C Master
C Master
Springer
Berlin Heidelberg NewYork
Barcelona Hong Kong
London Milan Paris
Tokyo
To my parents
Table of Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
VIII Table of Contents
Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
Introduction
Borcherds’ first proof of this result in [Bo1] was rather indirect. There
he writes down the infinite product, shows that it can be meromorphi-
cally continued with the right poles and zeros using the Hardy-Ramanujan-
Rademacher asymptotics, and proves that it is automorphic by looking at
the transformation behavior under suitable generators of Γ (L). Later the
physicists Harvey and Moore discovered that Borcherds’ lifting can be un-
derstood in the framework of the theta correspondence, if one regularizes the
wildly divergent integrals in the right way [HM, Kon]. This idea was used by
Borcherds in [Bo2] to develop a more conceptual account to his lifting and to
extend it in many ways. We briefly indicate the idea of the latter approach:
The groups SL2 (R) and O(2, l) form a dual reductive pair in the sense
of Howe [Ho]. Thus it is possible to lift automorphic forms on one group to
automorphic forms on the other by integrating against a certain kernel func-
tion, the Siegel theta function ΘL (τ, Z) of the lattice L. This theta function
transforms as a non-holomorphic modular form in the variable τ ∈ H and is
Γ (L)-invariant in the second variable Z ∈ Hl . If f is a nearly holomorphic
modular form of weight 1 − l/2 as above, then its theta lifting is formally
given by Z
dx dy
Φ(Z) = f (τ )ΘL (τ, Z)y ,
y2
F
where F = {τ = x + iy ∈ H; |x| ≤ 1/2, |τ | ≥ 1} denotes the standard
fundamental domain for the action of SL2 (Z) on H. Unfortunately, since f
grows exponentially as =(τ ) → ∞, the integral diverges. However, it can be
regularized by taking the constant term in the Laurent expansion at s = 0 of
the meromorphic continuation in s of
Z
dx dy
lim f (τ )ΘL (τ, Z)y 1+s ,
u→∞ y2
Fu
We find that
Y p24 (1+q(λ))
Ψ (Z) = e (%, Z) 1 − e((λ, Z))
λ∈K
λ>0
is a holomorphic modular form of weight 12 for the group Γ (L). Its divisor is
equal to H(−1). Here we are in the special situation that 12 is the singular
weight for O(2, 26). Hence all Fourier coefficients a(λ) of Ψ must vanish unless
q(λ) = 0. The latter coefficients can be easily determined. One finds that
X
Ψ (Z) = det(w)∆ (w(%), Z) ,
w∈W
where W is the reflection group of the lattice II1,25 . If we compare the two
equalities, we get the Weyl denominator formula for the fake monster Lie
algebra (see [Bo1]).
In the present book we are mainly interested in geometric aspects of the
Borcherds lifting. The fact that it gives us explicit relations between Heegner
divisors in the divisor class group Cl(XL ) of XL makes it a very useful tool for
the study of the geometry of these divisors. In this context it is convenient to
work with a slightly modified divisor class group f Cl(XL ), which is defined as
the quotient of the group of divisors on XL modulo the subgroup of divisors
coming from meromorphic modular forms for Γ (L) of rational weight r with
some multiplier system.
By Theorem 0.1 we know that if m<0 c(m)q m is a Fourier polynomial,
P
which is the principal part of a nearly holomorphic Pmodular form of weight 1−
l/2, then the corresponding linear combination m<0 c(m)H(m) of Heegner
divisors is 0 in f
Cl(XL ). It is natural to ask, which Fourier polynomials occur
as principal parts of nearly holomorphic modular forms.
In the above example it is clear that every Fourier polynomial occurs as
the principal part of a nearly holomorphic modular form of weight −12. To
see this one can for instance apply the Hecke operator T (n) in weight −12 to
1/∆, to get a nearly holomorphic modular form with principal part q −n . More
generally, this argument shows that every Heegner divisor is trivial in f Cl(XL ),
if there exists a nearly holomorphic modular form of weight 1 − l/2 with
principal part q −1 . However, the situation gets more complicated, if there
are
P non-zero holomorphic cusp forms of complementary weight 1 + l/2. If g =
n n
P
n a(n)q is such a cusp form and f = n c(n)q any nearly holomorphic
modular form of weight 1 − l/2, then
f (τ )g(τ )dτ
where C[h] means the constant term of a Laurent series h ∈ C((q)). By the
residue theorem this quantity has to vanish. We get a necessary condition for
the existence of nearly holomorphic modular forms. In particular we see that
there is no such form with principal part q −1 , if there are non-zero cusp forms
of weight 1 + l/2. As an application of Serre duality on Riemann surfaces,
Borcherds proved (more generally for vector valued modular forms) that the
above residue condition is also sufficient (see [Bo3] Theorem 3.1):
In the easy special case of scalar valued modular forms for SL2 (Z) this
can be proved directly by induction: If k ≥ −12 and k 6= −10, there are
no non-zero cusp forms of weight 2 − k. The quotient of a modular form of
weight k + 12 with Fourier expansion 1 + O(q) and ∆ is a nearly holomorphic
modular form of weight k with principal part q −1 . For instance the above
Hecke operator argument shows that any principal part can be realized.
P let k = −10 or k < −12. Suppose that the Fourier polynomial
Now
p = n<0 c(n)q n satisfies the condition of the proposition. Then the principal
part of the Laurent series p∆ clearly satisfies the condition in weight k + 12.
By induction assumption there is a nearly holomorphic modular form f˜ of
weight k + 12 with the same principal part as p∆. We now show that the
constant terms of p∆ and f˜ also agree. If E is a modular form of weight
2 − k − 12 with E = 1 + O(q), then C[p∆ − f˜] = C[p∆E] − C[f˜E]. But the
first quantity on the right hand side vanishes by assumption and the second
by the residue theorem. Hence f˜ = p∆ + O(q), and f = f˜/∆ is a nearly
holomorphic modular form of weight k with principal part p.
The proposition in particular tells us that whenever there are cusp forms
of weight 1+l/2, then there are many linear combinations of Heegner divisors,
which are not the divisor of a Borcherds product. Here the natural question
arises, whether all relations between Heegner divisors in f
Cl(XL ) are given by
Borcherds products. In other words we ask:
Question 0.3. Let F be any meromorphic modular form for Γ (L) (with some
multiplier system), whose divisor is a linear combination of Heegner divisors
H(m). Is f then a Borcherds product?
κ = 1 + l/2 and denote by Sκ the space of cusp forms of weight κ for SL2 (Z).
We write Aκ (Z) for the Z-submodule of the dual space Sκ∗ of Sκ generated
by the functionals
X
ar : g 7→ b(r), for g = b(n)q n ∈ Sκ .
n≥1
The fact that Sκ has a basis of modular forms with coefficients in Z implies
that Aκ (Z) ⊗ C = Sκ∗ . Combining the Borcherds lifting with the proposition,
we get the following result.
η : Aκ (Z) −→ f
Cl(XL ).
1 X
Fm (τ, s) = [Ms (4π|m|y)e(mx)] |k M,
2Γ (2s)
M ∈Γ∞ \ SL2 (Z)
Chapter 4 is the technical heart of this book. We consider the O(2, l)-
invariant Laplace operator Ω acting on functions on the generalized upper
Introduction 9
(cf. [Sn]). Thus the assertion formally follows from the self-adjointness of
the Laplacian ∆k and the fact that Fm (τ, s) is an eigenfunction of ∆k . But
since the theta integral Φm (Z, s) does not converge absolutely, we have to
prove step by step that this argument can still be justified for the regularized
integral. As a consequence we find for the regularized function Φm (Z) that
ΩΦm (Z) = c
Then the function G(Z) = log(|F (Z)|q(Y )r/2 ) is Γ (L)-invariant and has
logarithmic singularities along Heegner divisors. It satisfies the differential
equation ΩG(Z) = −rl/8, similarly as the functions Φm (Z). We are now
ready to state the main result of section 4.3, which is a first step towards
answering Question 0.3 (see Theorem 4.23).
Hence, loosely speaking, any modular form, whose divisor is a linear combina-
tion of Heegner divisors, is given by the regularized theta lifting of a Maass
wave form with singularity at ∞. One might expect that this Maass form
actually has to be holomorphic on H. However, this seems far from being
obvious. We will come back to this problem in section 5.2 and Theorem 0.8.
10 Introduction
1 ¯ X
Ω(Z, τ ) = ∂ ∂ log q(Y ) + h−n (Z)q n .
4 n>0
To this end we first observe that the function Ω(Z, τ ) converges normally
and is square integrable in Z. Hence it is clear that Ω(Z, τ ) lies in H1,1 (XL )
for fixed τ . Using the fact that the Fourier coefficients of the Poincaré series
Fm (τ, 1 − k/2) are equal to certain coefficients of holomorphic Poincaré se-
ries in Sκ , we may rewrite Ω(Z, τ ) such that it becomes a sum of Poincaré,
Eisenstein, and theta series in the space of modular forms Mκ of weight κ
for SL2 (Z). Hence for fixed Z the function Ω(Z, τ ) is contained in Mκ (The-
orem 5.8). The idea of rewriting an automorphic kernel function in terms of
Poincaré series appears already in Zagier’s work on the Doi-Naganuma lifting
[Za]. It is related to the “Zagier identity” in [RS]. We prove:
Theorem 0.7. By taking the Petersson scalar product in the τ -variable the
function Ω(Z, τ ) defines a lifting
ϑ : Sκ −→ H1,1 (XL ),
f 7→ f (τ ), Ω(Z, τ ) ,
with the following properties (see Theorem 5.9 for more details):
1. Let f = n c(n)q n be a cusp form in Sκ . Then the image of f has the
P
Fourier expansion
X X
ϑ(f )(Z) = ϑ0 (f )(Y ) − C(κ) |λ|−l nl−1 c(−q(λ)/n2 )
λ∈K n|λ
q(λ)<0
¯ κ (π|λ||Y |, π(λ, Y )) e((λ, X)),
× ∂ ∂V
where C(κ) = 2−κ π 1/2−κ is a constant, and ϑ0 (f )(Y ) the 0-th Fourier
coefficient, which can also be computed explicitly. Moreover, Vκ (A, B)
denotes the generalized Whittaker function defined in (3.25).
2. The function Ω(Z, τ ) is an automorphic kernel function for ϑ̃, that is the
diagram
ϑ /
SO κ H1,1 (XL )
Sκ∗
ϑ̃ /H
e 1,1 (XL )
commutes. Here the left vertical arrow denotes the identification via Pe-
tersson scalar product.
Then F is a Borcherds product in the sense of Theorem 0.1, i.e. there exists
a
Pnearly holomorphic modular form f of weight 1 − l/2 with principal part
m
m<0 c(m)q , and the Borcherds lifting of f equals F .
We consider the special O(2, 3)-case of the paramodular group of level
t as an example. By Theorem 0.7 we get a lifting from skew-holomorphic
Jacobi forms of index t and weight 3 to square integrable harmonic (1, 1)-
forms on the corresponding quotient, the moduli space of Abelian surfaces
with a (1, t)-polarization.
Except for Theorem 0.8, we prove all the above results under the mild
assumptions that l > 2, and that L splits two orthogonal hyperbolic planes
over Q. This is always true, if l > 4. (We do not require that L is unimodular.)
Then the discriminant group, the quotient L0 /L of the dual lattice L0 modulo
L, is usually non-trivial. This leads to several technical complications. For
instance, in general we have to replace the space Sκ of elliptic cusp forms by
a certain space Sκ,L of C[L0 /L]-valued cusp forms of (possibly half-integral)
weight κ for the metaplectic group Mp2 (Z). In the same way we have to
consider vector valued nearly holomorphic modular forms and Maass wave
forms. Therefore in section 1.2 and 1.3 we collect some basic facts on such
modular forms and carry out the standard constructions of Poincaré and
Eisenstein series. Moreover, many objects which are indexed by an integer in
the above exposition (as H(m) or Φm ) have to be replaced by objects which
are indexed by a tuple (β, m), where β ∈ L0 /L and m ∈ Z + q(β) (as H(β, m)
or Φβ,m ). For the group Γ (L) we need to take the discriminant kernel of the
orthogonal group of L.
Theorem 0.8 is proved under the assumption that L splits two orthogonal
hyperbolic planes over Z. It seems likely that (at least for large l) this result
Introduction 13
can also be extended to a more general situation. However, this might require
additional considerations as some newform theory or Hecke theory for the
space Sκ,L and the lifting ϑ.
In the following chapter we briefly recall some facts about the metaplectic
cover of SL2 (R), the Weil representation, and certain vector valued modular
forms. We essentially follow the terminology of Borcherds [Bo2]. Moreover,
we consider vector valued Poincaré and Eisenstein series in some detail.
√
7→ ^
a b
a b = a b
c d c d c d , cτ + d (1.1)
Let Mp2 (Z) be the inverse image of SL2 (Z) under the covering map
Mp2 (R) → SL2 (R). It is well known that Mp2 (Z) is generated by
11
T = ,1 ,
01
0 −1 √
S= , τ .
1 0
is the standard generator of the center of Mp2 (Z). We will often use the
abbreviations Γ1 := SL2 (Z), Γ∞ = {( 10 n1 ) ; n ∈ Z} ≤ Γ1 , and
1n
Γ̃∞ := hT i = ,1 ; n ∈ Z .
01
for the dual lattice. Then the quotient L0 /L is a finite Abelian group, the so-
called discriminant group. The modulo 1 reduction of q(x) is a Q/Z-valued
quadratic form on L0 /L whose associated bilinear form is the modulo 1 re-
duction of the bilinear form (·, ·) on L0 .
Recall that there is a unitary representation %L of Mp2 (Z) on the group
algebra C[L0 /L]. If we denote the standard basis of C[L0 /L] by (eγ )γ∈L0 /L ,
then %L can be defined by the action of the generators S, T ∈ Mp2 (Z) as
follows (cp. [Bo2]):
For γ, δ ∈ L0 /L and (M, φ) ∈ Mp2 (Z) we define the coefficient %γδ (M, φ) of
the representation %L by
Shintani proved the following explicit formula for %L (cf. [Sn], Prop. 1.6).
if c = 0, and by
√ (b− −b+ ) sgn(c)
i X a(α + r, α + r) − 2(β, α + r) + d(β, β)
p e ,
|c|(b− +b+ )/2 |L0 /L| r∈L/cL 2c
(1.6)
if c 6= 0. Here, δ∗,∗ denotes the Kronecker-delta.
R1
with Fourier coefficients c(γ, n) = 0
fγ (τ )e(−nτ ) dx. Using the abbreviation
eγ (τ ) := eγ e(τ ) we may write
X X
f (τ ) = c(γ, n)eγ (nτ ). (1.9)
γ∈L0 /L n∈Z−q(γ)
By means of the scalar product in C[L0 /L] the coefficients can be expressed
by
Z1
c(γ, n) = hf (τ ), eγ (nτ̄ )i dx. (1.10)
0
Moreover, if all c(γ, n) with n = 0 vanish, then f is called a cusp form. The
C-vector space of modular forms of weight κ with respect to %∗L and Mp2 (Z)
is denoted by Mκ,L , the subspace of cusp forms by Sκ,L .
(recall that Γ̃∞ = hT i). Since κ ≥ 5/2, the usual argument shows that
this series converges normally on H and thereby defines a Mp2 (Z)-invariant
holomorphic function H → C[L0 /L]. We will often omit the superscript L, if
it is clear from the context.
Theorem 1.4. The Poincaré series Pβ,m has the Fourier expansion
X X
Pβ,m (τ ) = pβ,m (γ, n)eγ (nτ )
γ∈L0 /L n∈Z−q(γ)
n>0
with
e−πi sgn(c)κ/2
^
X ab ma + nd
Hc∗ (β, m, γ, n) = %βγ e ,
|c| ∗
cd c
d(c)
a b ∈Γ∞ \Γ1 /Γ∞
c d
(1.13)
and Jν the Bessel function of the first kind ([AbSt] Chap. 9). In particular
Pβ,m ∈ Sκ,L .
20 1 Vector valued modular forms for the metaplectic group
Z1 * +
1 X
pβ,m (γ, n) = eβ (mτ ) |∗κ (M, φ), eγ (nτ̄ ) dx.
2
0 (M,φ)∈Γ̃∞ \ Mp2 (Z)
Since eβ (mτ ) is invariant under the action of Z 2 , this can be written in the
form
Z1 Z1
heβ (mτ ), eγ (nτ̄ )i dx + heβ (mτ ) |∗κ Z, eγ (nτ̄ )i dx
0 0
Z1 * X
+
|∗κ ^
a b , e (nτ̄ )
+ eβ (mτ ) c d γ dx
0 a b ∈Γ∞ \Γ1
c d
c6=0
−1
^
ab ab
%∗L ^
a b a b
c d eβ (m c d τ ), eγ (nτ̄ ) = %βγ e m τ e(−nτ ),
cd cd
and thereby
1.2 Poincaré series and Eisenstein series 21
where C is a positive real constant. We now use i−κ sgn(c)κ = e−πi sgn(c)κ/2
and the definition of the generalized Kloosterman sum (1.13) and find
X
pβ,m (γ, n) = δm,n (δβ,γ + δ−β,γ ) + 2π |c|1−κ Hc∗ (β, m, γ, n)
c6=0
C+i∞
Z
1 −κ 2πm
× w exp − 2 + 2πnw dw. (1.16)
2πi c w
C−i∞
hf (τ ), g(τ )iy κ
is Mp2 (Z)-invariant. (As before, h·, ·i denotes the scalar product on C[L0 /L].)
If f or g is a cusp form, then hf (τ ), g(τ )iy κ is bounded on the standard
fundamental domain
F = {τ = x + iy ∈ H; |x| ≤ 1/2, |τ | ≥ 1}
It is easily seen that the integral does not depend on the choice of the fun-
damental domain. The space Sκ,L endowed with the scalar product (·, ·) is a
finite dimensional Hilbert space.
The Petersson coefficient formula for Sκ,L has the following form.
