This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

On Kleinian mock modular forms

Claudia Alfes and Michael H. Mertens Universität Bielefeld, Fakultät für Mathematik, Postfach 100 131, 33501 Bielefeld, Germany, E-Mail: [email protected] Universität zu Köln, Department Mathematik/Informatik, Abteilung Mathematik, Weyertal 86–90, 50931 Köln, Germany, E-Mail: [email protected]
Abstract.

We give an explicit and computationally efficient construction of harmonic weak Maass forms which map to weight 22 newforms under the ξ\xi-operator. Our work uses a new non-analytic completion of the Kleinian ζ\zeta-function from the theory of Abelian functions.

The first author is partially supported by the Daimler and Benz Foundation, and funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) – SFB-TRR 358/1 2023 – 491392403

1. Introduction and Statement of Results

Harmonic weak Maass forms are real-analytic generalizations of classical modular forms which were introduced by Bruinier and Funke in [13]. By now harmonic weak Maass forms are ubiquitous in number theory and many other areas of mathematics and theoretical physics (see for instance [31, 8] and the references therein). A harmonic weak Maass form of weight kk\in\mathbb{Z} for a congruence subgroup Γ0(N)\Gamma_{0}(N) is a smooth function on the upper half-plane \mathfrak{H} which transforms like a usual (holomorphic) modular form of weight kk under Γ0(N)\Gamma_{0}(N). Rather than being holomorphic, it is annihilated by the weight kk hyperbolic Laplace operator

Δk=y2(2x2+2y2)+iky(x+iy),\Delta_{k}=-y^{2}\left(\frac{\partial^{2}}{\partial x^{2}}+\frac{\partial^{2}}{\partial y^{2}}\right)+iky\left(\frac{\partial}{\partial x}+i\frac{\partial}{\partial y}\right),

where we write τ=x+iy\tau=x+iy for zz\in\mathfrak{H}. In addition, they need to satisfy a certain growth condition at the cusps.

One of the central tools in the theory of harmonic weak Maass forms is the ξ\xi-operator defined by ξk:=2iykτ¯¯\xi_{k}:=-2iy^{k}\overline{\frac{\partial}{\partial\overline{\tau}}}. As Bruinier and Funke first showed in [13], this operator yields a surjective map from the space Hk(N)H_{k}(N) of harmonic weak Maass forms of weight kk for Γ0(N)\Gamma_{0}(N) to the space S2k(Γ0(N))S_{2-k}(\Gamma_{0}(N)) of (holomorphic) cusp forms of dual weight 2k2-k. The image of a harmonic weak Maass form under the ξ\xi-operator is called the shadow of its canonical holomorphic part, which is itself referred to as a mock modular form.

As there are infinitely many preimages of a given cusp form under the ξ\xi-operator, we would like to identify distinguished preimages. This can be achieved by employing Poincaré series [11, 9] or holomorphic projection [3, 19, 27]. In recent work Ehlen, Li, and Schwagenscheidt were able to construct so-called good preimages of CM forms using a certain theta lift. In particular, the holomorphic parts of such preimages have algebraic Fourier coefficients at \infty. Previous work of Bruinier, Ono, and Rhoades [12] guarantees the existence of such good preimages. Ehlen, Li, and Schwagenscheidt were able to determine the exact algebraic number field containing their Fourier coefficients.

For weight 22 newforms associated to rational elliptic curves, i.e. with rational Fourier coefficients, the first author together with Griffin, Ono, and Rolen [1] constructed distinguished preimages under the ξ\xi-operator extending earlier work of Guerzhoy [23]. This construction uses a lattice-invariant completion of the Weierstrass ζ\zeta-function from the classical theory of elliptic functions. When evaluated at the Eichler integral of the newform this gives essentially a harmonic weak Maass form of weight 0.

These results were recently extended to newforms with rational Fourier coefficients of positive weight by the authors in joint work with Funke and Rosu [2].

In this paper, we extend the construction from [1] to newforms with non-rational coefficients. This directly leads to the question for an analogue of the Weierstrass ζ\zeta-function in the context of Abelian functions, as we shall explain in the following paragraphs (see also Section 2 for further details).

Let K/K/\mathbb{Q} be a number field and let fS2(N)f\in S_{2}(N) be a newform of weight 22 for Γ0(N)\Gamma_{0}(N) with coefficients in KK with Galois conjugates f1,=f,f2,,frf_{1},=f,f_{2},\ldots,f_{r}. By VV we denote the associated component of the modular curve X0(N)X_{0}(N) and write ΛV=ωr+ωrr\Lambda_{V}=\omega\mathbb{Z}^{r}+\omega^{\prime}\mathbb{Z}^{r}\subset\mathbb{C}^{r} for the associated period lattice. As it turns out (see Section 4.1) we can choose ω,ω\omega,\omega^{\prime} such that Ω=ω1ωr\Omega=\omega^{-1}\omega^{\prime}\in\mathfrak{H}_{r} is in the Siegel upper half-space of genus rr (see (2.1)).

For uru\in\mathbb{C}^{r} and α,βr\alpha,\beta\in\mathbb{R}^{r} we define the Kleinian σ\sigma-function by

σ[αβ](u;ΛV)=exp(12utrω1ηu)θ[αβ](ω1u;ω1ω).\sigma\left[\begin{smallmatrix}\alpha\\ \beta\end{smallmatrix}\right](u;\Lambda_{V})=\exp\left(\frac{1}{2}u^{tr}\omega^{-1}\eta u\right)\theta\left[\begin{smallmatrix}\alpha\\ \beta\end{smallmatrix}\right]\left(\omega^{-1}u;\omega^{-1}\omega^{\prime}\right).

where θ[αβ](u;ω1ω)\theta\left[\begin{smallmatrix}\alpha\\ \beta\end{smallmatrix}\right]\left(u;\omega^{-1}\omega^{\prime}\right) is the Riemann theta function of characteristic [αβ]\left[\begin{smallmatrix}\alpha\\ \beta\end{smallmatrix}\right] (compare (2.2)) and η,η\eta,\eta^{\prime} denote the quasi-periods of VV. For brevity we sometimes write σ(u)=σ[αβ](u;ΛV)\sigma(u)=\sigma\left[\begin{smallmatrix}\alpha\\ \beta\end{smallmatrix}\right](u;\Lambda_{V}).

