28.04.2014 Views

Mock-modular forms of weight one - UCLA Department of Mathematics

Mock-modular forms of weight one - UCLA Department of Mathematics

Mock-modular forms of weight one - UCLA Department of Mathematics

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

MOCK-MODULAR FORMS OF WEIGHT ONE<br />

W. DUKE AND Y. LI<br />

Abstract. The object <strong>of</strong> this paper is to initiate a study <strong>of</strong> the Fourier coefficients <strong>of</strong><br />

a <strong>weight</strong> <strong>one</strong> mock-<strong>modular</strong> form and relate them to the complex Galois representations<br />

associated to the form’s shadow, which is a <strong>weight</strong> <strong>one</strong> newform. In this paper, our focus<br />

will be on <strong>weight</strong> <strong>one</strong> dihedral new<strong>forms</strong> <strong>of</strong> prime level p ≡ 3 (mod 4). In this case we give<br />

properties <strong>of</strong> the Fourier coefficients themselves that are similar to (and sometimes reduce<br />

to) cases <strong>of</strong> Stark’s conjectures on derivatives <strong>of</strong> L-functions. We also give a new <strong>modular</strong><br />

interpretation <strong>of</strong> certain products <strong>of</strong> differences <strong>of</strong> singular moduli studied by Gross and<br />

Zagier. Finally, we provide some numerical evidence that the Fourier coefficients <strong>of</strong> a mock<strong>modular</strong><br />

form whose shadow is exotic are similarly related to the associated complex Galois<br />

representation.<br />

1. Introduction<br />

Roughly speaking, a mock-<strong>modular</strong> form is the holomorphic part <strong>of</strong> a harmonic <strong>modular</strong><br />

form. For our purposes a harmonic <strong>modular</strong> form <strong>of</strong> <strong>weight</strong> k ∈ 1 2 Z is a Maass form for Γ 0(M)<br />

that is killed by the <strong>weight</strong> k Laplacian and that is allowed to have polar-type singularities<br />

in the cusps. Associated to such a form f is the <strong>weight</strong> 2 − k weakly holomorphic form<br />

(1) ξ k f(z) = 2iy k ∂¯z f(z).<br />

The operator ξ k is related to the <strong>weight</strong> k Laplacian ∆ k through the identity<br />

(2) ∆ k = ξ 2−k ξ k .<br />

A special class <strong>of</strong> harmonic <strong>forms</strong> have ξ k f holomorphic in the cusps and a Fourier expansions<br />

at ∞ <strong>of</strong> the shape<br />

(3) f(z) = ∑ n≥n 0<br />

c + (n)q n − ∑ n≥0<br />

c(n)β k (n, y)q −n .<br />

This expansion is unique and absolutely uniformly convergent on compact subsets <strong>of</strong> H, the<br />

upper half-plane. Here q = e 2πiz with z = x + iy ∈ H and β k (n, y) is given for n > 0 by<br />

β k (n, y) =<br />

∫ ∞<br />

y<br />

e −4πnt t −k dt<br />

while for k ≠ 1 we have β k (0, y) = y 1−k /(k − 1) and β 1 (0, y) = − log y. For such f the Fourier<br />

expansion <strong>of</strong> ξ k f is simply<br />

ξ k f(z) = ∑ c(n)q n .<br />

n≥0<br />

After Zagier, the function ∑ n c+ (n)q n is said to be mock-<strong>modular</strong> with shadow ∑ n≥0 c(n)qn .<br />

It is important to observe that a mock-<strong>modular</strong> form is only determined by its shadow up to<br />

the addition <strong>of</strong> a holomorphic form.<br />

Date: July 3, 2012.<br />

Supported by NSF grant DMS-10-01527.<br />

Supported by NSF.<br />

1


2 W. DUKE AND Y. LI<br />

Some (non-<strong>modular</strong>) mock-<strong>modular</strong> <strong>forms</strong> have Fourier coefficients that are well-known<br />

arithmetic functions. Let σ(n) = ∑ m|n<br />

m and H(n) be the Hurwitz class number. Then<br />

−8π ∑ ( )<br />

σ(n)q n σ(0) = −<br />

1<br />

24<br />

n≥0<br />

is mock-<strong>modular</strong> <strong>of</strong> <strong>weight</strong> k = 2 for the full <strong>modular</strong> group with shadow 1 and<br />

−16π ∑ ( )<br />

H(n)q n H(0) = −<br />

1<br />

12<br />

n≥0<br />

is mock-<strong>modular</strong> <strong>of</strong> <strong>weight</strong> 3/2 for Γ 0 (4) with shadow the Jacobi theta series θ(z) = ∑ n∈Z qn2 .<br />

(See [45]). Actually, these two examples will play a fundamental role in our work on the <strong>weight</strong><br />

<strong>one</strong> case.<br />

In general, the Fourier coefficients <strong>of</strong> mock-<strong>modular</strong> <strong>forms</strong> are not well understood. Especially<br />

in the case <strong>of</strong> <strong>weight</strong> 1/2, which includes the mock-theta functions <strong>of</strong> Ramanujan, there<br />

has been considerable progress made by Zwegers in his thesis [47], by Bringmann-Ono [6],<br />

and others (see [46] for a good exposition). The Fourier coefficients <strong>of</strong> mock-<strong>modular</strong> <strong>forms</strong><br />

<strong>of</strong> <strong>weight</strong> 1/2 whose shadows Shimura–lift to cusp <strong>forms</strong> attached to elliptic curves have also<br />

been shown to be quite interesting by Bruinier and Ono [10] and Bruinier [13]. We remark<br />

that mock-<strong>modular</strong> <strong>forms</strong> whose shadows are only weakly holomorphic are also <strong>of</strong> interest<br />

(see [17],[18]) but in this paper we only consider those with holomorphic shadows.<br />

The self-dual case k = 1 presents special features and difficulties. The Riemann-Roch<br />

theorem is without content when k = 2 − k, and the existence <strong>of</strong> cusp <strong>forms</strong> is a subtle<br />

issue. Furthermore, the infinite series representing the Fourier series <strong>of</strong> <strong>weight</strong> <strong>one</strong> harmonic<br />

Poincaré series are difficult to handle.<br />

The fact that interesting non-<strong>modular</strong> mock-<strong>modular</strong> <strong>forms</strong> <strong>of</strong> <strong>weight</strong> 1 exist follows from<br />

work <strong>of</strong> Kudla, Rapoport and Yang [30]. Suppose that M = p > 3 is a prime with p ≡ 3<br />

(mod 4) and that χ p (·) = ( · ) is the Legendre symbol. Let<br />

p<br />

(4) E 1 (z) = 1 2 H(p) + ∑ n≥1<br />

R p (n)q n<br />

be Hecke’s Eisenstein series <strong>of</strong> <strong>weight</strong> <strong>one</strong> for Γ 0 (p) with character χ p , where for n > 0<br />

(5) R p (n) = ∑ m|n<br />

χ p (m).<br />

It follows from [30] that Ẽ1(z) := ∑ n≥0 R+ p (n)q n is mock-<strong>modular</strong> <strong>of</strong> <strong>weight</strong> k = 1 with<br />

shadow E 1 (z), where for n > 0 we have<br />

(6) R p + (n) = −(log p)ord p (n)R p (n) − ∑<br />

log q(ord q (n) + 1)R p (n/q),<br />

χ p(q)=−1<br />

and where R p + (0) is a constant. 1 The associated harmonic form is constructed using the<br />

s-derivative <strong>of</strong> the non-holomorphic Hecke-Eisenstein series <strong>of</strong> <strong>weight</strong> 1. An arithmetic interpretation<br />

<strong>of</strong> its coefficients is given in [30]; it is shown that (−2R + (n) + 2 log pR p (n)) is the<br />

degree <strong>of</strong> a certain 0-cycle on an arithmetic curve made from elliptic curves with CM by the<br />

ring <strong>of</strong> integers in F = Q( √ −p).<br />

1 Explicitly, R + p (0) = −H(p)( L′ (0,χ p)<br />

L(0,χ p) + ζ′ (0)<br />

ζ(0) − c), where c = 1 2 γ + log(4π3/2 ) and γ is Euler’s constant.


MOCK-MODULAR FORMS OF WEIGHT ONE 3<br />

In this paper we will study the Fourier coefficients <strong>of</strong> mock-<strong>modular</strong> <strong>forms</strong> whose shadows<br />

are new<strong>forms</strong>. For simplicity we will continue to assume that M = p > 3 is a prime. To each<br />

A ∈ Cl(F ), the class group <strong>of</strong> F , <strong>one</strong> can associate a theta series ϑ A (z) defined by<br />

(7) ϑ A (z) := 1 + ∑<br />

q N(a) = ∑ r<br />

2 A (n)q n .<br />

n≥0<br />

a⊂O F<br />

[a]∈A<br />

Hecke showed that ϑ A (z) ∈ M 1 (p, χ p ), the space <strong>of</strong> <strong>weight</strong> <strong>one</strong> holomorphic <strong>modular</strong> <strong>forms</strong><br />

for Γ 0 (p) with character χ p . Let ψ be a character <strong>of</strong> Cl(F ) and consider g ψ (z) ∈ M 1 (p, χ p )<br />

defined by<br />

(8) g ψ (z) := ∑<br />

r ψ (n)q n .<br />

A∈Cl(F )<br />

ψ(A)ϑ A (z) = ∑ n≥0<br />

When ψ = ψ 0 is the trivial character, the form g ψ0 (z) is just E 1 (z) from (4), as a consequence<br />

<strong>of</strong> Dirichlet’s fundamental formula<br />

(9) R p (n) = ∑<br />

r A (n).<br />

A∈Cl(F )<br />

Otherwise, g ψ (z) is a newform in S 1 (p, χ p ), the subspace <strong>of</strong> M 1 (p, χ p ) consisting <strong>of</strong> cusp <strong>forms</strong>.<br />

Let H be the Hilbert class field <strong>of</strong> F with ring <strong>of</strong> integers O H . The following result shows<br />

that the Fourier coefficients <strong>of</strong> certain mock-<strong>modular</strong> <strong>forms</strong> <strong>of</strong> <strong>weight</strong> <strong>one</strong> with shadow g ψ (z)<br />

can be expressed in terms <strong>of</strong> logarithms <strong>of</strong> algebraic numbers in H.<br />

Theorem 1.1. Let p ≡ 3 (mod 4) be a prime with p > 3. Let ψ be a non-trivial character<br />

<strong>of</strong> Cl(F ), where F = Q( √ −p). Then there exists a <strong>weight</strong> <strong>one</strong> mock-<strong>modular</strong> form<br />

˜g ψ (z) = ∑ n≥n 0<br />

r + ψ (n)qn<br />

with shadow g ψ (z) such that the following hold.<br />

(i) When χ p (n) = 1 or n < − p+1 24 r+ ψ<br />

(n) equals to zero.<br />

(ii) The coefficients r + ψ<br />

(n) are <strong>of</strong> the form<br />

(10)<br />

∑<br />

r + ψ<br />

(n) = −β<br />

A∈Cl(F )<br />

ψ 2 (A) log |u(n, A)|,<br />

where β ∈ Q, u(n, A) are units in O H when n ≤ 0 and algebraic numbers in H when<br />

n > 0. In addition, β depends only on p and u(n, A) depends only on n and A.<br />

(iii) Let σ C ∈ Gal(H/F ) be the element associated to the class C ∈ Cl(F ) via Artin’s isomorphism.<br />

Then it acts on u(n, A) by<br />

(iv) N H/Q (u(n, A)) is an integer and satisfies<br />

for all non-zero integers n.<br />

σ C (u(n, A)) = u(n, AC −1 ).<br />

− β 2 log(|N H/Q(u(n, A))|) = R + p (n)<br />

Observe that parts (ii) and (iii) are very similar in form to Stark’s Conjectures for special<br />

values <strong>of</strong> derivatives <strong>of</strong> L-functions. In fact, they are consequences <strong>of</strong> known cases <strong>of</strong> his<br />

conjectures when n ≤ 0. Part (iv) can be viewed as a mock-<strong>modular</strong> version <strong>of</strong> Dirichlet’s<br />

identity (9). It is important to observe that ˜g ψ (z) is not necessarily unique.


4 W. DUKE AND Y. LI<br />

The pro<strong>of</strong> <strong>of</strong> this result relies heavily on the Rankin-Selberg method for computing heights<br />

<strong>of</strong> Heegner divisors as developed in [24], but entails the use <strong>of</strong> <strong>weight</strong> <strong>one</strong> harmonic <strong>forms</strong><br />

with polar singularities in cusps in place <strong>of</strong> <strong>weight</strong> <strong>one</strong> Eisenstein series and hence requires<br />

regularized inner products. In order to get to the individual mock-<strong>modular</strong> coefficients, it<br />

is necessary to consider <strong>modular</strong> curves <strong>of</strong> large prime level N and their Heegner divisors <strong>of</strong><br />

height zero. The relation <strong>of</strong> this part <strong>of</strong> the pro<strong>of</strong> to [24] has some independent interest.<br />

Zagier pointed out that his identity with Gross for the norms <strong>of</strong> differences <strong>of</strong> singular<br />

values <strong>of</strong> the <strong>modular</strong> j-function can be neatly expressed in terms <strong>of</strong> the coefficients R + p (n)<br />

given in (6). For simplicity, let −d < 0 be a fundamental discriminant not equal to −p. and<br />

set F ′ = Q( √ −d). As is well-known, the <strong>modular</strong> j-function is well-defined on ideal classes<br />

<strong>of</strong> F and F ′ and takes values in the rings <strong>of</strong> integers <strong>of</strong> their respective Hilbert class fields.<br />

Also, values <strong>of</strong> the j-function at different ideal classes are Galois conjugates <strong>of</strong> each other.<br />

For any A ∈ Cl(F ) define the quantity<br />

(11) a d,A := ∏<br />

A ′ ∈Cl(F ′ )<br />

(j(A) − j(A ′ )).<br />

The norm <strong>of</strong> a d,A to F is thus ∏ A∈Cl(F ) a d,A and is an ordinary integer. The result <strong>of</strong> Gross<br />

and Zagier [23, Theorem 1.3] is that this integer can be expressed in terms <strong>of</strong> R p + (n) as<br />

follows:<br />

(12) log ∏<br />

A∈Cl(F )<br />

|a d,A | 2/w d<br />

= − 1 4<br />

∑<br />

k∈Z<br />

δ(k) R + p ( pd−k2<br />

4<br />

),<br />

where w d is the number <strong>of</strong> roots <strong>of</strong> unity in F ′ and δ(k) = 2 if p|k and 1 otherwise.<br />

There are two pro<strong>of</strong>s <strong>of</strong> this factorization in [23]: <strong>one</strong> analytic and <strong>one</strong> algebraic. In fact,<br />

the algebraic approach, which is based on Deuring’s theory <strong>of</strong> supersingular elliptic curves<br />

over finite fields, gives the factorization <strong>of</strong> the ideal (a d,A ) in O H for each class A ∈ Cl(F ).<br />

To state it, suppose l is a rational prime such that χ p (l) ≠ 1. Then the ideal (l) factors in<br />

O H as<br />

(13) l = ∏<br />

l δ(l)<br />

A .<br />

A∈Cl(F )<br />

The l A ’s are primes in H above l uniquely labeled so that σ C (l A ) = l AC −1 for all C ∈ Cl(F )<br />

and complex conjugation sends l A to l A −1. Let A 0 be the principal class. It is shown in [23]<br />

that the order <strong>of</strong> a d,A0 at the place associated to the prime l A is given by<br />

∑<br />

(14) ord lA (a d,A0 ) = 1 δ(k) ∑ ( )<br />

pd−k<br />

r 2<br />

2<br />

A 2 .<br />

4l m<br />

m≥1<br />

k∈Z<br />

The Galois action then yields the prime factorization <strong>of</strong> the ideal (a d,A ) for any A.<br />

Our second main result gives a mock-<strong>modular</strong> interpretation <strong>of</strong> the individual values |a d,A |.<br />

It is convenient to give it as a twisted version <strong>of</strong> (12).<br />

Theorem 1.2. For any ˜g ψ (z) = ∑ n≥n 0<br />

r + ψ (n)qn given by Theorem 1.1 and −d < 0 any<br />

fundamental discriminant different from −p we have<br />

∑<br />

∑<br />

(15)<br />

ψ 2 (A) log |a d,A | 2/w d<br />

= − 1 δ(k) r + 4<br />

ψ ( pd−k2 )<br />

4<br />

A∈Cl(F )<br />

where a d,A is defined in (11).<br />

k∈Z


MOCK-MODULAR FORMS OF WEIGHT ONE 5<br />

Since p > 3 is a prime number, Cl(F ) has odd order H(p) and ψ 2 is only trivial when ψ is<br />

trivial. Thus we can express each logarithmic term in the sum on the left hand side <strong>of</strong> the<br />

following identity in terms <strong>of</strong> the coefficients <strong>of</strong> a mock-<strong>modular</strong> form with a theta series as<br />

shadow (see Corollary 9.3 below).<br />

As with Theorem 1.1, this is proved using the methods <strong>of</strong> [24] and not the analytic technique<br />

<strong>of</strong> [23], which uses the restriction to the diagonal <strong>of</strong> an Eisenstein series for a Hilbert <strong>modular</strong><br />

group. In particular, we will define a real-analytic function Φ(z), which trans<strong>forms</strong> with<br />

<strong>weight</strong> 3/2 and level 4, and use holomorphic projection to obtain an equation between a<br />

finite linear combination <strong>of</strong> r + ψ<br />

(n)’s and an infinite sum, similar to the <strong>one</strong> in [24]. We<br />

also make use <strong>of</strong> machinery from [29]. One interesting new feature is an elementary counting<br />

argument needed to construct a Green’s function evaluated at CM points. Actually, equation<br />

(15) is a particular example <strong>of</strong> a more general identity involving values <strong>of</strong> certain Borcherds<br />

lifts. Although we will not carry this out here, it will become clear that similar methods can<br />

be used to prove a level N version in this form.<br />

Let us illustrate our Theorems in the first non-trivial case, which occurs when p = 23.<br />

The class group Cl(F ) has size 3 and two non-trivial characters ψ and ψ. The Hilbert class<br />

field H is generated by X 3 − X − 1 over F . Let α = 1.32472 . . . be the unique real root<br />

<strong>of</strong> X 3 − X − 1, which is also the absolute value <strong>of</strong> the square <strong>of</strong> a unit in H. The space<br />

S 1 (23, χ 23 ) is <strong>one</strong> dimensional and spanned by the cusp form<br />

g ψ (z) = η(z)η(23z),<br />

where η(z) = q ∏ 1/24 n≥1 (1 − qn ) is the eta function. According to Theorem 1.1, there exists<br />

a mock-<strong>modular</strong> form ˜g ψ (z) having a simple pole and the following Fourier expansion<br />

˜g ψ (z) =<br />

∑<br />

r + ψ (n)qn .<br />

n≥−1<br />

χ 23 (n)≠1<br />

The condition on the principal part determines ˜g ψ (z) uniquely in this case, though this is<br />

not true in general. Using Stokes’ Theorem and Stark’s calculation at the end <strong>of</strong> [40, II], <strong>one</strong><br />

can show that<br />

r + ψ (−1) = 〈g ψ, g ψ 〉 = 3 log(α) and r + ψ<br />

(0) = − log(α).<br />

From the numerical calculations <strong>of</strong> r + ψ<br />

(n), we can predict the values <strong>of</strong> β and u(n, A) in<br />

Theorem 1.1. Let ˜β and ũ(n, A) denote the guessed values for β and u(n, A) respectively.<br />

The following table lists r + ψ<br />

(n), which are calculated numerically, and the guessed values<br />

˜β, ũ(n, A 0 ) for the principal class A 0 ∈ Cl(F ) for 1 ≤ n ≤ 23 and a few other values <strong>of</strong> n, all<br />

with χ 23 (n) ≠ 1. It also contains the norms <strong>of</strong> ũ(n, A 0 ), which agree with condition (iv) in<br />

Theorem 1.1.<br />

n r + ψ (n) ˜β ũ(n, A0 ) N H/Q (ũ(n, A 0 )) ˜β/2<br />

5 1.1001149692823391 2 π 5 5 2<br />

7 1.7161505040673007 2 π 7 7 2<br />

10 3.9614773685309742 2 5α −6 π5 −1 5 4<br />

11 0.052996471463740862 2 π 11 11 2<br />

14 1.6582443878082415 2 7α −4 π7 −1 7 4


6 W. DUKE AND Y. LI<br />

15 -1.9437136922512246 2 5απ5 −1<br />

5 4<br />

17 -4.2163115309750479 2 π 17 17 2<br />

19 -2.7119255841404505 2 π 19 19 2<br />

20 5.9051910607821988 2 5α −7 5 6<br />

21 6.7198367256215547 2 7α −10 π −1<br />

7 7 4<br />

22 -4.2709900863081686 2 11α 5 π −1<br />

11 11 4<br />

23 -3.8460181706191355 2 π 23 23<br />

28 4.2179936148444277 2 7α −5 7 6<br />

34 -2.532478252776036 2 17α 8 π −1<br />

17 17 4<br />

38 -9.942055260392833 2 19α 15 π −1<br />

19 19 4<br />

40 -14.92826076712649 2 5α 19 π 5 5 8<br />

Table 1. Coefficients <strong>of</strong> ˜g ψ (z) for g ψ (z) = η(z)η(23z) ∈ S 23 (p)<br />

Here, π l is a generator <strong>of</strong> the prime ideal l A0 in O H and is given below in terms <strong>of</strong> α.<br />

l<br />

π l<br />

5 2α 2 − α − 1<br />

7 α 2 + α − 2<br />

11 2α 2 − α<br />

17 2α 2 + 3α + 3<br />

19 3α 2 + α<br />

1<br />

23 √ −23<br />

(10α 2 + 8α + 1)<br />

Table 2. Generators <strong>of</strong> l A0 .<br />

To illustrate how Theorem 1.2 supports these predictions, consider the classical example<br />

−d = −7, which also appeared in [43]. First, we can combine equation (15) and Theorem<br />

1.1 to write (see Corollary 9.3)<br />

4 log |a d,A | 2/w d<br />

= β ∑ ( )<br />

∣ pd−k<br />

δ(k) log ∣u<br />

2<br />

, A<br />

4<br />

∣ .<br />

k∈Z<br />

For d = 7, p = 23 and A = A 0 , we can use the predicted values in Table 1 to rewrite the<br />

equation above as<br />

( )<br />

log |u(40, A0 )| + log |u(38, A 0 )| + log |u(34, A 0 )|+<br />

4 log |a 7,A0 | = 4<br />

log |u(28, A 0 )| + log |u(20, A 0 )| + log |u(10, A 0 )|<br />

= 4 log |5 3 · 7 · 17 · π17 −1 · 19 · π19 −1 · α 24 |.<br />

This agrees with the exact <strong>of</strong> value <strong>of</strong> a 7,A0 , which is 5 3 · 7 · 17 · π17 −1 · 19 · π19 −1 · α 24 . Applying<br />

the Galois action to these predicted u(n, A 0 ) shows that they also agree with Theorem 1.2.<br />

This and several other numerical examples, together with equation (14), motivate us to<br />

make a conjecture about the factorization <strong>of</strong> the fractional ideal generated by u(n, A) in<br />

Theorem 1.1. It is not hard to see that the following conjecture implies (14).<br />

Conjecture. In (ii) <strong>of</strong> Theorem 1.1 we have the following.<br />

(i) The number 2/β is an integer dividing 24H(p)h H , where h H is the class number <strong>of</strong> H.<br />

(ii) For B ∈ Cl(F ), let l B be a prime ideal above the rational prime l as in (13). Then the<br />

order <strong>of</strong> u(n, A) at the place <strong>of</strong> H corresponding to l B is<br />

∑ ( n<br />

)<br />

ord lB (u(n, A)) = 2 r<br />

β (A −1 B) 2 .<br />

l m<br />

m≥1


MOCK-MODULAR FORMS OF WEIGHT ONE 7<br />

At all other places <strong>of</strong> H, u(n, A) is a unit. In particular, u(n, A) ∈ O H for all n, A.<br />

In general (for prime level p ≡ 3 (mod 4)) , the Deligne-Serre Theorem [15] (see also [36])<br />

identifies the L-function <strong>of</strong> a newform f ∈ S 1 (p, χ p ) with the Artin L-function <strong>of</strong> an irreducible,<br />

odd, two-dimensional, complex representation ρ f <strong>of</strong> the Galois group <strong>of</strong> a normal<br />

extension K/Q. Such a ρ f gives rise to a projective representation ˜ρ f : Gal(K/Q) → PGL 2 (C)<br />

whose image is isomorphic to D 2n , S 4 , or A 5 , in which case we say that f is dihedral, octahedral,<br />

or icosahedral, respectively. The <strong>forms</strong> g ψ (z) are precisely the dihedral <strong>forms</strong> and for<br />

them K = H. Since g ψ1 (z) = g ψ2 (z) if and only if ψ 1 = ψ 2 or ψ 1 = ψ 2 , there are exactly<br />

(H(p) − 1)/2 such <strong>forms</strong>. Theorem 1.1 relates the Fourier coefficients <strong>of</strong> a mock-<strong>modular</strong><br />

form with dihedral shadow g ψ (z) to linear combinations <strong>of</strong> logarithms <strong>of</strong> algebraic numbers<br />

in the number field K = H determined by the Galois representation. Furthermore, the coefficients<br />

in these linear combinations are algebraic numbers in the field generated by the<br />

Fourier coefficients <strong>of</strong> the shadow.<br />

Non-dihedral new<strong>forms</strong> are <strong>of</strong>ten referred to as being “exotic” since their occurrence is<br />

rare and unpredictable. We have carried out some numerical calculations for mock-<strong>modular</strong><br />

<strong>forms</strong> whose shadows are certain exotic new<strong>forms</strong> and have observed that they also seem to<br />

be related to linear combinations <strong>of</strong> logarithms <strong>of</strong> algebraic numbers in the number field K<br />

determined by the associated Galois representation. For instance, when p = 283 the space<br />

S 1 (283, χ 283 ) contains a pair <strong>of</strong> octahedral new<strong>forms</strong> associated to a Galois representation <strong>of</strong><br />

Gal(K/Q), where K is the degree 2 extension <strong>of</strong> the normal closure <strong>of</strong> Q[X]/(X 4 − X − 1)<br />

with Gal(K/Q) ∼ = GL 2 (F 3 ). Similar to the dihedral case, the Petersson norms <strong>of</strong> these<br />

new<strong>forms</strong> come from the residue <strong>of</strong> a certain degree four L-function. As a consequence <strong>of</strong><br />

proven cases <strong>of</strong> Stark’s conjecture, this residue is the logarithm <strong>of</strong> a unit inside a subfield F 6<br />

<strong>of</strong> K. This determines the principal part <strong>of</strong> a mock <strong>modular</strong> form whose shadow is <strong>one</strong> <strong>of</strong><br />

these octahedral new<strong>forms</strong>. Using this hint and some numerical calculations, we noticed that<br />

the other coefficients also seem to be linear combinations <strong>of</strong> logarithms <strong>of</strong> algebraic numbers<br />

in the same subfield F 6 . Furthermore, these algebraic numbers generate ideals with nice<br />

factorizations. Thus, it is natural to expect a statement analogous to Theorem 1.1 and the<br />

Conjecture should hold for exotic new<strong>forms</strong>. In the final section <strong>of</strong> the paper we will give<br />

computational details when p = 283.<br />

We end the introduction with a brief outline <strong>of</strong> the paper. The first two sections concern<br />

general properties <strong>of</strong> <strong>weight</strong> <strong>one</strong> mock-<strong>modular</strong> <strong>forms</strong>: in §2 we deal with the existence <strong>of</strong><br />

<strong>weight</strong> <strong>one</strong> mock <strong>modular</strong> <strong>forms</strong>, while in §3 we will decompose the space <strong>of</strong> <strong>weight</strong> <strong>one</strong><br />

harmonic Maass <strong>forms</strong> into a plus space and minus space. In §4 we study the principal part<br />

at infinity <strong>of</strong> a mock-<strong>modular</strong> form with shadow g ψ (z). The pro<strong>of</strong> <strong>of</strong> Theorem 1.1 is in §5 −<br />

§7. In §5 and §6 , we will prove Theorem 5.1 and Proposition 6.1, which imply that different<br />

types <strong>of</strong> linear combinations <strong>of</strong> r + ψ<br />

(n)’s can be put into the form <strong>of</strong> the right hand side <strong>of</strong><br />

equation (10). In §7 we verify that each individual r + ψ<br />

(n) can be solved from these linear<br />

combinations. Theorem 1.2 is proved in §8 and §9. Finally, an analysis <strong>of</strong> the specific case<br />

p = 283 is given in §10.<br />

Acknowledgements. We thank Ö. Imamoḡlu for a remark made to <strong>one</strong> <strong>of</strong> us some time ago<br />

that the self-dual case <strong>of</strong> <strong>weight</strong> <strong>one</strong> mock <strong>modular</strong> <strong>forms</strong> should be especially interesting.<br />

Also, we thank J. Bruinier, S. Ehlen, A. Venkatesh, M. Viazovska, D. Zagier and S. Zwegers<br />

for their comments. After we obtained the results <strong>of</strong> this paper, we learned that there is<br />

some intersection with [42] and [19].


