<<

arXiv:1403.3567v2 [math.NT] 10 Aug 2020 oee,teoeaosdfie nta eeec aeautomor take reference that orthog in on defined forms operators automorphic the on However, operators differential considers oriae ewe h uedmi oe n h rniiefre cha transitive the the of and of because subgroup model tedious, parabolic domain very appropriate is tube an way the this operato between in such coordinates formula of explicit existence the their establish ing to order in considerations weight hmr–aßoeaosotie ntecase the in obtained operators Shimura–Maaß iee nti ae,a ela S4 o oegnrlztoso th example). of for invaria generalizations [Sh5], of some in investigation for appearing the [Sh4] operators or to as references related mentioned well strongly previously as is the paper, which this [Sh1], o in reference subject sidered the the e.g., representat therefore (see, prop as are papers related well and have as interpretations, which number-theoretic Lie-algebraic) diffe groups have concerns unitary [Sh3] operators on paper these following for functions The modular on Siegel [Sh2]. operators on in effect generalized similar was a which have which operators ferential om naohrtp fSiuavreis aeyqoinso Gr (2 of signature quotients of namely spaces varieties, vector Shimura of of ans type another on forms iglmdlrfrsapaigi h case the in appearing forms modular Siegel h qaeo h hmr–aßoeaosfo h case the from operators Shimura–Maaß the of square the aig(litc elaayi)mdlrfrso weight of forms modular real-analytic) (elliptic, taking h lsia hmr–aßoperators Shimura–Maaß classical The Introduction emn.Teiiilpr fti okwsspotdb h Mi the by supported was work this Gesellschaft). of part initial The Germany. ∗ u rtga st en iia prtr o oua o automor (or modular for operators similar define to is goal first Our egtCagn prtr o Automorphic for Operators Changing Weight hswr a are u hl a okn tteTechnical the at working was I while out carried was work This om nGasanasadDifferential and Grassmannians on Forms k and 2 + rpriso eti ht Lifts Theta Certain of Properties k − epciey nadto,[a]ad[a]cnie dif- consider [Ma2] and [Ma1] addition, In respectively. 2 uut1,2020 12, August hu Zemel Shaul b , SO − 1 .Teeaeotie yinterpolating by obtained are These ). ∂ ∂ + τ ( + V b − .W lormr ht[h]also [Sh1] that remark also We ). 2 ∗ k iy .OemyueLie-theoretic use may One 3. = b − and ,adteoeaosfor operators the and 2, = ev elwhp(Max-Planck- Fellowship nerva y 2 k ∂ ∂ nvriyo Darmstadt, of University τ b omdlrfrsof forms modular to − r elkonfor well-known are ,temultiple the 1, = o-hoei (or ion-theoretic ayresearch many f s u obtain- but rs, hcfrsof forms phic tdifferential nt h aecon- case the nlgroups. onal eut of results e s work a ms, cinof action e ris All erties. assmanni- rential g of nge phic) some weight (i.e., a representation of the maximal compact subgroup) ρ to au- tomorphic forms having weight ρ η for some b−-dimensional representation η, hence in particular take scalar-valued⊗ automorphic forms on Grassmannians to vector-valued functions. Moreover, since that reference works with the coordi- nates arising from the bounded model while we consider the tube domain model (since the explicit formulae for the theta functions are more neatly presented in this model), an appropriate change of coordinates must be employed. It is true that after this change of coordinates, using the natural bilinear form on the tangent space of the Grassmannian in the tube domain model we may indeed obtain differential operators which remain in the scalar-valued realm. Indeed, after some additional normalization we obtain the operators defined in this pa- per using this method. However, the calculations involved are very delicate, laborious, long, an unenlightening, for which reason we have chosen to state and prove the formulae for the operators directly. The second goal of this paper is to present two applications of these weight changing operators, in the theory of theta lifts. We recall the generalization, defined in [B], for the Doi–Naganuma lifting first introduced in [DN] and [Ng]. This map is given in [B] in terms of a singular theta lift, and takes weakly holomorphic elliptic modular forms to meromorphic modular forms on Grass- mannians. On the other hand, [Ze2] defines a similar theta lift, using the same theta functions with polynomials. The first result of this paper states that in the case of an even dimension, a power of our weight raising operator sends the theta lift from [Ze2] to the generalized Doi–Naganuma lifting of [B]. In addition, recall that the theta lift from Section 13 of [B] (which is also studied extensively in [Bru] and others) is a real function. No automorphic forms of non-zero weight can be real. As a second application for our operators we define a notion of m-real automorphic forms of positive weight m, and show that in case one applies the theta lift from Section 14 of [B] (or from [Ze2]) to a with real Fourier coefficients, then resulting theta lift is m-real. The first half of the paper contains numerous statements whose proofs are delayed to later sections. We choose this way of presentation since most of the proofs consist of direct calculations, which may divert the reader’s attention from the main ideas. Specifically, the paper is divided into 4 sections. In Section 1 we define the weight raising and weight lowering operators and state their properties. Section 2 presents the images of certain functions under the weight raising operators, and proves the main theorem. Section 3 presents the proofs for the assertions of Section 1, while Section 4 contains the missing proofs of Section 2. I would like to thank J. Bruinier for numerous suggestions and intriguing discussions regarding the results of this paper.

2 1 Weight Changing Operators for Automorphic Forms on Orthogonal Groups

In this Section we present automorphic forms on complex manifolds arising as orthogonal Shimura varieties of signature (2,b−), introduce the weight raising and weight lowering operators on such forms, and give some of their properties. The proofs of most assertions are postponed to Section 3. Let V be a real with a non-degenerate bilinear form of signature 2 (b+,b−). The pairing of x and y in V is written (x, y), and x stands for the norm (x, x) of x. For S V , S⊥ denotes the subspace of V which is perpendicular to S. The Grassmannian⊆ G(V ) of V is defined to be the set of all decompositions of V into the orthogonal direct sum of a positive definite space v+ and a negative definite space v−. In the case b+ = 2 (which is the only case we consider in this paper), it is shown in Section 13 of [B], Sections 3.2 and 3.3 of [Bru], or Subsection 2.2 of [Ze2] (among others), that G(V ) carries a complex structure and has several equivalent models, which we now briefly present. Let

2 P = Z = X + iY VC = V R C Z =0, (Z , Z ) > 0 . V V V ∈ ⊗ V V V  ZV VC lies in P if and only if XV and YV are orthogonal and have the same positive∈ norm. P has two connected components (which are interchanged by complex conjugation), and let P + be one component. The map P + G(V ),Z RX RY → V 7→ V ⊕ V is surjective, and C∗ acts freely and transitively on each fiber of this map by multiplication. This realizes G(V ) as the image of P + in the projective space 2 P(LC), which is an analytically open subset of the (algebraic) quadric ZV = 0, yielding a complex structure on G(V ). This is the projective model of G(V ). Let z be a non-zero vector in V which is isotropic, i.e., z2 = 0. The vector ⊥ space KR = z /Rz is non-degenerate and Lorentzian of signature (1,b− 1). ⊥ − Choosing some ζ V with (z, ζ) = 1 and restricting the projection z KR ⊥ ∈ → to z, ζ gives an isomorphism. We thus write V as KR R R, in which { } × × ⊥ 2 2 2 2 (α,a,b)= aζ + bz + α z, ζ = KR , (α,a,b) = α +2ab + a ζ . ∈{ } ∼  A (holomorphic) section s : G(V ) P + is defined by the pairing with z being 1. → Subtracting ζ from any s-image and taking the KC-image of the result yields a biholomorphism between G(V ) ∼= s G(V ) and the tube domain KR +iC, where C is a cone of positive norm vectors in the Lorentzian space KR. C is called the positive cone, and it is determined by the choice of z and the connected + component P . The inverse biholomorphism takes Z = X + iY KC to ∈ Z2 ζ2 Y 2 X2 ζ2 Z = Z, 1, − − = X, 1, − − + i Y, 0, (X, Y ) , V,Z  2   2  −  with the real and imaginary parts denoted XV,Z and YV,Z respectively. They are orthogonal and have norm Y 2 > 0. This identifies G(V ) with the tube domain

3 model KR + iC. Taking the other connected component of P corresponds to taking the other cone C to be the positive cone, and to the conjugate complex structure. − The subgroup O+(V ) consisting of elements of O(V ) preserving the ori- entation on the positive definite part acts on P + and G(V ), respecting the projection. Elements of O(V ) O+(V ) interchange the connected components of P . The action of O+(V ) (and\ also of the connected component SO+(V )) on G(V ) is transitive, with the stabilizer K (or SK SO+(V )) of a point being isomorphic to SO(2) O(n) (resp. SO(2) SO(n≤)). Therefore G(V ) is isom- rphic to O+(V )/K and× to SO+(V )/SK.× Given an isotropic z as above, the + + action of O (V ) transfers to KR + iC, and for M O (V ) and Z KR + iC we have ∈ ∈

MZ = J(M,Z)Z , with J(M,Z) = (MZ ,z) C∗. V,Z V,MZ V,Z ∈ J is a factor of automorphy, namely the equality

J(MN,Z)= J(M,NZ)J(N,Z)

+ holds for all Z KR + iC and M and N in O (V ). For such M we define the slash operator of∈ weight m, and more generally of weight (m,n), by

−m −n Φ[M]m,n(Z)= J(M,Z) J(M,Z) Φ(MZ), [M]m = [M]m,0.