1.2 Poincaré series and Eisenstein series 23
Proof. Let Γ 1 = Γ1 /{±1}. We use the usual unfolding argument and find
Z * +
1 X dx dy
(f, Pβ,m ) = f (τ ), eβ (mτ ) |∗κ (M, φ) y κ 2
2 y
Γ 1 \H (M,φ)∈Γ̃∞ \ Mp2 (Z)
Z X dx dy
= hf (M τ ), eβ (mM τ )i=(M τ )κ
y2
M ∈Γ∞ \Γ1
Γ 1 \H
Z
=2 hf (τ ), eβ (mτ )iy κ−2 dx dy.
Γ∞ \H
Let β ∈ L0 /L with q(β) ∈ Z. Then the vector eβ = eβ (0) ∈ C[L0 /L], consid-
ered as a constant function H → C[L0 /L], is invariant under the |∗κ -action of
T, Z 2 ∈ Mp2 (Z). The Eisenstein series
1 X
EβL (τ ) = eβ |∗κ (M, φ) (1.18)
2
(M,φ)∈Γ̃∞ \ Mp2 (Z)
with
δ + δ−β,γ , if n = 0,
β,γ
qβ (γ, n) = (2π)κ nκ−1 X (1.19)
|c|1−κ Hc∗ (β, 0, γ, n), if n > 0.
Γ (κ)
c∈Z−{0}
Proof. We proceed in the same way as in the proof of Theorem 1.4. Let
γ ∈ L0 /L and n ∈ Z − q(γ). By (1.10) we have
Z1 * +
1 X
qβ (γ, n) = eβ |∗κ (M, φ), eγ (nτ̄ ) dx.
2
0 (M,φ)∈Γ̃∞ \ Mp2 (Z)
Z∞ Z∞
−κ −κ
(cτ + d) e(−nτ ) dx = |c| sgn(c) κ
(τ + d/c)−κ e(−nτ ) dx
−∞ −∞
Z∞
−κ nd
= |c| sgn(c) eκ
τ −κ e(−nτ ) dx.
c
−∞
Z∞
(cτ + d)−κ e(−nτ ) dx
−∞
C+i∞
Z
−κ −κ nd 1
= 2πi |c| κ
sgn(c) e w−κ e2πnw dw,
c 2πi
C−i∞
(In fact, one can deform the path of integration and essentially obtains the
Hankel integral for 1/Γ (κ).) Hence we get
(2π)κ nκ−1
^
X
−κ −κ κ ab nd
qβ (γ, n) = |c| i sgn(c) %βγ e .
Γ (κ) cd c
c6=0
a b ∈Γ∞ \Γ1 /Γ∞
c d
For computational purposes the formula (1.19) for the coefficients qβ (γ, n)
is not very useful. An explicit formula, involving special values of Dirichlet
L-series and finitely many representation numbers modulo prime powers at-
tached to the lattice L, is derived in [BK].
To give an example we briefly consider the case that L is unimodular. Then
L0 /L is trivial and Mκ,L is just the space of elliptic modular forms of weight
1 + l/2. We have one Eisenstein series E with Fourier coefficients q(n) and
q(0) = 2. For n > 0 the above expression for the coefficients can be simplified
as follows. First note that l ≡ 2 (mod 8). This implies e−πi sgn(c)κ/2 = −1
and
1 X nd 1 X
Hc∗ (β, 0, γ, n) = − e =− µ(|c|/d)d.
|c| ∗
c |c|
d(c) d|(c,n)
d>0
e( nd
P
Here we have used the evaluation of the Ramanujan sum d(c)∗ c ) by
means of the Moebius function µ. Thus we find
26 1 Vector valued modular forms for the metaplectic group
∞
(2π)κ nκ−1 X −κ X
q(n) = −2 c µ(c/d)d
Γ (κ) c=1 d|(c,n)
d>0
(2π)κ nκ−1 X 1−κ
= −2 d
Γ (κ)ζ(κ)
d|n
(2π)κ
= −2 σκ−1 (n),
Γ (κ)ζ(κ)
where σκ−1 (n) denotes the sum of the (κ−1)-th powers of the positive divisors
κ
of n and ζ(s) the Riemann zeta function. Using ζ(κ) = − (2πi)
2κ! Bκ with the
κ-th Bernoulli number Bκ , and the fact that κ ≡ 2 (mod 4), we get
4κ
q(n) = − σκ−1 (n).
Bκ
Thus in this case the Eisenstein series E(z) is the classical Eisenstein series
for Γ1 , normalized such that its constant term equals 2.
Let us now turn back to the general case and determine the number of
linearly independent Eisenstein series Eβ . The |∗κ -invariance of Eβ under Z
and the identity eβ |∗κ Z = e−β imply
1 X
Eβ = Eβ |∗κ Z = eβ |∗κ Z |∗κ (M, φ) = E−β .
2
(M,φ)∈Γ̃∞ \ Mp2 (Z)
Eis
Thus the space Mκ,L spanned by all Eisenstein series Eβ equals the space
spanned by the Eα where α runs through a set of representatives of {β ∈
L0 /L; q(β) ∈ Z} modulo the action of {±1}. Comparing constant terms one
finds that the set of Eisenstein series Eα with α as above is already linearly
independent. We obtain
Eis
dim(Mκ,L ) = #{β ∈ L0 /L; q(β) ∈ Z, 2β = 0 + L}
1
+ #{β ∈ L0 /L; q(β) ∈ Z, 2β 6= 0 + L}. (1.20)
2
Now let f ∈ Mκ,L and write the Fourier expansion of f in the form
X X
f (τ ) = c(γ, n)eγ (nτ ).
γ∈L0 /L n∈Z−q(γ)
The invariance of f under Z and (1.4) imply that c(γ, n) = c(−γ, n). Hence
1 X
f− c(β, 0)Eβ
2 0
β∈L /L
q(β)∈Z
1.3 Non-holomorphic Poincaré series of negative weight 27
Eis
lies in Sκ,L . This shows that Mκ,L = Mκ,L ⊕ Sκ,L and (1.20) is a formula for
the codimension of Sκ,L in Mκ,L .
Let Mν, µ (z) and Wν, µ (z) be the usual Whittaker functions as defined
in [AbSt] Chap. 13 p. 190 or [E1] Vol. I Chap. 6 p. 264. They are linearly
independent solutions of the Whittaker differential equation
d2 w µ2 − 1/4
1 ν
+ − + − w = 0. (1.21)
dz 2 4 z z2
The functions Mν, µ (z) and Mν, −µ (z) are related by the identity
Γ (−2µ) Γ (2µ)
Wν, µ (z) = Mν, µ (z) + Mν, −µ (z) (1.22)
Γ ( 12 − µ − ν) Γ ( 12 + µ − ν)
([AbSt] p. 190 (13.1.34)). This implies in particular Wν, µ (z) = Wν, −µ (z). As
z → 0 one has the asymptotic behavior
(<(µ − ν) > −1/2, <(z) > 0) of the W -Whittaker function ([E1] Vol. I p. 274
(18)), we find for y < 0:
If we insert the definition of the incomplete Gamma function (cf. [AbSt] p. 81)
Z∞
Γ (a, x) = e−t ta−1 dt, (1.32)
x
we obtain for y < 0 the identity
Ms (4π|m|y)eβ (mx)
is invariant under the |k -operation of T ∈ Mp2 (Z) (as defined in (1.7)) and
an eigenfunction of ∆k with eigenvalue
s(1 − s) + (k 2 − 2k)/4.
L
Definition 1.8. With the above notation we define the Poincaré series Fβ,m
of index (β, m) by
L 1 X
Fβ,m (τ, s) = [Ms (4π|m|y)eβ (mx)] |k (M, φ),
2Γ (2s)
(M,φ)∈Γ̃∞ \ Mp2 (Z)
(1.35)
where τ = x + iy ∈ H and s = σ + it ∈ C with σ > 1. If it is clear from the
context, to which lattice L the series (1.35) refers, we simply write Fβ,m (τ, s).
By (1.23) and (1.27) the above series (1.35) has the local majorant
X
=(M τ )σ .
(M,φ)∈Γ̃∞ \ Mp2 (Z)
for any smooth function f : H → C[L0 /L] and any (M, φ) ∈ Mp2 (Z). We may
infer that Fβ,m is an eigenfunction of ∆k :
Theorem 1.9. The Poincaré series Fβ,m has the Fourier expansion
Γ (1 + k/2 − s)
Fβ,m (τ, s) = M1−s (4π|m|y)(eβ (mx) + e−β (mx))
Γ (2 − 2s)Γ (s + k/2)
X
+ b(γ, 0, s)y 1−s−k/2 eγ
γ∈L0 /L
q(γ)∈Z
X X
+ b(γ, n, s)Ws (4πny)eγ (nx) + Feβ,m (τ, s),
γ∈L0 /L n∈Z+q(γ)
n>0
e−πi sgn(c)k/2
^
X ab ma + nd
Hc (β, m, γ, n) = %−1
γβ e
|c| ∗
cd c
d(c)
a b ∈Γ∞ \Γ1 /Γ∞
c d
(1.38)
and Jν (z), Iν (z) the usual Bessel functions as defined in [AbSt] Chap. 9.
The sum in (1.38) runs over all primitive residues d modulo c and ac db is a
representative for the double coset in Γ∞ \Γ1 /Γ∞ with lower row (c d0 ) and
d0 ≡ d (mod c). Observe that the expression %−1 ^ a b e ma+nd does not
γβ c d c
depend on the choice of the coset representative. The fact that Hc (β, m, γ, n)
is universally bounded implies that the series for the coefficients b(γ, n, s)
converge normally in s for σ > 1.
1.3 Non-holomorphic Poincaré series of negative weight 31
Proof of Theorem 1.9. We split the sum in equation (1.35) into the sum over
1, Z, Z 2 , Z 3 ∈ Γ̃∞ \ Mp2 (Z) and the sum over (M, φ) ∈ Γ̃∞ \ Mp2 (Z), where
M = ac db with c 6= 0. The latter sum will be denoted by H(τ, s). Since
eβ (mx) |k Z = e−β (mx), we find that the first term equals
1
Ms (4π|m|y)(eβ (mx) + e−β (mx)).
Γ (2s)
Γ (1 + k/2 − s)
M1−s (4π|m|y)(eβ (mx) + e−β (mx))
Γ (2 − 2s)Γ (s + k/2)
Γ (1 + k/2 − s)
+ Ws (4πmy)(eβ (mx) + e−β (mx)). (1.39)
Γ (2s)Γ (1 − 2s)
We now calculate the Fourier expansion of H(τ, s). Since eβ (mx) is in-
variant under the action of Z 2 , we may write H(τ, s) in the form
^
1 X ab
[Ms (4π|m|y)eβ (mx)] |k .
Γ (2s) cd
a b ∈Γ∞ \Γ1
c d
c6=0
Let γ ∈ L0 /L and n ∈ Z + q(γ) and denote the (γ, n)-th Fourier coefficient of
H(τ, s) by c(γ, n, y) (cp. (1.36)). Then we have
Z1
c(γ, n, y) = hH(τ, s), eγ (nx)i dx
0
Z∞
X (cτ + d)−k
y
= Ms 4π|m| |cτ +d| 2
Γ (2s)
c6=0 −∞
a b ∈Γ \Γ /Γ
∞ 1 ∞
c d
× %−1 ^
a b e m< a b
L c d β c d τ , eγ (nx) dx.
and therefore
32 1 Vector valued modular forms for the metaplectic group
^
1 X ab ma
c(γ, n, y) = %−1
γβ e
Γ (2s) cd c
c6=0
a b ∈Γ \Γ /Γ
∞ 1 ∞
c d
Z∞
y m
× (cτ + d)−k Ms 4π|m| e −< − nx dx.
|cτ + d|2 2
c (τ + d/c)
−∞
√ √ p
If we use cτ + d = sgn(c) c τ + d/c and substitute x by x − d/c in the
integral, we obtain
^
1 X ab ma + nd
c(γ, n, y) = %−1
γβ e |c|−k sgn(c)k
Γ (2s) cd c
c6=0
a b ∈Γ∞ \Γ1 /Γ∞
c d
Z∞
−k y −mx
× τ Ms 4π|m| 2 2 e 2 2 − nx dx
c |τ | c |τ |
−∞
1 X 1−k k
= |c| i Hc (β, m, γ, n)
Γ (2s)
c6=0
Z∞
−k y −mx
× τ Ms 4π|m| 2 2 e 2 2 − nx dx.
c |τ | c |τ |
−∞
(4π|m|y)−k/2 X
|c|Hc (β, m, γ, n)
Γ (2s)
c6=0
Z∞ −k/2
τ y −mx
× M−k/2, s−1/2 4π|m| 2 2 e 2 2 − nx dx
−τ̄ c |τ | c |τ |
−∞
Z∞ −k/2
(4π|m|y)−k/2 X y − ix
= |c|Hc (β, m, γ, n)
Γ (2s) y + ix
c6=0 −∞
4π|m|y −mx
× M−k/2, s−1/2 e 2 2 − nx dx. (1.40)
c2 (x2 + y 2 ) c (x + y 2 )
Z∞ −k/2
1 − iu 4πB 2πiBu
I=y M−k/2, s−1/2 exp + 2πiAu du.
1 + iu u2 + 1 u2 + 1
−∞
The integral I/(yΓ (2s)) was evaluated by Hejhal in [He] p. 357. (Notice that
M−k/2, s−1/2 (y)/Γ (2s) is the same as Hejhal’s function G(y) with ` = −k/2.
Moreover, Wν, µ (y) equals e−y/2 y µ+1/2 Ψ (µ + 1/2 − ν, 1 + 2µ, y) in Hejhal’s
notation.) We find
p
2π |B/A| p
W−k/2, s−1/2 (4π|A|)J2s−1 4π |AB| , if A > 0,
Γ (s − k/2)
4π 1+s
I
= Bs, if A = 0,
yΓ (2s) (2s − 1)Γ (s + k/2)Γ (s − k/2)
p
2π |B/A|
p
Wk/2, s−1/2 (4π|A|)I2s−1 4π |AB| , if A < 0.
Γ (s + k/2)
Since ∆k Fβ,m (τ, 1 − k/2) = 0, the Poincaré series Fβ,m (τ, s) are in par-
ticular interesting in the special case s = 1 − k/2.
34 1 Vector valued modular forms for the metaplectic group
Proposition 1.10. For s = 1−k/2 the Fourier expansion of Fβ,m (τ, s) given
in Theorem 1.9 can be simplified as follows:
X X
Fβ,m (τ, 1 − k/2) = eβ (mτ ) + e−β (mτ ) + b(γ, n, 1 − k/2)eγ (nτ )
γ∈L0 /L n∈Z+q(γ)
n≥0
Proof. By virtue of (1.30) and (1.31) this immediately follows from Theorem
1.9. t
u
Definition 1.11. A function f : H → C[L0 /L] is called a nearly holomorphic
modular form of weight k (with respect to %L and Mp2 (Z)), if
i) f |k (M, φ) = f for all (M, φ) ∈ Mp2 (Z),
ii) f is holomorphic on H,
iii) f has a pole in ∞, i.e. f has a Fourier expansion of the form
X X
f (τ ) = c(γ, n)eγ (nτ ).
γ∈L0 /L n∈Z+q(γ)
n−∞
!
The space of these nearly holomorphic modular forms is denoted by Mk,L .
The Fourier polynomial
X X
c(γ, n)eγ (nτ )
γ∈L0 /L n∈Z+q(γ)
n<0
The invariance of f under Z ∈ Mp2 (Z) implies that c(γ, n) = c(−γ, n).
According to Proposition 1.10 we have Fγ,n (τ, 1 − k/2) = eγ (nτ ) + e−γ (nτ ) +
O(1) for all γ, n. Hence g(τ ) is bounded as y → ∞. But g also satisfies
%βγ ^ a −b
^ %−1 ^ −1 ^
a −b
a b =%
a b
c d βγ −c d and γβ c d = %γβ −c d .
Now the assertion immediately follows from the definition of Hc (β, m, γ, n).
t
u
Remark 1.14. This implies that the Fourier coefficients b(γ, n, 1 − k/2) of the
function Fβ,m (τ, 1 − k/2) are real numbers.
We now compare the Fourier coefficients of Fβ,m (τ, 1 − k/2) with the
coefficients of the holomorphic Poincaré series Pγ,n ∈ Sκ,L constructed in the
previous section.
Proof. We use the fact that (M̃ )−1 = M]−1 for any M ∈ Γ , and that κ =
1
2 − k. We find
e−πi sgn(c)k/2
^
X ab ma + nd
Hc (β, m, γ, n) = %−1
γβ c d e
|c| ∗
c
d(c)
a b ∈Γ∞ \Γ1 /Γ∞
c d
e−πi sgn(−c)(2−k)/2
^
X db md + na
=− %−1
γβ e
|c| ∗
ca c
d(c)
a b ∈Γ∞ \Γ1 /Γ∞
c d
e−πi sgn(−c)κ/2
^
X a −b −na − md
=− %γβ e
|c| ∗
−c d −c
d(c)
a b ∈Γ∞ \Γ1 /Γ∞
c d
∗
= −H−c (γ, −n, β, −m).
t
u
Proposition 1.16. Let β, γ ∈ L0 /L, n ∈ Z+q(γ), m ∈ Z+q(β), and m < 0.
Denote the (n, γ)-th Fourier coefficient b(γ, n, 1 − k/2) of Fβ,m (τ, 1 − k/2)
(see Proposition 1.10) by bβ,m (γ, n). If n < 0, then
1
bβ,m (γ, n) = − pγ,−n (β, −m),
Γ (1 − k)
where pγ,−n (β, −m) is the (β, −m)-th coefficient of the Poincaré series
Pγ,−n (τ ) ∈ Sκ,L (see 1.12). If n = 0, then
where qγ (β, −m) is the (β, −m)-th coefficient of the Eisenstein series Eγ (τ ) ∈
Mκ,L (see 1.19).