8 W. DUKE AND Y. LI<br />

2. Existence <strong>of</strong> Harmonic Maass Forms <strong>of</strong> Weight One<br />

Given a cusp form <strong>of</strong> <strong>weight</strong> <strong>one</strong>, we will show the existence <strong>of</strong> a <strong>weight</strong> <strong>one</strong> mock-<strong>modular</strong><br />

form with it as shadow in this section. There are several different approaches to proving this<br />

result, such as a geometric approach in [9]. For completeness and since it might have some<br />

independent interest, in this section we prove the existence by analytically continuing <strong>weight</strong><br />

<strong>one</strong> Poincaré series via the spectral expansion.<br />

We begin with some basic definitions. Let k ∈ Z. For any function f : H → C and<br />

γ ∈ GL + 2 (R), define the <strong>weight</strong> k slash operator | k γ by<br />

(f| k γ)(z) := (det(γ))k/2 f(γz),<br />

(cz+d) k<br />

where γz is the linear fractional transformation <strong>of</strong> z under γ. We will write f|γ for f| k γ<br />

when the <strong>weight</strong> <strong>of</strong> f is understood. For M ∈ Z + let Γ 0 (M) denote the usual congruence<br />

subgroup <strong>of</strong> SL 2 (Z) <strong>of</strong> level M, namely<br />

Γ 0 (M) = {( a c d b ) ∈ SL 2(Z) ( a c d b ) ≡ ( ∗ 0 ∗<br />

∗ ) mod M}.<br />

Let ν : Z/MZ → C × be a Dirichlet character such that ν(−1) = (−1) k and ν(γ) := ν(d) for<br />

γ = ( a c d b ) ∈ Γ 0(M). Let F k (M, ν) be the space <strong>of</strong> smooth functions f : H → C such that for<br />

all γ ∈ Γ 0 (M)<br />

(f | k γ)(z) = ν(γ)f(z).<br />

Recall from (1) and (2) the differential operator ξ k and the <strong>weight</strong> k hyperbolic Laplacian<br />

∆ k . Let z = x + iy. Then ∆ k can be written as<br />

−∆ k = −y 2 (<br />

∂ 2<br />

+ ∂2<br />

∂x 2<br />

) (<br />

+ iky<br />

∂y 2<br />

∂<br />

+ i ∂<br />

∂x ∂y<br />

Then f(z) ∈ F k (M, ν) is a <strong>weight</strong> k harmonic weak Maass form <strong>of</strong> level M and character ν<br />

(or more briefly, a weakly harmonic form) if it satisfies the following properties.<br />

(i) f(z) is real-analytic.<br />

(ii) ∆ k (f) = 0.<br />

(iii) The function f(z) has at most linear exp<strong>one</strong>ntial growth at all cusps <strong>of</strong> Γ 0 (M).<br />

From now on we fix k = 1 and write | for the slash operator | 1 . Let H ! 1(M, ν) be the<br />

space <strong>of</strong> weakly harmonic <strong>modular</strong> <strong>forms</strong> <strong>of</strong> <strong>weight</strong> <strong>one</strong>, level M and character ν. Denote<br />

by M ! 1(M, ν) , M 1 (M, ν) and S 1 (M, ν) the usual subspaces <strong>of</strong> weakly holomorphic <strong>modular</strong><br />

<strong>forms</strong>, holomorphic <strong>modular</strong> <strong>forms</strong> and cusp <strong>forms</strong>, respectively. Let M ! 1(M, ν) be the space<br />

<strong>of</strong> mock-<strong>modular</strong> <strong>forms</strong> with shadows in M 1 (M, ν) and M 1 (M, ν) ⊂ M ! 1(M, ν) the subspace<br />

<strong>of</strong> mock-<strong>modular</strong> <strong>forms</strong> whose shadows are in S 1 (M, ν).<br />

From (2), it is clear that the image <strong>of</strong> H ! 1(M, ν) under ξ 1 lies in M ! 1(M, ν). Let H 1 (M, ν)<br />

be the preimage <strong>of</strong> S 1 (M, ν) under ξ 1 . Then H 1 (M, ν) is canonically isomorphic to M 1 (M, ν).<br />

We want to show that the map ξ 1 is in fact a surjection onto S 1 (M, ν).<br />

To proceed, we will construct two families <strong>of</strong> Poincaré series P m (z, s), Q m (z, s), where the<br />

first family is similar to the <strong>one</strong> used in [6]. They are a priori defined for Re(s) > 1 and will<br />

be analytically continued to Re(s) > 0 through their spectral expansions. Unlike the cases<br />

k ≥ 2, this will only be a statement about existence, and not a formula that could be used<br />

to calculate the holomorphic coefficients <strong>of</strong> the preimage explicitly. To prove the analytic<br />

continuation, we will refer to results in [31] and [35].<br />

Given any positive integer m, define<br />

(16)<br />

(17)<br />

)<br />

.<br />

φ ∗ m(z, s) := e 2πimx (4π|m|y) −1/2 M sgn(m)<br />

,s− 1 (4π|m|y),<br />

2 2<br />

ϕ ∗ m(z, s) := e 2πimx (4π|m|y) s−1/2 e −2π|m|y .


MOCK-MODULAR FORMS OF WEIGHT ONE 9<br />

Here M µ,ν (y) is the M-Whittaker function defined by<br />

(18) M µ,ν (z) := z ν+1/2 e z/2 1F 1<br />

( 1<br />

2 + µ + ν; 1 + 2ν; −z) ,<br />

with 1 F 1 (α; β; z) being the generalized hypergeometric function. Averaging them over the<br />

coset representatives <strong>of</strong> Γ ∞ \Γ 0 (M), we can define the following Poincaré series<br />

∑<br />

P m (z, s, ν) :=<br />

ν(γ)(φ ∗ m | γ)(z, s),<br />

Q m (z, s, ν) :=<br />

γ∈Γ ∞\Γ 0 (M)<br />

∑<br />

γ∈Γ ∞\Γ 0 (M)<br />

ν(γ)(ϕ ∗ m | γ)(z, s).<br />

For convenience, we shall omit ν and write P m (z, s) and Q m (z, s) instead. When Re(s) ><br />

1, both families have Fourier expansions in z that converges absolutely and uniformly on<br />

compact subset <strong>of</strong> H (see [31, Satz 6.5] for the <strong>weight</strong> zero case). In that region, both<br />

families are absolutely convergent and define a holomorphic function in s. Also, P m (z, s) is<br />

an eigenfunction <strong>of</strong> −∆ 1 with eigenvalue (s − 1/2)((1 − s) − 1/2) and belongs to F 1 (M, ν).<br />

Let ∆ ′ k be the differential operator defined by<br />

( )<br />

−∆ ′ k := −y 2 ∂ 2<br />

+ ∂2 + iky ∂ .<br />

∂x 2 ∂y 2 ∂x<br />

It is related to ∆ k by the following equation<br />

Define the space D 1 (M, ν) by<br />

∆ ′ k + k 2<br />

(<br />

1 −<br />

k<br />

2)<br />

= y k/2 ∆ k y −k/2 .<br />

D 1 (M, ν) := {y 1/2 f(z) : f(z) ∈ F 1 (M, ν) is smooth with compact support on H}<br />

Let ˜D 1 (M, ν) be the completion <strong>of</strong> D 1 (M, ν) with respect to the Petersson norm<br />

∫<br />

||g|| 2 = 〈g, g〉 := |g(z)| 2 dxdy .<br />

y 2<br />

Γ 0 (M)\H<br />

Satz 3.2 in [35] implies that −∆ ′ k has a self-adjoint extension − ˜∆ ′ k from D 1(M, ν) to ˜D 1 (M, ν).<br />

Also, − ˜∆ ′ k has a countable system <strong>of</strong> Maass cusp <strong>forms</strong> {e n(z)} n∈N ⊂ ˜D 1 (M, ν) forming the<br />

discrete spectrum. Each Maass cusp form e n (z) has eigenvalue λ n and each eigenvalue has<br />

finite multiplicity. For any f ∈ ˜D 1 (M, ν), the discrete spectrum contributes the following<br />

sum to its spectral expansion<br />

∞∑<br />

〈f, e n 〉e n (z),<br />

n=1<br />

which converges absolutely and uniformly for all z ∈ H [35, Satz 8.1].<br />

Let ι be a cusp <strong>of</strong> Γ 0 (M), σ ι ∈ GL 2 (R) the scaling matrix sending ∞ to ι, and Γ ι ⊂ Γ 0 (M)<br />

the subgroup fixing ι. One can define the <strong>weight</strong> <strong>one</strong> real analytic Eisenstein series E ι (z, s)<br />

by<br />

E ι (z, s) := y1/2<br />

2<br />

∑<br />

γ∈Γ ι\Γ 0 (M)<br />

ν(γ)<br />

cz+d (Im(σ−1 ι γz)) s−1/2 .<br />

Selberg showed that the Eisenstein series has meromorphic continuation to the whole complex<br />

plane in s. For ι ranging over the cusps <strong>of</strong> Γ 0 (M), the Eisenstein series {E ι (z, s)} ι make up


10 W. DUKE AND Y. LI<br />

the continuous spectrum <strong>of</strong> − ˜∆ ′ 1. So for any f ∈ ˜D 1 (M, ν), the contribution <strong>of</strong> the Eisenstein<br />

series to the spectral expansion <strong>of</strong> f is<br />

1<br />

4π<br />

∑<br />

ι<br />

∫ ∞<br />

−∞<br />

〈f(·), E ι (·, 1 2 + ir)〉E ι(z, 1 2 + ir)dr.<br />

If f is smooth, then the integral in r converges absolutely and uniformly for z in any fixed<br />

compact subset <strong>of</strong> H [35, Satz 12.3]. Applying the completeness theorem [35, Satz 7.2 ], we<br />

have the spectral expansion for any smooth f ∈ ˜D 1 (M, ν) in the form<br />

∞∑<br />

(19) f(z) = e n 〉e n (z) +<br />

n=1〈f, ∑ ι<br />

1<br />

4π<br />

∫ ∞<br />

−∞<br />

〈f(·), E ι (·, 1 2 + ir)〉E ι(z, 1 2 + ir)dr,<br />

where the sum over n ∈ N and the integral in r both converge uniformly and absolutely for<br />

z in any compact set <strong>of</strong> H.<br />

By setting f(z) = y 1/2 Q m (z, s) and comparing P m (z, s) to Q m (z, s) as in [31], we can<br />

deduce the following proposition.<br />

Proposition 2.1. For any positive integer m, the Poincaré series Q m (z, s) and P m (z, s)<br />

have analytic continuation to Re(s) > 0. At s = 1/2, P m (z, s) has at most a simple pole.<br />

Furthermore, the residues <strong>of</strong> P m (z, s) at s = 1/2 generate S 1 (M, ν).<br />

Pro<strong>of</strong>. First, we will prove the analytic continuation Q m (z, s). For any m > 0, the function<br />

˜Q m (z, s) := y 1/2 Q m (z, s) is square integrable for Re(s) > 1, hence contained in ˜D 1 (M, ν). So<br />

we can write out its spectral expansion<br />

(20)<br />

˜Q m (z, s) = D(z, s) + C(z, s),<br />

D(z, s) :=<br />

∞∑<br />

〈 ˜Q m (·, s), e n (·)〉e n (z),<br />

n=1<br />

C(z, s) := 1<br />

4π<br />

∑<br />

ι<br />

∫ ∞<br />

−∞<br />

〈 ˜Q m (·, s), E ι (·, 1 2 + ir)〉E ι(z, 1 2 + ir)dr,<br />

where D(z, s) and C(z, s) are the contributions from the discrete and continuous spectrum<br />

<strong>of</strong> − ˜∆ ′ 1 respectively.<br />

By Satz 5.2 and Satz 5.5 in [35], e n (z) has eigenvalue λ n ∈ [1/4, ∞) under − ˜∆ ′ 1. If<br />

λ n = 1/4, then y −1/2 e n (z) is in S 1 (M, ν). Since each eigenvalue has finite multiplicity, we<br />

can let R ≥ 0 such that {y −1/2 e n (z) : 1 ≤ n ≤ R} is an orthonormal basis <strong>of</strong> S 1 (M, ν). Note<br />

S 1 (M, ν) could be empty, in which case R = 0.<br />

For n ∈ N, let t n = √ λ n − 1/4, s n = 1/2 + it n . We can use equation (66) in [20], the<br />

bound M µ,ν (y) = O µ,ν (e y ) as y → ∞ and the vanishing property <strong>of</strong> cusp <strong>forms</strong> at all cusps<br />

to write<br />

e n (z) =<br />

∞∑<br />

u=−∞,u≠0<br />

c n,u W sgn(u)<br />

,s 1 (4π|u|y)e 2πiux ,<br />

2 n−<br />

2<br />

where W µ,ν (z) is the W -Whittaker function and c n,u are constants. Note if t n = 0, i.e.<br />

y −1/2 e n (z) is a holomorphic cusp form, then W sgn(u)/2,sn−1/2(4π|u|y) = (4π|u|y) 1/2 e −2π|u|y and


MOCK-MODULAR FORMS OF WEIGHT ONE 11<br />

c n,u = 0 for u ≤ 0. Applying the Rankin-Selberg unfolding trick to 〈 ˜Q m (·, s), e n (·)〉 gives us<br />

∫ ∞<br />

〈 ˜Q m (·, s), e n (·)〉 = (4π|m|) 1/2 e −2π|m|y (4π|m|y) s−1 c n,m W sgn(m)<br />

0<br />

= (4π|m|) 1/2 Γ ( s − 1<br />

c − it ) (<br />

2 n Γ s −<br />

1<br />

+ it )<br />

2 n<br />

(21)<br />

n,m ( )<br />

Γ s − sgn(m)<br />

2<br />

2 ,s n− 1 2<br />

(4π|m|y) dy<br />

y<br />

The last step uses the Mellin transform <strong>of</strong> the W -Whittaker function [2, eq. (8b)] and<br />

the substitution s n = 1/2 + it n . When Re(s) > 1, the sum defining D(z, s) is absolutely<br />

convergent [35, Satz 8.1 ], since y 1/2 Q m (z, s) ∈ ˜D 1 (M, ν) for Re(s) > 1. When 0 < Re(s) ≤ 1,<br />

we can write D(z, s) as<br />

D(z, s) = (4π|m|) ∑ 1/2 Γ ( s + 1 − 1<br />

c − it ) (<br />

2 n Γ s + 1 −<br />

1<br />

+ it )<br />

2 n s − sgn(m)<br />

n,m (<br />

)<br />

2<br />

e n (z) ·<br />

n∈N<br />

Γ s + 1 − sgn(m)<br />

(s − 1<br />

2<br />

2 )2 + t 2 n<br />

Since t 2 n = λ n − 1/4 and ∑ n>M λ−2 n converges [35, Satz 8.1], we can apply Cauchy-Schwarz<br />

inequality to see that the sum on the right hand side above converges absolutely on compact<br />

subsets <strong>of</strong> {s ∈ C : Re(s) > 0, s ≠ 1/2}. At s = 1/2, the first R terms in the sum produce a<br />

simple pole since t n = 0 for all 1 ≤ n ≤ R. The rest <strong>of</strong> the sum still converges absolutely. So<br />

the right hand side above gives the analytic continuation <strong>of</strong> D(z, s) to Re(s) > 0.<br />

For the continuous spectrum, the contribution C(z, s) can be treated similarly. For any<br />

cusp ι, we can write the Fourier expansion <strong>of</strong> E ι (z, s) at infinity in the following form (for<br />

ι = ∞, see [20, eq. (76)’])<br />

E ι (z, s) = y s + ψ ι (s)y 1−s + ∑ ψ ι,m (s)W sgn(m)<br />

,s− 1 (4π|m|y)e 2πimx ,<br />

2 2<br />

m≠0<br />

where ψ ι (s) and ψ ι,m (s) are products <strong>of</strong> gamma factors and Selberg-Kloosterman zeta functions.<br />

It is well-known that The Eisenstein series can be analytically continued in s to the<br />

whole complex plane. When Re(s) > 1/2, the poles <strong>of</strong> E ι (z, s) are in the interval s ∈ (1/2, 1]<br />

[35, Satz 10.3 ]. On the line Re(s) = 1/2, E ι (z 0 , s) is holomorphic in s for any fixed z 0 ∈ H<br />

[35, Satz 10.4]. So both ψ ι (s) and ψ ι,m (s) admit analytic continuation to Re(s) > 0 and are<br />

holomorphic on Re(s) = 1/2. Using the same unfolding trick above, we can evaluate<br />

〈 ˜Q m (·, s), E ι (·, 1 + ir)〉 = ψ ( 1 2 ι,m + ir) (4π|m|) 1/2 Γ ( s − 1 − ir) Γ ( s − 1 2 ( ) + ir) 2 .<br />

2<br />

Γ s − sgn(m)<br />

2<br />

Then C(z, s) can be written as<br />

C(z, s) = ∑ ι<br />

(22)<br />

= ∑ ι<br />

∫<br />

1 ∞<br />

4π −∞<br />

∫<br />

1 ∞<br />

4π −∞<br />

(<br />

ψ 1<br />

ι,m + ir) (4π|m|) 1/2 Γ ( s − 1 − ir) Γ ( s − 1 2<br />

( ) + ir) 2<br />

E<br />

2 ι (z, 1 + ir)dr<br />

Γ s − sgn(m)<br />

2<br />

2<br />

(<br />

ψ 1<br />

ι,m + ir) (4π|m|) 1/2 Γ ( s + 1 − 1 − ir) Γ ( s + 1 − 1<br />

(<br />

2<br />

)<br />

+ ir) 2<br />

2<br />

Γ s + 1 − sgn(m)<br />

2<br />

(s − sgn(m) )<br />

2<br />

·<br />

(s − 1 − ir)(s − 1 + ir) · E ι(z, 1 + ir)dr.<br />

2<br />

2 2


12 W. DUKE AND Y. LI<br />

When Re(s) > 0, C(z, s + 1) is absolutely convergent. So when Re(s) ≠ 1/2, we can apply<br />

Cauchy-Schwarz to bound expression (22) by<br />

∫ ∣<br />

∞<br />

s − sgn(m) ∣∣∣∣<br />

2<br />

(23)<br />

∣(s − 1 dr.<br />

2 )2 + r 2<br />

−∞<br />

When Re(s) > 1/2, let ˜C(z, s) be the expression (22), which gives the analytic continuation<br />

<strong>of</strong> C(z, s) in this region. When Re(s) < 1/2, we can define ˜C(z, s) in a similar fashion as in<br />

§6 <strong>of</strong> [31]. We start with (22) and Re(s) ∈ (1/2, 1/2 + ɛ) for some small ɛ > 0, then deform<br />

the contour so that it goes above 1/2−s and below s−1/2 . In the process, the following residue<br />

i<br />

i<br />

is picked up<br />

(24)<br />

(π|m|) 1/2 Γ(2s − 1)<br />

( ) ∑<br />

Γ s − sgn(m)<br />

2<br />

ι<br />

ψ ι,m (s)E ι (z, s) + ψ ι,m (1 − s)E ι (z, 1 − s).<br />

Finally we can reduce the real part <strong>of</strong> s to less than 1/2 and change the contour back to<br />

the real line. So when Re(s) < 1/2, let ˜C(z, s) be the sum <strong>of</strong> (24) and (22), which are<br />

holomorphic. The bound (23) and m ≥ 1 guarantees the existence <strong>of</strong> the limits <strong>of</strong> ˜C(z, s) as<br />

Re(s) approaches 1/2 from the left and from the right. The procedure defining ˜C(z, s) shows<br />

that the limits agree. So we define ˜C(z, s) on Re(s) = 1/2 to be this limit. Then ˜C(z, s) gives<br />

the analytic continuation <strong>of</strong> C(z, s) to Re(s) > 0. Putting these together with the analytic<br />

continuation <strong>of</strong> D(z, s), we have the analytic continuation <strong>of</strong> Q m (z, s) to Re(s) > 0.<br />

Now to analytically continue P m (z, s), we can simply compare it to Q m (z, s). Applying<br />

the power series expansions <strong>of</strong> 1 F 1 (α; β; z) and the exp<strong>one</strong>ntial map to (16) and (17), we see<br />

that for y small and 0 < Re(s) < 2,<br />

|(4π|m|y) −s+1/2 (φ ∗ m(z, s) − ϕ ∗ m(z, s))| = O m (y).<br />

So the difference P m (z, s) − Q m (z, s) defined by term-wise subtraction is a holomorphic function<br />

in s for Re(s) > 0. That means the analytic continuation <strong>of</strong> P m (z, s) to Re(s) > 0<br />

follows from that <strong>of</strong> Q m (z, s). Furthermore, they have the same poles and residues in the<br />

region Re(s) > 0. Thus, to prove the second half <strong>of</strong> the proposition, it suffices to analyze the<br />

poles and residues <strong>of</strong> Q m (z, s) at s = 1/2.<br />

Since m > 0, expression (23) is bounded by an absolute constant for all s ∈ (1/2, 2]. So<br />

˜C(z, s) does not contribute to the pole at s = 1/2. For a Maass cusp form e n (z), the right<br />

hand side <strong>of</strong> (21) vanishes at s = 1/2 if t n ≠ 0. Otherwise, y −1/2 e n (z) ∈ S 1 (M, ν) and<br />

〈 ˜Q m (·, s), e n (·)〉 has a simple pole <strong>of</strong> residue (4π|m|) 1/2 c n,m . Thus, we have<br />

∑ R<br />

(25) Res s=1/2 P m (z, s) = Res s=1/2 Q m (z, s) = (4π|m|) 1/2 c n,m y −1/2 e n (z).<br />

Since {y −1/2 e n (z) : n = 1, . . . , R} is a basis <strong>of</strong> S 1 (M, ν), the matrix<br />

{c n,m : 1 ≤ n ≤ R, 1 ≤ m ≤ K}<br />

has rank equals to R for K sufficiently large. Therefore the residues at s = 1/2 <strong>of</strong> P m (z, s)<br />

generate S 1 (M, ν).<br />

□<br />

The following theorem is an immediate consequence <strong>of</strong> the proposition above.<br />

Theorem 2.2. Using the notations above, the following map is a surjection<br />

ξ 1 : H 1 (M, ν) → S 1 (M, ν),<br />

n=1


MOCK-MODULAR FORMS OF WEIGHT ONE 13<br />

i.e. for any cusp form h(z) ∈ S 1 (M, ν), there exists ˜h(z) ∈ M 1 (M, ν) with shadow h(z).<br />

Pro<strong>of</strong>. When k = 1, equation (2) becomes ∆ 1 = ξ 1 ◦ ξ 1 . So the Poincaré series P m (z, s)<br />

satisfies<br />

(26) ∆ 1 (P m (z, s)) = ( s − 1 2) 2<br />

Pm (z, s)<br />

when Re(s) > 1. Since the difference between both sides is holomorphic in s and P m (z, s)<br />

can be analytically continued to Re(s) > 0 as in Proposition 2.1, equation (26) is valid for<br />

Re(s) > 0. At s = 1/2, suppose P m (z, s) has the following Taylor series expansion in s<br />

P m (z, s) = g −1 (z) ( s − 1 2) −1<br />

+ g0 (z) + g 1 (z) ( s − 1 2)<br />

+ Oz<br />

( (s<br />

−<br />

1<br />

2<br />

) 2<br />

)<br />

,<br />

with g j (z) ∈ F 1 (M, ν) real-analytic for j = −1, 0, 1. Since ξ 1 commutes with the slash<br />

operator and ∆ 1 (g 1 (z)) = ξ1(g 2 1 (z)) = g −1 (z), we have ξ 1 (g 1 (z)) ∈ F 1 (M, ν) is real-analytic<br />

and a preimage <strong>of</strong> g −1 (z) under ξ 1 .<br />

By considering the Fourier expansion <strong>of</strong> g 1 (z), we know that if g 1 (z) has at most linear<br />

exp<strong>one</strong>ntial growth near the cusps, then ξ 1 (g 1 (z)) has the same property. Suppose Q m (z, s)<br />

has the Laurent expansion<br />

Q m (z, s) = g −1 (z) ( s − 1 2) −1<br />

+ g<br />

′<br />

0 (z) + g ′ 1(z) ( s − 1 2)<br />

+ Oz<br />

( (s<br />

−<br />

1<br />

2<br />

near s = 1/2. Then from the spectral expansion <strong>of</strong> Q m (z, s), i.e. y −1/2 (D(z, s) + C(z, s))<br />

as in (20), it is not hard to see that g 1(z) ′ has at most linear exp<strong>one</strong>ntial growth near the<br />

cusps. The difference P m (z, s) − Q m (z, s) is a Poincaré series defined for Re(s) > 0. So the<br />

coefficient <strong>of</strong> (s − 1/2) <strong>of</strong> its Laurent series expansion around s = 1/2, say g 1(z), ′′ has at most<br />

linear exp<strong>one</strong>ntial growth near the cusps. That means the sum <strong>of</strong> g 1(z) ′ and g 1(z), ′′ which is<br />

g 1 (z) by analytic continuation, also has this property. Thus, ξ 1 (g 1 (z)) is in H 1 (M, ν) and a<br />

preimage <strong>of</strong> g −1 (z) ∈ S 1 (M, ν) under ξ 1 . Since the functions {Res s=1/2 P m (z, s) : m ≥ 1} span<br />

the space S 1 (M, ν), the map ξ 1 : H 1 (M, ν) → S 1 (M, ν) is surjective.<br />

□<br />

3. Regularized Inner Products and the Weight One Plus Space<br />

From now on, we fix M = p to be a prime number congruent to 3 modulo 4 and ν = χ p . The<br />

spaces H 1(p, ! χ p ), H 1 (p, χ p ), M 1(p, ! χ p ), M 1 (p, χ p ), S 1 (p, χ p ), M ! 1(p, χ p ), M 1 (p, χ p ) are the same<br />

as before, and we will drop χ p in these notations. In this section, we will relate the regularized<br />

inner products between g(z) ∈ S 1 (p) and f(z) ∈ M 1(p) ! to linear combinations <strong>of</strong> coefficients<br />

<strong>of</strong> a mock-<strong>modular</strong> form ˜g(z), whose shadow is g(z), via Stokes’ theorem. This technique<br />

has been used in [9], [11] and [18]. Then, we will split up the space M ! 1(p) canonically into<br />

the plus and minus spaces to facilitate further analysis. Whenever convenient, we will switch<br />

between M ! 1(p) and H 1(p).<br />

!<br />

Given f(z) ∈ M 1(p) ! and g(z) ∈ S 1 (p), the usual Petersson inner product 〈f, g〉 can be<br />

regularized as follows. Since p is prime, ( Γ 0 (p) has two inequivalent cusps, 0 and ∞. They<br />

−1/<br />

are related by the scaling matrix σ 0 =<br />

√ p<br />

√ p<br />

). Take a fundamental domain <strong>of</strong> Γ 0 (p)\H,<br />

cut <strong>of</strong>f the portion with Im(z) > Y for a large Y and intersect it with its translate under σ 0 .<br />

We will call this the truncated fundamental domain F(Y ). Now, define the regularized inner<br />

product by<br />

(27) 〈f, g〉 reg := lim<br />

Y →∞<br />

∫<br />

F(Y )<br />

f(z)g(z)y dxdy<br />

y 2 .<br />

) 2<br />

)


14 W. DUKE AND Y. LI<br />

If f(z) ∈ M 1 (p), then this is the usual Petersson inner product. Now let ĝ(z) ∈ H 1 (p) be a<br />

preimage <strong>of</strong> g(z) under ξ 1 with the following Fourier expansions<br />

where W p = (<br />

−1<br />

p<br />

ĝ(z) = ∑ n∈Z<br />

c + ∞(n)q n − ∑ n≥1<br />

c(g, n)β 1 (n, y)q −n ,<br />

(ĝ| 1 W p )(z) = ∑ c + 0 (n)q n − ∑ c(g| 1 W p , n)β 1 (n, y)q −n ,<br />

n∈Z<br />

n≥1<br />

)<br />

is the Fricke involution and it acts on f(z) ∈ H<br />

!<br />

1 (p) by<br />

( )<br />

(f| 1 W p )(z) = √ 1<br />

pz<br />

f − 1 .<br />

pz<br />

The expression for (ĝ| 1 W p )(z) follows from the commutativity between ξ 1 and the slash<br />

operator. Note that ∑ n∈Z c+ ∞(n)q n and ∑ n∈Z c+ 0 (n)q n are mock-<strong>modular</strong> <strong>forms</strong> with shadows<br />

g(z) and (g| 1 W p )(z) respectively.<br />

Suppose f(z) ∈ M 1(p) ! has the Fourier expansion ∑ n∈Z c ∞(f, n)q n and ∑ n∈Z c 0(f, n)q n at<br />

the cusp infinity and 0 respectively. Then we can express 〈f, g〉 reg in terms <strong>of</strong> these Fourier<br />

coefficients.<br />

Lemma 3.1. Let f(z) ∈ M ! 1(p) and g(z) ∈ S 1 (p). In the notations above, we have<br />

(28) 〈f, g〉 reg = ∑ n∈Z<br />

c + ∞(n)c ∞ (f, −n) + c + 0 (n)c 0 (f, −n).<br />

Pro<strong>of</strong>. Since g(z) = ξ 1 (ĝ(z)) = −2iy ∂ĝ<br />

∂z<br />

(29) 〈f, g〉 reg = − lim<br />

Y →∞<br />

and dzdz = 2idxdy, we can rewrite equation (27) as<br />

∫<br />

f(z) ∂ĝ<br />

∂z dzdz.<br />

F(Y )<br />

As an application <strong>of</strong> Stokes’ Theorem, we have<br />

∫ (<br />

f(z) ∂ĝ<br />

∂z + ∂f ) ∮<br />

∂z ĝ(z) dzdz =<br />

F(Y )<br />

∂F(Y )<br />

f(z)ĝ(z)dz,<br />

where the line integral is in the counterclockwise direction. Using the <strong>modular</strong>ity <strong>of</strong> f(z)<br />

and ĝ(z), we can cancel most <strong>of</strong> the contour integral on the right hand side, except the<br />

contributions from the horizontal segment from − 1 + iY and 1 + iY near ∞ and the arc<br />

2 2<br />

center at 0 connecting the points σ 0 (± 1 ∂f<br />

+ iY ). Also, f(z) is holomorphic, hence = 0.<br />

2 ∂z<br />

Thus, the equation above becomes<br />

∫<br />

f(z) ∂ĝ ∫ −<br />

1<br />

F(Y ) ∂z dzdz = 2 +iY<br />

f(z)ĝ(z) + (f| 1 W p )(z)(ĝ| 1 W p )(z)dz.<br />

1<br />

2 +iY<br />

Now combining this with equations (29) and the bound |β 1 (n, y)| = O(e −4πny ), we obtain<br />

equation (28)<br />

□<br />

Remark. Notice that the right hand side <strong>of</strong> equation (28) depends on the choice <strong>of</strong> ĝ(z),<br />

whereas the left hand side only depends on g(z). So if we replace ĝ(z) with h(z) ∈ M ! 1(p),<br />

then Lemma 3.1 still holds and we obtain<br />

0 = ∑ n∈Z<br />

c ∞ (h, n)c ∞ (f, −n) + c 0 (h, n)c 0 (f, −n),<br />

where h(z) has Fourier expansions ∑ n∈Z c ∞(h, n)q n and ∑ n∈Z c 0(h, n)q n at the cusp infinity<br />

and 0 respectively.