2 The fact that (ZV , ZV )=2Y and the definition of J(M,Z) yield the equalities

2 2 Y (MZ) = and F (Y 2)t [M] = F [M] (Y 2)t (1) ℑ J(M,Z) 2 m,n m+t,n+t  | |  the latter holding for every m, n, t, and function F on KR + iC (see Lemma 3.20 of [Bru] for the first equality in Equation (1), and the second one follows immediately). dXdY The invariant measure on KR + iC is (see Section 4.1 of [Bru], but (Y 2)b− one can also prove this directly, using the generators of O+(V ) considered in Section 3 below). Note that this measure depends on the choice of a basis for KR +iC, but changing the basis only multiplies this measure by a positive global scalar. Let Γ be a discrete subgroup Γ of O+(V ) of cofinite volume. In most of the interesting cases Γ will be either the O+ or the SO+ part of the orthogonal of an even lattice L in V , or the discriminant kernel of such a group. Given m Z, an automorphic form of weight m with respect to Γ is defined to ∈ be a (complex valued) function Φ on KR + iC for which the equation

m Φ(MZ)= J(M,Z) Φ(Z), or equivalently Φ[M]m(Z)=Φ(Z), holds for all M Γ and Z KR + iC. Using the standard argument, such a function is equivalent∈ to a∈ function on P + which is m-homogenous (with respect to the action of C∗) and Γ-invariant, as considered,− for example, in [B].

4 We now consider some differential operators on functions on KR +iC. Given ∂ a basis for KR, we write ∂x for (for 1 k b−). Similarly, ∂y stands for k ∂xk k the coordinates of the imaginary part from≤C. The≤ notation for the derivatives ∂ = 1 (∂ i∂ ) and ∂ = 1 (∂ + i∂ ) will be further shortened to ∂ and zk 2 xk − yk zk 2 xk yk k ∂k respectively. The operator I = k xk∂xk multiplies a homogenous function on KR by its homogeneity degree, andP is thus independent of the choice of basis (indeed, it has an intrinsic Lie-theoretic description). The operators

∗ ∗ D = yk∂k and D = yk∂k Xk Xk from [Na] are intrinsic as well, and they are also invariant under translations in the real part of KR +iC. If the basis for KR is orthonormal, i.e., orthogonal with the first vector having norm 1 and the rest having norm 1, then the Laplacian 2 b− 2 − of KR, denoted ∆KR , is defined to be ∂ ∂ . It is independent of the x1 − k=2 xk choice of the orthonormal basis (though usingP a basis which is not orthonormal it takes different forms), and it is invariant under the action of O(KR) as well as under translations in KR. With complex coordinates it has three counterparts,

b− b− b− ∆h = ∂2 ∂2, ∆h = ∂2 ∂2, and ∆R = ∂ ∂ ∂ ∂ , KC 1 − k KC 1 − k KC 1 1 − k k kX=2 Xk=2 kX=2 which we call the holomorphic Laplacian of KC (of Hodge weight (2, 0)), the anti-holomorphic Laplacian of KC (of Hodge weight (0, 2)), and the real Lapla- cian of KC (of Hodge weight (1, 1)), respectively. These operators have the same invariance and independence properties as ∆KR . Note that the appropri- ate combinations appearing in [Bru] and [Na] can be identified as our operators 1 h 1 h R 2 ∆KC , 2 ∆KC , and ∆KC respectively, expressed in a basis which is not orthonor- mal. We shall indeed discuss and generalize the operators ∆1 and ∆2 of [Na] in Proposition 1.5 below. The weight changing operators and their defining property are given in

(b−) Theorem 1.1. For any integer m define Rm to be the operator

b− b− 2 1 2 +1 i(2m +2 b−) m(2m +2 b−) (Y ) 2 −m− ∆h (Y )m − 2 = ∆h − D∗ − . KC KC − Y 2 − 2Y 2 In addition, define

b− b− ( −) 2 2 2 +1 2 1 2 2 2 L b = (Y ) R = (Y ) 2 ∆h (Y ) − 2 = (Y ) ∆h + iY (2 b )D∗. 0 KC KC − − Then the equalities

(b−) (b−) (b−) (b−) (Rm F )[M]m+2 = Rm F [M]m), (L F )[M]m−2 = L F [M]m) 2 + hold for every function F on KR + iC and any M O (V ). C ∈

5 (b−) (b−) The different descriptions of Rm and L coincide by Lemma 3.1 below. Theorem 1.1 has the following standard

Corollary 1.2. If Φ is an automorphic form of weight m on G(V ) ∼= KR + iC (b−) (b−) then Rm Φ and L Φ are automorphic forms on KR + iC which have weights m +2 and m 2 respectively. − (b−) In correspondence with Theorem 1.1 and Corollary 1.2 we call Rm and L(b−) the weight raising operator of weight m and the weight lowering operator for automorphic forms on Grassmannians of signature (2,b−) respectively. As already mentioned in the Introduction, these operators may also be given a Lie-theoretic description (see Section 3 for more details). However, the explicit operators appearing in Theorem 1.1 are more useful for our applications. We shall make use of the operator

D∗ D∗ D∗D∗ = D∗D∗ + = y y ∂ ∂ , − 2i 2i k l k l Xk,l which we denote D∗ 2. Lemma 2.2 of [Ze2] shows that | | ∆(b−) =8 D∗ 2 4Y 2∆R 4imD∗ +4inD∗ +2n(2m b ) m,n | | − KC − − −

(b−) is the weight (m,n) Laplacian on KR +iC, and the weight m Laplacian ∆m is (b−) just ∆m,0 (this extends the corresponding assertion of [Na], since his operator (b−) ∆1 is our ∆0 divided by 8). The constants are normalized such that

(b−) 2 t 2 t (b−) ∆m,n (Y ) = (Y ) ∆m+t,n+t (2) holds for every m, n, and t (see the remark after Lemma 3.1 below). The (b−) (b−) relations between Rm , L , and the corresponding Laplacians are given by Proposition 1.3. The equalities

∆(b−) R(b−) R(b−)∆(b−) = (2b 4m 4)R(b−) m+2 m − m m − − − m and ∆(b−) L(b−) L(b−)∆(b−) = (4m 2b 4)L(b−) m−2 − m − − − hold for every m Z. ∈ We recall that an automorphic form of weight m on KR + iC is said to have (b−) (b−) eigenvalue λ if it is annihilated by ∆m + λ (i.e., eigenvalues are of ∆m ). Hence Proposition 1.3 has the following −

Corollary 1.4. If F is an automorphic form of weight m on KR +iC which has (b−) (b−) eigenvalue λ then the automorphic forms Rm F and L F have eigenvalues λ +4m 2b +4 and λ 4m +2b +4 respectively. − − − −

6 By evaluating compositions of the weight changing operators one shows Proposition 1.5. The combination

Ξ(b−) = (Y 2)2∆h ∆h iY 2(2m +2 b )D∗∆h + iY 2(2 b )D∗∆h + m KC KC − − − KC − − KC (2 b )(2m +2 b ) m(2m +2 b ) + − − − − Y 2∆R − − Y 2∆h 2 KC − 2 KC commutes with all the weight m slash operators as well as with the Laplacian (b−) ∆m . The commutator of the global weight raising operator and the weight lowering operator is

(b−) − m∆m mb (2m 2 b ) R(b−),L(b−) = R(b ) L(b−) L(b−)R(b−) = − − − − . m m−2 − m 2 − 4  

(b−) Proposition 1.5 provides another proof to Lemma 2.2 of [Ze2] about ∆m . It (b−) also implies that Ξm preserves the spaces of automorphic forms of weight m for all m Z and for every discrete subgroup Γ of cofinite volume in O+(V ). It also ∈ (b−) commutes with ∆m , hence preserves eigenvalues of such automorphic forms. By rank considerations, one can probably show that the ring of differential operators which commute with all the slash operators of weight m is generated (b−) (b−) by ∆m and Ξm , hence is a polynomial ring in two variables (if b− > 1). This assertion should also follow from part (3) of Theorem 3.3 of [Sh5] (since the rank of the symmetric space G(V ) is 2 if b− > 1), though I have not verified (b−) (b−) this in detail. As ∆0 is 8∆1 and Ξ0 is 16∆2 in the notation of [Na], Proposition 1.5 generalizes the main result of that reference to other weights. A similar argument yields results of the same sort for (m,n), where a possible (b−) 2 −n (b−) 2 n normalization for Ξm,n is (Y ) Ξm−n(Y ) for which an equality similar to Equation (2) holds. We shall not need these results in what follows. We now consider compositions of the weight raising operators. The natural (b−) lth power of Rm is the composition

(R(b−))l = R(b−) ... R(b−). m m+2l−2 ◦ ◦ m The general formulae for the resulting operator seems too complicated to write h ∗ 1 as a combination of ∆KC , D , and Y 2 with explicit coefficients. However, we can establish the properties given in the following

(b−) l Proposition 1.6. (i) The operator (Rm ) takes automorphic forms of weight m on G(LR) to automorphic forms of weight m +2l. (ii) In case the former automorphic form is an eigenfunction with eigenvalue λ, the latter is also an eigenfunction, and the corresponding eigenvalue is λ + l(4m +4l 2b−). (iii) (b−) l − The operator (Rm ) can be written as

l c (iD∗)c−a(∆h )l−c (R(b−))l = A(l) KC , m a,c ( Y 2)c Xc=0 Xa=0 −

7 (0) (l) (l+1) where A0,0 =1 and given the coefficients Aa,c for given l, the coefficient Aa,c of the next power l +1 is defined recursively as

a c s m +2l c +1 − A(l) + (2m +4l 2c +4 b ) A(l) + − A(l) . a s s,c − − −  a,c−1 2 a−1,c−1 Xs=0 − (l) (iv) For a =0 the coefficients A0,c are given by the explicit formula

l! 2c m + l b− A(l) = · − 2 . 0,c (l c)! c  −

The binomial symbol appearing in part (iv) of Proposition 1.6 is the extended binomial coefficient: Indeed, for two non-negative integers x and n we have

x 1 n−1 = (x j), n n! − jY=0 a formula which makes sense for x R (as well as x in any Q-algebra). Part (i) of Proposition 1.6 follows∈ immediately from Corollary 1.2. For part (ii) Corollary 1.4 shows that the application of R (for 0 r l 1) to an m+2r ≤ ≤ − eigenfunction adds 4m +8r +4 2b− to the eigenvalue, so the assertion follows from evaluating −

l−1 (4m +8r +4 2b )= l(4m +4l 2b ). − − − − Xr=0 The proofs of parts (iii) and (iv) are given in Section 3. We recall that M = a b SL (R) defines the holomorphic map c d ∈ 2  aτ + b M : τ = τ = x + iy C y > 0 , with j(M, τ)= cτ + d, ∈ H ∈ 7→ cτ + d h  i the latter being the factor of automorphy of this action. Modular forms of weight (k,l) (or just weight k if l = 0) with respect to a discrete subgroup Γ of dxdy SL2(R) with cofinite volume (with respect to the invariant measure y2 ) are functions f : C which are invariant under the corresponding weight (k,l) slash operatorsH for → elements of Γ. The weight (k,l) Laplacian is