Proof. We compare the formulas for the Fourier coefficients given in Propo-
sition 1.10 and Theorem 1.4 (resp. Proposition 1.10 and Theorem 1.6) and
apply Lemma 1.15. Note that k − 1 = 1 − κ. t
u
For γ ∈ L0 /L and n ∈ Z − q(γ) with n > 0 let aγ,n : Sκ,L → C denote the
∗
functional in the dual space Sκ,L of Sκ,L which maps a cusp form f to its
(γ, n)-th Fourier coefficient aγ,n (f ). According to Proposition 1.5, aγ,n can
be described by means of the Petersson scalar product as
(4πn)κ−1
aγ,n = ( · , Pγ,n ). (1.41)
2Γ (κ − 1)
The following theorem can also be proved using Serre duality (see [Bo3]
Thm. 3.1). We deduce it from Proposition 1.16.
1.3 Non-holomorphic Poincaré series of negative weight 37
!
Theorem 1.17. There exists a nearly holomorphic modular form f ∈ Mk,L
with prescribed principal part
X X
c(β, m)eβ (mτ )
β∈L0 /L m∈Z+q(β)
m<0
∗
equals 0 in Sκ,L .
Proof. First, suppose that there exists such a nearly holomorphic modular
form f . Then by Proposition 1.12
1 X X
f (τ ) = c(β, m)Fβ,m (τ, 1 − k/2),
2 0
β∈L /L m∈Z+q(γ)
m<0
and in particular
∂ X X
c(β, m)Feβ,m (τ, 1 − k/2) = 0.
∂ τ̄
β∈L0 /L m∈Z+q(γ)
m<0
Since
∂
W1−k/2 (4πny)eγ (nx) 6= 0,
∂ τ̄
this implies that
X X
c(β, m)bβ,m (γ, n) = 0
β∈L0 /L m∈Z+q(γ)
m<0
for all γ ∈ L0 /L and n ∈ Z + q(γ) with n < 0. (Here we have used the same
notation as in Proposition 1.16.) Applying Proposition 1.16 we find that
X X X X
c(β, m)pγ,−n (β, −m) = c(β, m)aβ,−m (Pγ,−n )
β∈L0 /L m∈Z+q(γ) β∈L0 /L m∈Z+q(γ)
m<0 m<0
=0
for all γ ∈ L0 /L and n ∈ Z + q(γ) with n < 0. Since the Poincaré series Pγ,−n
generate the space Sκ,L we obtain
X X
c(β, m)aβ,−m = 0. (1.42)
β∈L0 /L m∈Z+q(β)
m<0
38 1 Vector valued modular forms for the metaplectic group
Now assume that (1.42) holds. Then we may reverse the above argument
to infer that
X X
c(β, m)Feβ,m (τ, 1 − k/2) = 0.
β∈L0 /L m∈Z+q(γ)
m<0
and X
ΘL (τ, v; r, t) = eγ θγ (τ, v; r, t). (2.2)
γ∈L0 /L
If r and t are both 0 we will omit them and simply write θγ (τ, v) resp. ΘL (τ, v).
Let Λ be the lattice Z2 equipped with the standard symplectic form and
W = Λ ⊗ R (such that Sp(Λ) = SL2 (Z)). Then W ⊗ V is a symplectic vec-
tor space. The usual theta function attached to the lattice Λ ⊗ L ⊂ W ⊗ V
restricted to the symmetric subspace H × Gr(L) ⊂ Sp(W ⊗ V )/K (where
K denotes a maximal compact subgroup) equals the Siegel theta function
θL (τ, v).
40 2 The regularized theta lift
O(L) = {g ∈ O(V ); gL = L} .
Over Q we have
L ⊗ Q = (K ⊗ Q) ⊕ (Qz ⊕ Qz 0 ). (2.4)
Hence K has signature (b+ − 1, b− − 1). If n ∈ V = L ⊗ R then we write nK
for the orthogonal projection of n to K ⊗ R. It can be easily checked that nK
can be computed as
2.1 Siegel theta functions 41
If one uses the fact that |L0 /L| is given by the absolute value of the
Gram determinant of L, then the above proposition in particular shows that
|L0 /L| = N 2 |K 0 /K|.
Moreover, we may infer that γ 7→ γ − (γ, ζ)z/N defines an isometric
embedding K 0 → L0 . The kernel of the induced map K 0 → L0 /L equals
{γ ∈ K; (γ, ζ) ∈ N Z}.
Consider the sub-lattice
V = v ⊕ v ⊥ = w ⊕ Rzv ⊕ w⊥ ⊕ Rzv⊥ .
in V ∩ z ⊥ = (K ⊗ R) ⊕ Rz.
Lemma 2.3. (cp. [Bo2] Lemma 5.1.) Let γ ∈ L0 /L. We have
1 X X
θγ (τ, v) = p e (τ q(λw ) + τ̄ q(λw⊥ ))
2yzv2 λ∈γ+K⊕Zζ d∈Z
Then X X
θγ (τ, v) = g(λ, v; d).
λ∈γ+K⊕Zζ d∈Z
We may apply the Poisson summation formula to rewrite the inner sum. We
find X X
θγ (τ, v) = ĝ(λ, v; d),
λ∈γ+K⊕Zζ d∈Z
where ĝ(λ, v; d) denotes the partial Fourier transform with respect to the
variable d. We now compute ĝ(λ, v; d). Using q(zv ) + q(zv⊥ ) = 0 we get
= e(Ad2 + Bd + C),
Using q(λv ) = q(λw ) + (λ, zv )2 /2zv2 and q(λv⊥ ) = q(λw⊥ ) + (λ, zv⊥ )2 /2zv2⊥
we obtain
τ (λ, zv )2 − τ̄ (λ, zv⊥ )2
1
ĝ(λ, v; d) = p e + τ q(λ w ) + τ̄ q(λ w ⊥ )
2yzv2 2zv2
2
−d − 2d(τ (λ, zv ) + τ̄ (λ, zv⊥ )) − (τ (λ, zv ) + τ̄ (λ, zv⊥ ))2
×e .
2(τ − τ̄ )zv2
(2.9)
Since
which is equal to
|cτ + d|2
1 X
0 0
θL+γ (τ, v) = p e − − d(γ, z ) + cdq(z )
2yzv2 4iyzv2
c,d∈Z
c≡(γ,z) (N )
Remark 2.5. Regarding the definition (2.1) of ΘK we should write more cor-
rectly µK instead of µ (note that µK = µ − (µ, z 0 )z). Since µw = (µK )w = zw
0
λ = λK + cz 0 + f (c, γ)z
44 2 The regularized theta lift
1 X X
p e (τ q((λK + cz 0 )w ) + τ̄ q((λK + cz 0 )w⊥ ))
2yzv2 c,d∈Z λK ∈K+p(γ−cz 0 )
c≡(γ,z) (N )
dc 0
z , zv /2zv2 + zv⊥ /2zv2⊥
−
2
dc
zv /2zv2 + zv⊥ /2zv2⊥ , zv /2zv2 + zv⊥ /2zv2⊥ .
− (2.11)
2
Since q(z) = 0, the third term on the right hand side of (2.11) is 0. The
second term can be written as
2.1 Siegel theta functions 45
dc
−d λK + c(z 0 − zv /2zv2 − zv⊥ /2zv2⊥ )/2, −z 0 − (z 0 , z 0 ).
2
Thus the left hand side of (2.10) equals
t
u
with FL |k (M, φ) = FL for all (M, φ) ∈ Mp2 (Z). Let r, t be given integers.
We define a function
X
FK (τ ; r, t) = eγ fγ (τ ; r, t), (2.12)
γ∈K 0 /K
by putting
X
fγ (τ ; r, t) = e (−r(λ, z 0 ) − rtq(z 0 )) fL+λ+tz0 (τ )
λ∈L00 /L
p(λ)=γ
for γ ∈ K 0 /K. (This is just a different notation for the definition given in
[Bo2] on page 512.) Let γ ∈ K 0 . Then as a set of representatives for λ ∈ L00 /L
with p(λ) = γ +K we may take γ −(γ, ζ)z/N +bz/N where b runs modulo N .
A set of representatives for L00 /L is given by λ = γ − (γ, ζ)z/N + bz/N , where
γ runs through a set of representatives for K 0 /K and b runs modulo N . (Since
the index of L00 in L0 equals N , we find again that |L0 /L| = N 2 |K 0 /K|.)
Theorem 2.6. We use the same notation as above. Let (M, φ) ∈ Mp2 (Z),
M = ac db . Then FK (τ ; r, t) satisfies
K
Proof. i). Assume that σ ≥ 1 − k/2. From the definition of Fβ,m (τ, s; 0, 0),
L
the Fourier expansion of Fβ,m (τ, s) (see Theorem 1.9), and the asymptotic
property (1.26) of the Whittaker function Wν,µ (y) it follows that
K
Fβ,m (τ, s; 0, 0) = O(1)
K
for y → ∞. Since Fβ,m (τ, s; 0, 0) also satisfies
K K
Fβ,m (τ, s; 0, 0) |k (M, φ) = Fβ,m (τ, s; 0, 0), (M, φ) ∈ Mp2 (Z),
K 2
K
∆k Fβ,m (τ, s; 0, 0) = s(1 − s) + (k − 2k)/4 Fβ,m (τ, s; 0, 0),
we may infer that it vanishes identically (cf. [He] Chap. 9 Prop. 5.14c). Be-
K
cause Fβ,m (τ, s; 0, 0) is holomorphic in s, we find that it vanishes for all s.
ii) In the same way as in (i) it follows that
K
Fβ,m (τ, s; 0, 0)
Γ (1 + k/2 − s)
= M1−s (4π|m|y)(ep(β) (mx) + e−p(β) (mx)) + O(1)
Γ (2 − 2s)Γ (s + k/2)
for y → ∞, and thereby
K K
Fβ,m (τ, s; 0, 0) − Fp(β),m (τ, s) = O(1) (y → ∞).
σ0 = max{1, b+ /2 − k/2}.
where Fβ,m (τ, s) denotes the Maass-Poincaré series of index (β, m) defined in
(1.35) and ΘL (τ, v) the Siegel theta function (2.2). If it is clear from the con-
text, to which lattice L the function (2.13) refers, we simply write Φβ,m (v, s).
According to Theorem 2.1, the integrand
+
hFβ,m (τ, s), ΘL (τ, v)iy b /2
Fu = {τ = x + iy ∈ F; y ≤ u} (u ∈ R>0 ). (2.14)
Then for s = σ + it ∈ C with σ > σ0 we may define the regularized theta lift
of Fβ,m (τ, s) by
Z
+ dx dy
ΦL
β,m (v, s) = lim L
hFβ,m (τ, s), ΘL (τ, v)iy b /2 . (2.15)
u→∞ y2
Fu
Roughly speaking this means that we first carry out the integration over
x and afterwards the integration over y. We will show that Φβ,m (v, s) as
defined in (2.15) is holomorphic in s for σ > σ0 . It can be continued to a
holomorphic function on {s ∈ C; σ > 1, s 6= b+ /2 − k/2} with a simple
pole at s = b+ /2 − k/2. Hence for σ > 1 and s 6= b+ /2 − k/2 the function
Φβ,m (v, s) is given by holomorphic continuation. For s = b+ /2 − k/2 we may
define the regularized theta integral to be the constant term of the Laurent
expansion in s of Φβ,m (v, s) at the point s = b+ /2 − k/2.
This construction is very similar to the definition of the regularized theta
lift due to Harvey, Moore, and Borcherds [HM, Bo2]. Following Borcherds, one
could also define Φβ,m (v, s) to be the constant term of the Laurent expansion
in s0 of Z
+ 0 dx dy
lim hFβ,m (τ, s), ΘL (τ, v)iy b /2−s (2.16)
u→∞ y2
Fu
0 0
at s = 0. Here s is an additional complex variable. The above expression
makes sense if <(s0 ) 0. Note that the integrand in (2.16) is only invariant
under Mp2 (Z) if s0 = 0. We will prove in Proposition 2.11 that (2.15) and
(2.16) essentially yield to the same Φβ,m (v, s). However, for our purposes the
definition of the regularized theta lift given in (2.15) is more natural.
As in [Bo2] we will show that Φβ,m (v, s) is a real analytic function on
Gr(L) with singularities along smaller sub-Grassmannians. To make this more
precise, we define a subset H(β, m) of the Grassmannian Gr(L) by
[
H(β, m) = λ⊥ . (2.17)
λ∈β+L
q(λ)=m
Here λ⊥ means the orthogonal complement of λ in Gr(L), i.e. the set of all
positive definite b+ -dimensional subspaces v ⊂ V with v ⊥ λ.1 It is known
that for any subset U with compact closure U ⊂ Gr(L) the set
{s ∈ C; σ > 1, s 6= b+ /2 − k/2},
and has a pole of first order with residue b(0, 0, b+ /2 − k/2) at s = b+ /2 − k/2
(supposed that b+ /2 − k/2 > 1).
Proof. Recall that Fβ,m (τ, s) is holomorphic in s for σ > 1. The integral
Z
+ dx dy
hFβ,m (τ, s), ΘL (τ, v)iy b /2
y2
F1
over the compact set F1 is clearly holomorphic in s for σ > 1. Thus it suffices
to consider the function
Z∞ Z1
+
ϕ(v, s) = hFβ,m (τ, s), ΘL (τ, v)iy b /2−2
dx dy. (2.19)
y=1 x=0
According to Theorem 1.9 the Fourier expansion of Fβ,m (τ, s) can be written
more explicitly in the form
2.2 The theta integral 49
1
Fβ,m (τ, s) = Ms (4π|m|y)(eβ (mx) + e−β (mx)) + H(τ, s),
Γ (2s)
where
X X
H(τ, s) = b̃(γ, 0, s)y 1−s−k/2 + b̃(γ, n, s)Ws (4πny)eγ (nx),
γ∈L0 /L γ∈L0 /L
q(γ)∈Z n∈Z+q(γ)
n6=0
Γ (1 + k/2 − s)
b̃(γ, n, s) = b(γ, n, s) − δm,n (δβ,γ + δ−β,γ ).
Γ (2s)Γ (1 − 2s)
Observe that all Fourier coefficients b̃(γ, n, s) are holomorphic in s for σ >
1. (This is not true for b(±β, m, s). That is why we have introduced the
b̃(γ, n, s).) We find that ϕ(v, s) is given by
Z∞
+
b(0, 0, s) yb /2−k/2−s−1
dy
1
Z∞ X
+
+ b(λ, 0, s) exp(−4πyq(λv ))y b /2−k/2−s−1
dy
0
1 λ∈L −0
q(λ)=0
Z∞ X
+
+ b̃(λ, q(λ), s)Ws (4πq(λ)y) exp(2πyq(λv⊥ ) − 2πyq(λv ))y b /2−2
dy
0
1 λ∈L
q(λ)6=0
Z∞ X
2 +
+ Ms (4π|m|y) exp(2πym − 4πyq(λv ))y b /2−2
dy. (2.20)
Γ (2s)
1 λ∈β+L
q(λ)=m
The first integral obviously converges for σ > b+ /2 − k/2 and equals
1
.
s − b+ /2 + k/2
Thus the first summand in (2.20) has a holomorphic continuation to {s ∈
C; σ > 1, s 6= b+ /2 − k/2}. If b+ /2 − k/2 > 1 it has a simple pole at
s = b+ /2 − k/2 with residue b(0, 0, b+ /2 − k/2).
Therefore it suffices to show that the remaining terms in (2.20) converge
locally uniformly absolutely for σ > 1.
For the convergence of the second term, notice that q(λ) = 0 implies
q(λv ) = 21 q(λv ) − 12 q(λv⊥ ). The latter quadratatic form is a multiple of the
(positive definite) majorant associated with v. Thus the sum over λ is domi-
nated by a subseries of a theta series attached to a definite lattice.
To prove the convergence of the third term we also use the majorant
associated with v, and in addition the asymptotic property (1.26) of the
Whittaker function Wν,µ (y) and the estimates
50 2 The regularized theta lift
p
b(γ, n, s) = O exp(4π |mn| ) , n → +∞, (2.21)
b(γ, n, s) = O |n|σ+k/2−1 , n → −∞ (2.22)
(locally uniformly in s for σ > 1). These bounds can be easily obtained from
the asymptotic behavior of the Iν - resp. Jν -Bessel function and the fact that
Hc (β, m, γ, n) is universally bounded.
If b+ = 0, then L is a definite lattice, and λv = 0 for all λ. The convergence
of the last term in (2.20) follows from the asymptotic property (1.25) of the
M -Whittaker function.
To prove the convergence of the last term for b+ = 1, 2 we again use the
asymptotic property (1.25) and the following fact: For any C ≥ 0 and any
compact subset B ⊂ Gr(L) the set
is finite. By assumption v ∈ / H(β, m). Hence there exists an ε > 0 such that
q(λv ) > ε for all λ ∈ L0 with q(λ) = m. It follows that
X +
Ms (4π|m|y) exp(−4πyq(λv ) + 2πym)y b /2−2
λ∈β+L
q(λ)=m
X
exp(−4πyq(λv ))
λ∈β+L
q(λ)=m
X
e−2πεy e−π(m+q(λv )−q(λv⊥ ))
λ∈L0
q(λ)=m
uniformly for y ∈ [1, ∞) and locally uniformly in s. This implies the desired
convergence statement. t
u
If we compare this with the expression for ϕ(v, s) given in Proposition 2.8, we
immediately obtain the assertion in the case s 6= b+ /2 − k/2. For s = b+ /2 −
k/2 we find that the difference of our regularization Cs=b+ /2−k/2 [Φβ,m (v, s)]
and Borcherds’ regularization is given by
b(0, 0, s) +
h
0 −1
i
Cs=b+ /2−k/2 − b(0, 0, b /2 − k/2)Cs 0 =0 s
s − b+ /2 + k/2
= b0 (0, 0, b+ /2 − k/2).
t
u
+
For the rest of this section we assume that b = 1, 2. In Proposition 2.8
we considered Φβ,m (v, s) for a fixed v and varying s. Now we show that for
fixed s the function Φβ,m (v, s) is real analytic in v on Gr(L) − H(β, m).
Let U ⊂ Gr(L) be an open subset and f , g functions on a dense open
subset of U . Then we write
f ≈ g,
if f − g can be continued to a real analytic function on U . In this case we
will say that f has a singularity of type g.