We then define ζ(u):=ulogσ(u)\zeta(u):=\nabla_{u}\log\sigma(u). This function is analytic but not Abelian, i.e. invariant under translations by lattice points ΛV\ell\in\Lambda_{V}. Following an idea of Rolen [32] we find a non-meromorphic completion ζ^\widehat{\zeta} of the Kleinian ζ\zeta-function in Proposition 2.5, which indeed satisfies

ζ^(u+)=ζ^(u) for all =λω+μωΛV\widehat{\zeta}(u+\ell)=\widehat{\zeta}(u)\text{ for all }\ell=\lambda\omega+\mu\omega^{\prime}\in\Lambda_{V}

wherever it is defined.

We define the vector of Eichler integrals associated to f=(f1,,fr)\vec{f}=(f_{1},\ldots,f_{r}) as

(τ)=(2πiτif1(z)𝑑z,,2πiτifr(z)𝑑z).\vec{\mathcal{E}}(\tau)=\left(-2\pi i\int_{\tau}^{i\infty}f_{1}(z)dz,\ldots,-2\pi i\int_{\tau}^{i\infty}f_{r}(z)dz\right).

The obstruction to modularity of (τ)\vec{\mathcal{E}}(\tau) is an element of ΛV\Lambda_{V}, that is (γ.τ)(τ)ΛV\vec{\mathcal{E}}(\gamma.\tau)-\vec{\mathcal{E}}(\tau)\in\Lambda_{V}, for γΓ0(N)\gamma\in\Gamma_{0}(N). Evaluating the Kleinian ζ\zeta-function at (τ)\vec{\mathcal{E}}(\tau) gives a distinguished preimage of f\vec{f}.

Theorem 1.1.

Let the notation be as above. The function

^V:1×r,τζ^((τ))\widehat{\mathfrak{Z}}_{V}:\mathfrak{H}\to\mathbb{C}^{1\times r},\quad\tau\mapsto\widehat{\zeta}(\vec{\mathcal{E}}(\tau))

is defined for all τ\tau\in\mathfrak{H} such that the Kleinian σ\sigma-function σ((τ);ΛV)\sigma(\vec{\mathcal{E}}(\tau);\Lambda_{V}) does not vanish. There it is Γ0(N)\Gamma_{0}(N)-invariant and is annihilated by the hyperbolic Laplacian Δ0\Delta_{0}. Moreover, we have

ξ0^V(τ)=4π2ftrP1\xi_{0}\widehat{\mathfrak{Z}}_{V}(\tau)=4\pi^{2}{\vec{f}}^{tr}P^{-1}

where

P=12i(ω¯ωtrω¯ωtr)P=\frac{1}{2i}\left(\overline{\omega}\omega^{\prime tr}-\overline{\omega^{\prime}}\omega^{tr}\right)

is positive definite.

For later reference we denote the meromorphic part of ^V\widehat{\mathfrak{Z}}_{V} by

(1.1) V(τ)=ζ((τ))12(τ)tr(ω1η+ηtrωtr)+π(τ)trP1\displaystyle\mathfrak{Z}_{V}(\tau)=\zeta(\vec{\mathcal{E}}(\tau))-\frac{1}{2}{\vec{\mathcal{E}}(\tau)}^{tr}\left(\omega^{-1}\eta+\eta^{tr}\omega^{-tr}\right)+\pi\vec{\mathcal{E}}(\tau)^{tr}P^{-1}

so that

^V(τ)=V(τ)π(τ)¯trP1.\widehat{\mathfrak{Z}}_{V}(\tau)=\mathfrak{Z}_{V}(\tau)-\pi\overline{\vec{\mathcal{E}}(\tau)}^{tr}P^{-1}.

We call V(τ)\mathfrak{Z}_{V}(\tau) the (polar) Kleinian mock modular form associated to VV.

Remark 1.2.

For r=1r=1, i.e. a newform with rational coefficients, Theorem 1.1 yields

ξ0^V(τ)=4π2Im(ω¯ω)f(τ)=4π2vol(ΛV)f(τ),\xi_{0}\widehat{\mathfrak{Z}}_{V}(\tau)=\frac{4\pi^{2}}{\operatorname{Im}(\overline{\omega}\omega^{\prime})}f(\tau)=\frac{4\pi^{2}}{\operatorname{vol}(\Lambda_{V})}f(\tau),

recovering the corresponding result in [1].

The Kleinian mock modular form V(τ)\mathfrak{Z}_{V}(\tau) may have poles. This phenomenon also occurs in the setup of [1], where the authors show that there is a modular function which cancels the poles. We obtain the following result in this direction.

Theorem 1.3.

Assume the notation as above and choose the characteristic of the Riemann theta function as the Riemann characteristic of the base point \infty. Assume in addition that VV is an (n,s)(n,s)-curve, i.e. there is a model of the form

(1.2) yn=xs+i,j0in+js<nscijxiyj,cij and s>n are coprime\displaystyle y^{n}=x^{s}+\sum_{\begin{subarray}{c}i,j\geq 0\\ in+js<ns\end{subarray}}c_{ij}x^{i}y^{j},\quad c_{ij}\in\mathbb{Q}\text{ and }s>n\text{ are coprime}

and that for τ\tau\in\mathfrak{H} the vector (τ)\vec{\mathcal{E}}(\tau) is not contained in the theta divisor except for points in ΛV\Lambda_{V} (see (2.7)).