∆ =4y2∂ ∂ 2iky∂ +2ily∂ + l(k 1), k,l τ τ − τ τ − normalized such that ∆k = ∆k,0 annihilates holomorphic functions and the Laplacians commute with powers of y as in Equation (2). The Shimura–Maaß operators k δ = y−k∂ yk = ∂ + and y2∂ k τ τ 2iy τ

8 (note the different normalization from [Bru] and [Ze2]!) take modular forms of weight k to modular forms of weight k + 2 and k 2 respectively, or more precisely, satisfy an appropriate commutation relation− with the slash operators for all the elements of SL2(R). They also change Laplacian eigenvalues (again, 2 with respect to ∆k rather than ∆k): δk adds k to the eigenvalue, while y ∂τ subtracts k 2− from it. Moreover, the powers of the Shimura–Maaß operators are given by,− e.g., Equation (56) in [Za], stating that

l l! k + l 1 ∂l−r δl = δ ... δ = − τ k k+2l−2 ◦ ◦ k (l r)!  r (2iy)r Xr=0 − (for arbitrary k, not necessarily integral and non-negative). Theorem 1.1 and (b−) (b−) Proposition 1.3 show that our weight changing operators Rm and L have similar properties. However, our operators are differential operators of order 2 while the Shimura–Maaß operators are of order 1. This is why the results of 2 Propositions 1.5 and 1.6 are more complicated than the fact that δk−2y ∂τ is ∆k 2 k just 4 , the commutator [δ, y ∂τ ]k being simply 4 , and Equation (56) of [Za]. (b−) (b−) Nonetheless, the operators Rm and L for small values of b− are closely + related to the Shimura–Maaß operators. Indeed, for b− = 1 the group SO2,1 is PSL (R) and the tube domain KR + iC is just . We have 2 H 2 H (1) J(M, τ)= j (M, τ), hence [M]m = [M]2m and ∆m = ∆2m

(the same assertions hold for the operators involving anti-holomorphic weights). (1) (1) Our operators Rm and L are squares of the Shimura–Maaß operators, namely

(1) 2 (1) 2 2 Rm = δ2m = δ2m+2δ2m and L = (y ∂τ ) .

Note that in this case (∆ )2 m∆ Ξ(1) = 2m 2m C[∆(1) = ∆ ], m 16 − 8 ∈ m 2m in accordance with the rank of the group being 1 rather than 2 (in particular, ∆1 in the notation of [Na] we have ∆2 = 4 in this case). For b− > 1 many authors (including [Bru] and [Na]) take the basis for KR as two elements spanning a hyperbolic plane together with an orthogonal basis of elements of norm 2. In elements of the positive cone C, the first two − coordinates are positive. In particular, for b− = 2 we have KR + iC = , ∼ H ×+ H with τ = x + iy and σ = s + it being the two coordinates. The group SO2,2 is an order 2 quotient of SL (R) SL (R), acting on G(V ) = through 2 × 2 ∼ H × H (M,N) : (τ, σ) (Mτ,Nσ) with J (M,N), (τ, σ) = j(M, τ)j(N, σ). 7→  It follows that

H H (2) [M,N]m = [M]m,τ [N]m,σ and ∆m = 2∆m,τ + 2∆m,σ

9 (which extend to the operators with anti-holomorphic weights as well). Our operators are (2) (2) 2 2 (2) Rm =2δm,τ δm,σ,L =8y t ∂τ ∂σ and Ξm = ∆m,τ ∆m,σ.

In both cases b− = 1 and b− = 2 the assertions of this Section follow from 2 2 properties of the Shimura–Maaß operators (note that Y is 2y for b− = 1). When b− = 2 the special orthogonal group of a negative definite subspace is also SO(2), which makes the theory of automorphic forms more symmetric. Working with b− = 3 in this model yields another coordinate z = u + iv. The positivity of y, t, and yt v2 is equivalent to − τ z Π= being in = Π= X + iY M (C) Π = Πt, Y 0 . z σ H2 ∈ 2 ≫ 

Hence KR + iC is identified with the Siegel upper half-plane of degree 2. The + group SO2,3 is PSp4(R), with the symplectic action and the factor of automor- phy (hence the slash operators) from the theory of Siegel modular forms. In this case M R(3) = m ,L(3) = Y 2N , and ∆(3) =2T r(Ω ) m − Y 2 − 0 m m,0 in the notation of [Ma1] and [Ma2] for degree 2 (for weight (m,n) the latter (3) assertion extends to the modified Laplacian ∆m,n presented in Section 3). The operator ∆h is also a constant multiple of the operator D considered, for ex- KC e ample, in [CE] and [Ch].

(b−) (b ) 2 Images of Theta Lifts under Rm and L −

For natural r, s, t, and l we define the polynomials (µ,Z )r(µ, Z )t P (µ,Z)= V V and P (l) (µ,Z)= P (µ,Z)(µ2 )l. r,s,t (Y 2)s r,s,t r,s,t −

As a functions of µ V , the polynomial Pr,s,t(µ,Z), considered, e.g., in [Ze2], is homogenous of degree∈ (r + t, 0) with respect to the element of G(V ) represented (l) by Z, while Pr,s,t(µ,Z) has homogeneity degree (r +t, 2l). Equation (5) of [Ze2] (0) (l) extends from Pr,s,t = Pr,s,t to the more general polynomials Pr,s,t: The equality

(l) s−r s−t (l) −1 Pr,s,t(MZ,µ)= j(M,Z) j(M,Z) Pr,s,t(Z,M µ) (3) + holds for every µ V , Z KR + iC, M O (V ), and r, s, t, and l from N. ∈ ∈ ∈ We shall assume that V = LR for some fixed even lattice L (of signature (2,b−)), and consider the of L which is based on the polynomial (l) Pr,s,t. These are (vector-valued) functions of τ = x + iy and Z KR + iC, which are sums of expressions of the form ∈ H ∈ 2 2 (l) (l) µ µ F (τ,Z,µ)= e−∆v/8πy(P )(µ,Z)e τ + + τ − . (4) r,s,t r,s,t  2 2 

10 Here µ are the parts of µ V which lie in the spaces v according to the ele- ± ∈ ± ment of G(V ) corresponding to Z, ∆v is the Laplacian on V which corresponds to the majorant of that element (i.e., to the bilinear form in which the sign on 2πiw the pairing on v− is inverted to be positive as well), and e(w)= e for every complex w. A simple and direct calculation proves

2 Lemma 2.1. (i) We have the equality µ+ = P1,1,1(µ,Z). In addition, the following equalities hold:

2 l 2 l−1 (ii) ∆ P =4rtP . (iii) ∆ − (µ ) =2l(2l + b 2)(µ ) . v+ r,s,t r−1,s−1,t−1 v − − − −

Part (i) of Lemma 2.1 shows that we can write the exponent in Equation 2 µ −2πyP1,1,1 (4) as the constant e τ 2 (independent of Z) times e . Since the differences in the indices in part (ii) of Lemma 2.1 remain the same, l does not (l) affect the weight of modularity of Pr,s,t, and P1,1,1 is invariant (by Equation (3)), we find that replacing P by F in Equation (3) still yields a valid equation. Let L∗ = Hom(L, Z) be the dual lattice of L and L∗/L the (finite) discriminant (l) ∗ group of L. Then the theta function ΘL,r,s,t is the C[L /L]-valued function defined by

(l) (l) (l) (l) ΘL,r,s,t(τ,Z)= θγ+L,r,s,t(τ,Z)eγ, θγ+L,r,s,t(τ,Z)= Fr,s,t(τ,Z,µ) γ∈XL∗/L µ∈Xγ+L

(l) (this function is ΘL(τ, 0, 0; v, Pr,s,t) in the notation of [B], where v G(LR) (l) ∈ corresponds to Z KR + iC). The extension of Equation (3) to ΘL,r,s,t shows (l) ∈ that ΘL,r,s,t is automorphic of weight (s r, s t) as a function of Z KR +iC, and Theorem 4.1 of [B] shows that as a− function− of τ it is a vector-valued∈ b− ∈ H modular form of weight 1+r +t, 2l + 2 and the Weil representation ρL. The latter is a representation of the metaplectic double cover Mp2(Z) of SL2(Z), which is defined by sending the generators T and S of Mp2(Z) lying over the 1 1 0 −1 elements 0 1 and 1 0 of SL2(Z) respectively to   2 ρL(T )(eγ )= e(γ /2)eγ,

b−−b+ ζ8 ρL(S)(eγ)= e( (γ,δ))eδ √∆L ∗ − δ∈XL /L respectively. For the properties of ρL see [Ze1], as well as the reference cited there. The space C[L∗/L] comes with a Hermitian pairing , in which the h· ·iρL eγ are orthonormal, and ρL is a unitary representation with respect to this (l) pairing. The operation of complex conjugation on ΘL,r,s,t interchanges r and t and sends τ to τ (this is equivalent to multiplying the bilinear form on V by 1, but as we rather− stay in the signature (2,b ) setting, we prefer this anti- − − holomorphic operation on τ). It also replaces ρL by its dual representation, but