Theorem 2.12. i) For fixed s ∈ {s ∈ C; σ > 1, s 6= b+ /2 − k/2} the
function Φβ,m (v, s) is real analytic in v on Gr(L) − H(β, m).
ii) Let U ⊂ Gr(L) be an open subset with compact closure U ⊂ Gr(L).
The function Φβ,m (v) is real analytic on Gr(L) − H(β, m). On U it has a
singularity of type X
−2 log q(λv ),
λ∈S(β,m,U )
+
if b = 2, and of type √ X
−4 2π |λv |,
λ∈S(β,m,U )
if b+ = 1.
52 2 The regularized theta lift
Proof. We use the same argument as in [Bo2] Theorem 6.2. We only give a
proof for (ii), because (i) can be treated similarly.
Since ΘL (τ, v) is real analytic and Fβ,m (τ, s) is holomorphic in s, the
integral Z
+ dx dy
hFβ,m (τ, s), ΘL (τ, v)iy b /2
y2
F1
+
+ b(λ, 0, s) exp(−4πyq(λv ))y b /2−k/2−s−1
dy
b /2 − k/2 − s 0
1 λ∈L −0
q(λ)=0
Z∞ X
+
+ b(λ, q(λ), s)Ws (4πq(λ)y) exp(−2πyq(λv ) + 2πyq(λv⊥ ))y b /2−2
dy
0
1 λ∈L
q(λ)6=0
Z∞ X
2Γ (1 + k/2 − s)
+ M1−s (4π|m|y)
Γ (2 − 2s)Γ (s + k/2)
1 λ∈β+L
q(λ)=m
+
× exp(−4πyq(λv ) + 2πym)y b /2−2
dy. (2.23)
The first quantity does not depend on v. The second and the third term are
holomorphic in s near s = 1 − k/2 and real analytic in v ∈ Gr(L). Hence they
do not contribute to the singularity. The last term is holomorphic in s near
s = 1 − k/2 and real analytic for v ∈ Gr(L) − H(β, m). We find that
In the last line we have used Mk/2 (y) = ey/2 . The integrals in the last line
only become singular if q(λv ) = 0. Thus for v ∈ U we get
2.3 Unfolding against Fβ,m 53
Z∞
X +
Φβ,m (v) ≈ 2 exp(−4πyq(λv ))y b /2−2
dy
λ∈S(β,m,U ) 1
X +
≈2 (4πq(λv ))1−b /2
Γ (b+ /2 − 1, 4πq(λv )),
λ∈S(β,m,U )
Γ (a + 1, x) = aΓ (a, x) + xa e−x
In particular, from (2.24) and (2.25) we can also read off the singularities of
Γ (0, x) and Γ (−1/2, x).
uniformly in x.
Proof. This can be proved in the same way as the analogous statement for
holomorphic elliptic modular forms. u
t
where (a)n = Γ (a + n)/Γ (a) (cf. [AbSt] Chap. 15 or [E1] Vol. I Chap. 2).
The circle of convergence of the series (2.26) is the unit circle |z| = 1. Using
the estimate ([E1] Vol. I p. 57 (5))
one finds that (2.26) converges absolutely for |z| = 1, if <(c − a − b) > 0.
Then the value for z = 1 is given by ([E1] Vol. I p. 104 (46))
Γ (c)Γ (c − a − b)
F (a, b, c; 1) = (c 6= 0, −1, −2, . . . ). (2.27)
Γ (c − a)Γ (c − b)
The series converges normally for v ∈ Gr(L)−H(β, m) and σ > b− /4+b+ /4.
In particular, if (b+ , b− ) = (0, l − 2), then
8π|m|
Φβ,m (v, s) = · #{λ ∈ β + L; q(λ) = m}.
(s + l/4 − 3/2)Γ (s + 3/2 − l/4)
The second integral on the right hand side of (2.28) can be evaluated by the
usual unfolding trick. It equals
Z
2 + dx dy
Ms (4π|m|y)y b /2 e(mx)θβ (τ, v) 2 , (2.29)
Γ (2s) y
G
where
G = {τ ∈ H; |x| ≤ 1/2, |τ | ≤ 1} (2.30)
S
is a fundamental domain for the action of Γ∞ on H − M ∈Γ∞ M F. By virtue
of (1.23) and Lemma 2.13, the integral in (2.29) converges absolutely if σ >
1 + b− /4 + b+ /4, and therefore the unfolding is justified. Combining (2.29)
with the first integral on the right hand side of (2.28), we obtain
Z∞ Z1
2 +
Φβ,m (v, s) = Ms (4π|m|y)y b /2−2
e(mx)θβ (τ, v) dx dy.
Γ (2s)
y=0 x=0
Z∞ X
2 +
/2−2 −4πyq(λv )+2πym
Φβ,m (v, s) = Ms (4π|m|y)y b e dy
Γ (2s)
y=0 λ∈β+L
q(λ)=m
∞
2(4π|m|)−k/2 X Z −
/4+b+ /4−2
= M−k/2, s−1/2 (4π|m|y)y b
Γ (2s)
λ∈β+L 0
q(λ)=m
(4π|m|)s Γ (b− /4 + b+ /4 + s − 1)
(4π|q(λv⊥ )|)b− /4+b+ /4+s−1
× F b− /4 + b+ /4 + s − 1, s − b− /4 + b+ /4, 2s; m/q(λv⊥ ) . (2.32)
The coefficients c(γ, n; y, s) were calculated in section 1.3. For instance, for
γ ∈ L0 /L with q(γ) ∈ Z we found:
2 s−b+ /4−b− /4
1 2 2zv
√ K
Φβ,m (w, s) + √ Γ (s + 1/2 − b+ /4 − b− /4)
2|zv | π π
X X + −
× b(`z/N, 0, s) e(`n/N )nb /2+b /2−1−2s
` (N ) n≥1
√
2 X X X
+ e(n(λ, µ)) e(n(δ, z 0 ))
|zv |
λ∈K 0 −0 n≥1 δ∈L00 /L
p(δ)=λ+K
Z∞
−πn2
b+ /2−5/2
× c(δ, q(λ); y, s)y exp − 2πyq(λw ) + 2πyq(λw⊥ ) dy.
2zv2 y
0
(2.35)
Here ΦK
β,m (w, s) is defined by
(
0, if (β, z) 6≡ 0 (mod N ),
ΦK
β,m (w, s) =
ΦK
p(β),m (w, s), if (β, z) ≡ 0 (mod N ).
|cτ + d|2
Z
1 X X
0 0
fγ (τ, s) e − + d(γ, z ) − cdq(z )
4iyzv2
p
2zv2 0 c,d∈Z
F γ∈L /L
c≡(γ,z) (N )
+ dx dy
× θK+p(γ−cz0 ) (τ, w; dµ, −cµ)y (b −1)/2
y2
Z
1 X + dx dy
=p fγ (τ, s)θK+p(γ) (τ, w)y (b −1)/2
2
2zv y2
γ∈L0 /L
F
(γ,z)≡0 (N )
Z X
1 0 X
+p e (d(γ, z 0 ) − cdq(z 0 )) fγ (τ, s)
2zv2 γ∈L0 /L
F c,d∈Z
(γ,z)≡c (N )
|cτ + d|2
+
−1)/2 dx dy
×e − θK+p(γ−cz0 ) (τ, w; dµ, −cµ)y (b . (2.36)
4iyzv2 y2
X0
Here means that the sum runs over all pairs (c, d) ∈ Z2 with (c, d) 6=
c,d∈Z
K
(0, 0). If we substitute in the definition (2.12) of Fβ,m (τ, s; 0, 0), we obtain for
the first term on the right hand side of (2.36):
Z
1 + dx dy
K
Fβ,m (τ, s; 0, 0), ΘK (τ, w) y (b −1)/2 .
y2
p
2zv2
F
|cτ + d|2
Z X
1 0
K
e − Fβ,m (τ, s; −d, c), ΘK (τ, w; dµ, −cµ)
4iyzv2
p
2zv2
F c,d∈Z
+ dx dy
−1)/2
× y (b
y2
1
Z X X n2 |cτ + d|2
=p e −
2zv2 4iyzv2
F (c,d)=1 n≥1
K +
−1)/2 dx dy
× Fβ,m (τ, s; −nd, nc), ΘK (τ, w; ndµ, −ncµ) y (b . (2.37)
y2
By Theorem 2.1 and Theorem 2.6 we get
2.4 Unfolding against ΘL 59
−πn2
Z
1 X X
p exp 2
2zv2 2zv =(M τ )
F M ∈Γ∞ \Γ1 n≥1
K +
−1)/2 dx dy
× Fβ,m (M τ, s; −n, 0), ΘK (M τ, w; nµ, 0) =(M τ )(b .
y2
As in the proof of Theorem 2.14 we split this up into
√ Z
−πn2
2 X + dx dy
exp K
Fβ,m (τ, s; −n, 0), ΘK (τ, w; nµ, 0) y (b −1)/2 2
|zv | 2zv2 y y
F n≥1
√ Z
−πn2
2 X X
+ exp
|zv | 2zv2 =(M τ )
F M = a b ∈Γ∞ \Γ 1 n≥1
c d
c6=0
K +
−1)/2 dx dy
× Fβ,m (M τ, s; −n, 0), ΘK (M τ, w; nµ, 0) =(M τ )(b , (2.38)
y2
where Γ 1 = Γ/{±1}. The first integral on the right hand side of (2.38) is to
be understood in the regularized sense. As in Lemma 2.13 it can be seen that
K
Fβ,m (τ, s; −n, 0) = O(e2π|m|/y ) (2.39)
for y → 0. This implies that the second integral converges absolutely if zv2 <
1
4|m| . By the usual unfolding argument it can be written in the form
√ Z
−πn2
2 X + dx dy
exp 2
K
Fβ,m (τ, s; −n, 0), ΘK (τ, w; nµ, 0) y (b −1)/2 ,
|zv | 2zv y y2
G n≥1
(2.40)
with G as in (2.30). By (2.39) the latter integral also converges absolutely for
1
zv2 < 4|m| . We find that (2.38) equals
√ Z∞ Z1
−πn2
2 X b+ 5
exp 2
K
Fβ,m (τ, s; −n, 0), ΘK (τ, w; nµ, 0) y 2 − 2 dx dy.
|zv | 2zv y
0 0 n≥1
The identity
62 3 The Fourier expansion of the theta lift
(3.2)
Recall that the projection p is only defined on L00 /L.
1
Proposition 3.1. Let v ∈ Gr(L) − H(β, m) with zv2 < 4|m| . Then
1 √ X
ΦL
β,m (v) = √ ΦK
β,m + 4 2π|zv | b(`z/N, 0)B2 (`/N )
2|zv | ` (N )
√ X
+ 4 2π|zv | B2 ((λ, µ) + (β, z 0 ))
λ∈p(β)+K
q(λ)=m
√ X X
+ 4 2(π/|zv |)−k b(δ, q(λ))|λ|1−k
λ∈K 0 −0 δ∈L00 /L
p(δ)=λ+K
X
× n−k−1 e(n(λ, µ) + n(δ, z 0 ))K1−k (2πn|λ|/|zv |), (3.3)
n≥1
where ΦK K
β,m denotes the constant Φβ,m (w, 1 − k/2), and b(γ, n) = b(γ, n, 1 −
L
k/2) denote the Fourier coefficients of the Poincaré series Fβ,m (τ, 1 − k/2).
The third term on the right hand side is to be interpreted as 0 if (β, z) 6≡
0 (mod N ). As usual Kν is the modified Bessel function of the third kind
(cf. [AbSt] Chap. 9).
Proof. We apply Theorem 2.15. Since L is Lorentzian, the first and the second
term on the right hand side of (2.35) are holomorphic in s near 1 − k/2. We
find
3.1 Lorentzian lattices 63
√
1 2 2|zv | X X
ΦL
β,m (v) = √ ΦK
β,m + b(`z/N, 0) e(`n/N )n−2
2|zv | π
` (N ) n≥1
√
2 X X X
+ e(n(λ, µ)) e(n(δ, z 0 ))
|zv | 0 0
λ∈K −0 n≥1 δ∈L0 /L
p(δ)=λ+K
Z∞
−πn2
−2
× c(δ, q(λ); y, 1 − k/2)y exp + 2πyq(λ) dy. (3.4)
2zv2 y
0
2z 2
The first integral in (3.5) equals πnv2 . If we replace the sum over n ≥ 1 by a
sum over n 6= 0 and substitute in identity (3.1) for B2 (x), we find
1 √ X
ΦL
β,m (v) = √ ΦK
β,m + 4 2π|zv | b(`z/N, 0)B2 (`/N )
2|zv | ` (N )
√ X
+ 4 2π|zv | B2 ((λ, µ) + (β, z 0 ))
λ∈p(β)+K
q(λ)=m
√
2 X X X
+ e(n(λ, µ)) e(n(δ, z 0 ))
|zv |
λ∈K 0 −0 n≥1 δ∈L00 /L
p(δ)=λ+K
Z∞
−πn2
× b(δ, q(λ)) Γ (1 − k, 4π|q(λ)|y) exp y −2 dy. (3.6)
2zv2 y
0
64 3 The Fourier expansion of the theta lift
V1 −→ Gr(L), v1 7→ Rv1 .
zv = (z, v1 )v1 ,
|zv | = (z, v1 ).
After a short calculation one finds that v1 can be expressed in terms of (z, v1 )
and the projection µK = µ − (µ, z 0 )z ∈ K ⊗ R by
1 √ X
ΦL
β,m (v1 ) = √ ΦK
β,m + 4 2π(z, v1 ) b(`z/N, 0)B2 (`/N )
2(z, v1 ) ` (N )
√
X (λ, v1 )
+ 4 2π(z, v1 ) B2 + (β, z 0 )
(z, v1 )
λ∈p(β)+K
q(λ)=m
√ X X
+ 4 2(π/(z, v1 ))−k b(δ, q(λ))|λ|1−k
λ∈K 0 −0 δ∈L00 /L
p(δ)=λ+K
X (λ, v1 )
× n−k−1 e n + n(δ, z 0 ) K1−k (2πn|λ|/(z, v1 )).
(z, v1 )
n≥1
(3.11)
L L
Definition 3.3. Using the above notation we define two functions ξβ,m , ψβ,m
on Gr(L) by
√
ΦK
β,m 1 4 2π X
L
ξβ,m (v1 ) = √ − 2(z 0 , v1 ) + (λ, v1 )2
2 (z, v1 ) (z, v1 )
λ∈p(β)+K
q(λ)=m
√ X X
+ 4 2(π/(z, v1 ))−k b(δ, q(λ))|λ|1−k
λ∈K 0 −0 δ∈L00 /L
p(δ)=λ+K
X (λ, v1 )
× n−k−1 e n + n(δ, z 0 ) K1−k (2πn|λ|/(z, v1 )) (3.12)
(z, v1 )
n≥1
and
L
ψβ,m (v1 ) = ΦL L
β,m (v1 ) − ξβ,m (v1 ).
Note that this definition depends on the choice of the vectors z and z 0 . It
L L
is easily checked that ξβ,m = ξ−β,m .
By Theorem 2.14, Φβ,m is a real valued function and ΦK
L
β,m a real constant.
According to Lemma 1.13 the coefficients b(δ, q(λ)) are real numbers. Hence
L L
ξβ,m and ψβ,m are real-valued functions.
Using the boundedness of the coefficients b(δ, n) with n < 0 (cf. (2.22))
and the asymptotic property
r
π −y
Kν (y) ∼ e (y → ∞) (3.13)
2y
L
of the K-Bessel function, one can show that ξβ,m (v1 ) is real analytic on the
whole Gr(L).
L
Theorem 3.4. The function ψβ,m is the restriction of a continuous piece-
wise linear function on V . Its only singularities lie on H(β, m). For v1 ∈ V1
1
with (z, v1 )2 < 4|m| it equals
3.1 Lorentzian lattices 67
√ √ X
L
ψβ,m (v1 ) = 2(z 0 , v1 )ΦK
β,m + 4 2π(z, v1 ) b(`z/N, 0)B2 (`/N )
` (N )
√ (λ, v1 )2
X (λ, v1 ) 0
+ 4 2π(z, v1 ) B2 + (β, z ) − .
(z, v1 ) (z, v1 )2
λ∈p(β)+K
q(λ)=m
(3.14)
Proof. Equality (3.14) immediately follows from (3.11) and the definition of
L
ψβ,m (v1 ). Using the fact that
one finds that the quadratic terms in the sum on the right hand side of (3.14)
L 1
cancel out. Hence ψβ,m (v1 ) is piecewise linear on {v1 ∈ V1 ; (z, v1 )2 < 4|m| }.
L L
Since ξβ,m (v1 ) is real analytic on Gr(L), the only singularities of ψβ,m (v1 )
L
lie on H(β, m) by Theorem 2.12. Moreover, ψβ,m (v1 ) is continuous on Gr(L).
Let W1 and W2 be two different Weyl chambers of index (β, m) such that
the interior of W1 ∪W2 is connected. Then there is a λ ∈ β +L with q(λ) = m
and W1 ∩ W2 ⊂ λ⊥ . If Φ1 resp. Φ2 is a real analytic function on W1 ∪ W2 with
Φ1 |W1 = ΦL L
β,m |W1 resp. Φ2 |W2 = Φβ,m |W2 then Theorem 2.12 implies that
Definition 3.5. Let W be a Weyl chamber of index (β, m). Then we write
%β,m (W ) for the unique vector in V with the property
L
√
ψβ,m (v1 ) = 8 2π(v1 , %β,m (W ))
L
be a finite linear combination of the ξβ,m . If ξ L (v1 ) is a rational function on
V1 , then it vanishes identically.
68 3 The Fourier expansion of the theta lift
Proof of Theorem 3.6. We use the following notation: Let f and g be functions
on V1 . Then we write
.
f = g,
if f − g is the restriction of a piecewise linear function on V .
.
We first prove that ξ L (v1 ) = 0. Let
X
ΦL (v1 ) = cβ,m ΦL
β,m (v1 ).
β,m
.
By Theorem 3.4 we have ΦL (v1 ) = ξ L (v1 ). Hence it suffices to show that
.
ΦL (v1 ) = 0.