MOCK-MODULAR FORMS OF WEIGHT ONE 15<br />

Now, we want to split up the space H ! 1(p) and M ! 1(p) canonically into the direct sum <strong>of</strong><br />

two subspaces. This decomposition behaves nicely with respect to ξ 1 and regularized inner<br />

product. For ɛ = ±1, we define the following space<br />

(30) H !,ɛ<br />

1 (p) := {f(z) = ∑ n∈Z<br />

a(n, y)q n ∈ H ! 1(p) : a(n, y) = 0 whenever χ p (n) = −ɛ}.<br />

By imposing the same condition on the Fourier expansions, we can define H1(p), ɛ M !,ɛ<br />

1 (p),<br />

M1(p), ɛ S1(p), ɛ M !,ɛ<br />

1 (p) and M ɛ 1(p). Notice that M !,ɛ<br />

1 (p) is the same as the space A ɛ k (p, χ p)<br />

defined on page 51 <strong>of</strong> [8] for k = 1. By the definition <strong>of</strong> R p (n) in (5), it is clear that<br />

E 1 (z) ∈ M 1 + (p). Also, the non-holomorphic <strong>weight</strong> <strong>one</strong> Eisenstein series Ẽ1(z) belongs to<br />

M !,−<br />

1 (p). This can be checked from the factorization (6) and the fact that E 1 (z) ∈ M 1 + (p).<br />

Clearly, H !,ɛ<br />

1 (p) is a subspace <strong>of</strong> H 1(p) ! for ɛ = ±1. In fact, the space H 1(p) ! can be<br />

decomposed into direct sum <strong>of</strong> these subspaces using the operators defined as follows. Let<br />

U p be the standard operator on H 1(p) ! defined by<br />

(31) f(z)| 1 U p = √ 1<br />

p−1<br />

∑ (<br />

f| 1 jp<br />

)<br />

1 .<br />

p<br />

Its action on the Fourier expansion <strong>of</strong> f(z) = ∑ m a(m, y)qm ∈ H 1(p) ! is as follows<br />

(32) (f| 1 U p )(z) = ∑ ( )<br />

a pm, y q m .<br />

p<br />

m<br />

It is easy to see that U p preserves the space H 1(p). ! This is also true for the Fricke involution<br />

W p , since χ p is a real character. Applying W p twice produces a negative sign for odd <strong>weight</strong>,<br />

i.e. (f| 1 Wp 2 )(z) = −f(z). Define the following operator on H 1(p) ! for ɛ ∈ {±1}<br />

j=0<br />

(<br />

ɛi(f| 1 U p W p ) + f<br />

(33) pr ɛ (f) := 1 2<br />

Using these operators, we can decompose H 1(p) ! into the direct sum <strong>of</strong> H !,+<br />

1 (p) and H !,−<br />

1 (p).<br />

On M ! 1(p), <strong>one</strong> can define U p and W p by considering their action on the associated harmonic<br />

Maass form in H 1(p) ! and obtain the same decomposition <strong>of</strong> spaces.<br />

Lemma 3.2. For ɛ = ±1, the operator pr ɛ is the identity operator when restricted to the<br />

subspace H !,ɛ<br />

1 (p) and the zero operator when restricted to the subspace H !,−ɛ<br />

1 (p). Equivalently,<br />

we have<br />

(34) (f| 1 W p )(z) = −iɛ(f| 1 U p )(z),<br />

for all f(z) ∈ H !,ɛ<br />

1 (p). Also, we can write<br />

(35) H ! 1(p) = H !,+<br />

1 (p) ⊕ H !,−<br />

1 (p).<br />

Similarly, we have this decomposition for M ! 1(p).<br />

Remark. The decomposition above clearly holds for subspaces <strong>of</strong> H ! 1(p) or M ! 1(p), such as<br />

H 1 (p), M ! 1(p), M 1 (p), S 1 (p) and M 1 (p).<br />

Pro<strong>of</strong>. Let f(z) = ∑ n∈Z a(n, y)qn ∈ H 1(p) ! and h := f| 1 U p W p . First, we will show that<br />

f ∈ H !,ɛ<br />

1 (p) if and only if pr −ɛ (f) = 0. With some matrix calculations, such as in Lemma 3<br />

<strong>of</strong> [8], we obtain<br />

(36) h(z) = √ 1<br />

∑<br />

(f| 1 W p V p )(z) + ε p χ p (n)a(n, y)q n ,<br />

p<br />

n∈Z<br />

)<br />

.


16 W. DUKE AND Y. LI<br />

where<br />

V p =<br />

( {<br />

p 1, p ≡ 1 (mod 4)<br />

, ε<br />

1)<br />

p =<br />

i, p ≡ 3 (mod 4)<br />

Since p ≡ 3 (mod 4), we know that ε p = i. From the definition <strong>of</strong> pr −ɛ , we have<br />

(37)<br />

(pr −ɛ f)(z) = 1 (−ɛih(z) + f(z))<br />

2<br />

(<br />

= 1 −ɛi<br />

√ (f| 1 W p V p )(z) + ∑ )<br />

(1 − χ p (n)ɛ) a(n, y)q n ∈ H !,ɛ<br />

2 p<br />

1 (p).<br />

n∈Z<br />

Notice that the action V p produces an expansion with coefficients supported on multiples <strong>of</strong><br />

p. So if pr −ɛ (f) = 0, then a(n, y) = 0 whenever χ p (n) = −ɛ and f ∈ H !,ɛ<br />

1 (p) by definition.<br />

Conversely if f ∈ H !,ɛ<br />

1 (p), then<br />

pr −ɛ (f) ∈ H !,+<br />

1 (p) ∩ H !,−<br />

1 (p),<br />

which means the coefficients ξ 1 (pr −ɛ (f)) ∈ M ! 1(p) are supported on multiples <strong>of</strong> p. A lemma<br />

due to Hecke ([33, Lemma 6, p.32] ) implies that<br />

(38) M !,+<br />

1 (p) ∩ M !,−<br />

1 (p) = {0}.<br />

So ξ 1 (pr −ɛ (f)) = 0 and pr −ɛ (f) is holomorphic and also contained in M !,+<br />

1 (p) ∩ M !,−<br />

1 (p). By<br />

the same argument, pr −ɛ (f) vanishes. From this, we can deduce that H !,ɛ<br />

1 (p) is the eigenspace<br />

<strong>of</strong> iU p W p with eigenvalue ɛ, and pr ɛ restricted to H !,ɛ<br />

1 (p) is the identity. Using the relationship<br />

(f| 1 Wp 2 )(z) = −f(z), we can then deduce equation (34). It is easy to see that we have the<br />

following direct sum decomposition<br />

H 1(p) ! = H !,+<br />

1 (p) ⊕ H !,−<br />

1 (p)<br />

f = pr + (f) + pr − (f).<br />

Applying the same procedure to the harmonic Maass form associated to any mock-<strong>modular</strong><br />

form in M ! 1(p), we have the same decomposition <strong>of</strong> space<br />

M ! 1(p) = M !,+<br />

1 (p) ⊕ M !,−<br />

1 (p).<br />

The decomposition <strong>of</strong> the space H 1(p) ! above behaves well under the action <strong>of</strong> ξ 1 . By Theorem<br />

2.2, we can prove a statement about surjectivity <strong>of</strong> ξ 1 : H1(p) ɛ → S1 −ɛ (p). Furthermore,<br />

there is a nice characterization <strong>of</strong> the space S1(p) ɛ as a consequence <strong>of</strong> facts about <strong>weight</strong> <strong>one</strong><br />

<strong>modular</strong> <strong>forms</strong>.<br />

Lemma 3.3. For ɛ = ±1, we have a surjection<br />

(39) ξ 1 : H ɛ 1(p) → S −ɛ<br />

1 (p).<br />

In other words, for any g(z) ∈ S −ɛ<br />

1 (p), there exists a mock-<strong>modular</strong> form ˜g(z) ∈ M ɛ 1(p) whose<br />

shadow is g(z).<br />

In addition, S + 1 (p) contains all dihedral cusp <strong>forms</strong>. If there are 2d − non-dihedral <strong>forms</strong> in<br />

S 1 (p), then the spaces S + 1 (p) and S − 1 (p) have dimensions 1 2 (H(p)−1)+d − and d − respectively<br />

Pro<strong>of</strong>. By Theorem 2.2, the following map is surjective<br />

(40) ξ 1 : H 1 (p) → S −ɛ<br />

1 (p).<br />


MOCK-MODULAR FORMS OF WEIGHT ONE 17<br />

Since ξ 1 commutes with the action <strong>of</strong> U p and W p and conjugates their coefficients, it commutes<br />

with the operator pr ɛ as follows<br />

(41) ξ 1 (pr ɛ (f)) = pr −ɛ (ξ 1 (f)).<br />

The operator pr −ɛ is the identity when restricted to S1 −ɛ (p). So the preimage <strong>of</strong> S1 −ɛ (p) under<br />

ξ 1 lies in H1(p) ɛ and the map ξ 1 : H1(p) ɛ → S1 −ɛ (p) is surjective for ɛ = ±1.<br />

Now if f ∈ S 1 (p) is a dihedral newform, then it is a linear combination <strong>of</strong> theta series, whose<br />

n th coefficient is zero if χ p (n) = −1. So f ∈ S 1 + (p) by definition. If f(z) = ∑ n≥1<br />

c(f, n)qn<br />

is a octahedral or icosahedral newform, then f(z) and f(z) := f(z) = ∑ n≥1 c(f, n)qn are<br />

linearly independent new<strong>forms</strong>. When l ≠ p is a prime number, we have the relationship<br />

(42) c(f, l) = χ p (l)c(f, l).<br />

This is a consequence <strong>of</strong> the formula <strong>of</strong> the adjoint <strong>of</strong> the l th Hecke operator with respect to<br />

the Petersson inner product [34, p.21]. By the definition <strong>of</strong> S1(p), ɛ we have f + f ∈ S 1 + (p)<br />

and f − f ∈ S1 − (p). Since the number <strong>of</strong> octahedral and icosahedral new<strong>forms</strong> is set to be<br />

2d − , we obtain the formulae for the dimensions.<br />

□<br />

Remark: The spaces S ɛ 1(p) should be compared to the spaces M + and M − in (9.1.2) in<br />

[36]. If all octahedral and icosahedral new<strong>forms</strong> f(z) ∈ S 1 (p) satisfy<br />

(43) (f| 1 W p )(z) = −if(z),<br />

then M + is the span <strong>of</strong> the <strong>weight</strong> <strong>one</strong> Eisenstein series and S + 1 (p) and M − = S − 1 (p).<br />

Besides compatibility with ξ 1 , the decomposition also behaves nicely with respect to the<br />

regularized inner product. As a consequence <strong>of</strong> Lemma 3.1, we have the following proposition.<br />

Proposition 3.4. Let f(z) ∈ M !,ɛ<br />

1 (p), g(z) ∈ S ɛ′<br />

1 (p) and ˜g(z) ∈ M −ɛ′<br />

1 (p) with shadow g(z)<br />

and Fourier expansion ∑ n∈Z c+ (n)q n . If ɛ = ɛ ′ , then<br />

(44) 〈f, g〉 reg = ∑ n∈Z<br />

δ(n)c(f, −n)c + (n).<br />

Here δ(n) is 2 if p|n and 1 otherwise. If ɛ ≠ ɛ ′ , then 〈f, g〉 reg = 0.<br />

Pro<strong>of</strong>. Let ĝ(z) ∈ H −ɛ′<br />

1 (p) be the harmonic Maass form associated to ˜g(z). Then both<br />

assertions are immediate consequences <strong>of</strong> Lemma 3.1 and<br />

(f| 1 W p )(z) = −iɛ(f| 1 U p )(z), (ĝ| 1 W p )(z) = −iɛ ′ (ĝ| 1 U p )(z).<br />

Since the adjoint <strong>of</strong> W p (resp. U p ) is W −1<br />

p<br />

= −W p (resp. W p U p Wp<br />

−1 ) with respect to the<br />

regularized inner product, it is easy to check that pr ɛ is self-adjoint, which also proves the<br />

second claim.<br />

□<br />

When g(z) above is zero, i.e. ˜g(z) is <strong>modular</strong>, equation (44) reduces to an orthogonal<br />

relationship between Fourier coefficients <strong>of</strong> weakly holomorphic <strong>modular</strong> <strong>forms</strong>. In particular,<br />

we can take ˜g(z) to be cusp <strong>forms</strong> and obtain relations satisfied by {c(f, −n) : n ≥ 1}. These<br />

relations turn out to characterize the space M !,ɛ<br />

1 (p), which gives a nice characterization <strong>of</strong><br />

the space M ɛ 1(p).<br />

Proposition 3.5. Let g(z) ∈ S1(p). ɛ Then there exists a mock-<strong>modular</strong> form ˜g(z) ∈ M −ɛ<br />

1 (p)<br />

with shadow g(z) and prescribed principal part ∑ n


18 W. DUKE AND Y. LI<br />

Pro<strong>of</strong>. The necessity part follows from Proposition 3.4. Since h(z) is a cusp form, the summation<br />

in (44) is only over n < 0. When ˜g(z) ∈ M −ɛ<br />

1 (p) is <strong>modular</strong>, the sufficiency follows<br />

from an application <strong>of</strong> Serre duality [5, §3] (also see [8, Theorem 6]).<br />

When ˜g(z) is mock-<strong>modular</strong>, let ∑ n


MOCK-MODULAR FORMS OF WEIGHT ONE 19<br />

For convenience, we write E(τ, s) for E 1 (τ, s). Kr<strong>one</strong>cker’s first limit formula states that<br />

(49) ζ(2s)E(τ, s) = π<br />

s − 1 + 2π ( γ − log(2) − log( √ y|η(τ)| 2 ) ) + O(s − 1),<br />

where γ is the Euler constant. Using (49) and Rankin-Selberg unfolding trick, we can relate<br />

the inner product between dihedral new<strong>forms</strong> to logarithm <strong>of</strong> u A as follows.<br />

Proposition 4.1. Let ψ, ψ ′ be characters <strong>of</strong> Cl(F ), with ψ non-trivial. When ψ ′ = ψ or ψ,<br />

we have<br />

(50) 〈g ψ , g ψ ′〉 = 1 ∑<br />

ψ 2 (A) log |u A |.<br />

12<br />

Otherwise, 〈g ψ , g ψ ′〉 = 0.<br />

A∈Cl(F )<br />

Pro<strong>of</strong>. Since p is prime, the Eisenstein series E p (z, s) has a simple pole at s = 1 with residue<br />

, which is independent <strong>of</strong> z. So we have the relationship<br />

3<br />

π(p+1)<br />

3<br />

π(p + 1) · 〈g ψ, g ψ ′〉 = Res s=1 g ψ (z)g ψ ′(z)E p (z, s)y<br />

∫Γ dxdy .<br />

0 (p)\H<br />

y 2<br />

Now, we can use the Rankin-Selberg method to unfold the right hand side and obtain<br />

∫<br />

g ψ (z)g ψ ′(z)E p (z, s)y dxdy = Γ(s) ∑ r ψ (n)r ψ ′(n)<br />

y 2 (4π) s n s<br />

Γ 0 (p)\H<br />

Let ρ ψ : Gal(Q/Q) → GL 2 (C) be the representation induced from ψ. Then it is also the<br />

<strong>one</strong> attached to g ψ (z) via Deligne-Serre’s theorem. Up to Euler factors at p, the right hand<br />

side is L(s, ρ ψ ⊗ ρ ψ ′), the L-function <strong>of</strong> the tensor product <strong>of</strong> the representations ρ ψ and ρ ψ ′.<br />

From the character table <strong>of</strong> the dihedral group D 2H(p) , we see that<br />

ρ ψ ⊗ ρ ψ ′ = ρ ψψ ′ ⊕ ρ ψψ ′.<br />

So when ψ ′ ≠ ψ or ψ ′ , the L-function L(s, ρ ψ ⊗ρ ψ ′) is holomorphic at s = 1 and 〈g ψ , g ψ ′〉 = 0.<br />

Otherwise, we have<br />

∑ r ψ (n)r ψ (n)<br />

= ζ(s)L(s, χ p)L(s, ρ ψ 2)<br />

,<br />

n s ζ(2s)(1 + p −s )<br />

n≥1<br />

Putting these together, we obtain<br />

(51) 〈g ψ , g ψ 〉 = p<br />

2π L(1, χ p)L(1, ρ 2 ψ 2).<br />

From the theory <strong>of</strong> quadratic <strong>forms</strong>, we have<br />

√ ∑<br />

pL(s, ρψ 2) = ζ(2s) ψ 2 (A)E(τ A , s).<br />

A∈Cl(F )<br />

Since ψ is non-trivial and H(p) is odd, ψ 2 is non-trivial and (49) implies that<br />

L(1, ρ ψ 2) = −√ 2π ∑<br />

ψ 2 (A) log( √ y A |η(τ A )| 2 ),<br />

p<br />

A∈Cl(F )<br />

Along with the class number formula for p > 3<br />

L(1, χ p ) = 2πH(p)<br />

w p<br />

√ p<br />

n≥1<br />

= πH(p) √ p<br />

,


20 W. DUKE AND Y. LI<br />

we arrive at<br />

(52) 〈g ψ , g ψ 〉 = −H(p)<br />

∑<br />

A∈Cl(F )<br />

Since ψ 2 is non-trivial, we have equation (50).<br />

ψ 2 (A) log( √ y A |η(τ A )| 2 ).<br />

Remarks.<br />

i) By the same procedure, <strong>one</strong> can analyze the inner product between any <strong>weight</strong> <strong>one</strong><br />

new<strong>forms</strong>. In particular, if g(z), h(z) ∈ S 1 (p) arises from different types <strong>of</strong> Galois representations,<br />

then 〈g, h〉 = 0.<br />

ii) This proposition is really a proven case <strong>of</strong> Stark’s conjecture in the abelian, rank <strong>one</strong><br />

case.<br />

Next we prove the existence <strong>of</strong> the special pre-image.<br />

Proposition<br />

∑<br />

4.2. Let ψ be a non-trivial character <strong>of</strong> Cl(F ). Then there exists ˜g ψ (z) =<br />

n≥n 0<br />

r + ψ (n)qn ∈ M − 1 (p) with shadow g ψ (z) such that<br />

(i) When χ p (n) = 1 or n < − p+1 , the coefficient 24 r+ ψ<br />

(n) equals to zero.<br />

(ii) For n ≤ 0, the coefficients r + ψ<br />

(n) are <strong>of</strong> the form<br />

r + ψ<br />

(n) = −β<br />

∑<br />

A∈Cl(F )<br />

ψ 2 (A) log |u(n, A)|,<br />

where β ∈ Q depends only on p, and u(n, A) ∈ O × H<br />

depends only on n and A.<br />

(iii) Let σ C ∈ Gal(H/F ) be the element associated to the class C ∈ Cl(F ) via Artin’s isomorphism.<br />

Then it acts on the units u(n, A) by<br />

σ C (u(n, A)) = u(n, AC −1 ).<br />

Pro<strong>of</strong>. From Lemma 3.3, we know that the dimension <strong>of</strong> S + 1 (p), denoted by d + , is 1 2 (H(p) −<br />

1)+d − . Let {g 1 , g 2 , . . . , g d+ } be a basis <strong>of</strong> S + 1 (p). Suppose we have chosen n 1 , n 2 , . . . , n d+ > 0<br />

such that χ p (n k ) ≠ −1 and the matrix<br />

P := [c(g j , n k )] 1≤j,k≤d+<br />

is invertible. Then for any ψ, there exists ˜g ψ (z) ∈ M − 1 (p) with shadow g ψ (z) and principal<br />

part coefficients r + ψ (−n) = 0 for all positive integer n not in the set {n 1, n 2 , . . . , n d+ } by<br />

Proposition 3.5. Furthermore, these r + ψ<br />

(−n) can be uniquely determined from solving a<br />

d + × d + system <strong>of</strong> equations. We will show that such ˜g ψ (z) can be made to satisfy the<br />

proposition.<br />

First, we can choose {n j : 1 ≤ j ≤ d + } such that n j ≤ (p + 1)/24. Let {g 1 , g 2 , . . . , g d+ } to<br />

be a q-echelon basis. Then these n j ’s can be chosen to be bounded by the supremum <strong>of</strong> the<br />

set<br />

{ord ∞ h(z) : h(z) ∈ S 1 + (p)}.<br />

Now, the square <strong>of</strong> a cusp form in S 1 (p) is in S 2 (p), the space <strong>of</strong> <strong>weight</strong> 2 cusp form <strong>of</strong> trivial<br />

nebentypus on Γ 0 (p). This gives rise to a holomorphic differential form on the <strong>modular</strong> curve<br />

X 0 (p). Atkin showed in [1] that ∞ is not a Weiestrass point <strong>of</strong> the <strong>modular</strong> curve X 0 (p),<br />

i.e. ord ∞ f(z) is no more than the genus <strong>of</strong> X 0 (p). Using the Riemann-Hurwitz formula (see<br />

[37]), we know that the genus <strong>of</strong> X 0 (p) is bounded by (p + 1)/12. So the n j ’s can be chosen<br />

to be bounded by (p + 1)/24.<br />


MOCK-MODULAR FORMS OF WEIGHT ONE 21<br />

Let {g j (z) : 1 ≤ j ≤ (H(p) − 1)/2} be the dihedral new<strong>forms</strong>, and label the class group<br />

characters as {ψ j : 1 ≤ j ≤ H(p) − 1} such that g ψj (z) = g j (z) and ψ j = ψ H(p)−j for 1 ≤ j ≤<br />

(H(p) − 1)/2. When (H(p) + 1)/2 ≤ j ≤ d + , let g j (z) be linear combinations <strong>of</strong> non-dihedral<br />

new<strong>forms</strong> such that g j (z) has integral coefficients and the set {g j (z) : (H(p) + 1)/2 ≤ j ≤ d + }<br />

is linearly independent. Then for each non-trivial ψ, the values {r + ψ (−n k) : 1 ≤ k ≤ d + }<br />

satisfies the equation<br />

P · [δ(n k )r + ψ (−n k)] 1≤k≤d+ = R,<br />

where R is the d + × 1 matrix (〈g j , g ψ 〉) 1≤j≤d+<br />

. By Proposition 4.1, the inner product 〈g j , g ψ 〉<br />

∑<br />

is either 0 or 1<br />

12 A∈Cl(F ) ψ2 (A) log |u A |. So we have<br />

⎛<br />

⎞<br />

0.<br />

∑<br />

0<br />

1<br />

R =<br />

12 A ψ2 (A) log |u A |<br />

⎜<br />

⎟<br />

⎝<br />

0. ⎠<br />

0<br />

Let A 0 be the principal class in Cl(F ) and label the non-principal ( classes as A i for 1 ≤<br />

i ≤ H(p) − 1 such that A −1 = A H(p)−i . Let M ′ M1 0<br />

be the matrix 0 Id d−<br />

), where M 1 is an<br />

i<br />

1<br />

(H(p) − 1) × 1 (H(p) − 1) matrix defined by<br />

2 2<br />

[ψ j (A i ) + ψ j (A H(p)−i ) − 2] 1≤i,j≤(H(p)−1)/2<br />

and Id d− is the d − × d − identity matrix. The product M ′ P is the matrix<br />

⎛<br />

⎞<br />

H(p)(r A1 (n 1 ) − r A0 (n 1 )) · · · H(p)(r A1 (n d+ ) − r A0 (n d+ ))<br />

. .<br />

.. .<br />

M ′ H(p)(r<br />

P =<br />

A(H(p)−1)/2 (n 1 ) − r A0 (n 1 )) · · · H(p)(r A(H(p)−1)/2 (n d+ ) − r A0 (n d+ ))<br />

,<br />

c(g<br />

⎜<br />

(H(p)+1)/2 , n 1 ) · · · c(g (H(p)+1)/2 , n d+ )<br />

⎝<br />

. ⎟<br />

.<br />

.. .<br />

⎠<br />

c(g d+ , n 1 ) · · · c(g d+ , n d+ )<br />

which has integer coefficients. It is also non-singular since<br />

{ϑ Aj − ϑ A0 : 1 ≤ j ≤ (H(p) − 1)/2} ∪ {g j : (H(p) + 1)/2 ≤ j ≤ d + }<br />

is a basis <strong>of</strong> S 1 + (p). The product M ′ R is the d + × 1 matrix<br />

⎛ ∑<br />

A ψ2 (A) log |u A (A 1 )|<br />

where<br />

Also by (47), we have<br />

M ′ R =<br />

⎜<br />

⎝<br />

1<br />

12<br />

1<br />

12<br />

⎞<br />

∑<br />

.<br />

A ψ2 (A) log |u A (A (H(p)−1)/2 )|<br />

,<br />

⎟<br />

0.<br />

⎠<br />

0<br />

u A (B) := u A √ B u A √ B −1<br />

u 2 A<br />

σ C (u A (B)) = u AC −1(B)<br />

∈ O × H .


