex structure: Firstly, we extend the bilinear form on𝑉 to aC-bilinear form on the complexification 𝑉(C) = 𝑉 βŠ—Cof 𝑉, which we again denote by (Β·,Β·), and we let P(𝑉(C)) be the projective space over 𝑉(C).

We write 𝑍 ↦→ [𝑍] for the canonical projection of 𝑉(C)βˆ– {0} onto P(𝑉(C)). Now, the complex manifold

𝒦:={οΈ€

[𝑍]∈P(𝑉(C)) : (𝑍, 𝑍) = 0 and (𝑍, 𝑍)>0}οΈ€

(4.1.1)

has two connected components which we denote by 𝒦±, and it is easy to check that the two maps

𝒦± β†’Gr(𝑉), [𝑋+π‘–π‘Œ]↦→Rπ‘‹βŠ•Rπ‘Œ

are both bijections, inducing a complex structure on the Grassmannian Gr(𝑉).

Next we realize𝒦+as a so-called tube domain: Let𝐿be an even lattice in𝑉 and choose vectors 𝑒1 ∈ 𝐿 and 𝑒2 βˆˆπΏβ€² such that 𝑒1 is primitive with π‘ž(𝑒1) = 0 and (𝑒1, 𝑒2) = 1. We then define the sublattice

𝐾 :=πΏβˆ©π‘’βŠ₯1 βˆ©π‘’βŠ₯2

and let π‘ˆ := 𝐾 βŠ—Q. Then 𝐾 is an even lattice lying in the quadratic space (π‘ˆ, π‘ž) of signature (1, π‘›βˆ’1), and we have 𝑉 =π‘ˆ βŠ•Q𝑒1 βŠ•Q𝑒2. Further, we define the complex plane

β„‹:={𝑋+π‘–π‘Œ βˆˆπ‘ˆ(C) : π‘ž(π‘Œ)>0}

of vectors in π‘ˆ(C)with positive imaginary part, and for 𝑍 =𝑋+π‘–π‘Œ βˆˆπ‘ˆ(C)we set 𝑍𝐿:=𝑋𝐿+π‘–π‘ŒπΏ=π‘βˆ’(π‘ž(𝑍) +π‘ž(𝑒2))𝑒1+𝑒2.

Here𝑋𝐿=π‘‹βˆ’(π‘ž(𝑋)βˆ’π‘ž(π‘Œ) +π‘ž(𝑒2))𝑒1+𝑒2 and π‘ŒπΏ =π‘Œ βˆ’(𝑋, π‘Œ)𝑒1. Then(𝑍𝐿, 𝑍𝐿) = 0 and (𝑍𝐿, 𝑍𝐿) = 4π‘ž(π‘Œ)>0for 𝑍 =𝑋+π‘–π‘Œ ∈ β„‹, and thus the mapping

β„‹ β†’ 𝒦, 𝑍 ↦→[𝑍𝐿]

is well-defined. Moreover, one can check that this map is indeed biholomorphic, giving us an identification between the complex plane β„‹ and the complex manifold 𝒦. We let H𝑛 be the component of β„‹ which is mapped to 𝒦+, and we call H𝑛 a generalized upper half-plane. Note that the concrete realization of H𝑛 does depend on the choice of vectors 𝑒1 and 𝑒2.

We further introduce some notation. For𝑍 ∈H𝑛 we write 𝑍 = Re(𝑍) +𝑖Im(𝑍)

with Re(𝑍),Im(𝑍) ∈ π‘ˆ(R) and where π‘ž(Im(𝑍)) >0. Moreover, for 𝑍 ∈ H𝑛 with 𝑍𝐿 = 𝑋𝐿+π‘–π‘ŒπΏ and πœ†βˆˆπ‘‰ we denote the orthogonal projection of πœ† onto the positive definite subspaceRπ‘‹πΏβŠ•Rπ‘ŒπΏbyπœ†π‘. Correspondingly, we writeπœ†π‘βŠ₯ for the orthogonal projection ofπœ†onto the negative definite subspace(Rπ‘‹πΏβŠ•Rπ‘ŒπΏ)βŠ₯. Then an easy computation shows that

π‘ž(πœ†π‘) = |(πœ†, 𝑍𝐿)|2 4π‘ž(Im(𝑍)) (4.1.2)

for πœ†βˆˆπ‘‰ and 𝑍 ∈H𝑛. Finally, we introduce the notation

π‘žπ‘(πœ†) :=π‘ž(πœ†π‘)βˆ’π‘ž(πœ†π‘βŠ₯) = 2π‘ž(πœ†π‘)βˆ’π‘ž(πœ†) (4.1.3)

for πœ† βˆˆπ‘‰ and 𝑍 ∈H𝑛. Then π‘žπ‘ is a majorant of the quadratic formπ‘ž associated to 𝑍.