The assumption that ξ L is rational implies that
" #
K
. 1 X Φβ,m 8π X
ΦL (v1 ) = √ cβ,m − 2(z 0 , v1 )ΦK
β,m + (λ, v1 )2 .
2 (z, v1 ) (z, v1 )
β,m λ∈p(β)+K
q(λ)=m
(3.15)
This formula for ΦL (v1 ) depends on the choice of a primitive norm 0 vector
z ∈ L. We compute ΦL (v1 ) using a different norm 0 vector and compare the
result with (3.15).
02
The vector z 0 − z2 z ∈ L ⊗ Q has norm 0. Let a be the unique positive
integer such that
02
z̃ = a z 0 − z2 z
We now compare (3.15) and (3.17) in the upper half space model H. If v1 is
represented by (x0 , µ) as in (3.9), we have
a(x20 − µ2 )
(z̃, v1 ) = ,
2x0
µ2 1
= − 2(z 0 , v1 ). (3.18)
x0 (z, v1 )
We find
1 X X
cβ,m µ2 ΦK
β,m + 8π (λ, µ)2
x0 0
β,m λ∈K
. 2x0 X 8π X
= cβ,m ΦK̃
β,m + 2 (λK , µ)2 .
a(x20 − µ2 ) x0 0
β,m λ∈K̃
.
and thereby ξ L (v1 ) = 0.
In the upper half space model we have
1 X X
ξ L ((x0 , µ)) = √ cβ,m µ2 ΦK
β,m + 8π (λ, µ)2 . (3.19)
2x0 β,m λ∈K 0
.
Comparing coefficients in (3.19) and (3.10), we may infer that ξ L ((x0 , µ)) = 0
L
already implies ξ ((x0 , µ)) = 0. t
u
70 3 The Fourier expansion of the theta lift
µ2 µ2
∆L = (1 − l) + 2(l − 2)x0 . (3.21)
x0 x0
Regarding the second term we find
√ √
4 2π X
2 4 2π X
∆L (λ, µ) = (1 − l) (λ, µ)2
x0 x0
λ∈p(β)+K λ∈p(β)+K
q(λ)=m q(λ)=m
√
− 16 2π|m|x0 · #{λ ∈ p(β) + K; q(λ) = m}. (3.22)
2 X X X1
+√ b(δ, q(λ)) e(n(δ, z 0 ) + n(λ, µ))
π 0
n
λ∈K δ∈L00 /L n≥1
q(λ)<0 p(δ)=λ+K
Here b(γ, n) = b(γ, n, 1 − k/2) denote the Fourier coefficients of the Poincaré
L
series Fβ,m (τ, 1 − k/2) as in Proposition 1.10, and b0 (0, 0) means the deriva-
tive of b(0, 0, s) at s = 1 − k/2. The sum over λ in the second line is to be
interpreted as in (3.2).
1
Proof. We first assume zv2 < 4|m| and apply Theorem 2.15. The only term
on the right hand side of (2.35) that is not holomorphic in s near 1 − k/2 is
the second one
2 s−1/2−l/4 X X
h(s) = √ 2zv2 /π Γ (s − l/4) b(`z/N, 0, s) e(`n/N )nl/2−2s .
π
` (N ) n≥1
where ζ(s) denotes the Riemann zeta function. The remaining terms with
` 6≡ 0 (mod N ) converge for σ > l/4 and are therefore holomorphic. Their
value at s = 1 − k/2 = 1/2 + l/4 is
X
−2 b(`z/N, 0) log |1 − e(`/N )|. (3.28)
` (N )
`6≡0 (N )
(For the first expansion for instance see [E1] p. 34 (17); note that −Γ 0 (1) =
γ = Euler-Mascheroni constant.) We may infer that it is equal to
Γ 0 (1/2)
2
2zv
b(0, 0) log + √ − 2Γ 0 (1) + b0 (0, 0).
π π
√
From the duplication formula Γ (ν)Γ (ν√+ 1/2) = 21−2ν πΓ (2ν) for the
gamma function it follows that Γ 0 (1/2)/ π = Γ 0 (1) − 2 log 2. Thus the above
expression for the constant term in the Laurent expansion of (3.27) can be
simplified to
1
√ ΦK 2
β,m (w, 1 − k/2) + b(0, 0) log(zv ) + Cβ,m
2|zv |
√
2 X X X
+ e(n(λ, µ)) e(n(δ, z 0 ))
|zv | 0 0
λ∈K −0 n≥1 δ∈L0 /L
p(δ)=λ+K
Z∞
−πn2
× c(δ, q(λ); y, 1 − k/2) exp − 4πyq(λ w ) + 2πyq(λ) y −3/2 dy.
2zv2 y
0
(3.30)
We now insert the explicit formulas for the Fourier coefficients c(γ, q(λ); y, 1−
k/2) given in Proposition 1.10. We saw that c(γ, q(λ); y, 1 − k/2) equals
b(γ, q(λ))e−2πq(λ)y ,
if q(λ) ≥ 0,
b(γ, q(λ))W1−k/2 (4πq(λ)y), if q(λ) < 0, q(λ) 6= m,
(δβ,γ + δ−β,γ )e−2πmy + b(γ, q(λ))W1− k (4πq(λ)y), if q(λ) = m.
2
For instance by considering the limit A → 0 it can be shown that this formula
still holds for A = 0. We may conclude that
Z∞ √
−πn2
−3/2 2|zv | −2πn|λw |/|zv |
exp − 4πyq(λw ) y dy = e .
2zv2 y n
0
2zv2
we substitute u = πn2 y and find that it is equal to
√
2|z |
√ v V2−k (πn|λ|/|zv |, πn|λw |/|zv |) .
πn
We may rewrite (3.31) in the form
X X X1
2 b(δ, q(λ)) e(n(λ, µ) + n(δ, z 0 ))e−2πn|λw |/|zv |
0 0
n
λ∈K −0 δ∈L0 /L n≥1
q(λ)≥0 p(δ)=λ+K
X X1
+2 e(n(λ, µ) + n(±β, z 0 ))e−2πn|λw |/|zv |
n
λ∈±p(β)+K n≥1
q(λ)=m
2 X X X1
+√ b(δ, q(λ)) e(n(λ, µ) + n(δ, z 0 ))
π 0
n
λ∈K δ∈L00 /L n≥1
q(λ)<0 p(δ)=λ+K
In the zero-quadric
V = (K ⊗ R) ⊕ Rz 0 ⊕ Rz.
{Z = X + iY ∈ K ⊗ C; Y 2 > 0} (3.35)
The vector
zv z ⊥ zv z
µ = −z 0 + 2
+ v2 = −z 0 + 2 − 2
2zv 2zv⊥ zv 2zv
is given by
µ = (X, 0, −q(X) − q(z 0 ))
and its orthogonal projection µK by µK = µ − (µ, z 0 )z = X. The subspace
w ⊂ V , i.e. the orthogonal complement of zv in v, is equal to RYL . Recall
from section 2.1 that we have identified w with its orthogonal projection
wK ⊂ K ⊗ R. Obviously wK = RY , w1 = Y /|Y |, and thereby
λw = (λ, Y )Y /Y 2 ,
|λw | = |(λ, Y )||zv |.
The intersection
Γ (L) = O+ (V ) ∩ Od (L) (3.37)
of the discriminant kernel Od (L) with O (V ) is a discrete subgroup of O+ (V ).
+
L
Definition 3.11. We define a function ξβ,m : Hl → R by
L |Y | K 2 X X
ξβ,m (Z) = √ ξβ,m (Y /|Y |)−b(0, 0) log(Y 2 )+ √ b(δ, q(λ))
2 π
λ∈K 0 δ∈L00 /L
q(λ)<0 p(δ)=λ+K
X1
× e(n(δ, z 0 ) + n(λ, X))V2−k (πn|λ||Y |, πn(λ, Y ))
n
n≥1
L
and a function ψβ,m : Hl − H(β, m) → R by
L
ψβ,m (Z) = ΦL L
β,m (Z) − ξβ,m (Z).
K
Here ξβ,m is given by
(
K
0, if (β, z) 6≡ 0 (mod N ),
ξβ,m =
K
ξp(β),m , if (β, z) ≡ 0 (mod N ) (see Def. 3.3).
∂2
f =0
∂zi ∂ z̄j
for all 1 ≤ i, j ≤ l.
converge normally. The convergence of the second series for q(Y ) > |m| fol-
lows from the asymptotic behavior (2.21) of the coefficients b(γ, n).
It is easily seen that the first series converges normally for Z ∈ W . Now
suppose that U is an arbitrary open subset of Hl with compact closure U ⊂
Hl . Then there are only finitely many λ ∈ K 0 with q(λ) = m and λ + K =
±p(β) such that (λ, W ) > 0 and (λ, Y ) ≤ 0 for some Z = X + iY ∈ U .
Using this fact, one immediately obtains the convergence of the first series
on U . t
u
By Lemma 3.15 the function Ψβ,m (Z) is holomorphic on {Z ∈ Hl ; q(Y ) >
|m|} and satisfies
1 L
log |Ψβ,m (Z)| = − (ψβ,m (Z) − Cβ,m ) (3.40)
4
on the complement of H(β, m).
Theorem 3.16. The function Ψβ,m (Z) has a holomorphic continuation to
Hl , and (3.40) holds on Hl − H(β, m). Let U ⊂ Hl be an open subset with
compact closure U ⊂ Hl and denote by S(β, m, U ) the finite set
3.2 Lattices of signature (2, l) 81
Then Y
Ψβ,m (Z) (λ, ZL )−1 (3.41)
λ∈S(β,m,U )
On the non-empty intersection of the open sets U and {Z ∈ Hl ; q(Y ) > |m|}
we therefore have
Y 1
< Log Ψβ,m (Z) (λ, ZL )−1 − Cβ,m /4 = − <(f (Z)).
4
λ∈S(β,m,U )
But now − 41 f can only differ by an additive constant from the expression in
brackets on the left-hand side. We may assume that this constant equals zero
and find Y
Ψβ,m (Z) (λ, ZL )−1 = eCβ,m /4 e−f (Z)/4 .
λ∈S(β,m,U )
e + = {W ∈ V (C) − {0};
K [W ] ∈ K+ } ⊂ V (C) (3.43)
[W ]7→[σW ]
KO + / K+ (3.45)
O
Z7→[ZL ]
Hl
Z7→σZ / Hl
The function
j(σ, Z) = (σ(ZL ), z)
on O+ (V ) × Hl is an automorphy factor for O+ (V ), i.e. it satisfies the cocycle
relation
j(σ1 σ2 , Z) = j(σ1 , σ2 Z)j(σ2 , Z)
and does not vanish on Hl .
Let r be a rational number. If σ ∈ O+ (V ) and Z ∈ Hl , we define
where Log j(σ, Z) denotes a fixed holomorphic logarithm of j(σ, Z). There
exists a map wr from O+ (V ) × O+ (V ) to the set of roots of unity (of order
bounded by the denominator of r) such that
χ : Γ → S 1 = {t ∈ C; |t| = 1}
satisfying
q(Y )
q(=(σZ)) = .
|j(σ, Z)|2
Proof. We use the formula q(Y ) = 41 (ZL , Z L ) and (3.46). One has
1 σ(ZL ) σ(Z L )
q(=(σZ)) = ,
4 (σZL , z) (σZ L , z)
1
= 2
σ(ZL ), σ(Z L )
4|(σZL , z)|
q(Y )
= .
|j(σ, Z)|2
t
u
Lemma 3.21. (cp. [Bo2] Lemma 13.1.) Let r ∈ Q and Γ ≤ Γ (L) be sub-
group of finite index. Suppose that Ψ is a meromorphic function on Hl for
which |Ψ (Z)|q(Y )r/2 is invariant under Γ . Then there exists a multiplier sys-
tem χ of weight r for Γ such that Ψ is a meromorphic modular form of weight
r and multiplier system χ with respect to Γ .
3.4 Borcherds products 85
Ψ (σZ)
j(σ, Z)−r = χ(σ).
Ψ (Z)
is invariant under Γ (L). However, taking suitable finite products of the Ψβ,m
L
one can attain that the main parts of the ξβ,m cancel out. Thereby one finds
a new explanation of [Bo2] Theorem 13.3 and [Bo3] from a cohomological
point of view.
Let f : H → C[L0 /L] be a nearly holomorphic modular form for Mp2 (Z)
of weight k = 1 − l/2 with principal part
X X
c(γ, n)eγ (nτ ).
γ∈L0 /L n∈Z+q(γ)
n<0
the Weyl chambers of Gr(K) with respect to f . Let W be such a Weyl cham-
ber. For every γ ∈ L00 /L, n ∈ Z + q(γ) with n < 0 and c(γ, n) 6= 0 there is a
Weyl chamber Wγ,n of Gr(K) of index (p(γ), n) (as in section 3.1) such that
W ⊂ Wγ,n . Then \ \
W = Wγ,n .
γ∈L00 /L n∈Z+q(γ)
n<0
c(γ,n)6=0
86 3 The Fourier expansion of the theta lift
1 X X
%f (W ) = c(γ, n)%p(γ),n (Wγ,n ).
2
γ∈L00 /L n∈Z+q(γ)
n<0
Proof. Throughout the proof we write bβ,m (γ, n) instead of b(γ, n, 1 − k/2)
for the (γ, n)-th Fourier coefficient of the Poincaré series Fβ,m (τ, 1 − k/2) to
emphasize the dependence on (β, m). Note that
1 X X
f (τ ) = c(β, m)Fβ,m (τ, 1 − k/2)
2 0
β∈L /L m∈Z+q(γ)
m<0
In particular, by Lemma 1.13 the coefficients c(γ, n) are real numbers. Define
two functions
1 X X
Φ(Z) = c(β, m)ΦLβ,m (Z),
2 0 β∈L /L m∈Z+q(γ)
m<0
1 X X
L
ξ(Z) = c(β, m)ξβ,m (Z).
2 0β∈L /L m∈Z+q(γ)
m<0
1 X |Y | X
ξ(Z) = − log(Y 2 ) c(β, m)bβ,m (0, 0) + √ K
c(β, m)ξβ,m (Y /|Y |).
2 2 2
β,m β,m
K
If we apply the same argument to the Fourier expansion of the ξβ,m , we find
K
P
that the function β,m c(β, m)ξβ,m is rational. Hence, according to Theorem
3.6, it vanishes identically. Using (3.49) we finally obtain
and hereby
1 X X
Φ(Z) = −c(0, 0) log(Y 2 ) + L
c(β, m)ψβ,m (Z). (3.50)
2 0
β∈L /L m∈Z+q(γ)
m<0
Since Φ(Z) is invariant under Γ (L), the function |Ψ (Z)|q(Y )c(0,0)/4 is invari-
ant, too. Using Lemma 3.21 we may infer that Ψ (Z) is a modular form of
weight c(0,0)
2 with some multiplier system χ for Γ (L).
ii) This immediately follows from Theorem 3.16.
iii) It is a consequence of Theorem 3.16 and equation (3.49) that
1 X X
− c(β, m)q(β, −m)
2 0
β∈L /L m∈Z+q(γ)
m<0
3.4.1 Examples
Let H be the lattice Z2 with the quadratic form q((a, b)) = ab. Any lattice
isomorphic to H is called a hyperbolic plane.
!
in M0,L . Modular forms for the orthogonal group Γ (L) can be identified with
modular forms on H×H for the group SL2 (Z)×SL2 (Z). The Heegner divisors
H(m) = 21 H(0, m) are just the Hecke correspondences T (|m|) on H × H. In
particular H(−1) is given by the translates of the diagonal {(τ, τ ); τ ∈ H}.
In this case Theorem 3.22 (when extended to l = 2) tells us that there exists
a (on H × H) holomorphic modular form Ψ for SL2 (Z) × SL2 (Z) of weight 0
90 3 The Fourier expansion of the theta lift
with divisor T (1). It is easily checked that Ψ equals the function j(τ1 )−j(τ2 ).
We obtain the product expansion
Y
j(τ1 ) − j(τ2 ) = p−1 (1 − pm q n )c(mn) ,
m>0
n∈Z
where p = e(τ1 ) and q = e(τ2 ). The product converges for =(τ1 )=(τ2 ) > 1.
This is the famous denominator formula for the monster Lie algebra [Bo4].
It was proved independently in the 80’s by Koike, Norton, and Zagier.
φ12,1 (τ, z) X
φ0,1 (τ, z) = = c(n, r)q n ζ r
∆(τ )
n≥0
r∈Z
−1
= (ζ + 10 + ζ ) + q(10ζ −2 − 64ζ −1 + 108 − 64ζ + 10ζ 2 ) + O(q 2 )
in Je0,1 . Here we have used the notation of [EZ] §9, in particular ζ = e(z).
Modular forms for the orthogonal group Γ (L) can be identified with Siegel
modular forms for the symplectic group Sp(4, Z) of genus 2. The Heegner
divisor H(∗, m) only depends on m and is equal to the Humbert surface of
discriminant |4m| in the Siegel upper half plane HS of genus 2. For Z =
( τz τz0 ) ∈ HS we write q = e(τ ), r = e(z), and s = e(τ 0 ). According to
Theorem 3.22 there exists a holomorphic modular form Ψ (Z) of weight 5 for
the group Sp(4, Z), whose divisor is the Humbert surface of discriminant 1.
It is known that the product
(2)
Y
∆0 = ϑa,b (Z)
of the ten even theta constants is a Siegel modular form of weight 5 with the
same divisor (see [Fr1] chapter 3.1). Thus Ψ has to be a constant multiple of
(2)
∆0 . It easily checked that the right constant factor is 1/64. We obtain the
(2)
following product expansion for ∆0 :
1 (2) Y
∆ = (qrs)1/2 (1 − q i rj sk )c(ik,j) .
64 0
i,j,k∈Z
i−j/2+k>0
3.4 Borcherds products 91
The product converges on det(Y ) > 1/4, where Y = =(Z). Maass showed in
(2)
[Ma3] that ∆0 can also be constructed as a Saito-Kurokawa lifting.