22 W. DUKE AND Y. LI<br />

for any class A, B, C ∈ Cl(F ).<br />

Since M ′ P is a non-singular matrix with integer entries, its inverse, say (α i,j ) 1≤i,j≤d+ , has<br />

rational entries. Thus, r + ψ (−n i) can be written as<br />

δ(n i )r + ψ (−n i) = ∑ A<br />

(H(p)−1)/2<br />

∑<br />

ψ 2 (A) α i,j log |u A (A j )|<br />

j=1<br />

with α i,j ∈ Q for all 1 ≤ i, j ≤ d + . From this, we can choose β ∈ Q such that<br />

u(n i , A) =<br />

(H(p)−1)/2<br />

∏<br />

j=1<br />

u A (A j ) −α i,j/δ(n i )β ∈ O × H .<br />

α i,j<br />

2H(p)β ∈ Z and<br />

Then r + ψ (−n i) satisfies conditions (ii) and (iii) in the proposition. Furthermore, the unit<br />

u(n i , A) is an H(p) th power in O × H .<br />

Finally, apply Propositions 4.1 and 3.4 to the Eisenstein series E 1 (z) and the cusp form<br />

g ψ (z), we get<br />

∑<br />

δ(n)R p (n)r + ψ<br />

(−n) = 0.<br />

n≥0<br />

So we can write r + ψ<br />

(0) = −β log |u(0, A)| with<br />

u(0, A) =<br />

d +<br />

∏<br />

i=1<br />

which satisfies condition (iii) as well.<br />

u(n i , A) −δ(n i)R p(n i )/H(p) ∈ O × H ,<br />

□<br />

5. Rankin’s method and values <strong>of</strong> Green’s function I<br />

From the pro<strong>of</strong> <strong>of</strong> Proposition 4.2, we saw that information about r + ψ<br />

(n), n ≤ 0 can be<br />

retrieved from their various linear combinations coming from inner products. In this section,<br />

we will consider a different linear combination <strong>of</strong> the coefficients r + ψ<br />

(n). They can be viewed<br />

as the regularized inner product between g ψ (z) and certain weakly holomorphic <strong>forms</strong> in<br />

M !,+<br />

1 (p) lifted from <strong>modular</strong> functions <strong>of</strong> level <strong>one</strong>. Following the same procedure as in [24],<br />

we will show in Theorem 5.1 that these linear combination <strong>of</strong> r + ψ<br />

(n)’s can be expressed as the<br />

twisted sum <strong>of</strong> logarithms <strong>of</strong> the <strong>modular</strong> polynomial evaluated at certain CM points. This<br />

result is more or less a twisted version <strong>of</strong> the Gross-Zagier formula at level <strong>one</strong>, where there<br />

is no <strong>weight</strong> two cusp form or L-function. Also, the information provided by the theorem,<br />

as formulated in Corollary 5.2, will be a stepping st<strong>one</strong> to the pro<strong>of</strong> <strong>of</strong> Theorem 1.1. As a<br />

<strong>modular</strong> form on the degenerated Hilbert <strong>modular</strong> surface, the <strong>modular</strong> polynomial can be<br />

expressed in terms <strong>of</strong> a Borcherds lift <strong>of</strong> <strong>modular</strong> functions <strong>of</strong> level <strong>one</strong>. Thus, it is natural<br />

to phrase Theorem 5.1 in terms <strong>of</strong> a regularized inner product and this Borcherds lift. This<br />

prepares the way for §6, where we treat the level N case.<br />

For m ≥ 0, let j m (z) = q −m + O(q) be the unique <strong>modular</strong> function <strong>of</strong> level <strong>one</strong> with a<br />

pole <strong>of</strong> order m at infinity, e.g. j 0 (z) = 1 and j 1 (z) = j(z) − 744. For each B ∈ Cl(F ), choose<br />

an associated CM point τ B = x B + iy B ∈ H and define j lift,B<br />

m (z) ∈ M !,+<br />

1 (p) by<br />

jm<br />

lift,B (z) := j m (pz)ϑ B (z).


MOCK-MODULAR FORMS OF WEIGHT ONE 23<br />

Let M 2 (Z) be the space <strong>of</strong> 2 × 2 matrices with integer coefficients. For m ≥ 1, (τ 1 , τ 2 ) ∈<br />

(PSL 2 (Z)\H) 2 , define the function Ψ m (τ 1 , τ 2 ) by<br />

∏<br />

Ψ m (τ 1 , τ 2 ) :=<br />

(j(τ 1 ) − j(γτ 2 )) .<br />

γ∈SL 2 (Z)\M 2 (Z)<br />

det(γ)=m<br />

It is the value <strong>of</strong> the <strong>modular</strong> polynomial ϕ m (X, Y ) at X = j(τ 1 ), Y = j(τ 2 ), and also the<br />

Borcherds lift <strong>of</strong> j m (z) to a function on the degenerated Hilbert <strong>modular</strong> surface. The main<br />

result <strong>of</strong> this section is as follows.<br />

Theorem 5.1. Let ψ be a non-trivial character <strong>of</strong> Cl(F ) and m ≥ 1. Then<br />

∑<br />

( log |Ψm (τ<br />

(53) 〈jm<br />

lift,B , g ψ 〉 reg = −2 lim ψ 2 (A ′ A<br />

)<br />

′ B ′−1 + ɛ, τ A ′ B ′)| + )<br />

ɛ→0<br />

r B (m) log |y A ′ B ′−1| ,<br />

A ′ ∈Cl(F )<br />

where B ′ ∈ Cl(F ) is the unique class such that B ′2 = B.<br />

It will be clear in the pro<strong>of</strong> that equation (53) is independent <strong>of</strong> the choice <strong>of</strong> τ B . Before<br />

stating the pro<strong>of</strong>, we need a crucial result in [23]. For two distinct points τ j = x j + iy j ∈ H,<br />

the invariant hyperbolic distance d(τ 1 , τ 2 ) between them is defined by<br />

(54) cosh d(τ 1 , τ 2 ) = (x 1 − x 2 ) 2 + y1 2 + y2<br />

2 .<br />

2y 1 y 2<br />

Notice for d(τ 1 , τ 2 ) = d(γτ 1 , γτ 2 ) for all γ ∈ SL 2 (R). The Legendre function <strong>of</strong> the second<br />

kind Q s−1 (t) is defined by<br />

∫ ∞<br />

Q s−1 (t) = (t + √ t 2 − 1 cosh u) −s du, Re(s) > 1, t > 1,<br />

(55)<br />

0<br />

Q 0 (t) = 1 log ( )<br />

1 + 2<br />

2 t−1<br />

For two distinct points τ 1 , τ 2 ∈ PSL 2 (Z)\H, the following convergent series defines the automorphic<br />

Green’s function<br />

(56) G s (τ 1 , τ 2 ) := ∑<br />

g s (τ 1 , γτ 2 ), Re(s) > 1,<br />

where<br />

γ∈PSL 2 (Z)<br />

(57) g s (τ 1 , τ 2 ) := −2Q s−1 (cosh d(τ 1 , τ 2 )).<br />

Recall that E(τ, s) is defined in (48) and ϕ 1 (s) is the coefficient <strong>of</strong> y 1−s in the Fourier<br />

expansion <strong>of</strong> E(τ, s).<br />

Proposition 5.1 in [23] tells us that for distinct τ 1 , τ 2 ∈ PSL 2 (Z)\H, the values <strong>of</strong> the<br />

j-function is related to the values <strong>of</strong> the automorphic Green’s function by<br />

(58) log |j(τ 1 ) − j(τ 2 )| 2 = lim<br />

s→1<br />

(G s (τ 1 , τ 2 ) + 4πE(τ 1 , s) + 4πE(τ 2 , s) − 4πϕ 1 (s)) − 24.<br />

Apply the m th Hecke operator to τ 2 on both sides gives us [23, eq.(5.2)]<br />

( G<br />

m )<br />

(59) log |Ψ m (τ 1 , τ 2 )| 2 s (τ 1 , τ 2 ) + 4πσ 1 (m)E(τ 1 , s)+<br />

= lim<br />

s→1 4πm s σ 1−2s (m)E(τ 2 , s) − 4πσ 1 (m)ϕ 1 (s)<br />

where σ ν (m) = ∑ d|m dν and<br />

G m s (τ 1 , τ 2 ) =<br />

∑<br />

γ∈SL 2 (Z)\M 2 (Z)<br />

det(γ)=m<br />

g s (τ 1 , γτ 2 ).<br />

− 24σ 1 (m),


24 W. DUKE AND Y. LI<br />

Using this result and appropriate Rankin-Selberg unfolding trick as in [24], we can prove<br />

Theorem 5.1.<br />

Pro<strong>of</strong> <strong>of</strong> Theorem 5.1. Let ˜g ψ (z) ∈ M − 1 (p) be a mock-<strong>modular</strong> form as in Proposition 4.2<br />

with Fourier expansion<br />

˜g ψ (z) = ∑ r + ψ (n)qn<br />

n∈Z<br />

and ĝ ψ (z) the associated harmonic Maass form. For a class B ∈ Cl(F ), define<br />

Φ B (ĝ ψ , z) := Tr p 1(−i(ĝ ψ |W p )(z)ϑ B (z))<br />

= (ĝ ψ |U p )(z)ϑ B (z) + (ĝ ψ (z)ϑ B (z))|U p .<br />

This should be compared to ˜Φ s (z) in Proposition (1.2) <strong>of</strong> [24, §IV]. When N = 1, the<br />

derivative <strong>of</strong> ˜Φ s (z) at s = 1 is more or less our Φ B (ĝ ψ , z) for ψ trivial.<br />

Here, Φ B (ĝ ψ , z) trans<strong>forms</strong> like a level <strong>one</strong>, <strong>weight</strong> two <strong>modular</strong> form and has the following<br />

Fourier expansion at infinity<br />

Φ B (ĝ ψ , z) = ∑ n∈Z(b B (ĝ ψ , n, y) + a B (ĝ ψ , n))q n ,<br />

b B (ĝ ψ , n, y) = − ∑ k∈Z<br />

a B (ĝ ψ , n) = ∑ k∈Z<br />

δ(k)r ψ (k)β 1<br />

(k, y p<br />

δ(k)r + ψ (k) r B(pn − k).<br />

)<br />

r B (pn + k),<br />

By the choice <strong>of</strong> ˜g ψ (z), we have ord ∞ (˜g ψ (z)) ≥ − p+1 > −p. So the sum a 24 B(ĝ ψ , n) is zero for<br />

all n < 0. It is easy to see from the definition above that jm lift,B ∈ M !,+<br />

1 (p). So we can apply<br />

Proposition 3.4 to obtain<br />

(60)<br />

〈j lift,B<br />

m , g ψ 〉 reg = a B (ĝ ψ , m).<br />

In particular when m = 0, we have j lift,B<br />

0 = 1 lift,B = ϑ B (z) and<br />

〈ϑ B , g ψ 〉 = a B (ĝ, 0).<br />

Now, let Ẽ2(z) the non-holomorphic Eisenstein series <strong>of</strong> level <strong>one</strong>, <strong>weight</strong> two defined by<br />

Ẽ 2 (z) = 1 − 24 ∑ n≥1<br />

σ 1 (n)q n − 3<br />

πy .<br />

Then we can use Poincaré series to apply holomorphic projection to the function<br />

Φ ∗ B(ĝ ψ , z) := Φ B (ĝ ψ , z) − a B (ĝ ψ , 0)Ẽ2(z),<br />

since it satisfies the growth condition Φ ∗ B (ĝ ψ, z) = O(1/y) at the cusp infinity. For m > 0, let<br />

P m,1 (z, s) be the Poincaré series <strong>of</strong> level <strong>one</strong>, <strong>weight</strong> two defined by<br />

∑<br />

P m,1 (z, s) :=<br />

(y s e 2πimz )| 2 γ.<br />

γ∈Γ ∞\PSL 2 (Z)<br />

The limit as s goes to zero <strong>of</strong> the inner product between P m,1 (z, s) and Φ ∗ B (ĝ ψ, z) is the m th<br />

Fourier coefficient <strong>of</strong> a cusp form <strong>of</strong> <strong>weight</strong> two, level <strong>one</strong>. Since the space <strong>of</strong> such <strong>forms</strong> is<br />

trivial, we have<br />

(61) lim<br />

s→0<br />

〈Φ ∗ B(ĝ ψ , z), P m,1 (z, s)〉 = 0.


MOCK-MODULAR FORMS OF WEIGHT ONE 25<br />

Using the Fourier expansion <strong>of</strong> Φ B (ĝ ψ , z) and Rankin-Selberg unfolding, we can rewrite equation<br />

(61) as<br />

( )<br />

k<br />

(62) a B (ĝ ψ , m) + 24σ 1 (m)a B (ĝ ψ , 0) = lim δ(k)r B (pm + k)r ψ (k)φ s ,<br />

where<br />

φ s (µ) :=<br />

∫ ∞<br />

1<br />

s→0<br />

∑<br />

k≥1<br />

du<br />

(1 + µu) 1+s u .<br />

Comparing Q s−1 (t) and φ s−1 (µ), we see that φ 0 (µ) = 2Q 0 (1 + 2µ) and<br />

φ s−1 (µ) − 2Γ(2s) Q<br />

sΓ(s) 2 s−1 (1 + 2µ) = O(µ −1−s ).<br />

( ) ( )<br />

2<br />

That means we can substitute Q<br />

sΓ(s) 2 s−1 1 + 2k<br />

k<br />

for φ<br />

pm s−1 and in the limit as s approaches<br />

1. Together with (8) and (60), equation (61)<br />

pm<br />

becomes<br />

(63)<br />

〈j lift,B<br />

m , g ψ 〉 reg +24σ 1 (m)〈ϑ B , g ψ 〉 =<br />

− lim<br />

s→1<br />

∑<br />

A<br />

ψ(A) ∑ k≥1<br />

pm<br />

( ))<br />

δ(k)r B (pm + k)r A (k)<br />

(−2Q s−1 1 + 2k .<br />

pm<br />

Now the counting argument in Proposition (3.11) in [23] tells us that<br />

δ(k)r B (pm + k)r A (k) = ρ m (A, B, k),<br />

where ρ m (A, B, k) counts the number <strong>of</strong> γ ∈ M 2 (Z)/{±1} such that det(γ) = m and<br />

cosh d(τ √ AB −1 , γτ √ AB ) = 1 + 2k<br />

pm .<br />

In particular when k = 0, the number <strong>of</strong> γ ∈ M 2 (Z)/{±1} such that det(γ) = m and<br />

γτ √ AB = τ√ AB<br />

is exactly r −1 B (pm) = r B (m). Since k ≥ 1 in the summation, equation (63)<br />

becomes<br />

(64)<br />

〈jm<br />

lift,B ,g ψ 〉 reg + 24σ 1 (m)〈ϑ B , g ψ 〉 = − lim<br />

= − lim<br />

ɛ→0<br />

lim<br />

s→1<br />

∑<br />

A<br />

s→1<br />

∑<br />

A<br />

ψ(A)<br />

∑<br />

(<br />

g s τ<br />

√<br />

AB −1, γτ √ AB<br />

γ∈M 2 (Z)/{±1}<br />

det(γ)=m,<br />

γτ √ AB≠τ√ AB −1<br />

ψ(A) ( G m s (τ √ AB −1 + ɛ, τ √ AB ) − r B(m)g s (τ √ AB −1 + ɛ, τ √ AB −1 ) )<br />

where g s is defined in (57). From this expression, it is clear that the choice <strong>of</strong> these CM<br />

points do not matter.<br />

)<br />

Since g 1 (τ + ɛ, τ) = − log<br />

(1 + 4Im(τ)2 and ψ is non-trivial, we see that<br />

ɛ 2<br />

lim lim<br />

ɛ→0<br />

s→1<br />

∑<br />

A<br />

ψ(A)g s (τ √ AB −1 + ɛ, τ √ AB −1 ) = − ∑ A<br />

Also by equation (59), we have<br />

∑<br />

lim ψ(A)G m s (τ √<br />

s→1<br />

AB<br />

+ ɛ,τ √ −1 AB ) = ∑<br />

A<br />

A<br />

4πσ 1 (m) ∑ A<br />

ψ(A) log(y 2 √<br />

AB −1).<br />

ψ(A) log |Ψ m (τ √ AB −1 + ɛ, τ √ AB )|2 −<br />

ψ(A) lim<br />

s→1<br />

(E(τ √ AB −1 + ɛ, s) + E(τ √ AB , s)).<br />

)


26 W. DUKE AND Y. LI<br />

By (49) and the substitution A = A ′2 , we have<br />

lim<br />

s→1 4π ∑ A<br />

ψ(A)(E(τ √ AB −1 , s)+E(τ √ AB , s)) = −24(ψ(B)+ψ(B)) ∑ A ′ ψ 2 (A ′ ) log | √ y A ′η 2 (τ A ′)|.<br />

From (52) and<br />

H(p)ϑ B (z) = ∑ ψ ′<br />

ψ ′ (B)g ψ ′(z),<br />

it is easy to see that<br />

〈ϑ B , g ψ 〉 = −(ψ(B) + ψ(B)) ∑ A ′ ψ 2 (A ′ ) log | √ y A ′η 2 (τ A ′)|.<br />

After substituting these into equation (64) and changing A, B to A ′2 , B ′2 respectively, we<br />

obtain equation (53).<br />

□<br />

For any m ≥ 1, r ≥ 0 and τ, τ ′ ∈ H, define the level <strong>one</strong> <strong>modular</strong> function Ψ ∗ m(z; τ, τ ′ , r)<br />

by<br />

Ψ ∗ m(z; τ, τ ′ Ψ m (τ, z)<br />

, r) :=<br />

(j(τ ′ ) − j(z)) r<br />

In terms <strong>of</strong> Ψ ∗ m(z; τ, τ ′ , r), we can re-write the right hand side <strong>of</strong> (53) as<br />

−2<br />

∑<br />

ψ 2 (A ′ ) (log |Ψ ∗ m(τ A ′ B ′; τ A ′ B ′−1, τ A ′ B ′, r B(m))| + r B (m) log |y A ′ B ′j′ (τ A ′ B ′)|)<br />

A ′ ∈Cl(F )<br />

The quantity Ψ ∗ −1<br />

m(τ A ′ B ′; τA ′ B<br />

, τ ′ A ′ B ′, r B(m)) is well-defined and non-zero, since the order <strong>of</strong><br />

Ψ m (τ A ′ B ′−1, z) at z = τ A ′ B ′ is exactly r B(m) from the pro<strong>of</strong> above. By equation (47) and the<br />

fact that<br />

(j ′ (z)) 6 = j(z) 4 (j(z) − 1728) 3 ∆(z),<br />

we have<br />

σ C<br />

⎛<br />

⎝<br />

∏<br />

C ′ ∈Cl(F )<br />

y 6 A ′ B ′j′ (τ A ′ B ′)6<br />

y 6 C ′ ∆(τ C ′)<br />

⎞<br />

⎛<br />

⎠ = ⎝<br />

∏<br />

C ′ ∈Cl(F )<br />

y 6 C −1 A ′ B ′ j ′ (τ C −1 A ′ B ′)6<br />

y 6 C ′ ∆(τ C ′)<br />

⎞<br />

⎠ ∈ H.<br />

Similarly, the <strong>modular</strong> function Ψ ∗ m(z; τ A ′ B ′−1, τ A ′ B ′, r B(m)), which is defined over H, is sent<br />

to Ψ ∗ m(z; τ C −1 A ′ B ′−1, τ C −1 A ′ B ′, r B(m)) under σ C . So if we let<br />

⎛<br />

u m,B (A ′ ) := Ψ ∗ m(τ A ′ B ′; τ A ′ B ′−1, τ A ′ B ′, r B(m)) 6H(p) · ⎝<br />

⎞<br />

∏<br />

r B (m)<br />

yA 6 ′ B ′j′ (τ A ′ B ′)6 ⎠<br />

yC 6 ∆(τ ′ C ′)<br />

,<br />

then we can write<br />

with u m,B (A ′ ) ∈ H satisfying<br />

〈j lift,B<br />

m , g ψ 〉 reg = − 1<br />

3H(p)<br />

∑<br />

A ′ ∈Cl(F )<br />

C ′ ∈Cl(F )<br />

(65) σ C (u m,B (A ′ )) = u m,B (A ′ C −1 ).<br />

ψ 2 (A ′ ) log |u m,B (A ′ )|<br />

When m = 0, we have j lift,B<br />

0 (z) = ϑ B (z) and<br />

∑<br />

〈ϑ B , g ψ 〉 = − 1 ψ 2 (A ′ ) log |u<br />

12H(p)<br />

A ′ B ′u A ′ B ′−1|,<br />

A ′


MOCK-MODULAR FORMS OF WEIGHT ONE 27<br />

where u C ′ is defined in (46) for any C ′ ∈ Cl(F ). So we can let u m,B (A ′ ) = u A ′ B ′u A ′ B ′−1 when<br />

m = 0 and equation (65) still holds by (47). Since every <strong>modular</strong> function <strong>of</strong> <strong>weight</strong> <strong>one</strong> is<br />

a linear combination <strong>of</strong> j m (z), m ≥ 0, we can deduce the following corollary concerning the<br />

algebraic property <strong>of</strong> r + ψ<br />

(n) from (60) and Proposition 3.4.<br />

Corollary 5.2. Let ˜g ψ (z) = ∑ n∈Z r+ ψ (n)qn ∈ M − 1 (p) be a mock-<strong>modular</strong> form satisfying<br />

condition (i) in Proposition 4.2. For any <strong>modular</strong> function f(z) = ∑ n∈Z c(f, n)qn <strong>of</strong> level<br />

<strong>one</strong> with rational Fourier coefficients, there exists β f,B ∈ Q and u f,B (A) ∈ H × independent<br />

<strong>of</strong> the character ψ such that<br />

(66)<br />

〈f lift,B , g ψ 〉 reg = ∑ c(f, n) ∑ δ(k)r B (pn − k)r + ψ (k)<br />

n∈Z k∈Z<br />

∑<br />

= −β f,B ψ 2 (A) log |u f,B (A)|<br />

A∈Cl(F )<br />

and σ C (u f,B (A)) = u f,B (AC −1 ) for any C ∈ Cl(F ). If f has integral Fourier coefficients, then<br />

12H(p)β f,B ∈ Z.<br />

6. Rankin’s method and values <strong>of</strong> Green’s Function II<br />

In this section, we will work out the generalization <strong>of</strong> Corollary 5.2 to <strong>modular</strong> functions <strong>of</strong><br />

prime level N. From Corollary 5.2, we see that linear combinations <strong>of</strong> r + ψ<br />

(n) have the desired<br />

algebraic property. To prove Theorem 1.1, we need different linear combinations arising from<br />

regularized inner product between g ψ and other f ∈ M !,+<br />

1 (p). By considering analogues <strong>of</strong><br />

j lift,B for <strong>modular</strong> functions <strong>of</strong> level N, we can obtain these linear combinations, which will be<br />

sufficient to solve each r + ψ<br />

(n) (see §7). Other than being a step toward the pro<strong>of</strong> <strong>of</strong> Theorem<br />

1.1, this section is interesting on its own as the analogue <strong>of</strong> the Gross-Zagier formula with<br />

trivial Heegner point divisor in the Jacobian. Also, the interpretation <strong>of</strong> the regularized inner<br />

product in terms <strong>of</strong> Borcherds lifts continues to hold for <strong>modular</strong> functions <strong>of</strong> level N.<br />

Let N be a prime number different from p. For a <strong>modular</strong> function f(z) <strong>of</strong> level N and<br />

B ∈ Cl(F ), <strong>one</strong> can define f lift,N,B (z) ∈ M 1(p) ! by<br />

(67) f lift,N,B (z) := Tr Np<br />

p ((f|W N )(pz)ϑ B (Nz)),<br />

−1<br />

where W N = (<br />

N<br />

) is the Fricke involution and ϑ B (z) is the <strong>weight</strong> <strong>one</strong> theta series corresponding<br />

to the class B ∈ Cl(F ). Notice that the form f lift,N,B (z) is the same as f lift,B (z) from<br />

§5 when N = 1. The main result <strong>of</strong> this section is the following generalization <strong>of</strong> Corollary<br />

5.2.<br />

Proposition 6.1. Let f(z) is a <strong>modular</strong> function <strong>of</strong> prime level N ≠ p, with rational Fourier<br />

coefficients. There exist β f,B ∈ Q and u f,B (A) ∈ H independent <strong>of</strong> the character ψ such that<br />

∑<br />

(68) 〈f lift,N,B , g ψ 〉 reg = −β f,B ψ 2 (A) log |u f,B (A)|.<br />

A∈Cl(F )<br />

For any class C ∈ Cl(F ), the algebraic integer u f,B (A) satisfies<br />

σ C (u f,B (A)) = u f,B (AC −1 )<br />

where σ C ∈ Gal(H/F ) is associated to C ∈ Cl(F ) via Artin’s isomorphism. Furthermore, if<br />

f has integral principal part Fourier coefficients and satisfies equation (69) below, then the<br />

denominator <strong>of</strong> β f,B only depends on N and p.