Then the principal part at \infty of the scalar-valued function 𝔷V(τ):=V((τ))(1,,1)tr\mathfrak{z}_{V}(\tau):=\mathfrak{Z}_{V}(\mathcal{E}(\tau))\cdot(1,...,1)^{tr} has coefficients in KK and there exists a modular function FVF_{V} for Γ0(N)\Gamma_{0}(N) with algebraic Fourier coefficients, such that 𝔷V+FV\mathfrak{z}_{V}+F_{V} is a harmonic weak Maaß form.

Remark 1.4.
  1. (1)

    We note that every elliptic or hyperelliptic curve is in particular an (n,s)(n,s)-curve. Ogg [30] famously proved that the modular curve X0(N)X_{0}(N) is hyperelliptic if and only if

    N{22,23,26,28,29,30,31,33,35,37,39,41,46,47,50,59,71},N\in\{22,23,26,28,29,30,31,33,35,37,39,41,46,47,50,59,71\},

    so that Theorem 1.3 applies in those cases. Furthermore it is a well-known fact that every curve of genus 22 is hyperelliptic, wherefore our result also applies whenever the length rr of the Galois orbit of the considered newform is 2\leq 2.

  2. (2)

    Our proof of Theorem 1.3 relies on the algebraicity of the coefficients of the expansion of the Kleinian σ\sigma-function around 0 ([29, Theorem 3]). To the authors’ knowledge this is only known for (n,s)(n,s)-curves explaining this technical assumption on VV.

  3. (3)

    The additional assumption that (τ)\vec{\mathcal{E}}(\tau) lies in the theta divisor only if (τ)ΛV\vec{\mathcal{E}}(\tau)\in\Lambda_{V} is also a technical one. It allows us to only consider the expansion of σ\sigma around 0 rather than other points. The algebraic properties of expansions of σ\sigma around other points do not seem to have been investigated in the literature.

  4. (4)

    So-called Millson theta lifts of the functions 𝔷V+FV\mathfrak{z}_{V}+F_{V} fall into the framework of results of Bruinier and Ono [10], i.e. it is possible to relate the algebraicity of the Fourier coefficients of these lifts to the vanishing of the twisted central LL-derivatives of the associated newform of weight 22 (compare [1] for the case of newforms with rational Fourier coefficients).

The following theorem gives the expansion of the Kleinian mock modular form at other cusps than \infty.

Theorem 1.5.

In the situation of Theorem 1.1 let QQ be an exact divisor of NN (i.e. gcd(Q,N/Q))=1\gcd(Q,N/Q))=1) and WQW_{Q} the corresponding Atkin-Lehner involution. Denote by λQ{±1}\lambda_{Q}\in\{\pm 1\} the eigenvalue of the newform ff under WQW_{Q}, i.e. f|WQ=λQff|W_{Q}=\lambda_{Q}f. Further let

LQ(f):=2πiWQ1.f(t)dt.L_{Q}(f):=-2\pi i\int_{W_{Q}^{-1}.\infty}^{\infty}f(t)\mathrm{d}t.

Then we have

^V|WQ(τ)=ζ^(λQ((τ)LQ(f))).\widehat{\mathfrak{Z}}_{V}|W_{Q}(\tau)=\widehat{\zeta}(\lambda_{Q}(\vec{\mathcal{E}}(\tau)-L_{Q}(\vec{f}))).

The paper is organized as follows. In Section 2 we introduce the Riemann theta function, define the Kleinian σ\sigma and ζ\zeta-function, and construct the completed Kleinian ζ\zeta-function. We prove Theorem 1.1, 1.3, and 1.5 in Section 3. In Section 4 we present some computational examples.

Acknowledgments

We thank Eberhard Freitag, Jens Funke, Eugenia Rosu, Fredrik Strömberg, Bernd Sturmfels, Don Zagier, and David Zureick-Brown for interesting discussions, and Annika Burmester and Paul Kiefer for comments on an earlier version of this manuscript. The first author thanks the Hausdorff Institute where this work was initiated.

2. Construction of Kleinian Abelian functions

In this section we construct the analog of the lattice-invariant Weierstrass ζ\zeta-function in the Abelian case. We employ the approach using Riemann theta functions to construct the σ\sigma-function.

2.1. The Riemann θ\theta-function

We review the definition and some basic properties of the Riemann θ\theta-function.

We denote by

(2.1) g:={Ωg×g:Ωtr=Ω,Im(Ω)>0}\displaystyle\mathfrak{H}_{g}:=\{\Omega\in\mathbb{C}^{g\times g}\>:\>\Omega^{tr}=\Omega,\ \operatorname{Im}(\Omega)>0\}

the Siegel upper half-space of genus gg. Here, we write A>0A>0 to indicate that a real symmetric matrix AA is positive definite.

Let α,βg\alpha,\beta\in\mathbb{R}^{g}. For ugu\in\mathbb{C}^{g} and Ωg\Omega\in\mathfrak{H}_{g} we define the Riemann theta function of characteristic [αβ]\left[\begin{smallmatrix}\alpha\\ \beta\end{smallmatrix}\right] by

(2.2) θ[αβ](u;Ω):=mge((m+α)tr(u+β)+12(m+α)trΩ(m+α)),\theta\left[\begin{smallmatrix}\alpha\\ \beta\end{smallmatrix}\right](u;\Omega):=\sum_{m\in\mathbb{Z}^{g}}e\left((m+\alpha)^{tr}(u+\beta)+\frac{1}{2}(m+\alpha)^{tr}\Omega(m+\alpha)\right),

where e(x)=exp(2πix)e(x)=\exp(2\pi ix). We write θ(u;Ω)=θ[0](u;Ω)\theta(u;\Omega)=\theta[0](u;\Omega).