!
belongs to M−16,L . Write H(m) = 12 H(0, m). Theorem 3.22 implies that
there is a holomorphic modular form Ψ of weight 43092 for the orthogonal
group Γ (L) with divisor H(−2) + 528H(−1) and product expansion
Y
Ψ (Z) = e((%f (W ), Z)) (1 − e((λ, Z)))c(q(λ))
λ∈K
(λ,W )>0
The aim of this chapter is to prove Theorem 4.23. To this end we have to
exploit the Riemann structure of the quotient Hl /Γ (L).
and
iX
ω= hµν (Z)dzµ ∧ dz̄ν ,
2 µ,ν
where h(Z) = (hµν (Z)) is a positive definite Hermitean matrix for any Z ∈
Hl . In fact, since ω is O+ (V )-invariant it suffices to check this at the special
point I = i(1, 1, 0, . . . , 0) ∈ Hl . A somewhat lengthy but trivial computation
shows that1
1
1
1 2
h(I) = .
4 ..
.
2
Furthermore, ω is obviously closed. Hence it defines an O+ (V )-invariant
Kähler metric on Hl . Its underlying Riemann metric has to be a constant
multiple of g. It is easily seen that the attached invariant volume element
equals
dX dY
,
q(Y )l
where dX = dx1 · · · dxl and dY = dy1 · · · dyl . The invariant Laplace operator
is given by
−(d ∗ d ∗ + ∗ d ∗ d).
Here ∗ denotes the Hodge star operator for the Riemann metric g. Nakajima
[Nak] showed that, up to a constant multiple, the invariant Laplace operator
is equal to
l l
!
X 1 X
Ω= yµ yν ∂µ ∂¯ν − q(Y ) ∂1 ∂¯2 + ∂¯1 ∂2 − ∂µ ∂¯µ , (4.1)
µ,ν=1
2 µ=3
where
1
Zero matrix entries are often omitted.
4.1 The invariant Laplacian 95
∂ 1 ∂ ∂
∂µ = = −i ,
∂zµ 2 ∂xµ ∂yµ
∂ 1 ∂ ∂
∂¯µ = = +i .
∂ z̄µ 2 ∂xµ ∂yµ
(Note that there is a misprint in [Nak] in the definition of ∆1 .) The proof
relies on the fact that there is up to a constant multiple only one O+ (V )-
invariant second order differential operator on Hl . Therefore one only has to
verify the invariance of Ω.
We will now show that for σ > 1 − k/2 and Z ∈ / H(β, m) the function
ΦLβ,m (Z, s) is an eigenfunction of Ω. First, we prove some lemmas.
Hence Rk raises the weight, whereas Lk lowers it. The operator ∆k can be
expressed in terms of Rk and Lk by
∆k = Lk+2 Rk − k = Rk−2 Lk .
Z Z Z1/2
k−2 k
hf, giy k y=u dx.
hf, (Lk+2 g)iy dx dy − h(Rk f ), giy dx dy =
Fu Fu −1/2
∂Fu Fu
Z
∂ k ∂
= − y hf (τ ), g(τ )i − i y k hf (τ ), g(τ ) dx ∧ dy
∂y ∂x
Fu
Z
y k−2 hf, (Lk+2 g)i − y k h(Rk f ), gi dx ∧ dy.
=
Fu
In the integral over ∂Fu on the left hand side the contributions from SL2 (Z)-
equivalent boundary pieces cancel. Thus
Z Z1/2
k
k
y hf (τ ), g(τ )idτ̄ = y hf (τ ), g(τ )i y=u dx.
∂Fu −1/2
Z1/2 Z1/2
hf, (Lk g)iy k−2 y=u dx − h(Lk f ), giy k−2 y=u dx.
=
−1/2 −1/2
Proof. Since
Z1/2 h i
= h(Lk Fβ,m (τ, s)), ΘL (τ, v)y l/2 iy k−2 dx
y=u
−1/2
Z1/2 h i
− hFβ,m (τ, s), (Lk ΘL (τ, v)y l/2 )iy k−2 dx. (4.6)
y=u
−1/2
Thus it suffices to show that the integral on the right hand side tends to 0
as u → ∞. We have
Z1/2
h(Lk Fβ,m (τ, s)), ΘL (τ, v)y l/2 iy k−2 dx
−1/2
Z1/2
∂ ∂
=i y +i Fβ,m (τ, s), ΘL (τ, v) dx. (4.7)
∂x ∂y
−1/2
Using the explicit formulas for the coefficients c(γ, n; y, s) given in Theorem
1.9 and the asymptotic behavior of the Whittaker functions we may infer
that
∂ ∂
lim (4.7) = − lim y c(0, 0; y, s) = −b(0, 0, s) lim y y 1−s−k/2 .
y→∞ y→∞ ∂y y→∞ ∂y
Since σ > 1 − k/2 this is equal to 0. In the same way it can be seen that
Z1/2
lim hFβ,m (τ, s), (Lk ΘL (τ, v)y l/2 )iy k−2 dx = 0.
y→∞
−1/2
98 4 Some Riemann geometry on O(2, l)
Proposition 4.5. The Siegel theta function ΘL (τ, Z), considered as a func-
tion on H × Hl , satisfies the differential equation
This essentially follows from [Sn] Lemma 1.5 as we will now indicate. As far
as possible we use the same notation as in [Sn].
We choose a basis for L and identify V with Rl+2 . Denote by Q the Gram
matrix of L with respect to this basis, so that (x, y) = xt Qy. Let g 7→ r0 (g, Q)
be the Weil representation of Mp2 (R) × O(V ) in L2 (V ) as defined in [Sn].
For fixed τ ∈ H and Z ∈ Hl the function Fτ,Z is rapidly decreasing.
According to [Sn] Lemma 1.2 we have
−l
(r0 (σ, Q)Fτ,Z )(λ) = φ(τ )−2 φ(τ ) FM τ,Z (λ), σ = (M, φ) ∈ Mp2 (R),
−1
(r0 (g, Q)Fτ,Z )(λ) = Fτ,Z (g λ) = Fτ,gZ (λ), g ∈ O+ (V ).
It is easily seen that Fi,Z satisfies the condition (1.19) in [Sn] with m = 2 − l.
If we put 1/2 −1/2
y xy
στ = , 1 ∈ Mp2 (R)
0 y −1/2
for τ = x + iy ∈ H, then r0 (στ , Q)Fi,Z = y (2+l)/4 Fτ,Z .
The map (1.1) gives a locally isomorphic embedding of SL2 (R) into
Mp2 (R). Thus for any element u of the universal enveloping algebra U of
the Lie algebra of SL2 (R) we may consider r0 (u, Q) as a differential operator
on functions V → C. Let C be the Casimir element in U and CQ = r0 (C, Q).
Then by [Sn] Lemma 1.4 we may infer
−4∆k y l/2 Fτ,Z (λ) = (1 − l2 /4)y l/2 Fτ,Z (λ) + y −k/2 (r0 (στ , Q)CQ Fi,Z ) (λ).
(4.8)
Let % denote the representation of O+ (V ) on C ∞ (Hl ) given by
Let D be the Casimir element of the universal enveloping algebra of the Lie
algebra of O+ (V ). Then DQ = r0 (D, Q) is a differential operator on functions
V → C, and %(D) a differential operator on functions Hl → C. According
to [Hel] p. 451 ex. 5 the operator %(D) equals (up to the sign) the Laplace
operator on Hl . By (4.9) we have %(D)(Fτ,Z (λ)) = (−DQ Fτ,Z )(λ). Hence
there exists a non-zero constant cl , only depending on l, such that
for all τ ∈ H and λ ∈ V . According to [Sn] Lemma 1.5 for any smooth rapidly
decreasing function F on V the identity
for any λ ∈ V .
To determine the constant cl we consider the above identity (4.12) for
λ = d ∈ K. In this case we have
x2 + y12
y l/2 Fτ,Z (d) = y l/2 exp −πy 1 ,
q(Y )
and also (4.12) becomes much easier. (Here we have to be careful with our
notation: y = =(τ ) and Y = (y1 , . . . , yl ) = =(Z).) A straightforward compu-
tation shows that cl = 8. This completes the proof. t
u
Remark 4.8. Theorem 4.6 and Theorem 4.7, together with a regularity result
for elliptic differential operators on analytic Riemann manifolds, also imply
that the functions Φβ,m (Z, s), Φβ,m (Z), and ξβ,m (Z) are real analytic (see
[Rh] §34).
{σ ∈ Γ ; σgSt ∩ St 6= ∅}
is finite.
ii) There exist a t > 0 and finitely many rational transformations
g1 , . . . , ga ∈ O+
Q (L) such that
S = g1 St ∪ · · · ∪ ga St
Proof. This can be deduced from the general reduction theory for algebraic
groups (cf. [Bl1], [Bl2]). We briefly indicate the argument. See [Me] for a more
detailed treatment.
For any algebraic group A we denote by A+ the component of the identity.
There is a basis e3 , . . . , el of D ⊗ Q such that the restriction of q to D ⊗ Q
has the diagonal Gram matrix C = diag(−c3 , . . . , −cl ) with positive rational
˜ z̃ of L ⊗ Q, the
numbers cj . Then, with respect to the basis z, d, e3 , . . . , el , d,
quadratic form q has the Gram matrix
0 0 J
F = 0 C 0 ,
J 0 0
H4 = H0−1 , (4.19)
CH3 + H1t JH4 = 0, (4.20)
H4t JH2 + H2t JH4 + H3t CH3 = 0. (4.21)
H4 JH3t C JH2t J
H −1 = 0 E C −1 H1t J .
0 0 H0
If we write
1 h12 h13 . . . h1l h1,l+1 h1,l+2
H0 = , H1 = , H2 = ,
0 1 h23 . . . h2l h2,l+1 h2,l+2
then H is completely determined by h12 , . . . , h1,l+1 , h23 , . . . , h2l , and these
may be freely chosen. Thus UR is homeomorphic to R2l−2 . Equality (4.21)
implies in particular that
1 −1 2
h2,l+1 = (c h + · · · + c−1 2
l h2l ).
2 3 23
For any t > 0 the set
n Pl o
1 −1 2
ωt = H ∈ U ; h212 + h21,l+1 + 2
2
j=3 cj h1j < t and h2,l+1 < t
2
St = K · At · ωt
is finite. We have
Z
I t (y1 y2 )p−l dy1 dy2 dYD
y1 >1/t
y2 /y1 >1/t2
|q(YD )|<t2 y12
Z
t (y1 y2 )p−l y1l−2 dy1 dy2
y1 >1/t
y2 /y1 >1/t2
Z
t y12p−l−1 up−l dy1 du.
y1 >1/t
u>1/t2
for any ν ∈ K ⊗ R.
Proof. Using Minkowski’s inequality and Lemma 4.11, we find that it suffices
to prove that St y2p q(Y )−l dX dY < ∞. This can be done as in Lemma 4.11.
R
t
u
h(ν, Y ) = 2(ν, Y )2 − Y 2 ν 2
E = E(d, YD /y1 ),
˜ YD /y2 ).
Ẽ = E(d,
ν1 2 ν2 2
q(Y ) (νD − y1 YD ) + q(Y ) (νD − y2 YD )
ν1 2 ν2 2
≥ −εq(Y ) (νD − y1 YD ) + (νD − y2 YD )
2 !
2 ν1 y 2 + ν2 y 1 ν1 ν2 2
= −εq(Y ) νD + νD − YD −2 Y
y1 y2 y1 y2 D
2
≥ εq(Y )|νD | − 2ε|ν1 ν2 ||YD2 |.
t4
It is easily seen that |q(YD )| < t4 +1 y1 y2 for Y ∈ Rt . This implies
1
h(ν, Y ) ≥ (y 2 ν 2 + y12 ν22 ) + εq(Y )|νD 2
|
2(t4 + 1)2 2 1
1
y 2 ν 2 − 8ε(t4 + 1)2 y1 y2 |ν1 ν2 | + y12 ν22 .
+
2(t4 + 1)2 2 1
1
Thus for ε < 4(t4 +1)2 we finally obtain
t
u
Corollary 4.14. There exists an ε > 0 with the following property: For any
ν ∈ K 0 with q(ν) ≥ 0 and any Y ∈ Rt the inequality
for all ν ∈ K 0 with q(ν) ≥ 0 and all Y ∈ Rt . This implies the assertion. t
u
Proof. The first assertion can be proved in the same way as the Koecher
principle for Siegel modular forms (cf. [Fr1] chapter I Hilfssatz 3.5). Note
that the assumption l ≥ 3 is crucial. The group Γ (L) ∩ O(K) plays the role
of SL(n, Z) in [Fr1]. Then the second assertion can be deduced using Lemma
4.13 or Corollary 4.14. t
u
over the translated Siegel domain gSt . It is convenient to use the identification
Hl → K+ , Z 7→ [Z + z̃ − q(Z)z], and to rewrite I as the integral
Z
|F (W )|p dω,
g S̃t
Hl (Λ) → K+ , Z 7→ [Z + λ2 − q(Z)λ1 ].
holds.
In order to prove that F ∈ Lp (K+ /Γ ), it suffices to show that for any
t > 0 and any admissible index-tuple Λ the integral
Z
|F (W )|p dω (4.27)
S̃t (Λ)
is finite.
4.2 Reduction theory and Lp -estimates 109
Proof. Write Λ0 = (λ1 , λ02 , λ03 , λ04 ) and put u = λ2 − λ02 . Then u ∈ L ⊗ Q with
(u, λ1 ) = 0. Consider the Eichler transformation E(λ1 , u) : L ⊗ Q → L ⊗ Q,
E(λ1 , u)(λ1 ) = λ1 ,
E(λ1 , u)(λ2 ) = λ02 ,
E(λ1 , u)(k) = k − (k, λ02 )λ1 , k ∈ K(Λ).
Hence there is a t > 0 such that E(λ1 , u)St0 (Λ) ⊂ St (Λ). This implies
S̃t0 (Λ0 ) = S̃t0 (E(λ1 , u)Λ) = E(λ1 , u)S̃t0 (Λ) ⊂ S̃t (Λ).
t
u
The function f (Z) = log(|F (Z)|q(Y )r/2 ) is Γ (L)-invariant and has logarith-
mic singularities along Heegner divisors. In this section we show that up to
an additive constant f is equal to the regularized theta integral
1 X X
− c(β, m)Φβ,m (Z).
8 0
β∈L /L m∈Z+q(β)
m<0
Proof. In the following proof we use some ideas of [Rh] §34, to simplify the
argument in [Ro].
i) Since M is complete, there exists a proper C ∞ function % : M → [0, ∞)
and a constant C > 0 such that
|d%(x)| < C
σr (x) = µ(%(x)/r) has its values in [0, 1]. Moreover, it satisfies σr (x) = 1 for
x ∈ Br , σr (x) = 0 for x ∈
/ B2r , and
(σr du, σr du) ≤ (∆u, uσr2 ) + 12 (σr du, σr du) + 2(udσr , udσr )
(σr du, σr du) ≤ 2(∆u, uσr2 ) + 4(udσr , udσr )
≤ 2(∆u, uσr2 ) + 4D2 r−2 (u, u).
In the last line we have used (4.28). As r → ∞ we obtain (du, du) ≤ 2(∆u, u).
So du is square integrable.
ii) In the same way as in (i), since σr has compact support, we have
By (4.28), as r → ∞, the second term on the right hand side tends to zero.
This implies (∆u, v) = (du, dv). t
u
Theorem 4.20. Let u be a harmonic form on M . Suppose that u ∈ Lp (M )
for some p > 1. Then u is closed and coclosed.
Proof. See [Yau] Proposition 1 on p. 663. The case p = 2 is also treated in
[Rh] Theorem 26. u
t
A continuous function u on M is called subharmonic, if it satisfies the local
harmonic maximum principle, i.e. if for any point a ∈ M and any smooth
harmonic function h defined in a neighborhood of a, the function u − h has a
local maximum at a only if it is constant in a neighborhood of a (cf. [GW]).
If u is smooth, then it is subharmonic if and only if ∆u ≥ 0 everywhere on
M . The following theorem is also due to S.-T. Yau (see [Yau] Theorem 3 on
p. 663 and Appendix (ii)).
Theorem 4.21. Let u be a non-negative continuous subharmonic function
on M . Suppose that u ∈ Lp (M ) (p > 1). Then u is constant.
Note that Theorem 4.20 for functions immediately follows from Theorem
4.21. If u is a harmonic function, then |u| is subharmonic in the above sense.
Corollary 4.22. Let M be a complete connected Riemann manifold of finite
volume. Let f ∈ Lp (M ) (p > 1) be a smooth function that satisfies ∆f = c,
where c is a real constant. Then f is constant.
112 4 Some Riemann geometry on O(2, l)
Then f − Φ is constant.
Proof. Let us temporarily assume l ≥ 4. First, we will prove that f − Φ ∈
L2 (K+ /Γ (L)). It suffices to show that
Z
|f − Φ|2 dω < ∞
S̃t (Λ)
for any t > 0 and any admissible index-tuple Λ. According to Lemma 4.16–
4.18 we may assume that Λ = (z, z̃, d, d),˜ where z ∈ L is a primitive norm 0
vector, z ∈ L with (z , z) = 1, z̃ = z − q(z 0 )z, and d is a primitive norm 0
0 0 0 0
⊥
vector in K = L ∩ z ⊥ ∩ z 0 , d0 ∈ K 0 with (d0 , d) = 1, and d˜ = d0 − q(d0 )d.
This is precisely the situation we considered in the previous sections. We may
view F , f , and Φ as functions on Hl = Hl (Λ) and use the Fourier expansion
of Φ which was determined in section 3.2. We have to prove that
Z
dX dY
|f (Z) − Φ(Z)|2 < ∞.
q(Y )l
St
4.3 Modular forms with zeros and poles on Heegner divisors 113
L
In view of the Fourier expansion (3.38) of the ψβ,m we find
of G̃ into (4.30) and compare the constant terms on both sides, we find C = 0
and thereby
G̃(σZ) = G̃(Z).
But now, by the Koecher principle (Proposition 4.15), G̃ is bounded on St .
We obtain
on St , where
1 X X
B=− c(β, m)bβ,m (0, 0).