In particular, the quadratic form π‘žπ‘ is positive definite.

In the following, we introduce an automorphy factor for the action of an orthogonal group on the generalized upper half-plane defined above. Recall that 𝑂(𝑉) is the or-thogonal group of the quadratic space(𝑉, π‘ž). We write SO(𝑉)for the special orthogonal group of π‘ž, i.e., for the subgroup of 𝑂(𝑉) of elements of determinant1. It is well-known that for 𝑛 β‰₯ 1 the group SO(𝑉) is not connected. As usual we denote the connected component of the identity in SO(𝑉)by SO+(𝑉).

The action of𝑂(𝑉)on𝑉(R)naturally induces actions on the GrassmannianGr(𝑉)and on the complex manifold 𝒦which are by construction compatible with the identification [𝑋+π‘–π‘Œ]↦→Rπ‘‹βŠ•Rπ‘Œ, and the subgroupSO+(𝑉)preserves the two connected components 𝒦± of 𝒦, whereas SO(𝑉)βˆ–SO+(𝑉) interchanges them. Since the mapping 𝑍 ↦→ [𝑍𝐿] defines a bijection between the generalized upper half-planeH𝑛and the manifold𝒦+, the action of the group SO+(𝑉)on𝒦+ further induces an action ofSO+(𝑉)onH𝑛 which we denote by𝜎.𝑍for𝜎∈SO+(𝑉)and𝑍 ∈H𝑛in order to avoid confusion with the operation of SO+(𝑉) on𝑉(C) which we simply denote by 𝜎(𝑍). By construction we then have

[(𝜎.𝑍)𝐿] = [𝜎(𝑍𝐿)]

74

for 𝜎 ∈ SO+(𝑉) and 𝑍 ∈ H𝑛, i.e., there is πœ‡βˆˆ CΓ— such that πœ‡Β·(𝜎.𝑍)𝐿 = 𝜎(𝑍𝐿) where we understand (𝜎.𝑍)𝐿 and 𝜎(𝑍𝐿) as elements of the cone over 𝒦+ given by

𝐢(𝒦+) := {οΈ€

π‘Š βˆˆπ‘‰(C)βˆ– {0}: [π‘Š]∈ 𝒦+}οΈ€

.

Using that H𝑛 βŠ† π‘ˆ(C) it is easy to check that the factor πœ‡ from above is given by the scalar product (𝜎(𝑍𝐿), 𝑒1), i.e., we have

(𝜎(𝑍𝐿), 𝑒1)Β·(𝜎.𝑍)𝐿=𝜎(𝑍𝐿)

for all 𝜎 ∈SO+(𝑉) and 𝑍 ∈H𝑛. We give this factor a name, i.e., we define the complex valued function

𝑗(𝜎, 𝑍) := (𝜎(𝑍𝐿), 𝑒1)

for 𝜎 ∈ SO+(𝑉) and 𝑍 ∈H𝑛. By construction 𝑗(𝜎, 𝑍) is non-vanishing and satisfies the so-called cocycle relation

𝑗(𝜎1𝜎2, 𝑍) = 𝑗(𝜎1, 𝜎2.𝑍)𝑗(𝜎2, 𝑍)

for 𝜎1, 𝜎2 ∈SO+(𝑉)and 𝑍 ∈H𝑛. Thus we can call the function 𝑗 an automorphy factor for the group SO+(𝑉). We will later see that it indeed generalizes the usual automorphy factor 𝑗((οΈ€π‘Ž 𝑏

𝑐 𝑑

)οΈ€, 𝜏) = π‘πœ +𝑑 defined onSL2(R)Γ—H at the beginning of Section 2.4.