This function is well-known to satisfy the following transformation properties (see e.g. [28, pp. 123, 194, 195] or [6, Theta Transformation Formula 8.6.1])

(2.3) θ[αβ](u+λΩ+μ;Ω)\displaystyle\theta\left[\begin{smallmatrix}\alpha\\ \beta\end{smallmatrix}\right](u+\lambda\Omega+\mu;\Omega) =e(12λtrΩλutrλλtrβ+μtrα)θ[αβ](u;Ω),λ,μg,\displaystyle=e\left(-\frac{1}{2}\lambda^{tr}\Omega\lambda-u^{tr}\lambda-\lambda^{tr}\beta+\mu^{tr}\alpha\right)\theta\left[\begin{smallmatrix}\alpha\\ \beta\end{smallmatrix}\right](u;\Omega),\quad\lambda,\mu\in\mathbb{Z}^{g},
(2.4) θ(u;Ω+S)\displaystyle\theta(u;\Omega+S) =θ(u;Ω),if Ssymg×g even,\displaystyle=\theta(u;\Omega),\quad\text{if }S\in\mathbb{Z}^{g\times g}_{sym}\text{ even},
(2.5) θ(Ω1u;Ω1)\displaystyle\theta(\Omega^{-1}u;-\Omega^{-1}) =(detiΩ)1/2e(12utrΩ1u)θ(u;Ω).\displaystyle=(\det-i\Omega)^{1/2}e\left(\frac{1}{2}u^{tr}\Omega^{-1}u\right)\theta(u;\Omega).

These transformations imply that θ(u;Ω)\theta(u;\Omega) is a Siegel-Jacobi form of weight and index 1/21/2 (see [38]).

2.2. Kleinian ζ\zeta-functions and their completions

Following the classical treatments by Klein and Baker [25, 4] and the more modern discussions, e.g. in [14, 21], we construct Abelian functions via the Riemann theta function.

Let VV be a non-singular algebraic curve of genus gg with period matrices ω,ωg×g\omega,\omega^{\prime}\in\mathbb{C}^{g\times g}, so that the Jacobian of VV is isomorphic to the quotient g/ΛV\mathbb{C}^{g}/\Lambda_{V} where

ΛV={ωm+ωn:m,ng}.\Lambda_{V}=\{\omega m+\omega^{\prime}n\>:\>m,n\in\mathbb{Z}^{g}\}.

Here and throughout, we choose the period matrices ω,ω\omega,\omega^{\prime} such that Ω=ω1ωg\Omega=\omega^{-1}\omega^{\prime}\in\mathfrak{H}_{g} lies in the Siegel upper half-space. Further let η,ηg×g\eta,\eta^{\prime}\in\mathbb{C}^{g\times g} denote the quasi-period matrices of VV. For an arbitrary characteristic [αβ]2g\left[\begin{smallmatrix}\alpha\\ \beta\end{smallmatrix}\right]\in\mathbb{R}^{2g} and u=(u1,,ug)trgu=(u_{1},...,u_{g})^{tr}\in\mathbb{C}^{g} we define the Kleinian σ\sigma-function of characteristic [αβ]\left[\begin{smallmatrix}\alpha\\ \beta\end{smallmatrix}\right] by

(2.6) σ[αβ](u)=σ[αβ](u;ΛV)=exp(12utrω1ηu)θ[αβ](ω1u;ω1ω).\sigma\left[\begin{smallmatrix}\alpha\\ \beta\end{smallmatrix}\right](u)=\sigma\left[\begin{smallmatrix}\alpha\\ \beta\end{smallmatrix}\right](u;\Lambda_{V})=\exp\left(\frac{1}{2}u^{tr}\omega^{-1}\eta u\right)\theta\left[\begin{smallmatrix}\alpha\\ \beta\end{smallmatrix}\right]\left(\omega^{-1}u;\omega^{-1}\omega^{\prime}\right).
Remark 2.1.
  1. (1)

    The definition of the Kleinian σ\sigma-function above differs from that in [14, 21] by a constant depending on the curve VV.

  2. (2)

    In [14, 21], the characteristic is fixed as the Riemann characteristic of the base point \infty. In particular, this implies that α,β\alpha,\beta are half-integral. We usually choose the characteristic with half-integer entries, but not necessarily corresponding to the base point \infty.

  3. (3)

    If V=EV=E is an elliptic curve, then the Kleinian σ\sigma-function of characteristic [1/21/2]\left[\begin{smallmatrix}1/2\\ 1/2\end{smallmatrix}\right] coincides, up to a constant factor, with the classical Weierstrass σ\sigma-function.

We require the following result on the zeros of the Kleinian σ\sigma-function (see e.g.  [28, Corollary II.3.6] and [21, p.1661]).

Lemma 2.2.

Choosing [αβ]122g\left[\begin{smallmatrix}\alpha\\ \beta\end{smallmatrix}\right]\in\frac{1}{2}\mathbb{Z}^{2g} as the Riemann characteristic of the base point \infty, we have that σ(u)=0\sigma(u)=0 if and only if

(2.7) u=±(P1𝑑u++Pg1𝑑u)+,\displaystyle u=\pm\left(\int_{\infty}^{P_{1}}du+...+\int_{\infty}^{P_{g-1}}du\right)+\ell,

where (P1,,Pg1)Symg1(V)(P_{1},...,P_{g-1})\in\operatorname{Sym}^{g-1}(V), where Symk(V)\operatorname{Sym}^{k}(V) denotes the kkth symmetric power of VV, and Λ\ell\in\Lambda.

From now on we suppress the characteristic from the notation when it can be chosen arbitrarily.