28 W. DUKE AND Y. LI<br />

The pro<strong>of</strong> <strong>of</strong> Proposition 6.1 follows the same steps as in the pro<strong>of</strong> <strong>of</strong> Theorem 5.1 and<br />

Corollary 5.2. First, we will express 〈f lift,N,B , g ψ 〉 reg in terms <strong>of</strong> other regularized inner products<br />

and the limit <strong>of</strong> an infinite sum. Then, via Green’s function, we will relate the limit <strong>of</strong><br />

the infinite sum to the height pairings between Heegner divisors on J 0 (N), the Jacobian <strong>of</strong><br />

X 0 (N). Since the Heegner divisor is trivial on J 0 (N), the height pairing is the same as the<br />

value <strong>of</strong> some <strong>modular</strong> function evaluated at a Heegner point. From there, we can deduce<br />

equation (68). Before presenting the pro<strong>of</strong>, we need some notations and facts about level N<br />

<strong>modular</strong> <strong>forms</strong>. For this part, we only require N to be a prime number.<br />

Let M 2(N), ! resp. M 0(N), ! be the space <strong>of</strong> weakly holomorphic <strong>weight</strong> 2 <strong>modular</strong> <strong>forms</strong>,<br />

resp. <strong>modular</strong> functions, <strong>of</strong> level N, and trivial nebentypus. Denote the subspaces <strong>of</strong> M 2(N)<br />

!<br />

<strong>of</strong> cusp <strong>forms</strong> and holomorphic <strong>modular</strong> <strong>forms</strong> by M 2 (N), S 2 (N) respectively. Let M !,new<br />

0 (N)<br />

be the subspace <strong>of</strong> M 0(N) ! containing weakly holomorphic <strong>modular</strong> <strong>forms</strong> f(z) satisfying<br />

(69) (f|W N )(z) = −N(f|U N )(z).<br />

The following lemma shows that M ! 0(N) can be be decomposed into the direct sum <strong>of</strong><br />

M !,new<br />

0 (N) and level <strong>one</strong> <strong>modular</strong> functions.<br />

Lemma 6.2. Let f(z) be a <strong>modular</strong> function <strong>of</strong> prime level N. Then it can be written<br />

uniquely as<br />

f(z) = f 1 (Nz) + f 2 (z),<br />

such that f 1 (z) is a <strong>modular</strong> function <strong>of</strong> level 1 and f 2 (z) ∈ M !,new<br />

0 (N).<br />

Pro<strong>of</strong>. Let a, b ∈ Z. With some matrix calculations, we have<br />

( ( ))<br />

f| UN W N + a N (z) =<br />

a−1<br />

f(z) + N TrN 1 (f)(Nz)<br />

( ( ) ( ))<br />

f| UN W N + a UN W<br />

N<br />

N + b N (z) =<br />

(a−1)(b−1)<br />

f(z) + a+b+N−1 Tr N N 2<br />

N 1 (f)(Nz)<br />

Set f j (z) as follows and we are d<strong>one</strong><br />

f 1 (z) = N<br />

N+1 TrN 1 (f)(z),<br />

f 2 (z) = − N N+1 NW N ) (z) − f(z)) .<br />

Alternatively, <strong>one</strong> can characterize the space M ! 0(N) via the space S 2 (N).<br />

Definition 6.3. A set <strong>of</strong> numbers {λ m : m ≥ 1} is called a relation for S 2 (N) if<br />

(i) λ m ∈ C is non-zero for all but finitely many m.<br />

(ii) For any cusp form h(z) = ∑ m≥1 c(h, m)qm ∈ S 2 (N), the numbers {λ m } satisfy<br />

∑<br />

(70)<br />

δ N (m)λ m c(h, m) = 0.<br />

m≥1<br />

where δ N (m) = N + 1 if N|m and 1 otherwise.<br />

A relation {λ m } m≥1 is called integral if λ m ∈ Z for all m ≥ 1. Denote Λ N the set <strong>of</strong> all<br />

relations for S 2 (N).<br />

For any f(z) ∈ M !,new<br />

0 (N), <strong>one</strong> knows that the principal part coefficients, {c(f, −m) :<br />

m ≥ 1}, is a relation for S 2 (N) from works by Petersson. By considering the Fourier expansion<br />

<strong>of</strong> vector-valued Poincaré series, (See [7], [27], [32]), <strong>one</strong> can also show that the converse<br />

is true. So given λ = {λ m } ∈ Λ N , the expression<br />

∑<br />

λ m q −m<br />

m≥1<br />


MOCK-MODULAR FORMS OF WEIGHT ONE 29<br />

is necessarily the principal part <strong>of</strong> the Fourier expansion at infinity <strong>of</strong> a unique function<br />

f λ (z) ∈ M !,new<br />

0 (N). Alternatively, this statement follows essentially from an application <strong>of</strong><br />

Serre duality [5, §3]. Suppose λ ∈ Λ N is integral. Then the n th Fourier coefficient <strong>of</strong> f λ (z) is<br />

integral when n ≠ 0 since f λ is a rational function <strong>of</strong> j(z) and j(Nz).<br />

Let g N be the genus <strong>of</strong> Γ 0 (N). From [1], we know that for all h(z) ∈ S 2 (N), ord ∞ (h(z)) ≤<br />

dim(S 2 (N)) = g N ≤ (N + 1)/12. Using matrix computations similar to that in the pro<strong>of</strong> <strong>of</strong><br />

Lemma 6.2, <strong>one</strong> can show that<br />

(71) h| 2 U N = −h| 2 W N<br />

for all h(z) ∈ S 2 (N), implying that U N ◦ U N is the identity operator on S 2 (N). Combining<br />

these facts together gives us the following lemma about M !,new<br />

0 (N).<br />

Lemma 6.4. The space M !,new<br />

0 (N) is spanned by <strong>modular</strong> functions<br />

∞∑<br />

f m (z) = q −m + c m (n)q n ,<br />

n=−g N<br />

with m ≥ g N + 1. For these m, the coefficient c m (n) are integers when n ≠ 0. Furthermore,<br />

for all m ≥ g N + 1,<br />

f N 2 m(z) − N+1<br />

δ N (m) f m(z) = q −N 2m − N+1<br />

δ N (m) q−m + O(1).<br />

There is a similar duality statement dictating the existence <strong>of</strong> <strong>forms</strong> in M ! 2(N).<br />

Lemma 6.5. For every n ≥ 0, there exists P n,N (z) = ∑ m∈Z c(P n,N, m)q m ∈ M ! 2(N) such<br />

that<br />

(i) At the cusp infinity, P n,N (z) = q −n + O(q).<br />

(ii) (P n,N |W N )(z) = −(P n,N |U N )(z).<br />

Furthermore, if f(z) = ∑ m∈Z c(f, m)qn ∈ M !,new<br />

0 (N), then<br />

∑<br />

δ N (m)c(f, m)c(P n,N , −m) = 0.<br />

m∈Z<br />

Now, we will recall some facts about Heegner points on the <strong>modular</strong> curve from [22] and<br />

their height pairings on the Jacobian. Let X 0 (N) be the natural compactification <strong>of</strong> Y 0 (N),<br />

the open <strong>modular</strong> curve over Q classifying pairs <strong>of</strong> elliptic curves (E, E ′ ) with an order N<br />

cyclic isogeny φ : E → E ′ . The complex points <strong>of</strong> X 0 (N) has a structure <strong>of</strong> a compact<br />

Riemann surface and can be identified with H ∪ P 1 (Q) modulo the action <strong>of</strong> Γ 0 (N). Heegner<br />

points on X 0 (N) corresponds to pair <strong>of</strong> elliptic curves (E, E ′ ) having complex multiplication<br />

by the same ring O in some imaginary quadratic field F . This occurs only when there exists<br />

an O-ideal n dividing N such that O/n is cyclic <strong>of</strong> order N. In this case, E(C) is isomorphic<br />

to C/a with a an invertible O-submodule in F . Since this a can be chosen independent <strong>of</strong><br />

homothety by elements in F × , we only need to consider [a], the class <strong>of</strong> a in Pic(O). So<br />

a Heegner point on X 0 (N) can be expressed as the triplet (O, n, [a]). To find the image <strong>of</strong><br />

such a Heegner point in H, choose an oriented basis 〈ω 1 , ω 2 〉 <strong>of</strong> a such that an −1 has basis<br />

〈ω 1 , ω 2 /N〉. Then τ [a],n := ω 1 /ω 2 ∈ Γ 0 (N)\H is the image <strong>of</strong> this Heegner point.<br />

For our purpose, F = Q( √ −p) and O = O F . Then we require χ p (N) = 1 for Heegner<br />

points to exist. In that case, N splits into nn in F with<br />

n = ZN + Z bn+√ −p<br />

2<br />

.


30 W. DUKE AND Y. LI<br />

Let A ∈ Pic(O F ) = Cl(F ). A point τ ∈ H corresponding to the Heegner point (O, n, A)<br />

must satisfy an equation<br />

Aτ 2 + Bτ + C = 0,<br />

with B 2 − 4AC = −p, N|A, B ≡ b n (mod 2N) and gcd(A/N, B, NC) = 1. We will denote<br />

this image by τ(A, n). Notice τ(A, n) is well-defined up to the action <strong>of</strong> Γ 0 (N).<br />

Let J 0 (N) be the Jacobian <strong>of</strong> X 0 (N). Its complex points is a compact Riemann surface<br />

and can be viewed as the set <strong>of</strong> divisors <strong>of</strong> degree zero modulo the set <strong>of</strong> principal divisors<br />

on X 0 (N). Let 〈, 〉 C be the height symbol on X 0 (N)(C). It is a unique real-valued function<br />

defined on the set <strong>of</strong> divisors <strong>of</strong> degree zero and satisfies<br />

〈∑<br />

ni (x i ), b〉<br />

= ∑ n i log |f(x i )| 2<br />

C<br />

i<br />

if b = (f) is a principal divisor.<br />

For n ≥ 0 and a class B ∈ Cl(F ), denote the inner product<br />

(72) ρ N,B,ψ (n) := 〈j lift,B[n]<br />

n<br />

+ j lift,B[n]−1<br />

n , g ψ 〉 reg .<br />

In particular for j 0 (z) = 1, the inner product ρ N,B,ψ (0) becomes 〈T N (ϑ B ), g ψ 〉, where T N is<br />

the N th Hecke operator on M ! 1(p). Also, notice that<br />

(73) T N (ϑ B )(z) = ϑ B (Nz) + (ϑ B |U N )(z) = ϑ B[n] (z) + ϑ B[n] −1(z).<br />

By Corollary 5.2, we know that ρ N,B,ψ (n) can be written in the form <strong>of</strong> the right hand<br />

side <strong>of</strong> (68) and satisfies the conditions in Proposition 6.1. The following lemma relates<br />

〈f lift,N,B , g ψ 〉 reg to ρ N,B,ψ (n) and an infinite sum analogous to the right hand side <strong>of</strong> (62).<br />

Lemma 6.6. Suppose f(z) ∈ M !,new<br />

0 (N). In the notations above, we have<br />

(74) 〈f lift,N,B , g ψ 〉 reg = ∑ m ′ ≥1<br />

c(f, −Nm ′ )ρ N,B,ψ (m ′ ) + C f ρ N,B,ψ (0) − Σ f,N,B,ψ ,<br />

where<br />

∑<br />

Σ f,N,B,ψ := − lim<br />

s→1<br />

C f =<br />

m≥1<br />

N∤m<br />

(N + 1) lim<br />

(<br />

c(f, −m) ∑ k≥1<br />

s→1<br />

∑<br />

m ′ ≥1<br />

c(f, −Nm ′ ) ∑ k ′ ≥1<br />

Nc(f, 0) + 24 ∑ m ′ ≥1<br />

c(f, −Nm ′ )σ 1 (m ′ )<br />

)<br />

δ(k)r B (pm + Nk)r ψ (k)2Q s−1<br />

(1 + 2kN +<br />

pm<br />

)<br />

δ(k ′ )r B (k ′ )r ψ (Nk ′ − pm ′ )2Q s−1<br />

(−1 + 2k′ N<br />

,<br />

pm ′<br />

)<br />

.<br />

Pro<strong>of</strong>. Let ˜g ψ (z) ∈ M − 1 (p) be a mock-<strong>modular</strong> form as in Proposition 4.2 with Fourier expansion<br />

˜g ψ (z) = ∑ n∈Z r+ ψ (n)qn and ĝ ψ (z) the associated harmonic Maass form. By Theorem<br />

5.1, we have the following identities<br />

)<br />

ρ N,B,ψ (n) = −24σ 1 (n)ρ N,B,ψ (0) + lim δ(k)c(T N (ϑ B ), pm + k)r ψ (k)2Q s−1<br />

(1 + 2k ,<br />

pm<br />

ρ N,B,ψ (n) = ∑ k∈Z<br />

s→1<br />

∑<br />

k≥1<br />

r + ψ (k) ( (<br />

r pn−k<br />

)<br />

B N + rB (pNn − Nk) ) δ(k).


Similar to the pro<strong>of</strong> <strong>of</strong> Theorem 5.1, we define<br />

MOCK-MODULAR FORMS OF WEIGHT ONE 31<br />

Φ N,B (ĝ, z) := Tr Np<br />

N (−i(ĝ|W p)(Nz)ϑ B (z))<br />

= (ĝ|U p )(Nz)ϑ B (z) + (ĝ(z)ϑ B (Nz))|U p .<br />

When N = 1, this is the same as Φ B (ĝ ψ , z) defined in the pro<strong>of</strong> <strong>of</strong> Theorem 5.1. It trans<strong>forms</strong><br />

like a <strong>modular</strong> form <strong>of</strong> level N, <strong>weight</strong> two and has the following Fourier expansion at infinity<br />

Φ N,B (ĝ ψ , z) = ∑ m∈Z(b N,B (ĝ ψ , m, y) + a N,B (ĝ ψ , m))q m ,<br />

b N,B (ĝ ψ , m, y) = − ∑ k∈Z<br />

a N,B (ĝ ψ , m) = ∑ k∈Z<br />

Suppose f(z) has the Fourier expansion<br />

at infinity. Define the sum Σ ′ f,N,B,ψ by<br />

δ(k)r ψ (k)β 1<br />

(<br />

k, Ny<br />

p<br />

r + ψ (k) r B(pm − Nk)δ(k).<br />

f(z) = ∑ m∈Z<br />

c(f, m)q m<br />

(75) Σ ′ f,N,B,ψ := ∑ m∈Z<br />

c(f, −m)δ N (m)a N,B (ĝ ψ , m).<br />

)<br />

r B (pm + Nk),<br />

Since f lift,N,B (z) ∈ M !,+<br />

1 (p), we can apply Proposition 3.4 to obtain<br />

⎛ ∑<br />

〈f lift,N,B , g ψ 〉 reg = ∑ r + ψ (k)r ⎞<br />

B(pm − Nk)δ(k)−<br />

c(f, −m) ⎜<br />

k∈Z<br />

⎝<br />

m∈Z N ∑ r + ψ (k)r ( pm−Nk<br />

)<br />

⎟<br />

⎠<br />

B N δ(k) 2<br />

k∈Z<br />

= −N ∑ c(f, −Nm ′ ) ∑ ( ( )<br />

)<br />

r + ψ (k) pm<br />

r ′ −k<br />

B + r<br />

N B (pNm ′ − Nk) δ(k)<br />

m ′ ∈Z<br />

k∈Z<br />

+ ∑ m∈Z<br />

δ N (m)c(f, −m)a N,B (ĝ ψ , m)<br />

(76)<br />

= −N ∑ m ′ ≥0<br />

c(f, −Nm ′ )ρ N,B,ψ (m ′ ) + Σ ′ f,N,B,ψ<br />

The sum over m ′ ∈ Z changes to be over only m ′ ≥ 0, since r + ψ<br />

(k) = 0 for all k ≤ −p by the<br />

choice <strong>of</strong> ˜g ψ (z).<br />

For n ≥ 0, let P n,N (z) = q −n + O(q) ∈ M 2(N) ! be the weakly holomorphic <strong>modular</strong> form<br />

as in Lemma 6.5. When n = 0, we can take P 0,N (z) ∈ M 2 (N) to be<br />

P 0,N (z) =<br />

Consider Φ ∗ N,B (ĝ ψ, z), defined by<br />

(77)<br />

Φ ∗ N,B(ĝ ψ , z) := Φ N,B (ĝ ψ , z) − ∑ n≥0<br />

(N−1)Ẽ2(Nz) N − (N−1)Ẽ2(z).<br />

1<br />

a N,B (ĝ ψ , −n)P n,N (z) + Nρ N,B,ψ(0)<br />

N 2 − 1<br />

(Ẽ2(Nz) − Ẽ2(z)).<br />

Now it is easy to check that Φ ∗ N,B (ĝ ψ, z) is O(1/y) at the cusp infinity. At the cusp 0, some<br />

calculations show that<br />

Φ N,B (ĝ ψ , z)|W N = −Φ N,B (ĝ ψ , z)|U N + Φ 1,B[n] (ĝ ψ , z) + Φ 1,B[n] −1(ĝ ψ , z)


32 W. DUKE AND Y. LI<br />

From the pro<strong>of</strong> <strong>of</strong> Theorem 5.1, we know that Φ 1,B[n] (ĝ ψ , z) + Φ 1,B[n] −1(ĝ ψ , z) has a constant<br />

term ρ N,B,ψ (0) and no pole. So Φ ∗ N,B (ĝ ψ, z) is also O(1/y) at the cusp 0.<br />

For m ≥ 1, let P m,N (z, s) be the Poincaré series <strong>of</strong> level <strong>one</strong>, <strong>weight</strong> two defined by<br />

∑<br />

P m,N (z, s) := (y s e 2πimz )| 2 γ.<br />

It is characterized by the property that<br />

γ∈Γ ∞\Γ 0 (N)<br />

lim 〈h(z), P m,N(z, s)〉 = mc(h, m)<br />

s→0<br />

for any h(z) = ∑ m≥1 c(h, m)qm ∈ S 2 (N). Since {c(f, −m) : m ≥ 1} ∈ Λ N , we know that<br />

〈<br />

(78) lim h(z), ∑ 〉<br />

δ N (m)<br />

c(f, −m)P<br />

s→0 m m,N(z, s) = 0,<br />

m≥1<br />

δ N (m)<br />

m c(f, −m)P m,N(z, s) ∈ S 2 (N) is 0.<br />

for any h(z) ∈ S 2 (N). So lim s→0<br />

∑m≥1<br />

Now we can consider the inner product between Φ ∗ N,B (ĝ ψ, z) and this linear combination<br />

<strong>of</strong> Poincaré series and obtain<br />

∑<br />

δ<br />

4π lim N (m)<br />

c(f,<br />

s→0 m −m)〈Φ∗ N,B(ĝ ψ , z), P m,N (z, s)〉 = 0.<br />

m≥1<br />

Since Φ ∗ N,B (ĝ ψ, z) is O(1/y) at both cusps, we can apply Rankin-Selberg unfolding as in the<br />

pro<strong>of</strong> <strong>of</strong> Theorem 5.1. The limit on the left hand side above then breaks up into two parts.<br />

The first part is<br />

⎛<br />

a N,B (ĝ ψ , m) − ∑ a N,B (ĝ ψ , −n)c(P n,N , m)<br />

∑<br />

⎜<br />

n≥0<br />

⎟<br />

(79)<br />

m≥1<br />

δ N (m)c(f, −m) ⎜<br />

⎝<br />

− 24Nρ N,B,ψ(0)<br />

N 2 − 1<br />

(<br />

σ1<br />

( m<br />

N<br />

⎞<br />

)<br />

− σ1 (m) ) ⎟<br />

⎠<br />

Since (f|W N )(z) = −N(f|U N )(z) and c(P n,N , 0) = 0 for n ≥ 1, Lemma 6.5 tells us that for<br />

n ≥ 0<br />

∑<br />

δ N (m)c(f, −m)c(P n,N , m) = −δ N (n)c(f, n).<br />

m≥1<br />

So expression (79) becomes<br />

Σ ′ f,N,B,ψ − 24Nρ N,B,ψ(0)<br />

N 2 − 1<br />

∑<br />

δ N (m)c(f, −m) ( (<br />

σ m<br />

)<br />

1 N − σ1 (m) ) .<br />

m≥1<br />

The second part involves the limit <strong>of</strong> an infinite sum, which can be evaluated as<br />

Σ f,N,B,ψ − (N + 1) ∑ m ′ ≥1<br />

c(f, −Nm ′ )(ρ N,B,ψ (m ′ ) + 24σ 1 (m ′ )ρ N,B,ψ (0)).<br />

Adding the last two expressions together, which yields 0, and substituting into (76), we<br />

obtain equation (74). The constant C f appears since<br />

∑ ( (<br />

(80) c(f, 0) = 24<br />

m<br />

)<br />

N 2 Nσ1<br />

−1<br />

N − σ1 (m) ) δ N (m)c(f, −m)<br />

m≥1<br />

from applying Lemma 6.5 to P 0,N (z) ∈ M 2(N) ! and f(z) ∈ M !,new<br />

0 (N). Notice if c(f, −m) ∈ Z<br />

for all m ≥ 1, then the denominator <strong>of</strong> c(f, 0) is independent <strong>of</strong> f.<br />


MOCK-MODULAR FORMS OF WEIGHT ONE 33<br />

Pro<strong>of</strong> <strong>of</strong> Proposition 6.1. If f(z) = f 1 (Nz) with f 1 (z) a <strong>modular</strong> function <strong>of</strong> level <strong>one</strong>, then<br />

(f|W N )(z) = (f|U N )(z) and<br />

f lift,N,B (z) = f lift,B<br />

1 (z) = f 1 (pz)ϑ B (z).<br />

This case <strong>of</strong> the proposition follows from Corollary 5.2. By Lemma 6.2, it is enough to prove<br />

the proposition for f(z) ∈ M !,new<br />

0 (N). In that case, some calculations show that<br />

(81) f lift,N,B (z) = −N(f|U N )(pz)ϑ B (Nz) + (f(pz)ϑ B (z)) |U N .<br />

Let χ p (N) = ɛ. It is easy to see that f lift,N,B (z) ∈ M !,ɛ<br />

1 (p). Since g ψ (z) ∈ S + 1 (p), Proposition<br />

3.4 tells us that 〈f lift,N,B , g ψ 〉 reg = 0 when ɛ = −1. So we can suppose χ p (N) = ɛ = 1.<br />

By Lemma 6.4, it is easy to see that M !,new<br />

0 (N) is generated by the set S 1 ∪ S 2 , where<br />

(82)<br />

S 1 := {f m (z) : m ≥ 1 + g N , N 2 ∤ m}<br />

S 2 := {f N 2 m(z) − N+1<br />

δ N (m) f m(z) : m ≥ 1 + g N }.<br />

Let f(z) = f N 2 m(z) − N+1<br />

δ N (m) f m(z) ∈ S 2 . We can uniquely write the integer m as N r m ′ with<br />

r ≥ 0, m ′ ≥ −1 and gcd(N, m ′ ) = 1. Then at the cusp infinity, we have<br />

f(z) = q −N r+2 m ′ − N+1<br />

δ N (N r m ′ ) q−N r m ′ + O(1).<br />

From equation (81), it follows that the −n th Fourier coefficient <strong>of</strong> f lift,N,B (z) can be written<br />

as<br />

(<br />

( ))<br />

( )<br />

pN r m ′<br />

c(f lift,N,B pN<br />

, −n) = −N r r+1 m ′ −n<br />

B − N+1 r N<br />

−n<br />

N<br />

δ N (N r m ′ ) B N<br />

(<br />

)<br />

+ r B (−Nn + pN r+2 m ′ ) − N+1 r δ N (N r m ′ ) B(−Nn + pN r m ′ ) .<br />

When r = 0, we can rewrite the equation above as<br />

c(f lift,N,B , −n) = −(N + 1) ( ( )<br />

r B pm ′ − n N + rB (pm ′ − Nn) )<br />

( )<br />

)<br />

pNm<br />

+<br />

(r ′ −n<br />

B + r<br />

N<br />

B (N(pNm ′ − n))<br />

(<br />

)<br />

= −(N + 1)c T N (j lift,B<br />

m<br />

), −n + c (j ′<br />

Nm ′(pz)T N (ϑ B )(z), −n) .<br />

Notice that j pNm ′(pz)T N (ϑ B )(z) = j lift,B[n]<br />

Nm<br />

(z) + j lift,B[n]−1<br />

′<br />

Nm<br />

(z). Since this is true for all n ≥ 1<br />

′<br />

and the principal part uniquely determines a form in M !,+<br />

f lift,N,B (z) = −(N + 1)T N<br />

(<br />

j lift,B<br />

m ′<br />

)<br />

(z)<br />

+ j lift,B[n]<br />

Nm ′<br />

1 (p) up to <strong>forms</strong> in M + 1 (p), we have<br />

(z) + j lift,B[n]−1 (z) + g f (z)<br />

for some g f (z) ∈ M + 1 (p) with integral linear combination <strong>of</strong> c(f, n)’s as Fourier coefficients.<br />

Since T N is self-adjoint with respect to the regularized inner product and<br />

T N (g ψ )(z) = (ψ([n]) + ψ([n]))g ψ (z)<br />

for N prime with χ p (N) = 1, we can rewrite 〈f lift,N,B , g ψ 〉 reg as<br />

〈f lift,N,B , g ψ 〉 reg = −(N + 1)(ψ([n]) + ψ([n]))〈j lift,B<br />

m ′ , g ψ 〉 reg + ρ N,B,ψ (Nm ′ ) + 〈g f , g ψ 〉.<br />

Nm ′


34 W. DUKE AND Y. LI<br />

When r ≥ 1, the situation is similar. In this case, we have<br />

( )<br />

f lift,N,B (z) = − (N + 1)T N j lift,B<br />

N r m<br />

(z) + (Nj ′ N r−1 m ′(pz) + j N r+1 m ′(pz))T N(ϑ B )<br />

+ g f (z),<br />

〈f lift,N,B , g ψ 〉 reg = − (N + 1)(ψ([n]) + ψ([n]))〈j lift,B<br />

m ′ , g ψ 〉 reg + Nρ N,B,ψ (N r−1 m ′ )<br />

+ ρ N,B,ψ (N r+1 m ′ ) + 〈g f , g ψ 〉,<br />

for some g f (z) ∈ M 1 + (p) with integral linear combination <strong>of</strong> c(f, n)’s as Fourier coefficients.<br />

By Proposition 4.1 and Corollary 5.2, the proposition holds for f(z) ∈ S 2 .<br />

Now, we will prove the proposition when f(z) is in the integral span <strong>of</strong> S 1 and show that<br />

the denominator <strong>of</strong> the constant β f,B in (68) is independent <strong>of</strong> f. This will finish the pro<strong>of</strong><br />

<strong>of</strong> Proposition 6.1. By Lemma 6.2 and Lemma 6.6, it is enough to show that Σ f,N,B,ψ can<br />

also be written in the form <strong>of</strong> the right hand side <strong>of</strong> equation (68). To accomplish that,<br />

we will relate Σ f,N,B,ψ to the automorphic Green’s function using the counting arguments in<br />

Proposition (3.11) from [24]. In their notation, let R N be a subset <strong>of</strong> M 2 (Z) containing all<br />

matrices whose lower left entry is divisible by N. For z, z ′ ∈ H and k ≥ 0, let ρ m N (z, z′ , k) be<br />

the number <strong>of</strong> γ = ( a c d b ) ∈ R N/{±1} such that det(γ) = m and<br />

cosh d(z, γz ′ ) = 1 + 2Nk<br />

pm .<br />

Define the automorphic Green’s function <strong>of</strong> level N to be<br />

∑<br />

G N,s (z, z ′ ) :=<br />

g s (z, γz ′ ) +<br />

γ∈Γ 0 (N)/{±1}<br />

z≠γz ′<br />

where<br />

∑<br />

γ∈Γ 0 (N)/{±1}<br />

z=γz ′<br />

g s (z) := lim<br />

w→z<br />

(g s (z, w) − log |2πiη(z) 4 (z − w)| 2<br />

g s (z),<br />

= − log |2π(z − z)η(z) 4 | 2 + 2 Γ′ Γ′<br />

(s) − 2 (1).<br />

Γ Γ<br />

When gcd(N, m) = 1, the m th Hecke operator T m acts on z ′ in G N,s (z, z ′ ) and defines<br />

G m N,s (z, z′ ). It can be written as<br />

∑<br />

G m N,s(z, z ′ ) :=<br />

g s (z)<br />

γ∈R N /{±1}<br />

det γ=m<br />

= −2 ∑ k>0<br />

γ∈R N /{±1}<br />

det γ=m<br />

z=γz ′<br />

z≠γz ′ g s (z, γz ′ ) + ∑<br />

ρ m N(z, z ′ , k)Q s−1<br />

(1 + 2Nk<br />

pm<br />

)<br />

+ ρ m N(z, z ′ , 0)g s (z).<br />

Notice that ρ m N (z, z′ , k) could be non-zero when k is not an integer. When z, z ′ ∈ H are<br />

Heegner points <strong>of</strong> discriminant −p, then ρ m N (z, z′ , k) is necessarily supported on integral k’s.<br />

For j = 1, 2, let A j ∈ Cl(F ) and τ j ∈ Γ 0 (N)\H be the image <strong>of</strong> the Heegner points<br />

x j = (O F , A j , [n]). Let a j be integral ideals in the class A j such that n|a j , N(a j ) = A j . Then<br />

by Proposition (3.11) in [24], the counting function ρ m N (τ 2, τ 1 , k) satisfies<br />

ρ m N(τ 2 , τ 1 , k) =#{(α, β) ∈ (a −1<br />

1 a −1<br />

2 × a −1<br />

1 a −1<br />

2 n)/{±1} | N F/Q (α) = Nk+pm<br />

A 1 A 2<br />

,<br />

N F/Q (β) = Nk<br />

A 1 A 2<br />

, A 1 A 2 (α − β) ≡ 0 (mod d)},<br />

=r A1 A −1(Nk<br />

+ pm)r A<br />

2 1 A 2 [n] −1(k)δ(k)


MOCK-MODULAR FORMS OF WEIGHT ONE 35<br />

for k ≥ 0, where d = ( √ −p) is the different <strong>of</strong> F . The first equality is established by the<br />

bijection<br />

γ = ( a c d b ) ∈ R N/{±1} ↦→ α = cτ 1τ 2 + dτ 2 − aτ 1 − b,<br />

β = cτ 1 τ 2 + dτ 2 − aτ 1 − b.<br />

The Fricke involution W N sends x j to the Heegner point (O F , A j [n] −1 , [n] −1 ). Set A ′ 1 =<br />

A 1 [n] −1 and a ′ 1 ∈ A ′ 1 a class such that n|a ′ 1, then the same counting argument establishes<br />

ρ m′<br />

N (τ 2 , τ ′ 1, k ′ ) =#{(α, β) ∈ ((a ′ 1) −1 ā −1<br />

2 n × (a ′ 1) −1 a −1<br />

2 )/{±1} | N F/Q (α) = Nk′<br />

A 1 A 2<br />

,<br />

N F/Q (β) = Nk′ −pm ′<br />

A 1 A 2<br />

, A 1 A 2 (α − β) ≡ 0 (mod d)},<br />

=r A ′<br />

1 A 2<br />

(Nk ′ − pm)r A ′<br />

1 A −1<br />

2 [n]−1 (k ′ )δ(k ′ )<br />

for k ′ ≥ pm ′ /N. Apply these counting arguments to G m N,s (τ 2, τ 1 ) and G m′<br />

N,s (τ 2, τ 1), ′ we obtain<br />

G m N,s(τ 2 , τ 1 ) = − 2 ∑ ( )<br />

r A1 A −1(Nk<br />

+ pm)r A<br />

2 1 A 2 [n] −1(k)δ(k)Q s−1 1 + 2Nk +<br />

pm<br />

(83)<br />

k≥1<br />

r A1 A −1(m)g<br />

s(τ 2 ),<br />

2<br />

∑<br />

( )<br />

G m′<br />

N,s(τ 2 , τ 1) ′ = − 2 r A ′<br />

1 A 2<br />

(Nk ′ − pm ′ )r A ′<br />

1 A −1 (k ′ )δ(k ′ )Q<br />

2 [n]−1 s−1 −1 + 2Nk′ +<br />

pm ′<br />

(84)<br />

k ′ >pm ′ /N<br />

r A ′<br />

1 A −1<br />

2 [n]−1 ( m′<br />

N )g s(τ 2 ).<br />

Since N ∤ m ′ in the definition <strong>of</strong> G m′<br />

N,s (z, z′ ), the term r A ′<br />

1 A −1 ( m′ )g<br />

2 [n]−1 N s(τ 2 ) above vanishes.<br />

By Proposition (2.23) in [24, p.242], we can relate the Green’s function to the height pairing<br />

between (x 2 ) − (∞), (x 1 ) − (0) and W N ((x 1 ) − (0)) on the Jacobian as follows<br />