11 we shall consider the effect of complex conjugation only for the automorphy in the Z variable. (b−) (b−) We are interested in the action of the operators Rm and L on theta kernels, and the resulting differential properties of the associated theta lifts. Several proofs will involve comparisons of these actions on theta kernels with 2 the actions of the operators δk and y ∂τ on these theta kernels (multiplied by the appropriate powers of y). The latter are given (in a more general context) (l) 2 l in Equations (6a) and (6b) of [Ze2]. As Pr,s,t = Pr,s,t(µ−) , Lemma 2.1 shows that in our case these equations take the form

b− b− b− +2 (l) +2 (l) il(2l + b− 2) +2 2 (l−1) δ y 2 lΘ = πiy 2 lΘ + − y 2 l− Θ k L,r,s,t L,r+1,s+1,t+1 8π L,r,s,t (5) (where k =1 b− + r + t 2l) and − 2 −

b− b− b− 2 +2 (l) +2 +2 (l+1) irt +2 (l) y ∂ y 2 lΘ = πiy 2 l Θ + y 2 lΘ (6) τ L,r,s,t L,r,s,t 4π L,r−1,s−1,t−1 (note again the different normalization of these operators). Recall that given a modular form F of weight 1 + r + t b− 2l and − 2 − representation ρL, possibly with exponential growth at the cusps, its theta lift (l) with respect to the polynomial Pr,s,t is defined in [B], [Ze2], and others as follows. For w> 1 let

D = τ τ 1/2, τ 1, τ w , w ∈ H |ℜ |≤ | |≥ ℑ ≤  and assume that

1+r+t−σ dxdy lim y F (τ), ΘL(τ,v,pv) ρ w→∞ L 2 ZDw h i y exists for σ 0 and defines a of σ on some right half- plane, whichℜ may≫ be extended to a meromorphic function of σ for all σ C. (l) ∈ Then the theta lift ΦL,r,s,t(F,Z) is the constant term of the expansion of this meromorphic function at σ = 0. Now, the modular form F has a Fourier expansion of the sort

n F (τ)= cn,γ(y)q eγ, (7) γ∈XL∗/L nX∈Q

n where q denotes e(nτ) and the cn,γ are smooth functions of y = τ, which 2 ℑ vanish unless n γ + Z. The modular forms which are usually considered also ∈ 2 satisfies the condition that cn,γ = 0 unless n (for F which is holomorphic on this means at most a pole at the cusp,≫ and −∞ no essential singularity). HThe relations between the action of the (classical) Shimura–Maaß operators on the lifted modular form and the action of these operators on the theta kernel used for the theta lift are given in the following

12 Lemma 2.2. Let F be a modular form of weight k 2=1+ r + t b− 2l 2 ± ± − 2 − ± and representation ρL, with Fourier expansion as in Equation (7), and assume (l) (l) 2 that the the regularized theta lifts ΦL,r,s,t(Z,δkF−) and ΦL,r,s,t(Z,y ∂τ F+) are εy well-defined. Assume that the growth condition cγ,n(y)= o(e ) as y holds T → ∞ for every γ, n, and ε> 0, and that c0,0(y) is o(y ) as y for some T . Then (l) → ∞ the theta lift ΦL,r,s,t(Z,δkF−) coincides, up to an additive constant which may appear only if r = t, with the value at Z of the theta lift of F− with respect to y2∂ Θ(l) . The same assertion holds for Φ(l) (Z,y2∂ F ) and the theta − τ L,r,s,t L,r,s,t τ + lift of F with respect to δ Θ(l) . + − k L,r,s,t Proof. See Lemmas 3.4 and 3.6 of [Ze2] as well as the argument proving Lemma 2.7 of that reference. Note the factors of 2i distinguishing our operators here from those of [Ze2], and observe that the theta function is conjugated in the integral defining the theta lift.

(l) The complex conjugation of ΘL,r,s,t in the definition of the theta lift implies (l) that ΦL,r,s,t(Z, F ) is automorphic of weight (s t,s r). We shall thus consider only the case r = s, where the automorphy− in (the− corresponding) Equation (3) involves only j(M,Z) and not its complex conjugate. As with Pr,s,t, we may omit the superscript (l) in case l = 0. In the same manner as in Section 1, we shall postpone most of the (calculational) proofs to Section 4. Only the assertions about theta lifts will be proved here. The first assertion we are interested in is b (b−) − Proposition 2.3. The action of Rm takes y 2 ΘL,m,m,0 to 4πi times the b− 2 complex conjugate of y ∂τ y 2 ΘL,m+2,m+2,0 . We remark that Proposition 2.3 may be formulated in terms of comparing the actions of elements from the universal enveloping algebras of sl2(R) and so so (V ) ∼= 2,b− on the theta kernel. However, unlike Proposition 2.3 of [Ze2] (and Proposition 4.5 of [Bru]), which compare the action of order 2 elements of so both universal enveloping algebras, here the one from the algebra of 2,b− has order 2 while the element from sl2 has order 1. We can now prove establish the first property of the theta lift from [Ze2].

Theorem 2.4. Assume that b− is even, and let f be a weakly holomorphic b− modular form of weight 1 2 m and representation ρL. Consider the mod- 1 m − − b− ular form F = (2πi)m δ b− f, of weight k = 1 2 + m, and its theta lift 1− 2 −m − ΦL,m,m,0(Z, F ) considered in Theorem 3.9 of [Ze2]. The image of the latter au- 1 (b−) b−/2 tomorphic form under 2 (Rm ) is a meromorphic automorphic form (8π2)b−/ of weight m + b− on KR + iC, whose singularities are poles of order m + b− along special divisors. Proof. Proposition 2.3 yields the equality

b− b− 1 (b−) i R y 2 Θ (τ,Z)= y2∂ y 2 Θ (τ,Z) 8π2 m L,m,m,0 2π τ L,m+2,m+2,0

13 for every τ and Z KR + iC. As F (as well as its images under any power ∈ H ∈ of δk) satisfies the conditions of Lemma 2.2, we establish the equality

1 1 R(b−)Φ (Z, F )=Φ Z, δ F . 8π2 m L,m,m,0 L,m+2,m+2,0 2πi k 

Repeating this argument, we get

1 1 (R(b−))lΦ (Z, F )=Φ Z, δl F (8π2)l m L,m,m,0 L,m+2l,m+2l,0 (2πi)l k  for any l N. Consider now the case l = b− . Then F = 1 δb−/2F is ∈ 2 (2πi)b−/2 k 1 δm+b−/2f with f weakly holomorphic of weighte 1 b− m (which (2πi)m+b−/2 k−2m − 2 − 1 m+b−/2 is integral since b− is even). But then m+b /2 δ is just the operator (2πi) − k−2m − ∂τ m+b /2 n m+b−/2 n 2πi (which takes q from a Fourier expansion to n q —this is the  b− reason for our normalization), so that the weight 1+ 2 + m modular form F is again weakly holomorphic. Theorem 14.3 of [B] now shows that our automorphic e form of weight m + b−, which we write as ΦL,m+b−,m+b−,0(Z, F ), is meromor- phic on K + iC, with poles of order m + b along rational quadratic divisors R − e associated with negative norm vectors in L∗ whose corresponding coefficients in Equation (7) do not vanish. This completes the proof of the theorem. We remark that in case the modular form f is a harmonic weak Maaß form then the modular form F from the proof of Theorem 2.4 is again weakly holo- morphic. Moreover, in case the image of f under the operator ξ of [BF] e k−2m does not have a pole at the cusp, the theta lift has no additional singulari- ties, and the result of Theorem 2.4 extends to this case. However, in the theta lift ΦL,m,m,0(Z, F ) itself one can still distinguish the case where f is weakly holomorphic from the one where F is such a harmonic weak Maaß form. For the weight lowering operator L(b−), we do not have a nice equivalent to Proposition 2.3. However, we do have an interesting result concerning its mth power. We begin with

Lemma 2.5. The image of Θ(l) ( τ,Z) under L(b−) is L,k,n,n − b 4π2y2Θ(l+1) ( τ,Z)+ n 2l + − Θ(l) ( τ,Z)+ L,k+2,n,n −  2  L,k+1,n−1,n−1 −

n(n 1)l l 1+ b− + − − 2 Θ(l−1) ( τ,Z). 4π2y2  L,k,n−2,n−2 −

Lemma 2.5 allows us to establish the following

14 (b−) s Proposition 2.6. For any s N, the image of ΘL,m,m,0 under (L ) attains, on τ and Z, the value ∈

b− 2 2 h s Γ s + 2 m!(4π y ) (h) ΘL,s+h,m−s+h,m−s+h( τ,Z). hΓh + b−  (m s + h)! − Xh 2 − 

The case s = m in Proposition 2.6 is of particular importance, as is shown in the following

b− (b−) m Proposition 2.7. The expression (L ) y 2 ΘL,m,m,0 equals the complex b− (m) m m 2 +2m conjugate of ( 4πi) δ b− y ΘL,0,0,m(τ,Z). − 1− 2 −m,τ Automorphic forms of non-zero weight can never be real-valued, because complex conjugation yields an automorphic form with a different weight. How- ever, multiplying the complex conjugate automorphic form by a power of Y 2 leads to an object which is comparable with the image of our automorphic form under the appropriate power of a weight changing operator, as these two func- tions do have the same weight. We shall thus say that an automorphic form Φ, of positive weight m, is m-real if its image under the mth power of the weight lowering operators L(b−) coincides with its complex conjugate multiplied by a positive multiple of (Y 2)m. We now show that the theta lifts from Theorem 3.9 of [Ze2] are m-real, or more generally: Theorem 2.8. Let F be as in Theorem 2.4 (but without the restriction on the parity of b−) , and assume that F is an eigenfunction with respect to (minus) the Laplacian of weight 1 b− +m, with eigenvalue λ = mb− . Assume further − 2 − 2 that the Fourier coefficients cγ,n of F appearing in Equation (7) are real. Then m (b−) m i applying the operator (L ) to 2 ΦL,m,m,0(Z, F ) yields the complex conjugate im b− 2 m b− of 2 ΦL,m,m,0(Z, F ) multiplied by m!Γ m + 2 (Y ) /Γ 2 .

m   i (b−) m Proof. By Proposition 2.7, the image of 2 ΦL,m,m,0 under (L ) coincides im with 2 times the regularized integral of F paired with the function

b− (m) m m 2 +2m ( 4πi) δ b− y ΘL,0,0,m(τ,Z). − 1− 2 −m,τ

On the other hand, the fact that the first index in P0,0,m vanishes allows us to use Equation (6) successively m times and write

b− b− +2 (m) 2 ( πi)my 2 mΘ (τ,Z) as just ( y ∂ )my 2 Θ (τ,Z). − L,0,0,m − τ L,0,0,m