We define the iith Kleinian ζ\zeta-function by

ζi(u)=uilogσ(u)\zeta_{i}(u)=\partial_{u_{i}}\log\sigma(u)

and the Kleinian ζ\zeta-function by

ζ(u):=(ζ1(u),,ζg(u))=ulogσ(u).\zeta(u):=(\zeta_{1}(u),...,\zeta_{g}(u))=\nabla_{u}\log\sigma(u).

The Kleinian \wp-functions ij(u):=ujζi(u)\wp_{ij}(u):=-\partial_{u_{j}}\zeta_{i}(u) are then Abelian functions with respect to the lattice ΛV\Lambda_{V}, i.e. they satisfy the transformation law

ij(u+)=(u),ΛV,\wp_{ij}(u+\ell)=\wp(u),\quad\ell\in\Lambda_{V},

wherever they are defined. The ζ\zeta-functions however are not Abelian but rather satisfy

(2.8) ζ(u+ωm+ωn)=ζ(u)+(mtrη+ntrη).\displaystyle\zeta(u+\omega m+\omega^{\prime}n)=\zeta(u)+(m^{tr}\eta+n^{tr}\eta^{\prime}).

Again, for V=EV=E an elliptic curve this reduces to the well-known transformation law of the Weierstrass ζ\zeta-function.

It is a classical fact going back to Eisenstein that the Weierstrass ζ\zeta-function admits a non-analytic completion which is invariant under translation by lattice points [37]. We adapt Rolen’s proof of this property [32] to the setting of Kleinian ζ\zeta-functions.

The following Lemma follows by a straightforward computation.

Lemma 2.3.

Let F:g×gF:\mathbb{C}^{g}\times\mathfrak{H}_{g}\to\mathbb{C} be a smooth function which satisfies the following functional equation

F(u+Ωλ+μ;Ω)=e(2mλtru+f(Ω,λ,μ))F(u;Ω),λ,μg,F(u+\Omega\lambda+\mu;\Omega)=e\left(-2m\lambda^{tr}u+f(\Omega,\lambda,\mu)\right)F(u;\Omega),\qquad\lambda,\mu\in\mathbb{Z}^{g},

for some real number mm and some fixed function f:g×g×gf:\mathfrak{H}_{g}\times\mathbb{R}^{g}\times\mathbb{R}^{g}\to\mathbb{C}.

Then the function defined by

F~(u;Ω):=1F(u;Ω)(iu4πmIm(u)trIm(Ω)1)F(u;Ω)\widetilde{F}(u;\Omega):=\frac{1}{F(u;\Omega)}\left(i\nabla_{u}-4\pi m\operatorname{Im}(u)^{tr}\operatorname{Im}(\Omega)^{-1}\right)F(u;\Omega)

satisfies

F~(u+Ωλ+μ;Ω)=F~(u;Ω),for all λ,μg,\widetilde{F}(u+\Omega\lambda+\mu;\Omega)=\widetilde{F}(u;\Omega),\quad\text{for all }\lambda,\mu\in\mathbb{Z}^{g},

wherever F(u,Ω)0F(u,\Omega)\neq 0.

Remark 2.4.

The differential operator

(2.9) 𝒴+:=iu4πmIm(u)trIm(Ω)1\displaystyle\mathcal{Y}_{+}:=i\nabla_{u}-4\pi m\operatorname{Im}(u)^{tr}\operatorname{Im}(\Omega)^{-1}

in Lemma 2.3 looks similar to the raising operator

Y+=iz4πmImzImτY_{+}=i\partial_{z}-4\pi m\frac{\operatorname{Im}z}{\operatorname{Im}\tau}

in the theory of Jacobi forms (see e.g. [5]). This operator maps Jacobi forms of weight kk and index mm to non-holomorphic Jacobi forms of weight k+1k+1 and index mm. An extension of this operator in the context of Siegel-Jacobi forms is given in [38]. However, the operator 𝒴+\mathcal{Y}_{+} in (2.9) does not respect the action of the symplectic group in the same way as (2.5). For the purpose of this paper, this is not required and the operators in [38] are not suitable for our setup.

We obtain the following result which is analogous to that in [32].

Proposition 2.5.

The function

ζ^(u)=ζ(u)12utr(ω1η+ηtrωtr)+2πiIm(ω1u)trIm(ω1ω)1ω1\widehat{\zeta}(u)=\zeta(u)-\frac{1}{2}u^{tr}\left(\omega^{-1}\eta+\eta^{tr}\omega^{-tr}\right)+2\pi i\operatorname{Im}(\omega^{-1}u)^{tr}\operatorname{Im}(\omega^{-1}\omega^{\prime})^{-1}\omega^{-1}

is a non-meromorphic Abelian function for the lattice ΛV\Lambda_{V}, i.e. for any =λω+μωΛV\ell=\lambda\omega+\mu\omega^{\prime}\in\Lambda_{V} we have

ζ^(u+)=ζ^(u)\widehat{\zeta}(u+\ell)=\widehat{\zeta}(u)

wherever both sides are defined.

Proof.

For simplicity we let

ϑ(u):=θ[αβ](ω1u;ω1ω).\vartheta(u):=\theta\left[\begin{smallmatrix}\alpha\\ \beta\end{smallmatrix}\right]\left(\omega^{-1}u;\omega^{-1}\omega^{\prime}\right).

It follows immediately from Lemma 2.3 that the function

(2.10) 1ϑ(u)uϑ(u)+2πiIm(ω1u)trIm(ω1ω)1ω1\displaystyle\frac{1}{\vartheta(u)}\nabla_{u}\vartheta(u)+2\pi i\operatorname{Im}(\omega^{-1}u)^{tr}\operatorname{Im}(\omega^{-1}\omega^{\prime})^{-1}\omega^{-1}

is a non-meromorphic Abelian function where it is defined. From (2.6) we obtain that

ζ(u)=u(12utrω1ηu+logϑ(u))=12utr(ω1η+ηtrωtr)+1ϑ(u)u(ϑ(u)),\zeta(u)=\nabla_{u}\left(\frac{1}{2}u^{tr}\omega^{-1}\eta u+\log\vartheta(u)\right)=\frac{1}{2}u^{tr}(\omega^{-1}\eta+\eta^{tr}\omega^{-tr})+\frac{1}{\vartheta(u)}\nabla_{u}(\vartheta(u)),

so the claim follows. ∎

3. Proofs of the main results

In this section we prove the main results of the paper.