( )<br />

G<br />

m<br />

N,s (τ 2 , τ 1 ) + 4πσ 1 (m)E N (W N τ 2 , s)<br />

〈(x 2 ) − (∞), T m ((x 1 ) − (0))〉 C = lim<br />

s→1 + 4πm s σ 1−2s (m)E N (τ 1 , s) + σ 1(m)κ N<br />

s−1<br />

−σ 1 (m)(λ N + 2κ N ),<br />

( )<br />

G<br />

m<br />

N,s(τ ′<br />

2 , τ 1) ′ + 4πσ 1 (m ′ )E N (τ 2 , s)<br />

〈(x 2 ) − (∞), T m ′W N ((x 1 ) − (0))〉 C = lim<br />

s→1<br />

+ 4π(m ′ ) s σ 1−2s (m ′ )E N (W N τ 1 , s) + σ 1(m ′ )κ N<br />

s−1<br />

−σ 1 (m ′ )(λ N + 2κ N ).<br />

Here E N (z, s) is the Eisenstein series defined in (48) and κ N , λ N are explicit constants dependent<br />

on N.<br />

For A ′ , B ∈ Cl(F ), substitute A 1 = A ′√ B[n], A 2 = A ′√ B −1 [n] and A ′ 1 = A ′√ B[n] −1 into<br />

equations (83) and (84). Denote τ 1 = τ A1 ,n, τ 2 = τ A2 ,n and τ 1 ′ = τ ′ A 1 [n] −1 ,¯n by τ 1(A ′ , B), τ 2 (A ′ , B)<br />

and τ 1(A ′ ′ , B) respectively. Summing together equations (83), (84) over m, m ′ and A ′ with a<br />

character ψ 2 gives us<br />

Σ f,N,B,ψ = ∑ m≥1<br />

N∤m<br />

c(f, −m)<br />

∑<br />

A ′ ∈Cl(F )<br />

− (N + 1) ∑ m ′ ≥1<br />

c(f, −Nm ′ )<br />

ψ 2 (A ′ )(G m N,s(τ 2 (A ′ , B), τ 1 (A ′ , B)) − r B (m)g s (τ 2 (A ′ , B)))<br />

∑<br />

A ′ ∈Cl(F )<br />

ψ 2 (A ′ )G m N,s(τ 2 (A ′ , B), τ ′ 1(A ′ , B))<br />

We can rewrite the sum above in terms <strong>of</strong> height pairings as follows<br />

(85) Σ f,N,B,ψ = ∑<br />

ψ 2 (A ′ )〈(x 2 (A ′ , B)) − (∞), T f ((x 1 (A ′ , B)) − (0))〉 C + U f,B,ψ ,<br />

A ′ ∈Cl(F )


36 W. DUKE AND Y. LI<br />

where x j (A ′ , B) is the preimage <strong>of</strong> the Heegner point τ j (A ′ , B) on X 0 (N) and<br />

⎛<br />

⎞<br />

∑<br />

T f := ⎜<br />

⎝ c(f, −m)T m − (N + 1) ∑ c(f, −Nm ′ )T m ′W N<br />

⎟<br />

⎠ ,<br />

m≥1<br />

m ′ ≥1<br />

N∤m<br />

N∤m ′ (<br />

U f,B,ψ = c f,1ψ(B[n]) + c f,2 ψ(B −1 [n]) + c f,3 ψ(B[n] −1 ) + c f,4 ψ(B −1 [n] −1 ) ∑<br />

ψ 2 (A ′ ) log |u<br />

12H(p)(N 2 A ′|)<br />

− 1)<br />

A ′<br />

for some constants c f,j ∈ Q determined by f(z). Since f(z) ∈ S 1 , the coefficient c(f, −m)<br />

vanishes for all m ≥ 1 satisfying N 2 |m. So we have N ∤ m ′ in the summation above. Using<br />

(49) and equation (2.16) in [24, p. 240], <strong>one</strong> can show that the c f,j ’s are integral linear<br />

combinations <strong>of</strong> c(f, −m), m ≥ 1. Since f is in the integral span <strong>of</strong> S 1 , Lemma 6.4 implies<br />

that c(f, −m) ∈ Z for all m ≥ 1. Then the denominator <strong>of</strong> constant in the definition <strong>of</strong> U f,B,ψ<br />

is independent <strong>of</strong> f.<br />

Given any newform h(z) ∈ S 2 (N), we know that (h|W N )(z) = −(h|U N )(z). It is then<br />

easy to check that T f (h(z)) = 0. Since N is prime, S 2 (N) is spanned by new<strong>forms</strong>. So T f<br />

annihilates any h(z) ∈ S 2 (N). Then T f ((x 1 (A, B)) − (0)) is a trivial divisor on J 0 (N) defined<br />

over H, since the actions <strong>of</strong> the Hecke operators and Fricke involution on the Jacobian are<br />

the same as those on S 2 (N). So it is the divisor <strong>of</strong> a rational function φ f,A,B on X 0 (N),<br />

defined over H. By the axioms defining height pairings, we can rewrite Σ f,N,B,ψ as<br />

Σ f,N,B,ψ =<br />

∑<br />

ψ 2 (A) log |u f,B (A)| + U f,B,ψ ,<br />

where<br />

A∈Cl(F )<br />

u f,B (A) = φ f,A,B(τ 2 (f, A, B))<br />

∈ H.<br />

φ f,A,B (∞)<br />

Given any σ C ∈ Gal(H/F ) associated to C ∈ Cl(F ), it sends the Heegner point x 1 (A, B) to<br />

the Heegner point x 1 (AC −1 , B). Since the Galois action commutes with the Hecke action, we<br />

see that<br />

φ σ C<br />

f,A,B (z)<br />

φ f,AC −1 ,B(z) ∈ H×<br />

is a constant. So σ C (u f,B (A)) = u f,B (AC −1 ). Thus, the proposition holds for all f(z) in the<br />

integral span S 1 .<br />

□<br />

7. Pro<strong>of</strong> <strong>of</strong> Theorem 1.1<br />

From §5 and §6, we know that the regularized inner product between f lift,N,B (z) ∈ M !,+<br />

1 (p)<br />

and g ψ (z) can be put into a form quite similar to (10) and satisfies condition (iii) in Theorem<br />

1.1, whenever f(z) ∈ M 0(N) ! has rational Fourier coefficients and N is either 1 or a prime<br />

number different from p. For a mock-<strong>modular</strong> form ˜g ψ (z) = ∑ n∈Z r+ ψ (n)qn as in Proposition<br />

4.2, (76) gives rise to the following equation involving a linear combination <strong>of</strong> r + ψ (n)’s<br />

∑<br />

δ(k)r B (pm − Nk)r + ψ (k) =〈f lift,N,B , g ψ 〉 reg<br />

(86)<br />

m∈Z<br />

δ N (m)c(f, −m) ∑ k∈Z<br />

+ N ∑ m ′ ≥0<br />

c(f, −Nm ′ )ρ N,B,ψ (m ′ ).


MOCK-MODULAR FORMS OF WEIGHT ONE 37<br />

In this section, we will show that these linear combinations <strong>of</strong> r + ψ<br />

(n)’s are sufficient to determine<br />

˜g ψ (z) up to some element in S1 − (p). Then, <strong>one</strong> can choose an appropriate ˜g ψ (z)<br />

satisfying Theorem 1.1. The following lemma is key to the pro<strong>of</strong> <strong>of</strong> Theorem 1.1. Recall that<br />

δ(n) is the same as in equation (12) and δ N (m) appears in Definition 6.3(ii).<br />

Lemma 7.1. Let N 0 = 1 and {N j : 1 ≤ j ≤ (p − 1)/2} be a set <strong>of</strong> primes congruent to 1<br />

modulo N. Suppose {d(n)} n≥1 is a set <strong>of</strong> complex numbers such that d(n) = 0 for all n with<br />

χ p (n) = 1 and they satisfy the following equations<br />

∑<br />

m∈Z<br />

δ N (m)c(f, −m) ∑ n≥1<br />

∑<br />

δ(n)d(n)r B (pm − n) = 0,<br />

n≥1<br />

δ(n)d(n)r B (pm − Nn) = 0,<br />

for all classes B ∈ Cl(F ) and f(z) ∈ M !,new<br />

0 (N j ), 1 ≤ j ≤ (p − 1)/2. Then the power series<br />

D(z) := ∑ n≥1 d(n)qn is a <strong>weight</strong> <strong>one</strong> cusp form in S − 1 (p).<br />

Pro<strong>of</strong>. Define the sum s j,B (m) for m ≥ 1 and the series φ j,B (z) by<br />

s j,B (m) := ∑ n≥1<br />

φ j,B (z) := ∑ m≥1<br />

s j,B (m)q m .<br />

δ(n)d(n)r B (pm − N j n),<br />

First, it is not hard to see that φ j,B (z) ∈ S 2 (N j ). When j = 0, this is clear since φ j,B (z) is identically<br />

0. When 1 ≤ j ≤ (p−1)/2, N j is a prime and a power series h(z) = ∑ m≥1 c(h, m)qm ∈<br />

CJqK is in S 2 (N j ) if and only if<br />

∑<br />

δ Nj (m)c(f, −m)c(h, m) = 0,<br />

m∈Z<br />

for all f(z) = ∑ m∈Z c(f, m)qm ∈ M !,new<br />

0 (N j ). The sufficiency is a consequence <strong>of</strong> Lemma 6.4.<br />

Next, it is easy to see that<br />

(87)<br />

φ j,B (z) = (ϑ B (z)D(N j z))|U p + ϑ B (z)(D|U p )(N j z),<br />

= 1 p<br />

k=0<br />

p−1<br />

∑ ( (ϑ B<br />

) )<br />

z+k<br />

+ ϑ<br />

p B (z) D<br />

(<br />

Nj z+k<br />

p<br />

)<br />

where U p acts formally on power series in CJqK and we used N j ≡ 1 (mod p) to reach the<br />

second step. Set<br />

M :=<br />

(p+1)/2<br />

∏<br />

j ′ =1<br />

M j := M N j<br />

.<br />

N j ′,


38 W. DUKE AND Y. LI<br />

Then φ j,B (pM j z) is a <strong>modular</strong> form <strong>of</strong> <strong>weight</strong> two, level pM for all 0 ≤ j ≤ (p − 1)/2. Define<br />

the following matrices with entries in CJqK by<br />

( )<br />

)<br />

L :=<br />

(ϑ B M j z + k + ϑ<br />

p B (pM j z)<br />

( ( ))<br />

X := D Mz + k p<br />

0≤k≤p−1<br />

R := (pφ j,B (pM j z)) 0≤j≤(p−1)/2<br />

.<br />

0≤j≤(p−1)/2<br />

0≤k≤p−1.<br />

The dimensions <strong>of</strong> L, X and R are (p + 1)/2 × p, p × 1 and (p + 1)/2 × 1 respectively. Then<br />

we can rewrite equation (87) in the following matrix expression<br />

(88) L · X = R.<br />

Notice that matrix L has rank (p + 1)/2. This can be seen by considering the product<br />

(<br />

)<br />

( )<br />

∑<br />

1 /p<br />

L ·<br />

p e−2πikk′ = δ(k ′ ) r B (n)q M jn<br />

,<br />

0≤k,k ′ ≤p−1<br />

n≡k ′ mod p<br />

0≤j≤(p−1)/2<br />

0≤k ′ ≤p−1.<br />

which, after permuting columns, is composed <strong>of</strong> a (p + 1)/2 × (p − 1)/2 zero submatrix and<br />

a (p + 1)/2 × (p + 1)/2 submatrix <strong>of</strong> the form<br />

(<br />

)<br />

∑<br />

L ′ := δ(n ν ) r B (n)q M jn<br />

.<br />

n≡n ν mod p<br />

0≤j,ν≤(p−1)/2<br />

Here n 0 = 0 and 0 < n 1 < n 2 < · · · < n (p−1)/2 < p are all the residues modulo p. The<br />

determinant <strong>of</strong> this submatrix is a nonzero power series in q with leading term<br />

⎛<br />

⎞<br />

(p−1)/2<br />

∏<br />

⎝ r B (n B,ν ) ⎠ q ∑ 1≤j≤(p−1)/2 M jn B,j<br />

,<br />

ν=1<br />

where {n B,ν } 1≤ν≤(p−1)/2 is a set <strong>of</strong> positive integers such that<br />

(i) 0 < n B,1 < n B,2 < · · · < n B,(p−1)/2 ,<br />

(ii) r B (n B,ν ) ≠ 0 for all 1 ≤ ν ≤ (p − 1)/2,<br />

(iii) r B (n) = 0 if n ≡ n B,ν ′ (mod p) for some ν ′ and n < n B,ν ′.<br />

So the submatrix L ′ has full rank in M (p+1)/2,(p+1)/2 (CJqK) and the null space <strong>of</strong> L is spanned<br />

by the column vectors<br />

(<br />

1<br />

p e−2πikk′ /p<br />

)<br />

0≤k≤p−1<br />

with k ′ running over the (p − 1)/2 non-residues<br />

modulo p. This also shows that L has rank (p + 1)/2 in M (p−1)/2,(p−1) (C) after substituting<br />

q = e 2πiz for all z in a dense subset U ∈ H.<br />

Let γ = ( )<br />

a b<br />

pMc d ∈ Γ0 (pM), then applying | 2 γ to both sides <strong>of</strong> equation (87) gives us a new<br />

equation with the same right hand side. Let c denote the multiplicative inverse <strong>of</strong> c modulo<br />

p and we have the following standard transformation <strong>of</strong> ϑ B (z)<br />

{<br />

(<br />

)<br />

) (<br />

ϑ B<br />

(M j z + k | a b<br />

) χ<br />

p 1 pMc d = p (a + ck)ϑ B M j z + (a+ck)dk , if a + ck ≢ 0 mod p<br />

p<br />

i √ pχ p (−c)ϑ B (pM j z),<br />

if a + ck ≡ 0 mod p<br />

(<br />

ϑ B (pM j z) | a b<br />

)<br />

χ √ )<br />

1 pMc d = p(−c)i p<br />

ϑ<br />

p B<br />

(M j z + cd .<br />

p<br />

Using these properties, equation (88) trans<strong>forms</strong> into<br />

L · T · X ′ = R,


where<br />

T := (T kk ′) 0≤k,k ′ ≤p−1,<br />

T kk ′ := χp(−c)i √ p<br />

(δ cd(k) ′ − gp(k) ) +<br />

p<br />

δ ′ α(β) :=<br />

{<br />

1, if α = β<br />

0, otherwise<br />

MOCK-MODULAR FORMS OF WEIGHT ONE 39<br />

p−1<br />

∑<br />

{ p<br />

g p (k) := 1 (1 + χ<br />

2<br />

p (n))e 2πink/p =<br />

, if p | k<br />

2<br />

i √ pχ p(k)<br />

, if p ∤ k<br />

n=0<br />

2<br />

( ( ) )<br />

X ′ := D Mz + k′ |<br />

p 1 γ<br />

0≤k ′ ≤p−1<br />

{<br />

χp (a + ck ′ )(δ ′ (a+ck ′ )dk ′(k′ ) − gp(k) ),<br />

p<br />

if k ≢ −ac mod p<br />

iχ p(−c)<br />

2 √ g p p(k), if k ≡ −ac mod p<br />

So when q = e 2πiz for any z ∈ H, the column vector X − T · X ′ with entries in C is in)<br />

the null<br />

space <strong>of</strong> L. If z ∈ U, the null space is spanned by the column vectors<br />

with k ′ ∈ Z and χ p (k ′ ) = −1.<br />

Now define a 1 × p row vector V by<br />

(<br />

1<br />

V := (g p p(k) + 1) .<br />

2<br />

)0≤k≤p−1<br />

It is orthogonal to the null space <strong>of</strong> L when z ∈ U. So for all z ∈ U,<br />

Simple calculation shows that<br />

V · X = V · T · X ′ .<br />

V · T = χ p (d)V,<br />

V · X = χ p (d)D(Mz),<br />

V · X ′ = χ p (d)(D(Mz)| 1 γ).<br />

(<br />

1<br />

p e−2πikk′ /p<br />

0≤k≤p−1<br />

So we know that (D(Mz)| 1 γ) = χ p (d)D(Mz) for all z ∈ U. Since U ⊂ H is dense and both<br />

sides are holomorphic, we have (D(Mz)| 1 γ) = χ p (d)D(Mz) for all z ∈ H. This is true for<br />

any γ ∈ Γ 0 (pM), or equivalently<br />

for all ( )<br />

a Mb<br />

pc d ∈ Γ0 (p), b ∈ Z.<br />

Let Γ 0 (p, M) := {γ = ( a b<br />

(<br />

(D| a Mb<br />

1 pc d<br />

)<br />

)(z) = χp (d)D(z)<br />

c d ) : p|a, M|b}. It is not too hard to see that Γ 0(p) is generated<br />

by Γ 0 (p, M) and T = ( 1 1 1 ) ∈ Γ 0 (p). Since D(z) is also invariant under the action <strong>of</strong> T , it<br />

has level p. By the shape <strong>of</strong> the Fourier expansion <strong>of</strong> D(z), it is in S1 − (p).<br />

□<br />

Corollary 7.2. Let N 0 = 1 and N j be distinct primes congruent to 1 modulo p for 1 ≤ j ≤<br />

(p − 1)/2. Then the space M !,+<br />

1 (p) is spanned by the set<br />

{f lift,N j,B (z) : B ∈ Cl(F ), f(z) ∈ M !,+<br />

0 (N j ), 0 ≤ j ≤ (p − 1)/2}.<br />

Pro<strong>of</strong>. Denote the subspace <strong>of</strong> M !,+<br />

1 (p) spanned by the set above by M !,+,lift<br />

1 (p). Suppose<br />

there exists f(z) ∈ M !,+<br />

1 (p) not in M !,+,lift<br />

1 (p). Then by the perfect pairing between M !,+<br />

1 (p)<br />

and S1 − (p), there exists<br />

D(z) = ∑ d(n)q n ∈ CJqK<br />

n≥1<br />

such that


40 W. DUKE AND Y. LI<br />

(i) d(n) = 0 for all n with χ p (n) = 1,<br />

(ii) ∑ n∈Z c(H, n)d(n)δ(n) = 0 for all H(z) = ∑ n∈Z c(H, n)qn ∈ M !,+,lift<br />

1 (p),<br />

(iii) D(z) ∉ S1 − (p).<br />

In particular for H(z) = f lift,N,B (z), condition (ii) can be rewritten as<br />

(89) ∑ m∈Z<br />

c(f, −m) ∑ k∈Z<br />

Also, when H(z) = j lift,B[n]<br />

m ′<br />

∑<br />

δ(k)d(k)<br />

δ(k)d(k)r B (pm − Nk) = N ∑ m ′ ∈Z<br />

k∈Z<br />

So equation (89) can be written as<br />

∑<br />

c(f, −Nm ′ ) ∑ k∈Z<br />

(z) + j lift,B[n]−1<br />

m<br />

(z), we have<br />

(r ′<br />

( )<br />

)<br />

B + r B (pNm ′ − Nk) = 0.<br />

pm ′ −k<br />

N<br />

m∈Z<br />

δ N (m)c(f, −m) ∑ k∈Z<br />

δ(k)d(k)r B (pm − Nk) = 0.<br />

δ(k)d(k)r B<br />

(<br />

By Lemma 7.1, D(z) ∈ S − 1 (p). This contradicts condition (iii) and we must have<br />

M !,+,lift<br />

1 (p) = M !,+<br />

1 (p).<br />

)<br />

pm ′ −k<br />

.<br />

N<br />

With Proposition 6.1 and Lemma 7.1, we are ready to give the pro<strong>of</strong> <strong>of</strong> Theorem 1.1.<br />

Pro<strong>of</strong> <strong>of</strong> Theorem 1.1. Denote the dimension and the q-echelon basis <strong>of</strong> S − 1 (p) by d − and the<br />

basis by<br />

{h t (z) ∈ S − 1 (p) : h t (z) = q nt + O(q n d − +1<br />

), 1 ≤ t ≤ d − }.<br />

Choose a mock-<strong>modular</strong> form ˜g ψ (z) = ∑ n≥n 0<br />

r + ψ (n)qn such that it has a fixed principal part as<br />

in Proposition 4.2 and r + ψ (n t) satisfies conditions (ii) and (iii) in Theorem 1.1 for 1 ≤ t ≤ d − .<br />

Note that if S1 − (p) = ∅, then there is only <strong>one</strong> ˜g ψ (z) with a fixed principal part. We claim<br />

that this ˜g ψ (z) is a desired choice.<br />

Consider all f(z) ∈ M !,new<br />

0 (N j ), 1 ≤ j ≤ (p − 1)/2 with rational Fourier coefficients and<br />

j m (z) for all m ≥ 0. They generate the complex vector space M !,new<br />

0 (N j ) and M 0(1) ! respectively.<br />

They also give rise to equation (86) and equation (66), which form a system <strong>of</strong> equations<br />

satisfied by {r + ψ (n) : n ≥ n 0}. Suppose there is another set <strong>of</strong> solutions {c + (n) : n ≥ n 0 }<br />

with r + ψ (n) = c+ (n) whenever n ≤ 0, χ p (n) = 1, or n = n t for t = 1, . . . , d − . Then by Lemma<br />

7.1, the power series<br />

D(z) := ∑ n≥1<br />

(r + ψ (n) − c+ (n))q n<br />

is a cusp form in S1 − (p). However, since r + ψ (n t) = c + (n t ) for all 1 ≤ t ≤ d − , the cusp<br />

form D(z) must be zero in S1 − (p). So with {r + ψ (n) : n ≤ 0 or n = n t, 1 ≤ t ≤ d − } fixed,<br />

{r + ψ (n) : n ≥ d − + 1} is the unique solution to this system <strong>of</strong> equations. From Corollary<br />

7.2 and Proposition 3.5, it is clear that r + ψ<br />

(n) can be solved from this system <strong>of</strong> equations,<br />

i.e. the r + ψ<br />

(n)’s are a rational linear combination <strong>of</strong> these regularized inner products. By<br />

Corollary 5.2 and Proposition 6.1, we can write<br />

r + ψ (n) = −β ∑<br />

n ψ 2 (A) log |u(n, A)|,<br />

A∈Cl(F )<br />


MOCK-MODULAR FORMS OF WEIGHT ONE 41<br />

with β n ∈ Q, u(n, A) ∈ H independent <strong>of</strong> ψ and σ C (u(n, A)) = u(n, AC −1 ) for all C ∈ Cl(F ).<br />

Now to prove that we can choose β n independent <strong>of</strong> n, it is enough to show that there is a<br />

choice <strong>of</strong> β n whose denominators do not depend on n.<br />

When n > p · g N /N, it is not hard to see from equation (86) that we can explicitly express<br />

r + ψ (n) in terms <strong>of</strong> these regularized inner products and the r+ ψ<br />

(k)’s with k < n. If n = pn′<br />

with n ′ ∈ Z, then substituting f = j n ′(z) into equation (66) gives us<br />

r + ψ (n) = 〈f lift,B , g ψ 〉 reg − ∑ k l such that kN + l ≡ 0<br />

(mod p) and set m = (l + Nk)/p. Notice that the conditions on N and m forces N ∤ m and<br />

χ p (N) = 1. With such a choice <strong>of</strong> N and m, let f m ∈ S 1 as in (82). Then δ(n)δ N (m)r B (pm −<br />

Nn) = 1 and equation (86) becomes<br />

r + ψ (n) = − ∑ (<br />

gN<br />

)<br />

∑<br />

δ(k) δ N (m ′′ )c(f m , −m ′′ )r B (pm ′′ − Nk) r + ψ (k)<br />

k>0 m ′′ =1<br />

− ∑ ( )<br />

∑<br />

δ(k) δ N (m ′′ )c(f m , −m ′′ )r B (pm ′′ − Nk) r + ψ (k)<br />

k≤0 m ′′ ≤g N<br />

+ 〈f lift,N,B<br />

m , g ψ 〉 reg + Nc(f m , 0)ρ N,B,ψ (0).<br />

When k > 0 in the summation on the right hand side, the coefficient <strong>of</strong> r + ψ<br />

(k) is an integer<br />

by Lemma 6.4. Similarly when k ≤ 0, the denominator <strong>of</strong> the coefficient <strong>of</strong> r + ψ<br />

(k) divides<br />

(N 2 − 1). Thus after choosing β n for n ≤ p · g N /N , we can choose the other β n such that<br />

their denominators are independent <strong>of</strong> n.<br />

Now choose some β and u(n, A) for all n, A such that r + ψ<br />

(n) is expressed in terms <strong>of</strong> β and<br />

u(n, A) for all non-trivial character ψ as in equation (10). Define r + A<br />

(n) by<br />

(<br />

r + 1<br />

A<br />

(n) := R + H(p) p (n) +<br />

∑<br />

)<br />

ψ(A)r + ψ (n)<br />

= 1<br />

H(p)<br />

= 1<br />

H(p)<br />

ψ non-trivial<br />

⎛<br />

⎞<br />

∏<br />

⎝R p + (n) + β log<br />

u(n, A ′ )<br />

⎠ − β log |u(n, √ A)|<br />

∣A ′ ∈Cl(F ) ∣<br />

(<br />

R<br />

+<br />

p (n) + β log |N H/F (u(n, A))| ) − β log |u(n, √ A)|,<br />

where the last step follows from the transitive action <strong>of</strong> Gal(H/F ). Clearly, the generating<br />

series ∑ n∈Z r+ A (n)qn is a mock-<strong>modular</strong> form with shadow ϑ A (z). When n ≠ 0, there exists<br />

some b ∈ Q and c n ∈ Q independent <strong>of</strong> A such that<br />

1<br />

H(p)<br />

(<br />

R<br />

+<br />

p (n) + β log |N H/F (u(n, A))| ) = b log |c n |.<br />

This follows from equation (6). Thus, we could choose a different β ′ ∈ Q and u ′ (n, A) ∈ H<br />

such that they satisfy condition (iii) and<br />

−β ′ log |u ′ (n, A)| = b log |c n | − β log |u(n, A)|.<br />

Then r + A (n) = −β′ log |u ′ (n, √ A)| and it is not hard to see that β ′ and u ′ (n, A) satisfies<br />

condition (ii) and (iii) in Theorem 1.1. When ψ is a non-trivial character, ψ 2 is non-trivial


42 W. DUKE AND Y. LI<br />

and we have<br />

r + ψ (n) = −β ∑ A<br />

ψ 2 (A) log |u(n, A)| = −β ′ ∑ A<br />

ψ 2 (A) log |u ′ (n, A)|.<br />

By the choice <strong>of</strong> β ′ and u ′ (n, A), we also have<br />

(90)<br />

∑<br />

R p + (n) = r + A (n)<br />

A∈Cl(F )<br />

∑<br />

= −β ′ log |u ′ (n, A)|<br />

A∈Cl(F )<br />

= − β′<br />

2 log |N H/Q(u ′ (n, A))|.<br />

This is an analogue <strong>of</strong> equation (9) for mock-<strong>modular</strong> <strong>forms</strong> as menti<strong>one</strong>d in the introduction.<br />

□<br />

8. Counting CM points<br />

In preparation for the pro<strong>of</strong> <strong>of</strong> Theorem 1.2, we will count the number <strong>of</strong> CM points on a<br />

hyperbolic circle in terms <strong>of</strong> the representation numbers <strong>of</strong> positive definite binary quadratic<br />

<strong>forms</strong>. As before, such a counting argument is needed to construct a Green’s function at<br />

special points. This construction follows ideas <strong>of</strong> [23], but in the counting argument given<br />

there <strong>one</strong> also sums over all classes <strong>of</strong> a given discriminant. Even when <strong>one</strong> discriminant is<br />

prime, which we are assuming, their pro<strong>of</strong> involves quite an intricate application <strong>of</strong> algebraic<br />

number theory. Surprisingly, the refined version we need for a fixed class has a completely<br />

elementary pro<strong>of</strong> using the classical theory <strong>of</strong> composition <strong>of</strong> binary quadratic <strong>forms</strong>. It<br />

has the further advantage that it applies without extra effort when the other discriminant<br />

is not fundamental. Although we will not give details here, the argument adapts to give a<br />

corresponding refinement <strong>of</strong> the level N case in [24].<br />

Good references for the theory <strong>of</strong> binary quadratic <strong>forms</strong> are the books by Buell [3] and<br />