Lemma 4.1.1. Let 𝜎 ∈SO+(𝑉). Then π‘ž(Im(𝜎.𝑍)) = π‘ž(Im(𝑍))

|𝑗(𝜎, 𝑍)|2 and π‘ž(𝜎(πœ†)𝑍) = π‘ž(πœ†πœŽβˆ’1.𝑍).

for 𝑍 ∈H𝑛 and πœ†βˆˆπ‘‰. Proof. We find that

π‘ž(Im(𝜎.𝑍)) = ((𝜎.𝑍)𝐿,(𝜎.𝑍)𝐿)

4 = (𝜎(𝑍𝐿), 𝜎(𝑍𝐿))

4|𝑗(𝜎, 𝑍)|2 = (𝑍𝐿, 𝑍𝐿)

4|𝑗(𝜎, 𝑍)|2 = π‘ž(Im(𝑍))

|𝑗(𝜎, 𝑍)|2, which proves the first equality. Further, we obtain

π‘ž(𝜎(πœ†)𝑍) = |(𝜎(πœ†), 𝑍𝐿)|2

4π‘ž(Im(𝑍)) = |(πœ†, πœŽβˆ’1(𝑍𝐿))|2

4|𝑗(πœŽβˆ’1, 𝑍)|2π‘ž(Im(πœŽβˆ’1.𝑍)) =π‘ž(πœ†πœŽβˆ’1.𝑍) using equation (4.1.2).

Finally, we recall that𝑂(𝐿) ={𝜎 βˆˆπ‘‚(𝑉) : 𝜎(𝐿) =𝐿}, and we let 𝑂𝑑(𝐿)be the finite index subgroup of 𝑂(𝐿) consisting of all elements that act trivially on the discriminant group 𝐿′/𝐿. We then define the group

Ξ“(𝐿) := SO+(𝑉)βˆ©π‘‚π‘‘(𝐿).

(4.1.4)

Moreover, we set SO+(𝐿) := SO+(𝑉)βˆ©π‘‚(𝐿). One can show that the two groups Ξ“(𝐿) and SO+(𝐿) are discrete subgroups ofSO+(𝑉).

Definition 4.1.2. Let π‘˜ ∈Z and let Γ≀Γ(𝐿) be a finite index subgroup. We say that a function𝐹: H𝑛→C is modular of weight π‘˜ with respect to Ξ“ if

𝐹(𝛾.𝑍) =𝑗(𝛾, 𝑍)π‘˜πΉ(𝑍) for all 𝛾 βˆˆΞ“ and 𝑍 ∈H𝑛.

Lemma 4.1.3. Let 𝐺: 𝐢(𝒦+) β†’ C be homogeneous of degree βˆ’π‘˜ ∈ Z and let Ξ“ be a finite index subgroup ofΞ“(𝐿). Then𝐺is invariant under the action ofΞ“, i.e.,𝐺(𝛾(π‘Š)) = 𝐺(π‘Š)for all𝛾 βˆˆΞ“andπ‘Š ∈𝐢(𝒦+), if and only if the corresponding function𝐹: H𝑛→C defined by 𝐹(𝑍) =𝐺(𝑍𝐿) is modular of weight π‘˜ with respect to Ξ“.

Proof. Suppose that 𝐺 is invariant under the action ofΞ“. Then 𝐹(𝛾.𝑍) = 𝐺((𝛾.𝑍)𝐿) =𝐺(οΈ€

𝑗(𝛾, 𝑍)βˆ’1𝛾(𝑍𝐿))οΈ€

=𝑗(𝛾, 𝑍)π‘˜πΊ(𝛾(𝑍𝐿)) = 𝑗(𝛾, 𝑍)π‘˜πΉ(𝑍) for 𝛾 ∈ Ξ“ and 𝑍 ∈ H𝑛. Conversely, given π‘Š ∈ 𝐢(𝒦+) there is 𝑍 ∈ H𝑛 such that [𝑍𝐿] = [π‘Š]in 𝒦+, i.e., we find πœ‡βˆˆCΓ— with π‘Š =πœ‡Β·π‘πΏ. Hence

𝐺(𝛾(π‘Š)) =πœ‡βˆ’π‘˜πΊ(𝛾(𝑍𝐿)) = πœ‡βˆ’π‘˜π‘—(𝛾, 𝑍)βˆ’π‘˜πΊ((𝛾.𝑍)𝐿) =πœ‡βˆ’π‘˜π‘—(𝛾, 𝑍)βˆ’π‘˜πΉ(𝛾.𝑍)

for𝛾 βˆˆΞ“. Assuming that𝐹 is modular of weight π‘˜ with respect to Ξ“we therefore obtain 𝐺(𝛾(π‘Š)) = πœ‡βˆ’π‘˜πΉ(𝑍) =πœ‡βˆ’π‘˜πΊ(𝑍𝐿) = 𝐺(πœ‡Β·π‘πΏ) = 𝐺(π‘Š)

for all Ξ“βˆˆΞ“. This proves the claimed equivalence.