Proof of Theorem 1.1.

We first note that ζ(u)\zeta(u) is defined whenever uu does not lie in the divisor of the Kleinian σ\sigma-function.

We now prove the Γ0(N)\Gamma_{0}(N)-invariance. Let γΓ0(N)\gamma\in\Gamma_{0}(N) and τ\tau\in\mathfrak{H}. We then have

(γ.τ)=2πiγ.τf(t)dt=2πi(ττγ.τ)f(t)dt.\vec{\mathcal{E}}(\gamma.\tau)=-2\pi i\int_{\gamma.\tau}^{\infty}\vec{f}(t)\mathrm{d}t=-2\pi i\left(\int_{\tau}^{\infty}-\int_{\tau}^{\gamma.\tau}\right)\vec{f}(t)\mathrm{d}t.

Since all components of f\vec{f} are holomorphic cusp forms, we have

τγ.τf(t)dt=𝔞γ.𝔞f(t)dt\int_{\tau}^{\gamma.\tau}\vec{f}(t)\mathrm{d}t=\int_{\mathfrak{a}}^{\gamma.\mathfrak{a}}\vec{f}(t)\mathrm{d}t

for an arbitrary cusp 𝔞\mathfrak{a} of Γ0(N)\Gamma_{0}(N) (see e.g. [33, Proposition 10.5]). Since a path from 𝔞\mathfrak{a} to γ.𝔞\gamma.\mathfrak{a} lies in the cuspidal homology of X0(N)X_{0}(N), we find by definition of the period lattice ΛV\Lambda_{V} that =2πiτγ.τf(t)𝑑tΛV\ell=-2\pi i\int_{\tau}^{\gamma.\tau}\vec{f}(t)dt\in\Lambda_{V}. The completed Kleinian ζ\zeta-function ζ^(u;ΛV)\widehat{\zeta}(u;\Lambda_{V}) is invariant under translations by points in ΛV\Lambda_{V} by Proposition 2.5. Therefore, it follows that the function

^V(γ.τ)=ζ^((γ.τ))=ζ^((τ))=ζ^((τ))=^V(τ)\widehat{\mathfrak{Z}}_{V}(\gamma.\tau)=\widehat{\zeta}(\vec{\mathcal{E}}(\gamma.\tau))=\widehat{\zeta}(\vec{\mathcal{E}}(\tau)-\ell)=\widehat{\zeta}(\vec{\mathcal{E}}(\tau))=\widehat{\mathfrak{Z}}_{V}(\tau)

is indeed Γ0(N)\Gamma_{0}(N)-invariant wherever it is defined.

We now proceed to show the properties under the action of the differential operators Δ0\Delta_{0} and ξ0\xi_{0}. Using Proposition 2.5 and (2.10) we can write

ζ^(u)\displaystyle\widehat{\zeta}(u) =1θ(ω1;Ω)(uθ)(ω1u;Ω)ω1+2πiIm(ω1u)trIm(Ω)1ω1\displaystyle=\frac{1}{\theta(\omega^{-1};\Omega)}\left(\nabla_{u}\theta\right)(\omega^{-1}u;\Omega)\omega^{-1}+2\pi i\operatorname{Im}(\omega^{-1}u)^{tr}\operatorname{Im}(\Omega)^{-1}\omega^{-1}
=1θ(ω1u;Ω)(uθ)(ω1;Ω)ω1+πutrωtrIm(Ω)1ω1\displaystyle=\frac{1}{\theta(\omega^{-1}u;\Omega)}\left(\nabla_{u}\theta\right)(\omega^{-1};\Omega)\omega^{-1}+\pi u^{tr}\omega^{-tr}\operatorname{Im}(\Omega)^{-1}\omega^{-1}
(3.1) πu¯trω¯trIm(Ω)1ω1.\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\quad-\pi\overline{u}^{tr}\overline{\omega}^{-tr}\operatorname{Im}(\Omega)^{-1}\omega^{-1}.

This immediately implies that

ξ0(^V(τ))=4π2ftrωtrIm(Ω)1ω1.\xi_{0}\left(\widehat{\mathfrak{Z}}_{V}(\tau)\right)=4\pi^{2}\vec{f}^{tr}\omega^{-tr}\operatorname{Im}(\Omega)^{-1}\omega^{-1}.

A straightforward computation gives

ωIm(Ω)ωtr=12i(ω¯ωtrω¯ωtr)=P.\omega\operatorname{Im}(\Omega)\omega^{tr}=\frac{1}{2i}(\overline{\omega}\omega^{\prime tr}-\overline{\omega}^{\prime}\omega^{tr})=P.

Since Ω=ω1ωr\Omega=\omega^{-1}\omega^{\prime}\in\mathfrak{H}_{r} we have that

2Im(Ω)=i(Ω¯trΩ)=i(ω¯trω¯trω1ω)2\operatorname{Im}(\Omega)=i(\overline{\Omega}^{tr}-\Omega)=i(\overline{\omega^{\prime}}^{tr}\overline{\omega}^{-tr}-\omega^{-1}\omega^{\prime})

is positive definite. Therefore the same is true for

iω(ω¯trω¯trω1ω)ω¯tr=2P¯i\omega(\overline{\omega^{\prime}}^{tr}\overline{\omega}^{-tr}-\omega^{-1}\omega^{\prime})\overline{\omega}^{tr}=2\overline{P}

and hence PP is positive definite.