8 0
β∈L /L m∈Z+q(β)
m<0
rl l
Ω(f (Z) − Φ(Z)) = − − B
8 4
is constant. Since Hl /Γ (L) is a complete Riemann manifold of finite volume,
we may apply Corollary 4.22. We find that f − Φ is constant.
In the case l = 3 the same argument shows that f − Φ ∈ Lp (K+ /Γ (L))
for p < 2. Again, by Corollary 4.22 it has to be constant. t
u
Corollary 4.24. Let F be a meromorphic modular form of weight r with
respect to Γ (L). Suppose that its divisor is a linear combination of Heegner
divisors as in (4.29). Then
1 X X
r=− c(β, m)q(β, −m),
4 0
β∈L /L m∈Z+q(β)
m<0
where the q(γ, n) denote the Fourier coefficients of the Eisenstein series E
with constant term 2e0 in Mκ,L .
Proof. We use the same notation as in the proof of Theorem 4.23. By Theo-
rem 4.23 we know that f − Φ is constant. Thus
r
<(G̃(Z)) − 2π(λ0 , Y ) − ξ(Z) + 2 log(q(Y )) (4.34)
of (4.34) is constant.
This implies in particular that the function
|Y | X X
K
−2π(λ0 , Y /|Y |) + √ c(β, m)ξβ,m (Y /|Y |)
8 2 β∈L0 /L m∈Z+q(β)
m<0
is invariant under the action of the real orthogonal group O+ (K ⊗ R). Hence
it is constant.
We may infer that there are constants C, C 0 ∈ R such that
C|Y | + ( 2r + B) log(q(Y )) = C 0 .
r
But this is only possible, if C = 0 and 2 = −B. Using Proposition 1.16, we
obtain the assertion. t
u
Remark 4.25. Let β ∈ L0 /L and m ∈ Z + q(β) with m < 0. If there is no
λ ∈ L0 with (λ, q(λ)) = (β, m), then H(β, m) = 0 in the divisor group of
Hl /Γ (L). In this case Theorem 4.23 implies that Φβ,m is constant.
5 Chern classes of Heegner divisors
of irreducible closed
S analytic subsets Y of codimension 1 such that the sup-
port supp(D) = nY 6=0 Y is a closed analytic subset of everywhere pure
codimension 1. Then for any compact subset K ⊂ X there are only finitely
many Y with Y ∩ K 6= ∅ and nY 6= 0. We denote the group of divisors on X
by D(X).
If Γ is a group of biholomorphic transformations of X acting properly
discontinuously, we may consider the inverse image π ∗ (D) of a divisor D
on X/Γ under the canonical projection π : X → X/Γ . For any irreducible
component Y of the inverse image of supp(D) the multiplicity of Y with
respect to π ∗ (D) equals the multiplicity of π(Y ) with respect to D. Then
π ∗ (D) is a Γ -invariant divisor on X. (So we do not take account of possibly
occurring ramification in the definition of π ∗ .)
Now let X = Hl be the generalized upper half plane as in section 3.2, and
let Γ be the orthogonal group Γ (L) or a subgroup of finite index. If β ∈ L0 /L
and m ∈ Z + q(β) with m < 0, then
X
λ⊥
λ∈β+L
q(λ)=m
Moreover, we write Cl(X/Γ ) for the quotient of D(X/Γ ) modulo the sub-
group of divisors coming from meromorphic modular forms of weight 0 with
trivial character. If Γ acts freely, then Cl(X/Γ ) coincides with the usual
notion of the divisor class group.
Let ζN be a primitive N -th root of unity. It is easily seen that the co-
efficients of %L are contained in Q(ζN ). Indeed, it suffices to check this on
the generators T and S. For T the assertion is trivial. For S one uses the
evaluation of the Gauss sum
X p √ b+ −b−
e(q(γ)) = |L0 /L| i
γ∈L0 /L
due to Milgram (cf. [Bo2] Corollary 4.2). The following Lemma is a slight
improvement of [Bo3] Lemma 4.2.
Lemma 5.1. If κ is integral, then the space Sκ,L has a basis of cusp forms
whose Fourier coefficients all lie in Z[ζN ]. If κ is half-integral, then Sκ,L has
a basis of cusp forms whose coefficients all lie in Z[ζN 0 ], where N 0 is the least
common multiple of N and 8.
Proof. First, assume that κ is integral. We use the following well known
results on integral weight modular forms:
i) The space Sκ (N ) has a basis of cusp forms with coefficients in Z[ζN ]
(cf. [Sh] section 5 or [DI] Corollary 12.3.9).
ii) If f ∈ Sκ (N ) with coefficients in Z[1/N, ζN ] and (M, ϕ) ∈ Γ (N ),
then f |κ (M, ϕ) also lies in Sκ (N ) and has coefficients in Z[1/N, ζN ]. Here
5 Chern classes of Heegner divisors 119
where cσ (γ, n) denotes the conjugate of c(γ, n). If κ is integral (resp. half-
integral) and σ ∈ Gal(C/Q(ζN )) (resp. σ ∈ Gal(C/Q(ζN 0 ))), then Lemma
5.1 implies that f σ is also a cusp form in Sκ,L . For general σ ∈ Gal(C/Q) we
only have f σ ∈ Sκ (N ) ⊗ C[L0 /L]. In this context it would be interesting to
know, if Sκ,L always has a basis with rational Fourier coefficients.
Let
Sκ,L = Gal(C/Q) · Sκ,L
120 5 Chern classes of Heegner divisors
be the space of all Galois conjugates of the elements of Sκ,L and analogously
Mκ,L = Gal(C/Q) · Mκ,L . If R is a subring of C then we write Sκ,L (R)
resp. Sκ,L (R) for the R-module of cusp forms in Sκ,L resp. Sκ,L , whose coef-
ficients all lie in R. Observe that Lemma 5.1 implies
(
[Gal(Q(ζN )/Q) · Sκ,L (Q(ζN ))] ⊗ C, for κ ∈ Z,
Sκ,L =
[Gal(Q(ζN 0 )/Q) · Sκ,L (Q(ζN 0 ))] ⊗ C, for κ ∈ 21 Z.
In particular Sκ,L is finite dimensional.
By construction Sκ,L has a basis of cusp forms whose coefficients are
rational integers. Hence Sκ,L (Z) ⊗Z C = Sκ,L . We will also consider the dual
∗
Z-module Sκ,L (Z) of Sκ,L (Z). Special elements are the functionals
X
aβ,m : Sκ,L (Z) → Z; f= c(γ, n)eγ (nτ ) 7→ aβ,m (f ) = c(β, m)
γ,n
(β ∈ L0 /L and m ∈ Z − q(β), m > 0). We write Aκ,L P (Z) for the submodule
∗
of Sκ,L (Z) consisting of finite linear combinations β,m c(β, m)aβ,m , where
c(β, m) ∈ Z and c(β, m) = c(−β, m). It is easily seen that Aκ,L (Z) has finite
∗
index in Sκ,L (Z) and
∗
Aκ,L (Z) ⊗Z C = Sκ,L ,
∗
where Sκ,L denotes the C-dual space of Sκ,L .
1
Theorem 5.2. Let XL = Hl /Γ (L). The assignment aβ,m 7→ 2 H(β, −m)
defines a homomorphism
η : Aκ,L (Z) −→ f
Cl(XL ). (5.1)
Proof. We have to show that η is well defined. Suppose that
X X
a= c(β, m)aβ,m
β∈L0 /L m∈Z−q(β)
m>0
c : Cl(XL ) −→ H 2 (XL , C)
into the second cohomology. Using the results of the previous chapters, we
may determine the images of the divisors H(β, m) explicitly.
Before stating the theorem let us give an easy example. Let F : Hl → C
be a modular form of weight r with respect to Γ (L). Then |F (Z)|q(Y )r/2
is invariant under Γ (L). We may view F as a trivialization of the inverse
image L(π ∗ ((F ))) of the sheaf attached to (F ) ∈ D(XL ), and q(Y )−r/2 as
a Hermitean metric on L((F )). Hence the Chern class of the divisor (F ) is
given by
r
c((F )) = − ∂ ∂¯ log q(Y ),
2
a constant multiple of the Kähler class.
122 5 Chern classes of Heegner divisors
|Y | K b(0, 0) ¯
hβ,m (Z) = ∂ ∂¯ √ ξβ,m (Y /|Y |) − ∂ ∂ log(Y 2 )
4 2 4
1 X X
+ √ b(δ, q(λ))
2 π 0 0
λ∈K δ∈L0 /L
q(λ)<0 p(δ)=λ+K
X1
× ¯ 2−k (πn|λ||Y |, πn(λ, Y )) e(n(λ, X)).
e(n(δ, z 0 ))∂ ∂V
n
n≥1
Here b(γ, n) = b(γ, n, 1 − k/2) denote the Fourier coefficients of the Poincaré
L
series Fβ,m (τ, 1 − k/2) as in Proposition 1.10.
Remark 5.4. The above theorem, together with Theorem 2.12 on the singu-
larities of Φβ,m (Z), and Theorem 3.9 on its Fourier expansion, imply that
Φβ,m (Z) is something as a Green current for the divisor H(β, m) in the sense
of Arakelov geometry (see [SABK]). More precisely this function should define
a Green object with log-log-growth in the extended arithmetic intersection
theory due to Burgos, Kramer, and Kühn [BKK]. It will be interesting to
investigate this connection in future.
of the second cohomology H 2 (XL , C) and the span of the Kähler class
∂ ∂¯ log q(Y ). The Chern class map Cl(XL ) → H 2 (XL , C) induces a homo-
morphism
5.1 A lifting into the cohomology 123
Then Ω(τ ) is a priori a formal power series with coefficients in the subspace
of H̃ 2 (XL , C) spanned by the Heegner divisors. It is a consequence of the
existence of the Borcherds lift (Theorem 3.22), Serre duality (as in Theorem
1.17), and Lemma 5.1 that Ω(τ ) is in fact a modular form in Mκ,L with
values in H̃ 2 (XL , C). This argument is due to Borcherds (see [Bo3]). Hence
Ω(τ ) can be viewed as a kernel function for the map (5.2). The image of a
∗
functional a ∈ Sκ,L is given by
The purpose of this section is to refine the map (5.2) and to describe it
more precisely. First we will replace the image space H̃ 2 (XL , C) by a certain
space of harmonic (1, 1)-forms.
Let fClH (XL ) denote the subgroup of f Cl(XL ) generated by the classes of
the Heegner divisors H(β, m). Write
for the vector space of square integrable harmonic (1, 1)-forms on XL . Ac-
cording to [Bl3] (Theorem 6.2 and §7) H1,1 (XL ) is a finite dimensional space
of automorphic forms in the sense of [Ha]. In particular its elements are Z-
finite, where Z denotes the center of the universal enveloping algebra of the
Lie algebra of O(V ). By Theorem 4.20 any square integrable harmonic form
on a complete Riemann manifold is closed and thereby defines a cohomology
class via de Rham isomorphism.
Recall the basic facts on the Riemann geometry of Hl summarized in
section 4.1. We saw that the l × l matrix
∂2
1
h(Z) = h(Y ) = − log(q(Y ))
4 ∂yi ∂yj i,j
124 5 Chern classes of Heegner divisors
The Kähler form ∂ ∂¯ log q(Y ) is harmonic by (4.3). We leave it to the reader
to show that it is square integrable. We put
e 1,1 (XL ) = H1,1 (XL ) C∂ ∂¯ log q(Y ).
H (5.5)
Notice that the natural map H e 1,1 (XL ) → H̃ 2 (XL , C) is in general not injec-
tive.
The following strengthening of Theorem 5.3 is crucial for our argument.
commutes.
5.1 A lifting into the cohomology 125
¯ L (Z) = ∂ ∂Φ
Proof. i) Since ∂ ∂ξ ¯ L (Z), the (1, 1)-form ∂ ∂ξ ¯ L (Z) is in-
β,m β,m β,m
L
variant under the action of Γ (L). By Theorem 4.7 we know that Ωξβ,m is
constant. Because XL is Kählerian, the Laplace operator commutes with ∂
¯ Thus ∂ ∂ξ
and ∂. ¯ L is harmonic.
β,m
We now prove that ∂ ∂ξ¯ L is square integrable. In the same way as in the
β,m
proof of Theorem 4.23 it suffices to show that ∂ ∂ξ¯ L is square integrable on
β,m
any Siegel domain St (Λ) (where Λ is an admissible index tuple and t > 0).
By Lemmas 4.16–4.18 we may assume that Λ = (z, z̃, d, d), ˜ where z ∈ L is
a primitive norm 0 vector, z ∈ L with (z , z) = 1, z̃ = z 0 − q(z 0 )z, and d
0 0 0
⊥
is a primitive norm 0 vector in K = L ∩ z ⊥ ∩ z 0 , d0 ∈ K 0 with (d0 , d) = 1,
⊥
and d˜ = d0 − q(d0 )d. We put D = K ∩ d⊥ ∩ d0 and use the coordinates on
Hl = Hl (Λ) introduced in section 4.1 on page 93.
We consider the Fourier expansion of ∂ ∂ξ ¯ L (see Theorem 5.3). The
β,m
K
constant term involves the function ξβ,m (Y /|Y |) (see Definition 3.3). Here the
contribution ΦDβ,m can be evaluated by Theorem 2.14. If we put everything
¯ L (Z) is equal to
together we find that ∂ ∂ξ β,m
2 X X X1
√ b(δ, q(λ)) e(n(δ, z 0 ))
π 0
n
λ∈K δ∈L00 /L n≥1
q(λ)<0 p(δ)=λ+K
¯ κ (πn|λ||Y |, πn(λ, Y )) e(n(λ, X))
× ∂ ∂V
λ2 YD2
¯ 2
X
¯ 1 2
− b(0, 0)∂ ∂ log(Y ) + 4 π∂ ∂ (λ, YD ) −
0
y1 l−2
λ∈D
λ+D=p2 (β)
q(λ)=m
κ−1
4 X X X 1
¯ πn|λ||Y |
+ b(δ, q(λ)) ∂ ∂y1
π n2 y1
λ∈D 0 −0 δ∈K00 /K n≥1
p(δ)=λ+D
2πn|λ||Y | (λ, YD )
× Kκ−1 e n + n(δ, d0 ) . (5.6)
y1 y1
Moreover, we will use the following estimates. Let ε > 0 and a, b ∈ N0 . Then
a b √
∂ ∂ A2 +B 2
Vκ (A, B) ε,a,b e−2(1−ε) (5.8)
∂A ∂B
√
uniformly on A2 + B 2 > ε (compare (3.26)), and
126 5 Chern classes of Heegner divisors
a
∂
Kκ−1 (y) ε,a e−y (5.9)
∂y
uniformly on y > ε (see [AbSt] chapter 9).
By means of (5.7), (5.8), and Lemma 4.13 one can show that the first sum
over λ ∈ K 0 in (5.6) is rapidly decreasing on St . Hence it is square integrable
over St . The term b(0, 0)∂ ∂¯ log(Y 2 ) is a multiple of the Kähler form. It is also
square integrable.
We now prove that the third term in (5.6) is square integrable over St . For
l = 3 it vanishes identically. Hence we may temporarily assume that l ≥ 4. It
suffices to show that
∂2 1 λ2 YD2
2
(λ, YD ) − dzi ∧ dz̄j
∂yi ∂yj y1 l−2
is square integrable for all i, j = 1, . . . , l. On St the estimate
(
∂2 1 2 2
λ Y 1/y1 , if i 6= 2 and j 6= 2,
(λ, YD )2 − D
∂yi ∂yj y1 l−2 0, if i = 2 or j = 2,
for all Z ∈ St and all λ ∈ D0 − 0. Using (4.22) we find that the right hand
side satisfies the estimate
b
y12
|P (Y, λ)|
e−π|λ||Y |/y1
y1r+1 |λ2 |Y 2
b−r b−r0 /2
1 |P (Y, λ)| y1 1
e−π|λ||Y |/y1
y1 y2r |λ|r0 y2 |λ2 |
b−r b−r0 /2
1 y1 1
e−π|λ||Y |/y1 .
y 1 y2 |λ2 |
Hence the series in (5.10) has (up to a multiplicative constant) the majorant
0
X X 1 1 y1 b−r 1 b−r /2
2 y 2 |λ2 |
e−πn|λ||Y |/y1
0
n 1 y 2 n
λ∈D −0 n≥1
The latter integral is finite for l ≥ 3 by Lemma 4.11. This concludes the proof
of the first assertion.
ii) In order to prove that the above homomorphism is well defined it
suffices to show: For any meromorphic modular form F whose divisor
1 X X
(F ) = c(β, m)H(β, m)
2 0
β∈L /L m∈Z+q(β)
m<0
e 1,1 (XL ).
equals zero in H
By Theorem 4.23 we have
1X
log |F (Z)|q(Y )r/2 = − c(β, m)ΦL
β,m (Z),
8
β,m
of the special representatives hβ,−m (Z) of the Chern classes of the Heegner
divisors. (The presence of the constant term will become clear in the proof of
Theorem 5.8.) By a similar argument as in the proof of Theorem 5.5 it can
be shown that Ω L (Z, τ ) converges normally and is square integrable in Z.
Hence we may consider it as an element of H1,1 (XL ) ⊗ C[L0 /L] for fixed τ .
In the rest of this section we shall show that for fixed Z the components
of Ω L (Z, τ ) are contained in the space Mκ,L . Thus Ω L (Z, τ ) can be viewed
as the kernel function of a lifting
from elliptic cusp forms to H1,1 (XL ). The idea of the proof is the same as
in [Za]: We write Ω L (Z, τ ) as an explicit linear combination of Poincaré and
Eisenstein series in Mκ,L . This description is crucial for our further argument.
It can be used to evaluate the scalar product of Ω L (Z, τ ) and a cusp form
f ∈ Sκ,L explicitly. Thereby one obtains the Fourier expansion of ϑ(f ) in
terms of the Fourier coefficients of f . The map ϑ is one possible generalization
of the Doi-Naganuma map [DN, Na, Za]. In the O(2, 3)-case of Siegel modular
forms of genus 2 such a generalization was given by Piateskii-Shapiro using
representation theoretic methods ([PS1, PS2], see also [Wei]).