(b−) m As in the proof of Theorem 2.4, we can write (L ) ΦL,m,m,0(Z, F ), using im m m 2 m Lemma 2.2, as the theta lift 2 ΦL,0,0,m Z, 4 δ b− ( y ∂τ ) F . Now, 1− 2 −m,τ − 2  as F is an eigenfunction and y ∂τ takes eigenfunctions to eigenfunctions, we can replace each combination 4δ y2∂ , starting from the inner pair, by the − l τ

15 2 r appropriate eigenvalue. As after applying ( y ∂τ ) the eigenvalue becomes − − λ r m r b , the modular form we plug inside the latter lift is just F − − − 2 multiplied by the scalar m−1 λ r m r b− . Substituting the value r=0 − − − 2 of λ, the rth multiplierQ becomes just (r m) r +b− , and the product is − 2 m b− b− b− 2 m b− ( 1) m!Γ m + 2 /Γ 2 . Division by m!Γ m + 2 (Y ) /Γ 2 thus gives − m (−i)     2 ΦL,0,m,m(Z, F ), so that we need to show why ΦL,0,m,m(Z, F ) is the com- plex conjugate of ΦL,m,m,0(Z, F ). As the Fourier coefficients of F are real, we obtain F (τ) = F ( τ). On the other hand, we have seen that complex conju- gation on our theta− function interchanges the indices r and t and replaces the variable τ by τ. The required assertion now follows from the fact that powers − dxdy of y and the measure y2 are both preserved by the change of variable τ τ. This completes the proof of the theorem. 7→ − We remark that the choice of λ = mb− in Theorem 2.8 is not crucial. Any − 2 choice of λ for which the number m−1 r m r b− λ is positive will be r=0 − − 2 − sufficient for Theorem 2.8 to holdQ (with the same proof).  However, we chose this eigenvalue as it is the eigenvalue of the theta lifts from [Ze2].

(b−) (b ) 3 Proofs of the Properties of Rm and L −

In this Section we include the proofs of the properties of the weight raising and weight lowering operators appearing in Section 1. We first introduce (following [Na]) a convenient set of generators for O+(V ). + For ξ KR we define the element p SO (V ) whose action is ∈ ξ ∈ 2 ⊥ ξ µ KR = z, ζ µ (µ, ξ)z, ζ ζ + ξ z, z z. ∈ { } 7→ − 7→ − 2 7→   ∗ Furthermore, given an element A O(KR) and a scalar a R such that a> 0 + ∈ + ∈ if A O (KR) and a< 0 otherwise, we let ka,A O (V ) be the element acting as ∈ ∈ 2 2 ⊥ ζ 1 ζ µ KR = z, ζ Aµ, ζ z ζ z , z az. ∈ { } 7→ − 2 7→ a − 2  7→   For any Z KR + iC we have ∈ 1 p Z = Z + ξ, J(p ,Z)=1, k Z = aAZ, and J(k ,Z)= . ξ ξ a,A a,A a Note that the relation between A and the sign of a is equivalent to preserving C rather than mapping Z into KR iC—it appears that [Na] ignored this − point. Choose now an element of G(KR) in which the positive definite space is + generated by the norm 1 vector u1, and consider the involution w SO (KR) defined by ∈

2 2 ⊥ ζ ζ µ KR = z, ζ µ 2(µ,u )u , ζ z z, z ζ z ∈ { } 7→ − 1 1 − 2 7→ − 7→ − − 2   

16 (w inverts the positive definite space Ru1). Its action on KR + iC is through

2 Z2 wZ = Z 2(Z,u )u with J(w,Z)= . Z2 − 1 1 2   ∗ The elements k with (a, A) in the index 2 subgroup of R O(KR) thus a,A × defined and pξ for ξ KR generate the stabilizer StO+(V )(Rz) of the isotropic space Rz in O+(V ) as∈ the semi-direct product of these groups. The fact that + adding w to StO+(V )(Rz) generates O (V ) is now easily verified by considering the action on isotropic 1-dimensional subspaces of V . Some useful relations are derived in the following

Lemma 3.1. Let KR be a non-degenerate vector space of dimension b−, fix α C, and let F be a 2 function which is defined on a neighborhood of a point ∈ C 2 Z = X + iY KC with Y > 0. Then the following relations hold: ∈ 2iα α(α 1+ b− ) (Y 2)−α∆h (Y 2)αF (Z) = ∆h F (Z) D∗F (Z) − 2 F (Z) KC KC − Y 2 − Y 2  and

2iα α(α 1+ b− ) (Y 2)−α∆h (Y 2)αF (Z) = ∆h F (Z)+ D∗F (Z) − 2 F (Z). KC KC Y 2 − Y 2 

We remark that Lemma 3.1 holds for KR of arbitrary signature (not neces- sarily Lorentzian), but not negative definite (for Y 2 > 0 to be possible). Proof. The proof is obtained by a straightforward calculation, using an or- thonormal basis for KR and the action of ∂k and ∂k on functions of Y alone. R We remark that the third operator ∆KC bears a property similar to Lemma 3.1, which is used implicitly in Section 3 of [Ze2] in order to prove Equation (2). We can now present the

(b−) Proof of Theorem 1.1. Multiply both sides of the desired assertion for Rm , as 2 m (b−) well as the function F there, by (Y ) . Lemma 3.1, the first definition of Rm , and Equation (1) show that this yields the equivalent equality

(b−) (b−) (R0 F )[M]2,−m = R0 F [M]0,−m). Observe that conjugating the latter equation and multiplying by (Y 2)2 yields the required equality for L(b−). Hence we are reduced to proving only this equality. (b−) Moreover, R0 involves only holomorphic differentiations, which means that it commutes with the power of J(M,Z) coming from the anti-holomorphic weights. Hence we can take m = 0, which implies that proving the equation

(b−) (b−) (R0 F )[M]2 = R0 F [M]0) 17 (b−) (which the assertion for R0 in the formulation of the theorem) suffices for proving the theorem. Writing the arguments as M −1(Z) in both sides and using the cocycle condition brings the latter equation to the form

− (b−) −1 2 (b−) M 1 (R0 F )(Z)J(M ,Z) = (R0 ) F (Z). (8) By a standard argument it suffices to verify Equation (8) for M −1 being one of + −1 the generators of O (V ) considered above. Equation (8) with M = pξ follows h ∗ from the invariance of both ∆KC and D under translations of X = Z and the −1 h 2ℜ ∗ fact that J(pξ,Z) = 1. The action of M = ka,A divides ∆KC by a , leaves D 2 2 invariant, and divides Y by a (since A O(KR)), which proves Equation (8) 1 −1 ∈ since J(ka,A,Z)= a . Finally, for M = w we have the equalities

2 2 2 2 2 h w Z h Z ∗ w Z ∗ 2iY (∆KC ) = ∆KC (b− 2) D, (D ) = 2 D 2 D  2  − − 2 Z − Z 1 h with D = k zk∂k from [Na] (the corresponding operator from [Na] is 2 ∆KC 2 P h Z2 Z Y 2 rather than ∆KC , while δ = 2 , δ = 2 , and d = 2 there). Using Equation −1 h i(b−−2) ∗ (1) we thus find that applying M = w to the sum of ∆KC and Y 2 D (b−) Z2 2 (which is R0 ) multiplies it by 2 (as the coefficients in front of D cancel), which establishes Equation (8) also for this case using the value of J(w,Z). This completes the proof of the theorem. In order to indicate what is the Lie-theoretic interpretation of the operators (b−) (b−) Rm and L , we recall the vector u1 we used for defining w above, and take a vectoru ˜ KR of norm 1 which is orthogonal to u (we assume here b > 1, ∈ − 1 − but for b− = 1 our operators are squares of the order 1 operators δ2m and 2 y ∂τ , whose Lie-theoretic interpretation is given, e.g., in [Ve]). These choices determine the parabolic subgroup of SO+(V ) appearing in the following

+ Proposition 3.2. Let HKR be the subgroup of SO (KR) consisting of those matrices which preserve the isotropic subspace R(u1 +u ˜) and whose action on ⊥ the quotient (u1 +˜u) /R(u1 +˜u) is trivial. Define H to be the group generated by all the elements p with ξ KR and by the elements k with a > 0 and ξ ∈ a,A A H R . Then the group H operates freely and transitively on KR + iC. ∈ K Let K = SO(2) SO(b ) be the stabilizer, in SO+(V ), of the element of ∼ × − G(V ) represented by Z = iu1, and let k be its Lie algebra. The action of a normalized generator of so(2) k on so(V )C decomposes the latter space into the eigenspaces with eigenvalue⊆ 0 (this is precisely k) and i (complex conjugate ± spaces of dimension b− each). Hence the action on the space of products of two elements of so(V ) (inside its universal enveloping algebra, say) decomposes into eigenspaces with eigenvalues 0 and 2i. One verifies that in each of the 2i- ± ± eigenspaces, precisely one combination commutes with the part so(b−) of k. As our automorphic forms correspond to functions on SO(V ) on which SO(2) K ⊆ operates according to a specific character and SO(b−) operate trivially (normal- ized suitably), these elements (of order 2) of the universal enveloping algebra

18 of so(V ) lead to weight raising and weight lowering operators. One may then evaluate, using the interplay between the operations of k and the Lie algebra of the group H from Proposition 3.2, the action of these operators, and find that (b−) (b−) they lead to our Rm and L . However, the change of coordinates between