The proof of Theorem 1.3 is in part analogous to that of the corresponding result in [1]. Since the proof given there is rather short, we give a more detailed version here.

Proof of Theorem 1.3.

By our assumption on \vec{\mathcal{E}} and Lemma Lemma 2.2 we see that by construction V\mathfrak{Z}_{V} has a pole in τ\tau if and only if (τ)ΛV\vec{\mathcal{E}}(\tau)\in\Lambda_{V}. Since ζ^(u)\widehat{\zeta}(u) is lattice invariant, it is therefore enough to consider the expansion of ζ(u)\zeta(u) around u=0u=0. By [29, Theorem 3] we have

σ(u)=Sλ(n,s)(u)+higher order terms,\sigma(u)=S_{\lambda(n,s)}(u)+\text{higher order terms},

where Sλ(n,s)S_{\lambda(n,s)} is the so-called Schur function associated to the curve VV (for a precise definition see p. 192 of loc.cit.). This polynomial has rational coefficients. Since each newform f1,,frf_{1},...,f_{r} has coefficients in KK, so do the functions 1,,r\mathcal{E}_{1},...,\mathcal{E}_{r}. Therefore, we see that upon plugging (τ)\vec{\mathcal{E}}(\tau) into ujσ(u)/σ(u)\partial_{u_{j}}\sigma(u)/\sigma(u), the principal part of ζ((τ))\zeta(\vec{\mathcal{E}}(\tau)) at \infty has coefficients in KK as well.

Next we show the existence of the meromorphic modular function FVF_{V} which cancels all the poles of 𝔷V\mathfrak{z}_{V} within the upper half-plane: It is well-known (see e.g. Section 2.2) that any partial derivative of the Kleinian ζ\zeta-function yields (up to sign) a Kleinian \wp-function, thus a meromorphic Abelian function. Therefore the function

𝔭V(τ)=j=1ruj2logσ((τ))\mathfrak{p}_{V}(\tau)=\sum_{j=1}^{r}\partial_{u_{j}}^{2}\log\sigma(\mathcal{E}(\tau))

is a meromorphic modular function with respect to Γ0(N)\Gamma_{0}(N) by the same argument employed in the proof of Theorem 1.1. The poles of this function within the upper half-plane are clearly at the same points as those of 𝔷V\mathfrak{z}_{V}, but strictly with higher order. As indicated in [1] we follow the proof of [18, Theorem 11.9], which states that every modular function for Γ0(N)\Gamma_{0}(N) is a rational function in j(τ)j(\tau) and j(Nτ)j(N\tau). Let γ1,,γι(N)\gamma_{1},...,\gamma_{\iota(N)}, ι(N)=[Γ0(1):Γ0(N)]\iota(N)=[\Gamma_{0}(1):\Gamma_{0}(N)], be a fixed set of representatives of Γ0(N)\SL2()\Gamma_{0}(N)\backslash\operatorname{SL}_{2}(\mathbb{Z}) and assume γ1=(1001)\gamma_{1}=\left(\begin{smallmatrix}1&0\\ 0&1\end{smallmatrix}\right). We consider the function

G(X,τ)=i=1ι(N)𝔭V(γiτ)ji(Xj(γjτ)).G(X,\tau)=\sum_{i=1}^{\iota(N)}\mathfrak{p}_{V}(\gamma_{i}\tau)\prod_{j\neq i}(X-j(\gamma_{j}\tau)).

This is clearly a polynomial in XX whose coefficients are meromorphic functions in τ\tau. In fact it is not hard to show that these coefficients are modular functions for SL2()\operatorname{SL}_{2}(\mathbb{Z}), whence they are all rational functions in j(τ)j(\tau). We may therefore write

G(X,j(τ))=k=0ι(N)1pk(j(τ))qk(j(τ))XkG(X,j(\tau))=\sum_{k=0}^{\iota(N)-1}\frac{p_{k}\left(j(\tau)\right)}{q_{k}\left(j(\tau)\right)}X^{k}

for certain polynomials pk,qk[Y]p_{k},q_{k}\in\mathbb{C}[Y].

In fact we can choose pk,qkp_{k},q_{k} with algebraic coefficients. By assumption VV is an (n,s)(n,s)-curve, so it follows from [29, Theorem 3] that the coefficients of the Taylor expansion of σ\sigma, and therefore of the Laurent expansions of both 𝔷V\mathfrak{z}_{V} and 𝔭V\mathfrak{p}_{V} are rational polynomials in the curve coefficients cijc_{ij} in (1.2), and hence algebraic. Since the newform ff has algebraic Fourier coefficients at all cusps, it also follows that the modular function 𝔭V\mathfrak{p}_{V} has algebraic Fourier coefficients at all cusps.

Let

Q=lcm(q1,,qι(N))==1M(Yα)Q=\operatorname{lcm}(q_{1},...,q_{\iota(N)})=\prod_{\ell=1}^{M}(Y-\alpha_{\ell})

for some α¯\alpha_{\ell}\in\overline{\mathbb{Q}} and M0M\in\mathbb{N}_{0}. Now arguing exactly as in the aforementioned proof of [18, Theorem 11.9], we find that we can write

(3.2) 𝔭V(τ)=k=0ι(M)1p~k(j(τ))j(Nτ)=1M(j(τ)α)m1(j(Nτ)j(Nγmτ)).\displaystyle\mathfrak{p}_{V}(\tau)=\frac{\sum_{k=0}^{\iota(M)-1}\tilde{p}_{k}(j(\tau))j(N\tau)}{\prod_{\ell=1}^{M}(j(\tau)-\alpha_{\ell})\cdot\prod_{m\neq 1}(j(N\tau)-j(N\gamma_{m}\tau))}.