For technical reasons we first introduce two linear operators between the
spaces Sκ,L and Sκ,K where K is a sublattice of L as in section 2.1.
5.1 A lifting into the cohomology 129
P
Lemma 5.6. Let F = γ∈K 0 /K eγ fγ be a modular form in Mκ,K . Define a
0
function F |↑L
K : H → C[L /L] by
X
F |↑L
K= eβ fp(β) .
β∈L00 /L
Then F |↑L
K ∈ Mκ,L .
0
Proof. Denote the components of F |↑L K by gβ (β ∈ L /L). It suffices to check
the transformation behavior for the generators T , S of Mp2 (Z). For T this is
immediately verified. Regarding S we have to show that
√ 2κ √ l−2 X
τ i
gβ (−1/τ ) = p e((β, δ))gδ (τ ).
0
|L /L| δ∈L0 /L
t
u
P
Lemma 5.7. Let F = β∈L0 /L eβ fβ be a modular form in Mκ,L . Define a
0
function F |↓L
K : H → C[K /K] by
X X
F |↓L
K= eγ fδ .
γ∈K 0 /K δ∈L00 /L
p(δ)=γ
Then F |↓L
K ∈ Mκ,K .
130 5 Chern classes of Heegner divisors
For the rest of this section we fix a decomposition of the lattice L. Let
z ∈ L be a primitive norm 0 vector and define z 0 , N , L00 , p as in section
2.1. Then the lattice K = L ∩ z 0⊥ ∩ z ⊥ has signature (1, l − 1) and L ⊗ R =
(K ⊗ R) ⊕ Rz 0 ⊕ Rz. We assume that there also exists a primitive norm 0
vector d ∈ K and define d0 , M , K00 , p analogous to z 0 , N , L00 , p. Then the
lattice D = K ∩ d0⊥ ∩ d⊥ is negative definite and K ⊗ R = (D ⊗ R) ⊕ Rd0 ⊕ Rd.
If λ ∈ L ⊗ R then we write λK resp. λD for the orthogonal projection of λ to
K ⊗ R resp. D ⊗ R. Recall that k = 1 − l/2 and κ = 1 + l/2.
where
1 X X
¯ |ξ K (Y /|Y |)eβ (mτ ).
Ω K (Y, τ ) = √ ∂ ∂|Y β,−m (5.15)
8 2 β∈K 0 /K m∈Z−q(β)
m>0
in Sκ,K resp. Sκ,L , and E L the Eisenstein series with constant term 2e0 in
Mκ,L (cf. section 1.2).
In particular, for fixed Z the components of Ω K (Y, τ ) lie in Sκ,K . Simi-
larly Ω L (Z, τ ) can be viewed as an element of H1,1 (XL ) ⊗ Mκ,L .
Proof. Throughout the proof we write bL β,m (γ, n) instead of b(γ, n, 1 − k/2)
L
for the (γ, n)-th Fourier coefficient of the Poincaré series Fβ,m (τ, 1 − k/2) to
emphasize the dependence on L and (β, m).
To prove (5.14) we insert the Fourier expansions of the differential forms
hβ,−m (Z) into (5.12) and exchange the order of summation. We find
1 X X
¯ |ξ K
Ω L (Z, τ ) = √ ∂ ∂|Y p(β),−m (Y /|Y |)eβ (mτ )
8 2 β∈L0 /L m∈Z−q(β)
0
m>0
1 1 X X
+ ∂ ∂¯ log(Y 2 )e0 − ∂ ∂¯ log(Y 2 ) bL
β,−m (0, 0)eβ (mτ )
4 8 0
β∈L /L m∈Z−q(β)
m>0
1 X X X1
¯ κ (πn|λ||Y |, πn(λ, Y ))
+ √ e(n(δ, z 0 ))∂ ∂V
4 π 0
n
λ∈K δ∈L00 /L n≥1
q(λ)<0 p(δ)=λ+K
X X
× e(n(λ, X)) bL
β,−m (δ, q(λ))eβ (mτ ). (5.17)
β∈L0 /L m∈Z−q(β)
m>0
p(δ)=λ+D
2πn|λ||Y | (λ, Y ) 0 K
× Kκ−1 e n + n(δ, d ) Pδ,−q(λ) (τ ). (5.18)
(d, Y ) (d, Y )
¯
Since the ∂ ∂-operator annihilates all terms which are linear in Y , the first
summand in (5.18) is equal to
" #
1 ¯ −1
X X
2 D
X
2
∂ ∂(d, Y ) YD Φp(β),−m + 8π (λ, YD ) eβ (mτ ).
16 0 0
β∈K0 /K m∈Z−q(β) λ∈D
m>0 λ+D=p(β)
q(λ)=−m
(5.19)
According to Theorem 2.14 we have
16πm
ΦD
p(β),−m = # {λ ∈ p(β) + D; q(λ) = −m} .
l−2
Thus (5.19) can be rewritten as
λ2 YD2
π ¯ X X X
∂ ∂(d, Y )−1 (λ, YD )2 − eβ (mτ )
2 l−2
β∈K00 /K m∈Z−q(β) λ∈D 0
m>0 λ+D=p(β)
q(λ)=−m
λ2 YD2
π ¯ X X
= ∂ ∂(d, Y )−1 (λ, YD )2 − eβ (−q(λ)τ )
2 l−2
β∈K00 /K λ∈p(β)+D
¯ Y )−1 ΘD (YD , τ ) |↑K .
= π∂ ∂(d, D
Putting this into (5.18) we get (5.16). The fact that ΘD (YD , τ ) is a cusp form
in Sκ,D follows from the usual theta transformation formula ([Bo2] Theorem
4.1). t
u
Theorem 5.9. The lifting
ϑ0 (f )(Y ) = f (τ ) |↓L K
K , Ω (Y, τ ) τ .
∗
Sκ,L / Sκ,L ϑ / H1,1 (XL ) (5.20)
O
∗
Sκ,L
η
/f
ClH ⊗C /H
e 1,1 (XL )
∗ ∗
commutes. Here Sκ,L → Sκ,L is given by the restriction of functionals, and
∗
Sκ,L is identified with Sκ,L by means of the Petersson scalar product (·, f ) 7→
f . The right vertical arrow denotes the canonical projection.
(4πm)κ−1
Pβ,m .
2Γ (κ − 1)
According to Theorem 5.8, for fixed Z the function Ω(Z, τ )− 18 E(τ )∂ ∂¯ log(Y 2 )
is a cusp form. Using Proposition 1.5 we find that the image of aβ,m in
H1,1 (XL ) is equal to
1 1
hβ,−m (Z) + bβ,−m (0, 0)∂ ∂¯ log(Y 2 ).
2 8
This implies the assertion. t
u
We now compare the map ϑ constructed above with the theta lifting from el-
liptic cusp forms to holomorphic cusp forms on the orthogonal group O+ (V )
due to Oda [Od] and Rallis-Schiffmann [RS]. We use the description of
134 5 Chern classes of Heegner divisors
Borcherds [Bo2], because it fits most easily into our setting. Let us first recall
how holomorphic cusp forms of weight l contribute to the cohomology of XL .
It is easily verified that the determinant of the Jacobi matrix of the trans-
formation Z 7→ σ(Z) (σ ∈ O+ (V )) at Z ∈ Hl is equal to
j(σ, Z)−l
(see [Fr1] chapter I Hilfssatz 1.6 for the case of Siegel modular forms). Thus,
if f is a holomorphic modular form of weight l for Γ (L), then
ωf = f dz1 ∧ · · · ∧ dzl
is a holomorphic differential form of type (l, 0), which is invariant under Γ (L).
If f is a cusp form, then ωf is square integrable.
If r ∈ Q, we denote by Sr (Γ (L)) the space of holomorphic cusp forms of
weight r for the group Γ (L). Moreover, we write Hl,0 (XL ) for the space of
square integrable holomorphic l-forms on the quotient XL .
Sl (Γ (L))
∼ / Hl,0 (XL ).
Proof. This can be proved in the same way as Satz 2.6 in chapter III of [Fr1].
t
u
Note that the natural map Hl,0 (XL ) → H l (XL , C) to the middle coho-
mology of XL is injective.
We see that in some cases an elliptic modular form f ∈ Sκ,L can contribute
to both, the middle cohomology and the second cohomology of XL . For in-
stance, if the level of L is 1 or 2, then %L = %−L and therefore Sκ,L = Sκ,−L .
In this section we show, how Theorem 5.9 can be used to answer the
question, whether F is already the Borcherds lift of a nearly holomorphic
modular form as in Theorem 3.22.
Recall that a hyperbolic plane is a lattice which is isomorphic to the lattice
Z2 with the quadratic form q((a, b)) = ab.
We define two subspaces of Sκ,L by
X
−
Sκ,L = {f = c(γ, n)eγ (nτ ) ∈ Sκ,L ; c(λ, −q(λ)) = 0 for all λ ∈ L0 },
γ,n
(5.21)
⊥
+ −
Sκ,L = Sκ,L . (5.22)
− +
If the lattice L splits a hyperbolic plane over Z, then Sκ,L = 0 and Sκ,L =
Sκ,L . It is an immediate consequence of Theorem 5.9 (i) that the lifting ϑ
−
vanishes identically on Sκ,L .
(c(λ) ∈ Z with c(λ) = c(−λ) and c(λ) = 0 for all but finitely many λ),
is a Borcherds product; i.e. there exists a nearly holomorphic modular form
!
f ∈ Mk,L with principal part
X
c(λ)eλ (q(λ)τ ) (5.24)
0
λ∈L
q(λ)<0
Proof. Let F be a meromorphic modular form for the group Γ (L) with divisor
(5.23). Then
1 X
c(λ)c(H(λ, q(λ))) = 0
2 0 λ∈L
−
is contained in Sκ,L . On the other hand, by Proposition 1.5 this linear com-
−
bination of Poincaré series is orthogonal to Sκ,L . Therefore it vanishes iden-
tically. But this is equivalent to saying
X
c(λ)aλ,−q(λ) = 0
λ∈L0
∗
in Sκ,L . By Theorem 1.17 there exists a nearly holomorphic modular form
!
f ∈ Mk,L with principal part (5.24). Let B(f ) be the Borcherds lift of f
as in Theorem 3.22. Then B(f )/F is a holomorphic modular form (with
a multiplier system of finite order) without any zeros on Hl . Hence it is
constant. t
u
Theorem 5.12. Let L be an even lattice of signature (2, l), that splits two
orthogonal hyperbolic planes over Z. Then every meromorphic modular form
for the group Γ (L), whose divisor is a linear combination of Heegner divisors,
is a Borcherds product as in Theorem 3.22.
for all λ ∈ K 0 with q(λ) < 0. By an easy inductive argument we find that
c(λ, −q(λ)) = 0 for all λ ∈ K 0 with q(λ) < 0.
Recall that there is an isomorphism between H1 ⊕ H2 and the lattice
M2 (Z) of integral 2 × 2 matrices such that the quadratic form q corresponds
138 5 Chern classes of Heegner divisors
to the determinant on M2 (Z). The group SL2 (Z) acts on M2 (Z) by multipli-
cation from both sides. This gives rise to a homomorphism
(cf. [FH] Lemma 4.4). Hence, by the theorem of elementary divisors for
SL2 (Z), for any λ ∈ L0 there is a σ ∈ O+ (H1 ⊕ H2 ) ⊂ Γ (L) such that
σ(λ) ∈ K 0 .
Thereby we obtain c(λ, −q(λ)) = 0 for all λ ∈ L0 with q(λ) < 0. t
u
It seems likely that Theorem 5.11 can also be used to prove a more general
version of Theorem 5.12. However, this seems to be a non-trivial problem,
which might require additional arguments like some newform theory or Hecke
theory for the space Sκ,L and the lifting ϑ.
Example 5.13. Fix a positive integer t. Let L be the lattice L = D ⊕H1 ⊕H2 ,
where H1 , H2 are hyperbolic planes and D denotes the lattice Z with the
negative definite quadratic form q(a) = −ta2 . Obviously L has signature
(2, 3), k = −1/2, and κ = 5/2. The discriminant group L0 /L ∼ = D0 /D has
order 2t. In the same way as in [EZ] §5 it can be seen that the space S5/2,L
cusp
is isomorphic to the space J3,t of skew-holomorphic Jacobi cusp forms of
weight 3 and index t (see [Sk], [Ko]).
The group Γ (L) is isomorphic to the paramodular group of level t (cf. [GN]
Lemma 1.9 and [GrHu]). Moreover, the quotient XL = Hl /Γ (L) is the moduli
space of Abelian surfaces with a (1, t)-polarization. The Heegner divisors
H(β, m) are known as Humbert surfaces. Since L splits two hyperbolic planes
over Z, any H(β, m) is a prime divisor in D(XL ) (cf. [FH] Lemma 4.4).
By Theorem 5.9 we have an injective map
cusp
ϑ : J3,t −→ H1,1 (XL )
cusp
from J3,t into the space of square integrable harmonic (1, 1)-forms on XL .
cusp
Observe that the 0-th Fourier coefficient ϑ0 (f )(Y ) of the lift of f ∈ J3,t
can be identified with the Shimura lift of f . According to Theorem 5.12 any
meromorphic modular form with respect to Γ (L), whose only zeros and poles
lie on Humbert surfaces, is a Borcherds product in the sense of Theorem 3.22.
In this case the classical theta lift is essentially the Maass lift. It gives rise
to an injective map
cusp
ι : J3,t −→ H3,0 (XL )
from holomorphic Jacobi cusp forms of weight 3 and index t into H3,0 (XL ).
We obtain a partial answer to a question raised by W. Kohnen ([Ko] Problem
(ii) in §2). Holomorphic Jacobi forms of weight 3 contribute via Maass lift
to the middle cohomology of XL , whereas skew-holomorphic Jacobi forms
contribute to the second cohomology. Here it is natural to ask, if there is a
similar connection in higher weights.
References
√
· The principal branch of the square root
f (·) g(·) |f (·)| ≤ C|g(·)|
f (·) ε g(·) |f (·)| ≤ C(ε)|g(·)|
f ≈g See p. 51
|κ The Petersson slash operator of weight κ with representation %L (p. 17)
|∗κ The Petersson slash operator of weight κ with representation %∗L (p. 17)
ha, bi The standard scalar product on C[L0 /L] for a, b ∈ C[L0 /L]
(a, b) The bilinear form on L for a, b ∈ L
(f, g) The Petersson scalar product of f, g ∈ Sκ,L (p. 22)
↑LK See p. 129
↓LK See p. 129
λ2 = (λ, λ) for λ ∈ V
|λ| = |(λ, λ)|1/2 for λ ∈ V
λ⊥ The orthogonal complement of λ
λv The orthogonal projection of λ ∈ V onto v
∗ ∗
aγ,n A certain functional in Sκ,L (p. 36) or Sκ,L (Z) (p. 120)
Aκ,L (Z) See p. 120
b(γ, n, s) The Fourier coefficients of Fβ,m (τ, s) (p. 30)
bβ,m (γ, n) = b(γ, n) = b(γ, n, 1 − k/2)
Br (x) The r-th Bernoulli polynomial (p. 61)
Br (x) See p. 62
c(D) The Chern class of the divisor D in the second cohomology
c(γ, n; y, s) The Fourier coefficients of Fβ,m (τ, s) (p. 57)
C The positive cone in K ⊗ R (p. 77)
Cβ,m A certain constant (p. 72)
Cl(X/Γ ) The divisor class group of X/Γ (p. 117)
Cl(X/Γ )
f The modified divisor class group of X/Γ (p. 117)
ClH (XL )
f The subgroup of f Cl(XL ) generated by the Heegner divisors H(β, m)
C The complex numbers
C[L0 /L] The group algebra of the discriminant group L0 /L
Cs=a [f (s)] The constant term of the Laurent expansion of f at s = a (p. 50)
χ A multiplier system or character
d Often a primitive isotropic vector in K
144 Notation
L
ψβ,m See p. 66 and p. 79
Ψβ,m (Z) See p. 80
q(a) = 21 (a, a), the quadratic form on L
qβ (γ, n) The Fourier coefficients of EβL (τ )
q(γ, n) The Fourier coefficients of E(τ )
Q The rational numbers
Rk A Maass differential operator (p. 95)
R The real numbers
Rt The set of Y defined by (4.18b)–(4.18d)
<(·) The real part
%L The Weil representation attached to (L0 /L, q) (p. 16)
%γδ (M, φ) A coefficient of the representation %L (p. 17)
%∗L The dual representation of %L
%β,m (W ) See p. 67
%f (W ) See p. 86
s A complex variable
√
S = 01 −1 0 , τ ∈ Mp2 (Z)
SL2 (R) The group of real 2 × 2-matrices of determinant 1
SL2 (Z) The group of integral 2 × 2-matrices of determinant 1
Sκ,L The subspace of cusp forms in Mκ,L
∗
Sκ,L The dual space of Sκ,L
−
Sκ,L A certain subspace of Sκ,L (p. 136)
+
Sκ,L A certain subspace of Sκ,L (p. 136)
supp(D) The support of the divisor D
St A Siegel domain in Hl (p. 101)
S̃t (Λ) A Siegel domain in K+ (p. 108)
S(β, m, U ) See p. 47 and p. 81
Sκ,L = Gal(C/Q) · Sκ,L (p. 119)
Sκ,L (Z) The module of cusp forms in Sκ,L whose coefficients all lie in Z
∗
Sκ,L (Z) The dual module of Sκ,L (Z)
σ Often the real part of s
T = (( 10 11 ) , 1) ∈ Mp2 (Z)
τ A variable in H
ΘL (τ, v; r, t) The Siegel theta function attached to L (p. 39)
θγ (τ, v; r, t) A component of ΘL (τ, v; r, t)
ϑ A lifting, see p. 128
ϑ0 (f )(Y ) The 0-th Fourier coefficient of ϑ(f ) (p. 133)
ΘD (YD , τ ) A theta series, see p. 130
v Usually an element of Gr(L)
V L⊗R
V (C) =V ⊗C
Vκ (A, B) A special function, see p. 71
w The orthogonal complement of zv in v
W A Weyl chamber (p. 61), or a complex variable
Notation 147