HKR and KR +iC in this evaluation is more tedious than one might have believe. We also indicate briefly the connection between our operators and those of × [Sh1]. That reference defines, for every representation ρ of C GLb− (C) (a subgroup of which we identify the complexification of the compact× subgroup × K, which is isomorphic to the product C SO(b−, C)), a differential operator that roughly sends (vector-valued) automorphic× forms with weight (i.e., repre- sentation) ρ to automorphic forms having representation ρ ω, where ω is the standard representation of that product on Cb− . This representation⊗ space is considered as the holomorphic cotangent space of G(V ), and the operator is, in fact, just the holomorphic differential map d, twisted by the image of a scalar η and a matrix ξ (both defined explicitly in [Sh1]) via the representation ρ. Starting with the 1-dimensional representation which is the mth power of C× (this is the representation associated with our automorphic forms of weight m) and repeating this operation twice, we obtain an automorphic form with repre- sentation involving ω⊗2. The idea is expressing the resulting automorphic form when ω is identified with KC, and using the bilinear form on the latter space in order to replace the ω⊗2-valued automorphic forms by scalar-valued ones. Now, we replace the coordinate denoted z in [Sh1] by u = √2z, considering it as lying in the complexified space (v−)C associated to some base point for

G(V ), and decompose it as some multiple uz of zv− plus a vector u⊥ which is perpendicular to zv− . Here z is again the isotropic vector we used for defining KR. Choosing the positive part of z appropriately (recall that the vector denoted p(z) in [Sh1] is not presented in the canonical form), we obtain that our norm t t 0 vector has pairing 1 + u u 2u zv− with z and its positive and negative KC t − coordinates are i(1 u u) and 2u⊥ respectively. It follows that the associated − 2 element Z of KC (which can be shown to be in KR + iC) satisfies (Z + ie+) = −4 1+utu−2utz− (where e+ is the generator of the positive part of KR), so that the inverse map sends Z to the vector obtained by multiplying the positive part of Z+ie+ 2 2 by i, and adding zv− to the result. Given an automorphic form F − (Z+ie+) of weight m on G(V ), a very lengthy, tedious, and involved calculation gives us the expression for the ω⊗2-valued automorphic form obtained from F under the operator mentioned in the previous paragraph, and after applying the pairing we 2m b− −2m obtain an expression closely related to (Z +ie+) Rm [(Z +ie+) F ]. Indeed, 16Y 2 the expression denoted by η in [Sh1] becomes 2 2 using our variable, so |(Z+ie+) | that multiplying by ηm before applying the operator and by η−m afterwards 2m b− corresponds to the operation involving Y appearing in the definition of Rm , 2m as well as the additional operation with (Z + ie+) . However, the details of this calculation are very long as well, and therefore we have chosen to state and prove Theorem 1.1 more directly.

19 For calculational purposes it turns out convenient to introduce the operator

∆(b−) = ∆(b−) 2n(2m b ), m,n m,n − − − on which complex conjugatione interchanges the indices m and n. The operator D∗ (D∗)2 = y y ∂ ∂ − 2i k l k l Xk,l will also show up, so we denote it (^D∗)2. We now turn to the

(b−) 2 m Proof of Proposition 1.3. Conjugating the desired equality for Rm by (Y ) , (b−) applying Equation (2), and taking the differences between the operators ∆m,n (b−) (b−) and ∆m,n into consideration, we see that the asserted equality for Rme is equivalent to

∆(b−) R(b−) R(b−)∆(b−) = (2b +4m 4)R(b−). 2,−m 0 − 0 0,m − − 0 Moreover, multiplyinge the complexe conjugate of the latter equation by (Y 2)2 (b−) (b−) (b−) and comparing ∆2,−m with ∆2,−m yields the required property for L (with the index m replaced by m). Hence, as in the proof of Theorem 1.1, we are e reduced to proving this single− equation. In addition, the dependence on m of the left hand side enters only through the difference 4imD∗ between the operators − − − ∆(b ) and ∆(b ) with l 0, 2 . As a simple calculation yields l,−m l ∈{ } e D∗ iD∗ D∗, ∆h = i∆h and D∗, = , KC KC  Y 2  Y 2   (b−) it suffices to prove the equality for m = 0 (i.e., the original assertion for R0 ):

∆(b−)R(b−) R(b−)∆(b−) = (2b 4)R(b−). 2 0 − 0 0 − − 0 (b−) (b−) The commutator of ∆0 and R0 is evaluated using the equalities ∆R D∗ 2, ∆h = iD∗∆h + iD∗∆R + KC , | | KC KC KC 2   D∗ 3i D∗ 2 i(^D∗)2 + D∗ b D∗ 2, = | | − , Y 2∆R , ∆h =2iD∗∆R + − ∆R , | | Y 2  2Y 2 KC KC KC 2 KC   ∗ ∗ 2 ^∗ 2 h R ∗ D 2i D 2i(D ) + i∆ + i∆ + (2 b−)D and Y 2∆R , = | | − KC KC −  KC Y 2  2Y 2 (which all follow from straightforward calculations). Applying the equalities

∗ ∗ 2 ∗ − − D 2 D iD ∆(b ) = ∆(b ) 8iD∗ and D∗ = | | − 2 0 − ◦ Y 2 2Y 2  and putting in the appropriate scalars now establishes the proposition.

20 Our next task is the

(b−) (b−) Proof of Proposition 1.5. We begin by evaluating Rm−2L written as

R(b−) (Y 2)2∆h +R(b−) i(2 b )Y 2D∗ = (Y 2)2R(b−)∆h +i(2 b )Y 2R(b−) D∗. m−2 KC m−2 − − m KC − − m−1 Using the equalities D∗ ∆h , D∗ = i∆R and D∗D∗ = D∗ 2 + KC − KC | | 2i   we establish the equation

− (2 b )(2m b ) R(b ) L(b−) = Ξ(b−) + − − − − ∆(b−), m−2 m 8 m

(b−) where Ξm is defined in the formulation of the proposition. We now decompose (b−) (b−) (b−) 2 2 (b−) (b−) Rm in L Rm (which is (Y ) R0 Rm ), yielding

− − m(2m +2 b ) − (Y 2)2R(b )∆h i(2m +2 b )Y 2R(b )D∗ − − Y 2R(b ). 0 KC − − − −1 − 2 −1 The formulae D∗ ∆h ,D∗ = i∆R and D∗D∗ = D∗ 2 KC KC | | − 2i   now show that b (2m +2 b ) mb (2m +2 b ) L(b−)R(b−) = Ξ(b−) − − − ∆(b−) + − − − . m m − 8 m 4 The required commutation relation follows. As Theorem 1.1 shows that the (b−) (b−) (b−) (b−) compositions Rm−2L and L Rm commute with all the slash operators of weight m, and Proposition 1.3 implies that these operators commute with ∆m, (b−) the assertion about Ξm is also established. This proves the proposition. Finally, we come to the Proof of parts (iii) and (iv) of Proposition 1.6. We prove part (iii) by induc- (b−) l tion (the case l = 0 being trivial). If (Rm ) is presented by the asserted (b−) l+1 (b−) (b−) l formula then (Rm ) , which is Rm+2l(Rm ) , equals

l c (iD∗)c−s(∆h )l−c R(b−) (iD∗)c−s(∆h )l−c R(b−) A(l) KC = A(l) m+2l−c KC . m+2l s,c ( Y 2)c s,c ( Y 2)c Xc=0 Xs=0 − Xs,c − ∗ D 1 (b−) For each c, the term involving Y 2 (resp. Y 2 ) in Rm+2l−c takes the term with indices c and s (for l) to a multiple of the term with corresponding to c + 1 and h s (resp. c + 1 and s +1) for l +1. For ∆KC we have c c s ∆h ,iD∗ = ∆h hence ∆h (iD∗)c−s = − (iD∗)c−a∆h , KC KC KC a s KC   Xa=s −

21 − (∆h )l c KC (b−) l+1 and we multiply the latter sum by (−Y 2)c . This shows that (Rm ) can be (l) (b−) l expressed by the asserted formula. Putting in the multipliers As,c from (Rm ) ∗ D 1 (b−) and the coefficients of Y 2 and Y 2 in Rm+2l−c, summing over c and s, and taking the coefficient in front of the term with indices c and a (and l + 1) in the result, we obtain the recursive relation asserted in part (iii). We now observe that for a = 0 the recursive formula reduces to

A(l+1) = A(l) + (2m +4l 2c +4 b )A(l) . 0,c 0,c − − − 0,c−1 (l) (l) (0) (0) Denote the asserted value of A0,c by B0,c. As A0,0 =1= B0,0, it suffices to show (l) that the numbers B0,c satisfy the latter recursive formula. But the equality b b 2(l c+1) m+l c − +1 +c(2m+4l 2c+4 b )=2(l+1) m+l − +1 −  − − 2  − − −  − 2 

− l!·2c 1 holds for every l and c (and m and b−), and multiplication by c(l+1−c)! and by b− m+l− 2 the binomial coefficient c−1 yields the required recursive relation for the (l)  numbers B0,c. This completes the proof of the proposition.

4 Actions on Theta Kernels—Proofs

The main technical lemma, which will be required for the evaluations in most of the following proofs, is based on

(b−) (b−) Lemma 4.1. Given µ LR, the operators R0 and L take the function ∈ b− b− P of Z KR + iC to P and P . 1,1,1 ∈ − 2 0,2,2 − 2 2,0,0 2 (b−) Proof. The commutation relation between powers of Y and the operators Rm obtained from the first definition of the latter operators in Theorem 1.1 and the fact that the latter operators involve only holomorphic differentiation allows us (b−) (b−) to write R0 P1,1,1 as P0,1,1R−1 (µ,ZV,Z ). Hence we must evaluate the oper- h ∗ ation of ∆KC and D on (µ,ZV,Z ). For the latter operator a simple calculation yields

2iD∗(µ,Z )=2i(µ, Y )+2Y 2(µ,z) = (µ,Z ) (µ, Z )+2Y 2(µ,z). V,Z V,Z V,Z − V,Z The former operator is pure of weight 2, hence its action gives a non-zero result Z2 only on the part 2 (µ,z), and using an orthonormal basis one finds that this result is just b −(µ,z). Combining these results, we find that − − ib b b R(m) = ∆h + − D∗ − (µ,Z )= − (µ, Z ),  −1 KC Y 2 − 2Y 2  V,Z −2Y 2 V,Z

(b−) (b−) from which the value of R0 P1,1,1 follows. The assertion about L P1,1,1 is a (b−) (b−) consequence of the value of R0 P1,1,1, since P1,1,1 is a real function and L

22 (b−) 2 2 is the operator which is complex conjugate to R0 , multiplied by (Y ) . This proves the lemma. Another useful evaluation appears in the following

Lemma 4.2. The holomorphic and anti-holomorphic Z-gradients of P1,1,1 have, 2 2 as vectors in KC, the norms P0,2,2µ− and P2,2,0µ− respectively.