Cox [14]. Let −d < 0 be a discriminant, not necessarily fundamental, and Q −d the set <strong>of</strong><br />

positive definite integral binary quadratic <strong>forms</strong><br />

Q(x, y) = ax 2 + bxy + cy 2<br />

with −d = b 2 − 4ac, where a, b, c ∈ Z. For any Q ∈ Q −d the associated CM point is defined<br />

to be<br />

τ Q = −b + √ −d<br />

∈ H,<br />

2a<br />

where H is the upper half plane. Clearly Q(τ Q , 1) = 0. The <strong>modular</strong> group Γ = PSL 2 (Z) acts<br />

on Q ∈ Q −d by a linear change <strong>of</strong> variables, which induces linear fractional transformation<br />

on τ Q . Let w Q be the number <strong>of</strong> stabilizers <strong>of</strong> Q. The number <strong>of</strong> equivalence classes <strong>of</strong><br />

quadratic <strong>forms</strong> is the Hurwitz class number H(d). Those classes represented by primitive<br />

<strong>forms</strong> (those with gcd(a, b, c) = 1) comprise a finite abelian group under composition, the<br />

class group. When −d is fundamental, this class group is canonically isomorphic to the ideal<br />

class group <strong>of</strong> Q( √ −d) by sending [Q] to the class A ∈ Cl(Q( √ −d)) containing the fractional<br />

ideal Z + Zτ Q . For primitive Q ∈ Q −d , let Q 2 denote a representative <strong>of</strong> the square class <strong>of</strong><br />

Q.<br />

Associated to Q is the counting function<br />

r Q (k) = 1 2 #{(x, y) ∈ Z2 | Q(x, y) = k},


MOCK-MODULAR FORMS OF WEIGHT ONE 43<br />

which is finite and can be non-zero only for non-negative integers k. Clearly r Q (k) is a class<br />

invariant. When −d is fundamental and A ∈ Cl(Q( √ −d)) is the class associated to [Q], the<br />

counting function r Q (k) is the same as r A (k), the number <strong>of</strong> integral ideals in the class A<br />

having norm equals to k.<br />

The invariant hyperbolic distance d(τ, τ ′ ) between two points τ = x + iy and τ ′ = x ′ + iy ′<br />

in H is defined by (54). The goal now is to count the CM points τ Q ′ <strong>of</strong> a given discriminant<br />

−d that are at a fixed hyperbolic distance from a fixed CM point τ Q <strong>of</strong> a (possibly) different<br />

negative discriminant −D. The number we will study is given by<br />

(91) ρ Q (k, d) = #{Q ′ ∈ Q −d | cosh d(τ Q ′, τ Q ) = k √<br />

dD<br />

}.<br />

Since the hyperbolic distance is invariant under linear fractional transformation, the quantity<br />

ρ Q (k, d) depends only on the class <strong>of</strong> Q. In case −D is the negative <strong>of</strong> a prime we will give<br />

a formula for ρ Q in terms <strong>of</strong> the counting function r Q 2 for the squared class <strong>of</strong> the quadratic<br />

form Q.<br />

Proposition 8.1. Let Q be a positive definite binary quadratic form <strong>of</strong> discriminant −D<br />

where D = p ≡ 3 (mod 4) is a prime. Suppose that −d < 0 is a discriminant and that k ≥ 0.<br />

Then<br />

(92) ρ Q (k, d) = δ(k) r Q 2( k2 −pd<br />

4<br />

).<br />

Our pro<strong>of</strong> <strong>of</strong> this result is an entirely elementary exercise in the classical theory <strong>of</strong> binary<br />

quadratic <strong>forms</strong>. It relies on a remarkable algebraic identity given in (97) below for a particular<br />

quadratic form in the square class, whose existence can be coaxed out <strong>of</strong> the work <strong>of</strong><br />

Dirichlet [16].<br />

Lemma 8.2. Every class <strong>of</strong> primitive positive definite binary quadratic <strong>forms</strong> <strong>of</strong> discriminant<br />

−D contains a representative <strong>of</strong> the form<br />

(93) Q(x, y) = Ax 2 + Bxy + ACy 2<br />

with −D = B 2 − 4A 2 C, A > 0 and gcd(A, B) = 1. Furthermore, the class <strong>of</strong> the square <strong>of</strong> Q<br />

is represented by the form<br />

(94) Q 2 (x, y) = A 2 x 2 + Bxy + Cy 2 .<br />

Pro<strong>of</strong>. It is a well–known result <strong>of</strong> Gauss ([21, §228], see also [3, Prop 4.2 p.50] ) that a<br />

primitive form <strong>of</strong> discriminant −D properly represents a positive integer A prime to −D and<br />

is thus, upon an appropriate change <strong>of</strong> variables, equivalent to a form <strong>of</strong> the shape (A, b, c)<br />

with b 2 − 4Ac = −D. In particular we have that gcd(A, b) = 1.<br />

Next, by means <strong>of</strong> a simple translation transformation x ↦→ x + ty, y ↦→ y we will arrange<br />

that (A, b, c) ∼ (A, B, AC) as desired. This transformation leaves A al<strong>one</strong> and changes b to<br />

B = b + 2At. We now choose t to force<br />

Since c = b2 +D<br />

4A<br />

B 2 = (b + 2At) 2 ≡ −D (mod 4A 2 ).<br />

this is equivalent to solving for t the congruence<br />

tb ≡ c<br />

(mod A),<br />

which is possible since gcd(b, A) = 1.<br />

The fact that (A 2 , B, C) represents the square class <strong>of</strong> (A, B, AC) is a consequence <strong>of</strong> a<br />

classical result <strong>of</strong> Dirichlet on the convolution <strong>of</strong> “united” <strong>forms</strong> (see [3, p.57]). □


Thus<br />

(An<br />

− (2A 2 x + By) )( An + (2A 2 x + By) ) ≡A 2 n 2 − (2A 2 x + By) 2<br />

44 W. DUKE AND Y. LI<br />

Turning now to the pro<strong>of</strong> <strong>of</strong> Proposition 8.1, it is easily checked using (54) that for<br />

Q ′ (x, y) = ax 2 + bxy + cy 2 ∈ Q −d<br />

and Q ∈ Q −D as in (93) above we have<br />

√<br />

(95)<br />

dD cosh d(τQ ′, τ Q ) = 2Ac + 2ACa − Bb.<br />

It follows that the statement <strong>of</strong> Theorem 8.1 is equivalent to the equality<br />

(96) #{(a, b, c) ∈ Z 3 | b 2 − 4ac = −d and 2Ac + 2ACa − Bb = k} = δ(k) r Q 2( k2 −dD<br />

4<br />

),<br />

when D = p. Note that a > 0 for any (a, b, c) in the set since k ≥ 0.<br />

In order to prove this we will establish a direct bijection between the solutions to the<br />

relevant equations. A calculation verifies the truth <strong>of</strong> the following key identity:<br />

(97) 4 Q 2 (x, y) = (2Ac + 2ACa − Bb) 2 − (B 2 − 4A 2 C)(b 2 − 4ac)<br />

where Q 2 is as in (94) above and<br />

x =c − Ca<br />

y =Ba − Ab.<br />

Thus every solution (a, b, c) from (96) gives rise to a solution (x, y) <strong>of</strong><br />

(98) Q 2 (x, y) = k2 − dD<br />

.<br />

4<br />

In fact, there is no need to assume that D = p for this part <strong>of</strong> the argument.<br />

This assumption simplifies our treatment <strong>of</strong> the converse, to which we now turn. Suppose<br />

we are given a solution (x, y) <strong>of</strong> (98). The result is trivial when d = pm 2 and k = pm for<br />

some integer m, so assume otherwise. Then (−x, −y) is another distinct solution. Observe<br />

that since p = 4A 2 C − B 2 we have<br />

4A 2 Q 2 (x, y) ≡ A 2 (4A 2 x 2 + 4Bxy + 4Cy 2 ) ≡ (2A 2 x + By) 2<br />

(mod p).<br />

Since Q is assumed to be primitive we know that p ∤ A.<br />

If p ∤ k we can find an integer a such that<br />

An = ±(2A 2 x + By) + pa<br />

≡A 2 n 2 − 4A 2 Q 2 (x, y) ≡ 0<br />

(mod p).<br />

in precisely <strong>one</strong> <strong>of</strong> the two cases <strong>of</strong> ±. Suppose as we may that An = 2A 2 x + By + pa. Thus<br />

we have<br />

a = An − 2A2 x − By<br />

.<br />

p<br />

Then since p = 4A 2 C − B 2 we have<br />

k = 2A(x + Ca) + 2ACa − B Ba − y<br />

A .<br />

Since gcd(A, B) = 1 we must have that A | (Ba − y) and we may define<br />

b =(Ba − y)/A<br />

c =x + Ca.


MOCK-MODULAR FORMS OF WEIGHT ONE 45<br />

Then we have 2Ac + 2ACa − Bb = k. A computation also shows that b 2 − 4ac = −d. If p | k<br />

then we get two distinct triples (a, b, c) for ±(x, y) which both satisfy these. Obviously the<br />

definition <strong>of</strong> (a, b, c) inverts the map we started with to get (x, y), at least for those pairs<br />

that correspond to an (a, b, c). This finishes the pro<strong>of</strong> <strong>of</strong> Proposition 8.1.<br />

To give a level N version <strong>of</strong> Proposition 8.1, we can start with the more general identity<br />

4N Q ∗ (x, y) = (2ANc + 2ACNa − Bb) 2 − (B 2 − 4A 2 NC)(b 2 − 4Nac)<br />

where Q ∗ (x, y) = A 2 Nx 2 + Bxy + Cy 2 and<br />

x =c − Ca<br />

y =Ba − Ab.<br />

9. Pro<strong>of</strong> <strong>of</strong> Theorem 1.2<br />

In this section we will prove Theorem 1.2. It will be deduced as a corollary <strong>of</strong> a more<br />

general identity for special values <strong>of</strong> certain Borcherds lifts <strong>of</strong> <strong>weight</strong> 1/2 weakly holomorphic<br />

<strong>forms</strong>. We again apply the Rankin method to <strong>forms</strong> with poles. We remark that the method<br />

also yields another pro<strong>of</strong> <strong>of</strong> (12), whose analytic pro<strong>of</strong> in [23] uses Hecke-Eisenstein series <strong>of</strong><br />

<strong>weight</strong> <strong>one</strong>. In fact, the method can be generalized to higher levels in order to obtain some<br />

refinements <strong>of</strong> the type we have been considering <strong>of</strong> certain results <strong>of</strong> [25] on height pairings<br />

<strong>of</strong> Heegner divisors.<br />

Let M<br />

1/2 ! be the space <strong>of</strong> weakly holomorphic <strong>modular</strong> <strong>forms</strong> <strong>of</strong> <strong>weight</strong> 1/2 and level 4<br />

satisfying Kohnen’s plus space condition. It has a canonical basis {f d } d≥0 with d ≡ 0, 3<br />

(mod 4) and Fourier expansions<br />

f d (z) = q −d + ∑ n≥1<br />

c(f d , m)q m .<br />

Let f(z) ∈ M<br />

1/2 ! be a weakly holomorphic form with integral Fourier coefficients c(f, n). In<br />

[4], Borcherds constructed an infinite product Ψ f (z) using c(f, n) as exp<strong>one</strong>nts, and showed<br />

that it is a <strong>modular</strong> form <strong>of</strong> <strong>weight</strong> c(f, 0) and some character. The divisors <strong>of</strong> Ψ f (z) are<br />

supported on cusps and imaginary quadratic irrationals. In particular, if τ is a quadratic<br />

irrational <strong>of</strong> discriminant −D < 0, then its multiplicity in Ψ f (z) is<br />

ord τ (Ψ f ) = ∑ k>0<br />

c(f, −Dk 2 ).<br />

For example, when f(z) = f d (z) with d > 0, the Borcherds product Ψ fd (z) equals to<br />

∏<br />

(99)<br />

(j(z) − j(τ q )) 1/wq ,<br />

q∈Q −d /Γ<br />

where Γ = PSL 2 (Z). Note that when −d is fundamental, w d , the number <strong>of</strong> roots <strong>of</strong> unity in<br />

Q( √ −d), is equal to 2w q for all q ∈ Q −d .<br />

Given f(z) ∈ M ! 1/2 , define a <strong>modular</strong> form f lift,θ (z) ∈ M ! 1(p) by<br />

(100) f lift,θ (z) := U 4 (f(pz)θ(z)),<br />

where U 4 is the standard U-operator. It is easy to verify that f lift,θ (z) ∈ M !,+<br />

1 (p) from its<br />

Fourier expansion.<br />

For ψ a non-trivial character <strong>of</strong> Cl(F ), let<br />

˜g ψ (z) = ∑ n∈Z<br />

r + ψ (n)qn ∈ M − 1 (p)


46 W. DUKE AND Y. LI<br />

be a mock-<strong>modular</strong> form with shadow g ψ (z). By Proposition 3.4, the regularized inner<br />

product 〈f lift,θ , g ψ 〉 reg can be expressed in terms <strong>of</strong> r + ψ<br />

(n) as<br />

(101) 〈f lift,θ , g ψ 〉 reg = ∑ ∑<br />

( )<br />

−pm − k<br />

r + 2<br />

ψ<br />

c(f, m)δ(k).<br />

4<br />

m∈Z<br />

k∈Z<br />

The following theorem relates this inner product, hence the linear combination <strong>of</strong> r + ψ<br />

(n), to<br />

the values <strong>of</strong> the Borcherds lift Ψ f (z).<br />

Theorem 9.1. Let f(z) ∈ M<br />

1/2 ! be a weakly holomorphic <strong>modular</strong> form with integral Fourier<br />

coefficients c(f, n), and Ψ f (z) its Borcherds lift. Suppose ψ is a non-trivial character <strong>of</strong><br />

Cl(F ). Then we have<br />

∑<br />

(102) 〈f lift,θ , g ψ 〉 reg = −2 lim ψ 2 (A) ( log |Ψ f (τ A + ɛ)| 2 + C f log |y A | ) ,<br />

ɛ→0<br />

A∈Cl(F )<br />

where C f = ∑ k∈Z c(f, −pk2 ) is the constant term <strong>of</strong> f lift,θ (z), and τ A is a CM point associated<br />

to the class A.<br />

Remark. Recall that jm<br />

lift,B (z) = j m (pz)ϑ B (z) for B ∈ Cl(F ). When B = A 0 is the principal<br />

class in Cl(F ), we can write<br />

2j lift,A 0<br />

m (z) = j m (pz)U 4 (θ(pz)θ(z))<br />

= (j m (4z)θ(z)) lift,θ<br />

= ∑ t∈Z<br />

f lift,θ<br />

4m−t<br />

(z). 2<br />

t 2 ≤4m<br />

Suppose r A0 (m) = 0, then m is not a perfect square and the right hand side <strong>of</strong> equation (53)<br />

becomes<br />

−2<br />

∑<br />

ψ 2 (A) log |Ψ m (τ A , τ A )| .<br />

A∈Cl(F )<br />

By Kr<strong>one</strong>cker’s identity, this is the same as<br />

⎛<br />

−2<br />

∑<br />

A∈Cl(F )<br />

∑<br />

ψ 2 (A) ⎜<br />

⎝ log |Ψ f4m−t 2<br />

(τ A )| ⎟<br />

⎠ .<br />

t∈Z<br />

t 2 ≤4m<br />

Thus, this case <strong>of</strong> Theorem 5.1 is a consequence <strong>of</strong> Theorem 9.1.<br />

The plan <strong>of</strong> the pro<strong>of</strong> goes as follows. First, notice that both sides <strong>of</strong> equation (102) are<br />

additive with respect to f(z). So it suffices to prove the theorem when f(z) = f d (z) for all<br />

d ≥ 0 and d ≡ 3 or 0 modulo 4. For convenience, we denote<br />

(103) a(d, ψ) := 〈f lift,θ<br />

d<br />

, g ψ 〉 reg .<br />

Then, we define a function Φ(z) analogous to G D (z) in [29], where the inner product between<br />

G D (z) and a <strong>weight</strong> k+1/2 level 4 eigenform g(z) gives special values <strong>of</strong> L-function associated<br />

to the Shimura lift <strong>of</strong> g(z). For us, Φ(z) is non-holomorphic, trans<strong>forms</strong> with <strong>weight</strong> 3/2 and<br />

level 4, and satisfies Kohnen’s plus space condition. Since there is no holomorphic cusp form<br />

<strong>of</strong> <strong>weight</strong> 3/2, level 4 satisfying the plus space condition, the inner product between the<br />

holomorphic projection <strong>of</strong> (almost) Φ(z) and the Poincaré series with an s-variable giving<br />

rise to such cusp <strong>forms</strong> is zero as s approaches zero. This gives us the identity relating finite<br />


MOCK-MODULAR FORMS OF WEIGHT ONE 47<br />

linear combinations <strong>of</strong> r + ψ<br />

(n) to the limit <strong>of</strong> an infinite sum. The finite linear combination<br />

will be exactly the right hand side <strong>of</strong> equation (101). Then, we will relate the limit <strong>of</strong> the<br />

infinite sum to Green’s function, hence Ψ fd (z), via the counting argument in §8. In some<br />

sense, the pro<strong>of</strong> is in the same spirit as Zagier’s pro<strong>of</strong> <strong>of</strong> Borcherds theorem in [44].<br />

The following lemma is analogous to Lemma 6.6 and relates a(d, ψ) with the limit <strong>of</strong> an<br />

infinite sum involving r ψ (n).<br />

Lemma 9.2. Let d ≥ 1 be an integer congruent to 0 or 3 modulo 4. Then we have<br />

(<br />

a(0, ψ)H(d) ∑<br />

( ) ( ) )<br />

(104) a(d, ψ) − = lim δ(k)ϱ s−1 −1 + k2 −pd + k<br />

2<br />

r ψ ,<br />

H(0) s→1 pd 4<br />

k∈Z<br />

where the function ϱ s (µ) is defined by<br />

ϱ s (µ) :=<br />

∫ ∞<br />

1<br />

du<br />

(µu + 1) 1+s<br />

2 u , µ > 0.<br />

Pro<strong>of</strong>. Let ˜g ψ (z) ∈ M − 1 (p) be a mock-<strong>modular</strong> form with shadow g ψ (z) and<br />

ord ∞˜g ψ (z) ≥ − p 4 .<br />

Forms satisfying these conditions exist by Proposition 4.2. The restriction on the order <strong>of</strong><br />

˜g ψ (z) at infinity simplifies the pro<strong>of</strong> and is not important. But the requirement that ˜g ψ (z)<br />

be in the minus space is crucial for the validity <strong>of</strong> the statement.<br />

Denote ĝ ψ (z) ∈ H − 1 (p) the associated harmonic Maass form and define the function Φ(z)<br />

by<br />

Φ(z) = Tr 4p<br />

4 ((ĝ ψ |U p )(4z)θ(pz)),<br />

where Tr 4p<br />

4 is the trace down operator from level 4p to level 4. The function Φ(z) trans<strong>forms</strong><br />

like a <strong>weight</strong> 3/2 <strong>modular</strong> form <strong>of</strong> level 4 and should be compared to G D (z) in [29], which<br />

was defined by applying the trace down operator Tr 4D<br />

4 to the product <strong>of</strong> the holomorphic<br />

<strong>weight</strong> k Eisenstein series and θ(|D|z). If <strong>one</strong> replaces the <strong>weight</strong> k holomorphic Eisenstein<br />

series by the <strong>weight</strong> <strong>one</strong> real-analytic Eisenstein series with an s variable, and consider the<br />

derivative with respect to s, then it is something quite similar to this Φ(z).<br />

Using the coset representatives<br />

Γ 0 (4p)\Γ 0 (4) =<br />

p−1<br />

⊔<br />

µ=0<br />

(<br />

1<br />

4 1<br />

) (<br />

1 µ<br />

1<br />

) ⊔<br />

(<br />

1<br />

1)<br />

,<br />

and the fact that (ĝ ψ |U p )(z) = −i(ĝ ψ |W p )(z), We can calculate the Fourier expansion <strong>of</strong><br />

Φ(z) = Tr 4p<br />

4 ((ĝ ψ |U p )(4z)θ(pz)) as<br />

Φ(z) = ∑ n∈Z(b(n, y) + a(n))q n ,<br />

b(n, y) = − ∑<br />

a(n) = ∑ k∈Z<br />

k 2 −4m=pn<br />

δ(k)r + ψ<br />

δ(k)r ψ (m)β 1<br />

(m, 4y p<br />

( pn − k<br />

2<br />

Since p ≡ 3 (mod 4), Φ(z) satisfies Kohnen’s plus space condition and n ≡ 0 or 3 (mod 4).<br />

Since ord ∞˜g ψ (z) ≥ −p/4, the right hand side <strong>of</strong> equation (101) equals to a(d) for f(z) = f d (z).<br />

4<br />

)<br />

.<br />

)<br />

,


48 W. DUKE AND Y. LI<br />

So we have<br />

(105) a(d) = a(d, ψ).<br />

Let F(z) be the <strong>weight</strong> 3/2 Eisenstein series studied in [28], which has the following Fourier<br />

expansion<br />

∞∑<br />

∑ ∞<br />

F(z) = H(n)q n + y −1/2 1<br />

16π β 3/2(m 2 , y)q −m2 ,<br />

n=0<br />

m=−∞<br />

and satisfies Kohnen’s plus space condition. For all n ∈ Z, notice that b(n, y)q n is exp<strong>one</strong>ntially<br />

decaying as y → ∞. Also, a(n) vanishes for all n < 0. Thus, the function<br />

Φ ∗ (z) := Φ(z) − a(0)F(z)<br />

H(0)<br />

is O(y −1/2 ) as y → ∞. The same decaying property holds at the other two cusps <strong>of</strong> Γ 0 (4) as<br />

well, since Φ ∗ (z) satisfies Kohnen’s plus space condition. So we can consider the holomorphic<br />

projection <strong>of</strong> Φ ∗ (z) to the Kohnen plus space S + 3/2 (Γ 0(4)). Define the <strong>weight</strong> 3/2 Poincaré<br />

series by<br />

∑<br />

(106)<br />

P d (z, s) :=<br />

j(γ, z) −3 e 2dπiγz Im(γz) s/2 ,<br />

where for γ ∈ Γ 0 (4)<br />

γ∈Γ ∞\Γ 0 (4)<br />

j(γ, z) := θ(γz)<br />

θ(z) .<br />

This series converges absolutely for Re(s) > 1 and can be analytically continued to Re(s) ≥<br />

0. As s → 0, the inner product 〈Φ ∗ (z), P d (z, s)〉 is the d th coefficient <strong>of</strong> a cusp form in<br />

S + 3/2 (Γ 0(4)), since Φ ∗ (z) is already in the plus space. Given S + 3/2 (Γ 0(4)) = {0}, we know the<br />

limit is zero and obtain the following equation after applying Rankin-Selberg unfolding,<br />

( Γ(<br />

1+s<br />

(107) lim<br />

) (<br />

2<br />

a(d) − a(0)H(d) ) ∫ ∞<br />

)<br />

+ b(d, y)e −4πdy 1/2+s/2 dy<br />

y = 0.<br />

s→0 (4πd) 1/2+s/2 H(0)<br />

y<br />

After some manipulations, we have<br />

∫ ∞<br />

After substituting µ = k2<br />

pd<br />

(108) −<br />

∫ ∞<br />

Here, we used the fact<br />

0<br />

0<br />

β 1 (d, µy)e −4πdy 1/2+s/2 dy<br />

y<br />

y<br />

− 1 and µ =<br />

pk2<br />

d<br />

b(d, y)e −4πdy 1/2+s/2 dy<br />

y<br />

y<br />

0<br />

=<br />

Γ(<br />

1+s<br />

2 )<br />

(4πd) 1/2+s/2 ϱ s (µ) .<br />

− 1, we arrive at the following expression<br />

∑<br />

( ) −pd + k<br />

2<br />

δ(k)r ψ ϱ s<br />

(−1 + k2<br />

4<br />

pd<br />

1+s<br />

Γ(<br />

2<br />

= )<br />

(4πd) 1+s<br />

2<br />

k∈Z<br />

β 1 (d, αy) = β 1 (dα, y)<br />

for all α, y, d ∈ R >0 . Now substituting (108) and (105) into (107) gives us equation (104)<br />

Pro<strong>of</strong> <strong>of</strong> Theorem 9.1. When d = 0, f d (z) = θ(z) is the Jacobi theta function and θ lift,θ (z) =<br />

2ϑ A0 (z) is twice the <strong>weight</strong> <strong>one</strong> theta series associated to the principal class A 0 . The<br />

Borcherds lift <strong>of</strong> θ(z) is η 2 (z) and equation (102) becomes<br />

〈2ϑ A0 , g ψ 〉 = −4 ∑<br />

ψ 2 (A) log ∣ √ y A η 2 (τ A ) ∣ A∈Cl(F )<br />

)<br />

.<br />


MOCK-MODULAR FORMS OF WEIGHT ONE 49<br />

This is justified by Proposition 4.1 and the facts that<br />

1 ∑<br />

〈ϑ A0 , g ψ 〉 =<br />

〈g ψ ′, g ψ 〉,<br />

H(p)<br />

ψ ′ :Cl(F )→C × ∣ √ y A η 2 (τ A ) ∣ = ∣ √ y A η 2 (τ A ) ∣ .<br />

When d > 0 is congruent to 0 or 3 modulo 4, we need to study the right hand side<br />

<strong>of</strong> equation (104), which will yield the Green’s function. The procedure here follows part<br />

<strong>of</strong> §5 in [23] and is analogous to the pro<strong>of</strong> <strong>of</strong> Theorem 5.1. Recall that Q s−1 (t) is the<br />

Legendre function <strong>of</strong> the second kind defined in (55). By comparing it with ϱ s (µ), we see<br />

that ϱ 0 (µ) = 2Q 0 ( √ µ + 1) and<br />

Q s−1 ( √ µ + 1) −<br />

( )<br />

Substituting 22−s Γ(2s)<br />

Q |k|<br />

sΓ(s) 2 s−1<br />

√ pd<br />

equation (104) yields<br />

a(d, ψ) −<br />

a(0, ψ)H(d)<br />

H(0)<br />

⎛<br />

= lim ⎝2<br />

s→1<br />

sΓ(s)2<br />

2 2−s Γ(2s) ϱ s−1(µ) = O(µ −1/2−s/2 ).<br />

( )<br />

for ϱ s−1 −1 + k2 in the limit on the right hand side <strong>of</strong><br />

∑<br />

A∈Cl(F )<br />

pd<br />

ψ(A) ∑<br />

k> √ pd<br />

δ(k)r A<br />

( −pd + k<br />

2<br />

4<br />

)<br />

2Q s−1<br />

( k<br />

√ pd<br />

) ⎞ ⎠ .<br />

The factor <strong>of</strong> 2 next to the summation comes from the contribution k < − √ pd.<br />

Let Q A ∈ Q −p be the quadratic form such that Q A (τ A , 1) = 0. By Proposition 8.1 we have<br />

( ) ( )<br />

ρ QA (k, d) = δ(k)r Q 2<br />

A<br />

= δ(k)r A 2 ,<br />

k 2 −pd<br />

4<br />

k 2 −pd<br />

4<br />

where ρ QA (k, d) is defined in (91) and counts the number <strong>of</strong> quadratic <strong>forms</strong> Q ′ ∈ Q −d such<br />

that cosh d(τ QA , τ Q ′) = √ k<br />

pd<br />

and is independent <strong>of</strong> the choice <strong>of</strong> τ A . So equation (104) becomes<br />

a(d, ψ) −<br />

a(0, ψ)H(d)<br />

H(0)<br />

= (−2) lim<br />

= (−2) lim<br />

∑<br />

s→1<br />

A∈Cl(F )<br />

∑<br />

s→1<br />

A∈Cl(F )<br />

= (−2) lim ⎜<br />

⎝ lim<br />

ɛ→0<br />

⎛<br />

⎛<br />

ψ(A 2 ) ⎝ ∑<br />

ψ 2 (A) ∑<br />

∑<br />

s→1<br />

A∈Cl(F )<br />

k> √ pd<br />

⎞<br />

( )<br />

k<br />

−2Q s−1<br />

√ pd<br />

ρ QA (k, d) ⎠<br />

g s (τ A , τ Q )<br />

Q∈Q −d<br />

τ Q ≠τ A<br />

ψ 2 (A) ∑<br />

⎞<br />

g s (τ A + ɛ, τ Q ) ⎟<br />

⎠<br />

Q∈Q −d<br />

τ Q ≠τ A<br />

where g s (τ 1 , τ 2 ) is defined in (57).<br />

As long as d is not <strong>of</strong> the form pm 2 for any integer m, the condition τ Q ≠ τ A is satisfied<br />

by all Q ∈ Q −d . Otherwise, there exists exactly <strong>one</strong> Q ∈ Q −d such that τ Q = τ A . Since<br />

c(f d , 0) = 0, the number <strong>of</strong> such Q ∈ Q −d can also be written as<br />