Proof. (µ, ZV,Z ) is anti-holomorphic, and the holomorphic gradients of (µ,ZV,Z ) 2 and Y are µKR (µ,z)Z and iY respectively, where µKR is the orthogonal − − ⊥ projection of µ LR onto KR = z, ζ . It follows that P1,1,1 has holomorphic gradient ∈ { } P Y 2(µ (µ,z)Z)+ i(µ,Z )Y . 0,2,1 KR − V,Z Now, the (easily evaluated) equalities  µ (µ,z)Z, Y = (µ, Y ) iY 2(µ,z) KR − V,Z −  and (µ,z)2Z2 2(µ,z)(µ ,Z)+2(µ,z)(µ,Z )=2(µ,z)(µ, ζ) ζ2(µ,z)2 − KR V,Z − reduce to the norm of the latter gradient P µ2 + 2(µ, ζ)(µ,z) ζ2(µ,z)2 P . 0,2,2 KR − − 1,1,1  2  But µ is µKR ,µz, (µ, ζ) ζ µz in the KR R R coordinates, so that the − × × 2 2 sum of the first three terms in the brackets is just µ . Subtracting P1,1,1 = µ+ completes the proof of the first assertion, and the second assertion follows from complex conjugation since the function P1,1,1 is real-valued. This proves the lemma.

For µ LR and τ = x + iy we denote the vector √2πyµ byµ ˜. Its norm 2 ∈ ∈ H is 2πyµ , and after choosing an element of G(LR), it decomposes intoµ ˜+ (of 2 2 norm 2πyµ+) andµ ˜− (whose norm is 2πyµ−). We now prove Proposition 4.3. Let f : R R be a smooth function. Then the images of −→ − the function f(˜µ ) under R(b ) and L(b−) are 2πyP µ˜2 f ′′(˜µ ) b f ′(˜µ ) + 0 0,2,2 − + − 2 + and 2πyP µ˜2 f ′′(˜µ ) b− f ′(˜µ ) respectively.   2,0,0 − + − 2 + Proof. Both operators consist of a first order operator D (a multiple of D∗ ∗ h h or of D ) and a second order operator ∆ (which equals ∆KC or ∆KC ). Then D f(T ) = DT f ′(T ), and ∆ f(T ) is the sum of ∆T f ′(T ) and an ex- · ′′ 2 · pression  involving f (T ). In our caseT =µ ˜+ = 2πyµ+ = 2πyP1,1,1, so that ′ (b−) (b−) the coefficient of f (T ) is just 2πy times R0 P1,1,1 and L P1,1,1, and the latter expressions are evaluated using Lemma 4.1. The coefficients of f ′′(T ) h 2 2 h coming from ∆ being ∆KC or (Y ) ∆KC are the norms (in KC) of the holomor- phic and anti-holomorphic gradients of T , the latter being multiplied by (Y 2)2. For T =2πyP1,1,1 these norms take the values given in Lemma 4.2, multiplied 2 2 by (2πy) . Gathering these results together and substituting the value ofµ ˜− completes the proof of the proposition.

23 We now turn to proving assertions concerning the images of theta lifts (or complex conjugates of theta functions), having only holomorphic weights of (b−) (b−) automorphy, under the operators Rm and L . This was seen to boil down (l) to the operation on the function Fr,s,t from Equation (4), with τ replaced by τ, under the additional assumption s = t. The exponent was seen, using part − µ2 2 (i) of Lemma 2.1 to be e τ e−µ˜+ , where the first multiplier is a constant − 2 (i.e., independent of Z). The polynomial part is evaluated in Lemma 4.4. (i) For any natural numbers k and n we have dk n −∆v+ /8πy −2πyP1,1,1 k n −T (2πy) Pn−k,0,0e (Pk,n,n)e = ( 1) k (T e ) . − dT 2 T =˜µ+

∆v− /8πy l 2 l l b− b− 2 p (ii) Applying e to (2πy) (µ−) yields p p Γ l+ 2 /Γ p+ 2 µ˜− . P     We allow the index n k appearing in Part (i) here to be negative, with the − natural extension of the definition of Pr,s,t to negative r. We remark that the expressions obtained in this part are just the generalized Laguerre polynomials (n−k) Lk , multiplied by the exponents, and normalized appropriately. Proof. Multiple applications of part (ii) of Lemma 2.1 show that

h ∆v k!n!P + P = k−h,n−h,n−h . h!( 8πy)h k,n,n (k h)!(n h)!h!( 2πy)h − − − − n Multiplying by (2πy) Pn−k,0,0 and summing over h, the left hand side of the equation in part (i) becomes just k n! ( 1)h(2πyP )n−he−2πyP1,1,1 . h(n h)! − 1,1,1 Xh − On the other hand, differentiating the product T ne−T k times with respect to T yields

k k k d k d k−h k n!T n−h T n e−T = ( 1)k−he−T , hdT  · dT  h(n h)! − hX=0 hX=0 − 2 and substituting T =µ ˜+ = 2πyP1,1,1 yields the same expression multiplied by ( 1)k. This establishes part (i). For part (ii), applying part (iii) of Lemma 2.1 successively− evaluates

4l−pl! Γ l + b− ∆l−p(µ2 )l = 2 (µ2 )p. v− − p! · b−  − Γ p + 2  Dividing this term by (8πy)l−p(l p)!, multiplying everything by (2πy)l, and 2 2 − substitutingµ ˜− = 2πyµ− gives the asserted expression. This completes the proof of the lemma.

24 2 2 2 Asµ ˜− =µ ˜ µ˜+, Lemma 4.4 implies that the dependence of the expression n+l − (l) (2πy) Pn−k,0,0Fk,n,n( τ,Z,µ) (or the corresponding theta function) on the − 2 variable Z is only through the quantityµ ˜+. For convenience, we gather these results in the following Corollary 4.5. Define the functions

dk f (w) (T ) = ( 1)k (T ne−T ) (w T )p, k,n,p − dT k · − where k, p, and n are natural numbers and w R. Then the theta function ∈ Θ(l) ( τ,Z) equals L,k,n,n − 2 b− (˜µ ) 2 2 l Γ l + f (˜µ+) µ 2 k,n,p e τ e . b− n+l µ+L ∗ p  (2πy) Pn−k,0,0  − 2  µX∈L Xp Γ p + 2 

−∆v/8πy (l) Proof. Just substitute the value of e (Pk,n,n), which equals the product of −∆v+ /8πy −∆v+ /8πy 2 l e (Pk,n,n) and e (µ−) , from Lemma 4.4 into the expression defining the theta function. 

We can now present the Proof of Proposition 2.3. As seen above, it suffices to consider the action of − (b ) −2πyP1,1,1 Rm only on the expression P0,m,m(µ,Z)e with fixed µ (recall that (b−) P0,m,m is harmonic). The holomorphicity of the differentiation in Rm shows − 2 (b ) −µ˜+ −T that the result is the same as P0,m,mR0 e . By putting f(T )= e , Propo- − 2 − (b ) −µ˜+ b −µ˜+ sition 4.3 evaluates R0 e as 2πyP0,2,2 µ˜− + 2 e , and multiplying by P0,m,m yields 

2 b− R(b−)P e−µ˜+ =4π2y2P µ2 + e−2πyP1,1,1 . m 0,m,m 0,m+2,m+2 − 4πy 

∆v− /8πy 2 But the expression in parentheses is e (µ−) by part (ii) of Lemma 4.4, and the harmonicity of P0,m+2,m+2 allows us to put it also into the action of e−∆v/8πy without affecting the resulting expression. Putting in the missing b− µ2 constant y 2 e τ e and summing over µ L∗ we establish the equality − 2 µ+L ∈  b− b− ( −) 2 2+ (1) R b y 2 Θ ( τ,Z)=4π y 2 Θ ( τ,Z). m L,0,m,m − L,0,m+2,m+2 −

But as Pm+2,m+2,0 is harmonic, Equation (6) shows that applying the operator b− 2 4πiy ∂τ to y 2 ΘL,m+2,m+2,0(τ,Z) yields the complex conjugate of the latter expression,− and complex conjugation inverts the sign of 4πi. This proves the proposition.

25 We now turn to the

(l) Proof of Lemma 2.5. Write the theta function ΘL,k,n,n( τ,Z) as in Corollary − 2 4.5. It suffices to fix µ L∗ and compare the coefficients of e τ µ e in ∈ − 2 µ+L both sides. Take some 0 p l, and apply Proposition 4.3 with the function (˜µ2) ≤ ≤ f = fk,n,p. The powers of 2πy and P1,0,0 from Corollary 4.5 and Proposition n+l−1 4.3 merge to (2πy) Pn−2−k,0,0 in the denominator, and the remaining part 2 (b−) (˜µ ) of L fk,n,p is

b− 2 2 l Γ l + 2 2 (˜µ ) ′′ 2 b− (˜µ ) ′ 2 µ˜− f (˜µ+) f (˜µ+) . pΓp + b−   k,n,p − 2 k,n,p  2    Asµ ˜2 =µ ˜2 µ˜2 , and as one easily evaluates − − + (f w )′(T )= pf w (T ) f w (T ), k,n,p − k,n,p−1 − k+1,n,p the part in brackets in latter expression equals

2 b 2 b 2 f (˜µ ) (˜µ2 )+ 2p + − f (˜µ ) (˜µ2 )+ p p + − 1 f (˜µ ) (˜µ2 ). (9) k+2,n,p+1 +  2  k+1,n,p +  2 −  k,n,p−1 +

We now write the denominator in the preceding constant as

p + b− 1 1 2 , , and Γ p +1+ b− Γ p + b− p 1+ b− Γ p 1+ b− 2 2 − 2 − 2     in front of the three terms in Equation (9) respectively, and after taking the sum over p and gathering the functions with the same index p together, we see b− b− that the quotient Γ l + 2 /Γ p + 2 multiplies   b l 2 b l 2 l 2 p 1+ − f (˜µ ) + 2p + − f (˜µ ) + (p + 1) f (˜µ )  − 2 p 1 k+2,n,p  2 p k+1,n,p p +1 k,n,p − 2 a a−1 (where we have omitted the variableµ ˜+). Using the identity b b = a b−1 we can write the latter expression as  

l 1 2 l 1 2 l 1 2 l − f (˜µ ) (˜µ2 )+2 − f (˜µ ) (˜µ2 )+ − f (˜µ ) (˜µ2 ) + p 2 k+2,n,p + p 1 k+1,n,p +  p  k,n,p +  − − b l 2 l 2 + − f (˜µ ) (˜µ2 )+ f (˜µ ) (˜µ2 ) . (10) 2 p 1 k+2,n,p + p k+1,n,p +  − Now, differentiating k times and multiplying by (w T )p takes the equality − (T ne−T )′ = (nT n−1 T n)e−T to f (w) (T ) f (w) (T )= nf (w) (T ). − k,n,p − k+1,n,p k,n−1,p

26 (˜µ2) (˜µ2) (˜µ2) One application of this relation replaces fk+1,n,p by fk+2,n,p + nfk+1,n−1,p, and we also obtain

2 2 2 2 f (˜µ ) = f (˜µ ) +2nf (˜µ ) + n(n 1)f (˜µ ) . k,n,p k+2,n,p k+1,n−1,p − k,n−2,p Each of the terms in Equation (10) thus contributes to the total coefficient in (˜µ2) front of fk+2,n,p, which using the classical properties of the binomial coefficients b− l+1 reduces to l + 2 p . Using the recursive property of the gamma function   l+1 b− b− again, we obtain the coefficient p Γ l +1+ 2 /Γ p + 2 , which together with    1 4π2y2 n+l−1 = n+l+1 Pn−2−k,0,0(2πy) Pn−2−k,0,0(2πy)

2 (˜µ ) 2 yields the coefficient appearing in front of fk+2,n,p(˜µ+) in the expansion of 2 2 (l+1) 4π y ΘL,k+2,n,n( τ,Z) in Corollary 4.5. The total coefficient in front of the 2 − (˜µ ) function fk+1,n−1,p in Equation (10) becomes (again, using binomial identities) b− l just 2l + 2 p , and the gamma quotient and the powers of 2piy and P1,0,0 complete the formula for the second asserted term. For the remaining term 2 n(n 1)l l−1 f (˜µ ) from Equation (10) we use the functional equation of the − p k,n−2,p gamma function  again to write Γ l + b− as l 1+ b− Γ l 1+ b− , and we 2 − 2 − 2 also decompose   

n+l−1 2 2 n−2+l−1 Pn−2−k,0,0(2πy) =4π y Pn−2−k,0,0(2πy) .

Corollary 4.5 then establishes the remaining asserted term in a similar manner. This completes the proof of the lemma. We go on to the Proof of Proposition 2.6. We prove the assertion by induction on s. The case s = 0 is trivial. Denote the asserted coefficient corresponding to the hth term (b−) s in the expression for the image under (L ) by as,h(y). We need to evaluate

(b−) (h) as,h(y)L ΘL,s+h,m+s−h,m+s−h, Xh and compare it with the asserted expression for s + 1. Lemma 2.5 shows that for each h the L(b−)-image of the corresponding theta function is a linear com- bination of three theta functions, which correspond to the index s + 1 and the indices h 1, h, and h + 1. After applying the appropriate summation index − (h) changes, the coefficient which we get in front of ΘL,s+1+h,m−s−1+h,m−s−1+h in (b−) s+1 (L ) ΘL,m,m,0 is

b 4π2y2a (y) + (m s + h) 2h + − a (y)+ s,h−1 −  2  s,h

27 (m s + h)(m s + h + 1)(h + 1) h + b− + − − 2 a (y). 4π2y2  s,h+1

Substituting the values of as,t for t being h 1, h, and h+1, one easily sees that −m!(4π2y2)h all three terms yield the same multiplier (m−s−1+h)! . Applying the functional equation for the gamma function in the first and third term, we obtain that the remaining expression equals

Γ s + b− b s b s s 2 h 1+ − + 2h + − + (h + 1) . Γh + b−   − 2 h 1  2 h h +1 2 −  (˜µ2) The same considerations we applied for evaluating the coefficient of fk+2,n,p in b− s+1 Lemma 2.5 show that the expression in brackets equals s + 2 h . Apply- ing the functional equation of the gamma function once more, this yields the asserted value of as+1,h. This completes the proof of the proposition. Finally, we come to the Proof of Proposition 2.7. We begin by proving that for any q N, the action of q b− (l) ∈ q 2 +2l the operator ( 4πi) δ b− sends y ΘL,r,s,t(τ,Z) to − 1− 2 +r+t−2l,τ

q b− q b− l!Γ l + (l−q+h) 2 h 2 +2l−2q+2h 2 (4π ) y ΘL,r+h,s+h,t+h(τ,Z). h (l q + h)!Γ l q + h + b− hX=0 − − 2  For q = 0 the assertion is trivially true. We write the asserted function of y preceding the theta function in the term corresponding to h in the sum arising q 2 h from the index q as h (4π ) bl−q+h(y). Given that this assertion holds for q, we apply Equation (5)  for the operator 4πiδ b− acting on each − 1− 2 +r+t−2l+2q term, and observe that the resulting theta functions correspond to the index q +1 and to the summation indices h+1 and h. Moreover, after the usual index change manipulations one sees that the total coefficient in front of the theta function with indices q + 1 and h is q q b b (y) (4π2)h b (y)+ (l q +h) l q +h+ − 1 l−q+h−1 . h 1 l−q+h−1 h −  − 2 −  y2  − q As the second term here is easily seen to be just h−1 bl−q+h−1(y), the inductive assertion follows from the classical property of the binomial coefficients. With r = s = 0 and t = l = q = m the general formula from above becomes

b− 2 2 h m Γ m + m!(4π y ) b− (h) 2 y 2 Θ (τ,Z).  h  b−  h! L,h,h,m+h Xh Γ h + 2

 b− On the other hand, Putting m = s in Proposition 2.6, multiplying by y 2 (which commutes with differential operators in the variable Z), and taking the complex conjugate of the result, yields precisely the same expression. This proves the proposition.

28 References

[B] Borcherds, R. E., Automorphic Forms with Singularities on Grass- mannians, Invent. Math., vol. 132, 491–562 (1998). [Bru] Bruinier, J. H., Borcherds Products on O(2,l) and Chern Classes of Heegner Divisors, Lecture Notes in , vol 1780, Springer– Verlag (2002). [BF] Bruinier, J. H., Funke, J., On Two Geometric Theta Lifts, Duke Math J., vol 125 no. 1, 45–90 (2004).

[Ch] Choie, Y. J. Multilinear differential operators on Siegel mod- ular forms of genus 1 and 2, J. Math. Anal. Appl., vol. 232, 34–44 (1999). [CE] Choie, Y. J., Eholzer, W., Rankin–Cohen operators of Jacobi and Siegel forms, J. , vol. 68 No. 2, 160–177 (1998). [DN] Doi, K., Naganuma, H., On the Functional Equation of Certain Dirichlet Series, Invent. Math., vol 9, 1–14 (1969). [G] Gritsenko, V. A., Jacobi functions of n variables, Zap. Nauch. Sem- inars of LOMI, vol 168, 32–54 (1988) in Russian, English translation in J. Soviet Math., vol 53 ,243-252 (1991). [Ma1] Maaß, H., Die Differenzialgleichungen in der Theorie der Siegelschen Modulfunktionen, Math. Ann., vol 126, 44–68 (1953). [Ma2] Maaß, H., Siegel’s Modular Forms and Dirichlet Series, Lecture Notes in Mathematics, vol 216, Springer–Verlag (1971). [Na] Nakajima, S., On Invariant Differential Operators on Bounded Symmetric Domains of Type IV, Proc. Japan Acad., vol 58 Ser. A, 235– 238 (1982). [Ng] Naganuma, H., On the Coincidence of Two Dirichlet Series As- sociated with Cusp Forms of Hecke’s ”Neben”-Type and Hilbert Modular Forms over a Real Quadratic Field, J. Math. Soc. Jpn., vol 25, 547–555 (1973). [Sh1] Shimura, G., The Arithmetic of Certain Zeta Functions and Au- tomorphic Forms on Orthogonal Groups, Ann. of Math., vol 111, 313–375 (1980). [Sh2] Shimura, G., Arithmetic of Differential Operators on Symmet- ric Domains, Duke Math. J., vol 48 no. 4, 813–843 (1981). [Sh3] Shimura, G., Differential Operators and the Singular Values of , Duke Math. J., vol 51 no. 2, 261–329 (1984).

29 [Sh4] Shimura, G., On Differential Operators Attached to Certain Representations of Classical Groups, Invent. Math., vol 77, 463–488 (1984). [Sh5] Shimura, G., Invariant Differential Operators on Hermitian Symmetric Spaces, Ann. of Math., vol 132 no. 2, 237–272 (1990). [Ve] Verdier, J. L., Sur les integrales´ attachees´ aux formes automor- phes (d’apres` ), S´em. BOURBAKI 216, vol 13 (1961). [Za] Zagier, D. et al., The 1-2-3 of Modular Forms, Universitext (2008). [Ze1] Zemel, S., A p-adic Approach to the Weil Representation of Discriminant Forms Arising from Even Lattices, Math. Ann. Quebec, vol 39 issue 1, 61–89 (2015). [Ze2] Zemel, S., A Gross–Kohnen–Zagier Type Theorem for Higher- Codimensional Heegner Cycles, Research in Number Theory, vol 1 issue, 1–44 (2015).

Einstein Institute of Mathematics, the Hebrew University of Jerusalem, Edmund Safra Campus, Jerusalem 91904, Israel E-mail address: [email protected]

30