1<br />

C 2 f d<br />

= ∑ c(f d , −pm 2 ) = 1.<br />

m>0


50 W. DUKE AND Y. LI<br />

To treat both cases simultaneously, we can write<br />

⎛<br />

⎞<br />

∑<br />

g s (τ A + ɛ, τ Q ) = ⎝ ∑<br />

g s (τ A + ɛ, τ Q ) ⎠ − 1C 2 f d<br />

g s (τ A + ɛ, τ A )<br />

Q∈Q −d<br />

Q∈Q −d<br />

disc(τ Q )≠−p<br />

=<br />

⎛<br />

⎝<br />

∑<br />

Q∈Q −d /Γ<br />

⎞<br />

G s (τ A + ɛ, τ Q ) ⎠ − 1C 2 f d<br />

g s (τ A + ɛ, τ A ),<br />

where G s (τ 1 , τ 2 ) is the automorphic Green’s function defined in (56). After substituting this<br />

into equation (109), it becomes<br />

⎛<br />

⎞<br />

a(0, ψ)H(d)<br />

∑ ∑<br />

a(d, ψ) − =(−2) lim ⎝lim ψ 2 (A) G s (τ A + ɛ, τ Q ) ⎠ +<br />

H(0)<br />

ɛ→0 s→1<br />

A∈Cl(F ) Q∈Q −d /Γ<br />

(109)<br />

⎛<br />

⎞<br />

C fd lim ⎝ ∑<br />

ψ 2 (A)g 1 (τ A + ɛ, τ A ) ⎠ .<br />

ɛ→0<br />

A∈Cl(F )<br />

The first summand above can be evaluated using the identity (58). Substituting τ 1 = τ A +ɛ,<br />

τ 2 = τ Q into equation (58) and summing them together over A ∈ Cl(F ) and Q ∈ Q −d /Γ with<br />

coefficient ψ 2 (A) gives us<br />

⎛<br />

⎞<br />

∑<br />

ψ 2 (A) log |Ψ fd (τ A + ɛ)| 2 = lim ⎝ ∑ ∑<br />

ψ 2 (A) G s (τ A + ɛ, τ Q ) + 4πE(τ A + ɛ, s) ⎠ .<br />

s→1<br />

A<br />

A<br />

By (49), we have<br />

⎛<br />

lim ⎝ ∑<br />

s→1<br />

A<br />

ψ 2 (A)<br />

So the first summand is<br />

⎛<br />

−2 ⎝lim<br />

∑<br />

Q∈Q −d /Γ<br />

∑<br />

ɛ→0<br />

A∈Cl(F )<br />

⎞<br />

Q∈Q −d /Γ<br />

4πE(τ A , s) ⎠ = −24H(d) ∑ A<br />

a(0, ψ)H(d)<br />

= −<br />

2H(0)<br />

ψ 2 (A) log |Ψ fd (τ A + ɛ)| 2 ⎞<br />

(<br />

1 + 4y2<br />

⎠ −<br />

ψ 2 (A) log ∣ ∣ √ y A η 2 (τ A ) ∣ ∣<br />

a(0, ψ)H(d)<br />

.<br />

H(0)<br />

)<br />

Substitute g 1 (τ + ɛ, τ) = − log with τ = τ<br />

ɛ 2 A into the second summand, we obtain<br />

⎛<br />

⎞<br />

a(0, ψ)H(d)<br />

∑<br />

a(d, ψ) − = − 2 ⎝lim ψ 2 (A) log |Ψ fd (τ A + ɛ)| 2 ⎠<br />

a(0, ψ)H(d)<br />

−<br />

H(0)<br />

ɛ→0<br />

H(0)<br />

− 2C fd<br />

A∈Cl(F )<br />

∑<br />

A∈Cl(F )<br />

ψ 2 (A) log |y A |.<br />

After canceling the term − a(0,ψ)H(d)<br />

H(0)<br />

, equation (109) becomes equation (102) for f = f d , d > 0.<br />


MOCK-MODULAR FORMS OF WEIGHT ONE 51<br />

We can now easily deduce Theorem 1.2. It is convenient to give a slightly more general<br />

version stated in terms <strong>of</strong> theta series and quadratic <strong>forms</strong>.<br />

Corollary 9.3. Let p be a prime number congruent to 3 modulo 4, and −d a discriminant<br />

not <strong>of</strong> the form −pm 2 for any integer m. For a class B ∈ Cl(F ), suppose ˜ϑ B 2(z) =<br />

∑<br />

n∈Z r+ B<br />

(n)q n ∈ M − 2 1 (p) has shadow ϑ B 2(z) and ord ∞ ( ˜ϑ B 2) > −p/4. Then<br />

∣<br />

(110) log<br />

∣<br />

∏<br />

q∈Q −d /Γ<br />

(j(τ B ) − j(τ q )) 1/wq ∣ ∣∣∣∣∣<br />

= − 1 4<br />

∑<br />

Remark. When −d is fundamental, w d = 2w q for all q ∈ Q −d .<br />

k∈Z<br />

δ(k) r + B 2 ( pd−k2<br />

4<br />

).<br />

Pro<strong>of</strong>. Let f(z) = f d (z) with d not <strong>of</strong> the form pm 2 . Using the Borcherds lift <strong>of</strong> f d (z) in (99),<br />

the right hand side <strong>of</strong> (102) can be rewritten as<br />

∣<br />

−4<br />

∑<br />

∏<br />

∣∣∣∣∣<br />

ψ 2 (A) log<br />

(j(τ A ) − j(τ q )) 1/wq .<br />

∣<br />

A∈Cl(F )<br />

q∈Q −d /Γ<br />

For each non-trivial character ψ, choose ˜g ψ (z) ∈ M − 1 (p) having shadow g ψ (z) and r + ψ (n) = 0<br />

for all n ≤ −p/4. For these ˜g ψ (z), equation (101) becomes<br />

〈f lift,θ<br />

d<br />

, g ψ 〉 reg = ∑ ( ) pd − k<br />

r + 2<br />

ψ<br />

δ(k).<br />

4<br />

k∈Z<br />

Combining these two together and dividing both sides by -4, we have<br />

∣<br />

∑<br />

( ) pd − k<br />

(111) − 1 r + 2<br />

4 ψ<br />

δ(k) =<br />

∑<br />

∣∣∣∣∣ ∏<br />

ψ 2 (A) log<br />

4<br />

(j(A) − j(τ q )) 1/wq ,<br />

∣<br />

k∈Z<br />

A∈Cl(F )<br />

which is precisely equation (15) when −d is fundamental.<br />

Consider the mock-<strong>modular</strong> form ˜F (z) ∈ M − 1 (p) defined by<br />

⎛<br />

˜F (z) := 1<br />

H(p)<br />

⎜<br />

⎝ E+ 1 (z) +<br />

∑<br />

ψ:Cl(F )→C ×<br />

ψ non-trivial<br />

q∈Q −d /Γ<br />

⎞<br />

ψ(B −2 )˜g ψ (z) ⎟<br />

⎠ − ˜ϑ B 2(z).<br />

∑<br />

Since ϑ B 2(z) = 1<br />

H(p) ψ ψ(B−2 )g ψ (z), ˜F (z) has shadow F (z) = 0, hence is <strong>modular</strong>. Also, it<br />

has Fourier expansion in the form<br />

˜F (z) =<br />

∑<br />

c + (n)q n .<br />

From the pro<strong>of</strong> <strong>of</strong> Lemma 9.2, we know that<br />

∑ (<br />

c +<br />

k∈Z<br />

pd−k 2<br />

4<br />

n>−p/4<br />

χ p(n)≠1<br />

)<br />

δ(k) = 〈f lift,θ<br />

d<br />

, F 〉 reg = 0.


52 W. DUKE AND Y. LI<br />

for any d. By the same argument in the pro<strong>of</strong> <strong>of</strong> Theorem 9.1, <strong>one</strong> can show that equation<br />

(12) holds when d ≠ pm 2 for any integer m. So in that case, the equation above becomes<br />

∣<br />

− 1 ∑<br />

( ) pd − k<br />

r + 2<br />

1 ∑<br />

B<br />

δ(k) = ψ(B −2 ) ∑ ∣∣∣∣∣<br />

∏<br />

ψ 2 (A) log<br />

4<br />

2 4<br />

H(p)<br />

− j(τ q ))<br />

k∈Z<br />

ψ<br />

A ∣q∈Q −d /Γ(j(A) 1/wq ∣ ∣∣∣∣∣ ∏<br />

= log<br />

(j(B) − j(τ q )) 1/wq .<br />

∣<br />

q∈Q −d /Γ<br />

10. Case p = 283: Some numerical calculations<br />

Finally, we will present another numerical example to demonstrate that statements similar<br />

to the Conjecture in §1 should be true for octahedral new<strong>forms</strong>. These calculations were<br />

conducted in SAGE [41]. When p = 283, Cl(F ) has order 3 and the space S 1 (p) is 3-<br />

dimensional spanned by dihedral newform h(z) and octahedral new<strong>forms</strong> f ± (z) = f 1 (z) ±<br />

√ −2f2 (z), where<br />

h(z) = q + q 4 − q 7 + q 9 + O(q 10 ),<br />

f 1 (z) = q − q 4 + 2q 6 − q 7 − q 9 + O(q 10 ),<br />

f 2 (z) = q 2 − q 3 − q 5 + O(q 10 ).<br />

By Lemma 3.3, h(z), f 1 (z) ∈ S + 1 (p) and f 2 (z) ∈ S − 1 (p). By Deligne-Serre Theorem, f ± (z)<br />

arises from ρ : Gal(K/Q) → GL 2 (C), where K is the degree 2 extension <strong>of</strong> the normal closure<br />

<strong>of</strong> Q[X]/(X 4 − X − 1) such that Gal(K/Q) ∼ = GL 2 (F 3 ).<br />

Let F 6 be the subfield <strong>of</strong> K fixed by a subgroup <strong>of</strong> Gal(K/Q) isomorphic to Z/8Z, and H 3<br />

the cubic subfield <strong>of</strong> F 6 . Explicitly, we can write<br />

H 3 = Q[Y ]/(Y 3 + 4Y + 1),<br />

F 6 = Q[X]/(X 6 − 3X 5 + 6X 4 − 7X 3 + 10X 2 − 7X + 6),<br />

□<br />

with the embedding<br />

H 3 ↩→ F 6<br />

Y ↦→ X 2 − X + 1.<br />

Fix a complex embedding <strong>of</strong> F 6 ↩→ C such that X = t, Y = θ with t, θ ∈ C having positive<br />

imaginary part. Since F 6 is totally complex, the unit group <strong>of</strong> F 6 has rank 2 and is generated<br />

by<br />

u 1 = t 2 − t + 1, u 2 = t 3 + t − 1.<br />

As in the case <strong>of</strong> dihedral new<strong>forms</strong>, <strong>one</strong> can express the Petersson norms <strong>of</strong> f + (z) and f − (z)<br />

as<br />

〈f + , f + 〉 = 〈f − , f − 〉 = p + 1<br />

12<br />

Res s=1<br />

ζ(s)L(s, F 6 )<br />

L(s, H 3 )ζ(2s)(1 + p −s ) .<br />

Notice the ratio L(s, F 6 )/L(s, H 3 ) is the L-function <strong>of</strong> the quadratic Hecke character <strong>of</strong> H 3<br />

associated to the Hilbert quadratic norm residue symbol <strong>of</strong> F 6 /H 3 . So it is holomorphic at<br />

s = 1 and the right hand side can be evaluated to be 8 log |u 1 u 2 2|.


MOCK-MODULAR FORMS OF WEIGHT ONE 53<br />

By Proposition 3.5, there exists mock <strong>modular</strong> <strong>forms</strong> ˜f 1 (z) ∈ M − 1 (p) and ˜f 2 (z) ∈ M + 1 (p)<br />

with shadows f 1 , f 2 respectively such that<br />

∑<br />

˜f 1 (z) = c + 1 (−4)q −4 + c + 1 (−1)q −1 + c + 1 (0) +<br />

c + 1 (n)q n ,<br />

˜f 2 (z) = c + 2 (−2)q −2 + c + 2 (0) +<br />

Using Proposition 3.4, we find that<br />

∑<br />

n>0,χ 283 (n)≠−1<br />

n>0,χ 283 (n)≠1<br />

c + 2 (n)q n .<br />

−c + 1 (−4) = c + 1 (−1) = c + 2 (−2) = 2 log |u 1 u 2 2|.<br />

Since M 1 + (p) ⊂ M + 1 (p) and M1 − (p) ⊂ M − 1 (p) are both non-empty, the principal part <strong>of</strong> ˜f j (z)<br />

does not determine it uniquely. However, once c + 1 (2) is chosen, then ˜f 1 is fixed. Similarly,<br />

once c + 2 (0), c + 2 (1) and c + 2 (4) are chosen, ˜f2 is fixed. Since we expect c + j (n) to be logarithms<br />

<strong>of</strong> absolute values <strong>of</strong> algebraic numbers in F 6 , it is natural to choose c + 1 (2), c + 2 (0), c + 2 (1) and<br />

c + 2 (4) this way and study the other coefficients. So we will write<br />

c + j (n) = β j log |u j (n)|,<br />

where β j ∈ Q and u j (n) ∈ C × is some complex number. Then we fix ˜f j (z) by letting<br />

β 1 = 2, β 2 = u 2 (0) = u 2 (1) = 1 and<br />

u 2 (4) = u 2 1(2) = − 7 16 t5 + 5 4 t4 − 7 8 t3 + 47<br />

16 t2 − 13<br />

16 t + 19<br />

8 .<br />

From the numerical calculations, we can make predictions <strong>of</strong> the algebraicity <strong>of</strong> u j (n). In the<br />

table below, we list c + j (l) for j = 1, 2 and various primes l. Also, we list the predicted values<br />

˜β j and fractional ideals generated by ũ j (n) in F 6 .<br />

When χ p (l) ≠ 1, the ideal (l) splits into L 0 L 1 in H 3 , and L 1 splits into l l,1 l l,2 in F 6 .<br />

When χ p (l) = 1 and c(f 1 , l) = ±1, the ideal (l) splits into l 1 l 2 in F 6 . When χ p (l) = 1 and<br />

c(f 1 , l) = 0, the ideal (l) splits into L l,0 L l,1 in H 3 and L l,1 splits into l l,1 l l,2 in F 6 , where<br />

l l,j has order 4 in Cl(F 6 ), the class group <strong>of</strong> F 6 . Since F 6 and H 3 have class numbers 8 and<br />

2 respectively, the fractional ideals in Table 3 and Table 4 are all principal. For example,<br />

the value we chose for u 2 1(2) generates the fractional ideal (l 2,1 /l 2,2 ) 4 , where l 2,1 and l 2,2 have<br />

order 8 in Cl(F 6 ). So it is necessary to make ˜β 1 = 1/2. The numerical pattern also justifies<br />

this choice <strong>of</strong> ˜f j (z).<br />

Table 3. Coefficients <strong>of</strong> ˜f 1 (z)<br />

l c(f 2 , l) c + 1 (l) ˜β1 (ũ 1 (l))<br />

2 1 1.2075349695016218 1/2 (l 2,1 /l 2,2 ) 4<br />

3 -1 -0.44226603950742649 1/2 (l 3,1 /l 3,2 ) 4<br />

5 -1 -3.9855512247433431 1/2 (l 5,1 /l 5,2 ) 4<br />

17 0 -3.2181607607379323 1/2 (l 17,1 /l 17,2 ) 4<br />

19 1 5.3481233955176073 1/2 (l 19,1 /l 19,2 ) 4<br />

31 1 -1.7192005338244623 1/2 (l 31,1 /l 31,2 ) 4<br />

37 0 0.32541651822318252 1/2 (l 37,1 /l 37,2 ) 4


54 W. DUKE AND Y. LI<br />

43 1 -4.6200896216743352 1/2 (l 43,1 /l 43,2 ) 4<br />

47 -1 -1.0203031328088645 1/2 (l 47,1 /l 47,2 ) 4<br />

53 0 5.8419201851710110 1/2 (l 53,1 /l 53,2 ) 4<br />

67 0 -3.9318486618330462 1/2 (l 67,1 /l 67,2 ) 4<br />

79 0 7.5720154112893967 1/2 (l 79,1 /l 79,2 ) 4<br />

107 0 -0.81774052769784944 1/2 (l 107,1 /l 107,2 ) 4<br />

109 -1 4.8808523053451562 1/2 (l 109,1 /l 109,2 ) 4<br />

283 0 6.86483405137284 1/2 (l 283,1 /l 283,2 ) 4<br />

Table 4. Coefficients <strong>of</strong> ˜f 2 (z)<br />

l c(f 1 , l) c + 2 (l) ˜β2 (ũ 2 (l))<br />

7 -1 -3.27983974462451 1 l 7,1 /l 7,2<br />

11 1 -2.56257986300244 1 l 11,1 /l 11,2<br />

13 1 -5.57196302179201 1 l 13,1 /l 13,2<br />

23 -1 1.01652189648251 1 l 23,1 /l 23,2<br />

29 -1 1.54494007675715 1 l 29,1 /l 29,2<br />

41 1 0.771808245755645 1 l 41,1 /l 41,2<br />

71 0 -4.99942007705695 1 (l 71,1 /l 71,2 ) 2<br />

73 0 -1.64986308549260 1 (l 73,1 /l 73,2 ) 2<br />

83 -2 8.97062724307569 1 1<br />

89 -1 -0.399183274865547 1 l 89,1 /l 89,2<br />

101 0 5.99108448704487 1 (l 101,1 /l 101,2 ) 2<br />

283 1 4.48531362153791 1 1<br />

643 2 2.32782185606303e-10 1 1<br />

773 2 5.08403073372794e-9 1 1<br />

859 -2 -8.97062719191082 1 1<br />

References<br />

1. Atkin,A.O.L. Modular Forms <strong>of</strong> Weight 1 and Supersingular Reduction U.S.-Japan Seminar on Modular<br />

Functions, Ann Arbor, 1975<br />

2. Buchholz, Herbert The confluent hypergeometric function with special emphasis on its applications. Translated<br />

from the German by H. Lichtblau and K. Wetzel. Springer Tracts in Natural Philosophy, Vol. 15<br />

Springer-Verlag New York Inc., New York 1969 xviii+238 pp<br />

3. Buell, Duncan A. Binary quadratic <strong>forms</strong>. Classical theory and modern computations. Springer-Verlag,<br />

New York, 1989. x+247 pp.<br />

4. Borcherds, Richard E. Automorphic <strong>forms</strong> on O s+2,2 (R) and infinite products. Invent. Math. 120 (1995),<br />

no. 1, 161-213.<br />

5. Borcherds, Richard E. The Gross-Kohnen-Zagier theorem in higher dimensions. Duke Math. J. 97 (1999),<br />

no. 2, 219-233.<br />

6. Bringmann, Kathrin; Ono, Ken, Lifting cusp <strong>forms</strong> to Maass <strong>forms</strong> with an application to partitions.<br />

Proc. Natl. Acad. Sci. USA 104 (2007), no. 10, 3725-3731<br />

7. Bruinier, Jan Hendrik Borcherds products on O(2, l) and Chern classes <strong>of</strong> Heegner divisors. Lecture<br />

Notes in <strong>Mathematics</strong>, 1780. Springer-Verlag, Berlin, 2002. viii+152 pp.<br />

8. Bruinier, Jan Hendrik; Bundschuh, Michael On Borcherds products associated with lattices <strong>of</strong> prime<br />

discriminant. Rankin memorial issues. Ramanujan J. 7 (2003), no. 1-3, 49-61.<br />

9. Bruinier, Jan Hendrik; Funke, Jens On two geometric theta lifts. Duke Math. J. 125 (2004), no. 1, 45-90.<br />

10. Bruinier, Jan Hendrik; Ono, Ken, Heegner Divisors, L-Functions and Harmonic Weak Maass Forms,<br />

Annals <strong>of</strong> Math. 172 (2010), 2135–2181<br />

11. Bruinier, Jan H.; Ono, Ken; Rhoades, Robert C. Differential Operators for harmonic weak Maass <strong>forms</strong><br />

and the vanishing <strong>of</strong> Hecke eigenvalues. Math. Ann. 342 (2008), no. 3, 673693.


MOCK-MODULAR FORMS OF WEIGHT ONE 55<br />

12. Bruinier, Jan Hendrik; Yang, Tonghai CM-values <strong>of</strong> Hilbert <strong>modular</strong> functions. Invent. Math. 163 (2006),<br />

no. 2, 229-288.<br />

13. Bruinier, Jan Hendrik Harmonic Maass <strong>forms</strong> and periods, preprint (2011)<br />

14. Cox, David A. Primes <strong>of</strong> the form x 2 + ny 2 . Fermat, class field theory and complex multiplication. A<br />

Wiley-Interscience Publication. John Wiley & Sons, Inc., New York, 1989. xiv+351 pp.<br />

15. Deligne, Pierre; Serre, Jean-Pierre Formes modulaires de poids 1. (French) Ann. Sci. cole Norm. Sup. (4)<br />

7 (1974), 507-530 (1975).<br />

16. L. Dirichlet, Werke II.<br />

17. Duke, W.; Imamoḡlu.Ö. ; Tóth, Á . Cycle integrals <strong>of</strong> the j-function and mock <strong>modular</strong> <strong>forms</strong>. Ann. <strong>of</strong><br />

Math. (2) 173 (2011), no. 2, 947-981.<br />

18. Duke, W.; Imamoḡlu.Ö. ; Tóth, Á . Real quadratic analogs <strong>of</strong> traces <strong>of</strong> singular moduli. Int. Math. Res.<br />

Not. IMRN 2011, no. 13, 3082-3094<br />

19. Ehlen, Stephan On CM values <strong>of</strong> Borcherds products and <strong>weight</strong> <strong>one</strong> harmonic weak Maass <strong>forms</strong>, in<br />

preparation.<br />

20. Fay, John D. Fourier coefficients <strong>of</strong> the resolvent for a Fuchsian group. J. Reine Angew. Math. 293/294<br />

(1977), 143-203.<br />

21. C.F. Gauss, Werke, I.<br />

22. Gross, Benedict H. Heegner points on X 0 (N). Modular <strong>forms</strong> (Durham, 1983), 87-105, Ellis Horwood<br />

Ser. Math. Appl.: Statist. Oper. Res., Horwood, Chichester, 1984.<br />

23. Gross, Benedict H.; Zagier, Don B. On singular moduli. J. Reine Angew. Math. 355 (1985), 191—220.<br />

24. Gross, Benedict H.; Zagier, Don B. Heegner points and derivatives <strong>of</strong> L-series. Invent. Math. 84 (1986),<br />

no. 2, 225-320.<br />

25. Gross, B.; Kohnen, W.; Zagier, D. Heegner points and derivatives <strong>of</strong> L-series. II. Math. Ann. 278 (1987),<br />

no. 1-4, 497-562.<br />

26. Hecke, Erich Mathematische Werke. (German) With introductory material by B. Schoeneberg, C. L.<br />

Siegel and J. Nielsen. Third edition. Vandenhoeck & Ruprecht, Gttingen, 1983. 960 pp.<br />

27. Hejhal, Dennis A. The Selberg trace formula for PSL(2, R). Vol. 2. Lecture Notes in <strong>Mathematics</strong>, 1001.<br />

Springer-Verlag, Berlin, 1983. viii+806 pp.<br />

28. Hirzebruch, F.; Zagier, D. Intersection numbers <strong>of</strong> curves on Hilbert <strong>modular</strong> surfaces and <strong>modular</strong> <strong>forms</strong><br />

<strong>of</strong> Nebentypus. Invent. Math. 36 (1976), 57-113.<br />

29. Kohnen, W.; Zagier, D. Values <strong>of</strong> L-series <strong>of</strong> <strong>modular</strong> <strong>forms</strong> at the center <strong>of</strong> the critical strip. Invent.<br />

Math. 64 (1981), no. 2, 175-198.<br />

30. Kudla, Stephen S.; Rapoport, Michael; Yang, Tonghai On the derivative <strong>of</strong> an Eisenstein series <strong>of</strong> <strong>weight</strong><br />

<strong>one</strong>. Internat. Math. Res. Notices 1999, no. 7, 347-385.<br />

31. Neunhöffer, H. Über die analytische Fortsetzung von Poincaréreihen. (German) S.-B. Heidelberger Akad.<br />

Wiss. Math.-Natur. Kl. 1973, 33-90.<br />

32. Niebur, Douglas A class <strong>of</strong> nonanalytic automorphic functions. Nagoya Math. J. 52 (1973), 133-145.<br />

33. Ogg, A. Survey on Modular functions <strong>of</strong> One Variable Notes by F. van Oystaeyen. Modular functions <strong>of</strong><br />

<strong>one</strong> variable, I (Proc. Internat. Summer School, Univ. Antwerp, Antwerp, 1972), pp. 1-35. Lecture Notes<br />

in Math. Vol. 320, Springer, Berlin, 1973.<br />

34. Ribet, K. Galois Representation Attached to Eigen<strong>forms</strong> with Nebentypus Modular functions <strong>of</strong> <strong>one</strong><br />

variable, V (Proc. Second Internat. Conf., Univ. Bonn, Bonn, 1976), pp. 17-51. Lecture Notes in Math.,<br />

Vol. 601, Springer, Berlin, 1977.<br />

35. Roelcke, Walter Das Eigenwertproblem der automorphen Formen in der hyperbolischen Ebene. I, II.<br />

(German) Math. Ann. 167 (1966), 292-337; ibid. 168 1966 261-324.<br />

36. Serre, J.-P. Modular <strong>forms</strong> <strong>of</strong> <strong>weight</strong> <strong>one</strong> and Galois representations. Algebraic number fields: L-functions<br />

and Galois properties (Proc. Sympos., Univ. Durham, Durham, 1975), pp. 193-268. Academic Press,<br />

London, 1977.<br />

37. Shimura, Goro Introduction to the arithmetic theory <strong>of</strong> automorphic functions. Reprint <strong>of</strong> the 1971 original.<br />

Publications <strong>of</strong> the Mathematical Society <strong>of</strong> Japan, 11. Kan Memorial Lectures, 1. Princeton University<br />

Press, Princeton, NJ, 1994. xiv+271 pp.<br />

38. Siegel, Carl Ludwig Lectures on advanced analytic number theory. Notes by S. Raghavan. Tata Institute<br />

<strong>of</strong> Fundamental Research Lectures on <strong>Mathematics</strong>, No. 23 Tata Institute <strong>of</strong> Fundamental Research,<br />

Bombay 1965 iii+331+iii pp.


56 W. DUKE AND Y. LI<br />

39. Stark, H. M. Class fields and <strong>modular</strong> <strong>forms</strong> <strong>of</strong> <strong>weight</strong> <strong>one</strong>. Modular functions <strong>of</strong> <strong>one</strong> variable, V (Proc.<br />

Second Internat. Conf., Univ. Bonn, Bonn, 1976), pp. 277-287. Lecture Notes in Math., Vol. 601, Springer,<br />

Berlin, 1977.<br />

40. Stark, H. M. L-functions at s = 1.I– IV. Advances in Math. 7 1971 301-343 (1971), 17 (1975), no. 1,<br />

60-92., 22 (1976), no. 1, 64-84, 35 (1980), no. 3, 197-235.<br />

41. Stein, William A. et al. Sage <strong>Mathematics</strong> S<strong>of</strong>tware (Version 5.0.1), The Sage Development Team, 2012,<br />

http://www.sagemath.org.<br />

42. Viazovska, Maryna. CM Values <strong>of</strong> Higher Green’s Functions, preprint 2011 (see arXiv:1110.4654)<br />

43. Zagier, Don. Letter to Gross<br />

44. Zagier, Don Traces <strong>of</strong> singular moduli. Motives, polylogarithms and Hodge theory, Part I (Irvine, CA,<br />

1998), 211-244, Int. Press Lect. Ser., 3, I, Int. Press, Somerville, MA, 2002.<br />

45. Zagier, Don. Nombres de classes et formes modulaires de poids 3/2. (French) C. R. Acad. Sci. Paris Sr.<br />

A-B 281 (1975), no. 21, Ai, A883-A886.<br />

46. Zagier, Don Ramanujan’s mock theta functions and their applications (after Zwegers and Ono-<br />

Bringmann). Seminaire Bourbaki. Vol. 2007/2008. Astrisque No. 326 (2009), Exp. No. 986, vii-viii,<br />

143-164 (2010).<br />

47. Zwegers , S.P., <strong>Mock</strong> Theta Functions, Utrecht PhD Thesis, (2002) ISBN 90-393-3155-3<br />

<strong>UCLA</strong> <strong>Mathematics</strong> <strong>Department</strong>, Box 951555, Los Angeles, CA 90095-1555<br />

E-mail address: wdduke@ucla.edu<br />

<strong>UCLA</strong> <strong>Mathematics</strong> <strong>Department</strong>, Box 951555, Los Angeles, CA 90095-1555<br />

E-mail address: yingkun@ucla.edu

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!