Note that the numerator in (3.2) is holomorphic in \mathfrak{H} and each factor in the denominator yields a simple pole of 𝔭V\mathfrak{p}_{V} in \mathfrak{H} (we ignore the slight technical complication of elliptic fixed points for the sake of simplicity). By multiplying through by all but one of the factors in the denominator (after canceling against potential zeros in the numerator), we obtain a modular function with algebraic Fourier coefficients with a simple pole precisely where 𝔷V\mathfrak{z}_{V} has a pole. Thus, we can cancel all the poles using only modular functions with algebraic coefficients. ∎

As the proof of Theorem 1.5 is almost literally the same as that of the analogous result in [1] (Theorem 1.2) we omit it here.

4. Examples

4.1. Computational aspects

We briefly outline how to compute the quantities required for the construction of the Kleinian mock modular forms.

Most of the facts in this section are by now fairly standard and more or less implemented in computer algebra systems like Sage [34], Magma [7], or Pari/Gp [35]. We loosely follow the accounts in [15, 36] and Kapitel VI of [22].

Let fS2(N)f\in S_{2}(N) be a newform whose coefficients lie in a number field K/K/\mathbb{Q} and let f1=f,f2,,frf_{1}=f,f_{2},...,f_{r} denote its Galois conjugates. The vector of all these conjugates is denoted by f=(f1,fr)tr\vec{f}=(f_{1},...f_{r})^{tr}. Suppose we have Fourier expansions

fj(τ)=n=1aj(n)qn,q=e2πiτ,aj(n)K.f_{j}(\tau)=\sum_{n=1}^{\infty}a_{j}(n)q^{n},\quad q=e^{2\pi i\tau},\quad a_{j}(n)\in K.

Then there is a component over \mathbb{Q} of the modular curve X0(N)X_{0}(N) associated to the Galois orbit of ff. Its Jacobian is given by r/Λf\mathbb{C}^{r}/\Lambda_{f} for the period lattice Λf.\Lambda_{f}. We can find a basis for this lattice by computing the integrals

2πiγf(z)dz,-2\pi i\int_{\gamma}\vec{f}(z)\mathrm{d}z,

where γ\gamma runs through a basis of the integral homology H1(X0(N),)H^{1}(X_{0}(N),\mathbb{Z}), which can in turn be determined using the available functions in Sage or Magma.

This may be achieved very efficiently by evaluating holomorphic Eichler integrals

j(τ):=2πiτfj(z)dz=n=1aj(n)nqn\mathcal{E}_{j}(\tau):=-2\pi i\int_{\tau}^{\infty}f_{j}(z)\mathrm{d}z=\sum_{n=1}^{\infty}\frac{a_{j}(n)}{n}q^{n}

at suitable points τ\tau in the upper half-plane.

It follows from work of Hida [24] together with standard linear algebra that we can choose a basis (a1,,ar,b1,,br)(a_{1},...,a_{r},b_{1},...,b_{r}) of H1(X0(N),)H^{1}(X_{0}(N),\mathbb{Z}) with the property that the cycles follow the intersection pattern

aiaj=0,bibj=0,aibj=eiδij,a_{i}\circ a_{j}=0,\quad b_{i}\circ b_{j}=0,\quad a_{i}\circ b_{j}=e_{i}\delta_{ij},

where δij\delta_{ij} denotes the usual Kronecker delta and e1e2ere_{1}\mid e_{2}\mid...\mid e_{r} are positive integers. With respect to this basis we obtain matrices

ω=2πi(aifj(z)dz)i,j=1,r,ω=2πi(bifj(z)dz)i,j=1,,r\omega=-2\pi i\left(\int_{a_{i}}f_{j}(z)\mathrm{d}z\right)_{i,j=1,...r},\quad\omega^{\prime}=-2\pi i\left(\int_{b_{i}}f_{j}(z)\mathrm{d}z\right)_{i,j=1,...,r}

with the property that Ω:=ω1ωr\Omega:=\omega^{-1}\omega^{\prime}\in\mathfrak{H}_{r} lies in the Siegel upper half-space. An algorithm to compute this basis was found by Merel [26] and is implemented e.g. in Magma.

Note that since we have

ζ^(u)=1ϑ(u)uϑ(u)+πutrP1πu¯trP1\widehat{\zeta}(u)=\frac{1}{\vartheta(u)}\nabla_{u}\vartheta(u)+\pi u^{tr}P^{-1}-\pi\overline{u}^{tr}P^{-1}

with ϑ(u)=θ[αβ](ω1u,ω1ω)\vartheta(u)=\theta\left[\begin{smallmatrix}\alpha\\ \beta\end{smallmatrix}\right](\omega^{-1}u,\omega^{-1}\omega^{\prime}) and PP as in Theorem 1.1 by (3.1), we do not require the quasi-periods η,η\eta,\eta^{\prime} to compute the Kleinian mock modular form.

4.2. Level 27

We consider the unique newform f=η(3τ)2η(9τ)2S2(27)f=\eta(3\tau)^{2}\eta(9\tau)^{2}\in S_{2}(27) associated to the elliptic curve

y2+y=x37(LMFDB label 27.a3).y^{2}+y=x^{3}-7\quad\text{(LMFDB label 27.a3)}.

It has rational coefficients and complex multiplication by (3)\mathbb{Q}(\sqrt{-3}).

Since ff has rational coefficients, the results in [1] apply and we can compute a mock modular form whose shadow is ff (up to a constant multiple). Alternatively, we can apply the strategy of this paper and find that the period lattice of ff is generated by

ω=0.2944390.509984iandω=1.01996i.\omega=-0.294439-0.509984i\quad\text{and}\quad\omega^{\prime}=-1.01996i.

Using the Kleinian zeta function with characteristic α=β=1/2\alpha=\beta=1/2 we we employ Theorem 1.1 to construct the function

V(τ)=q1+12q27015q5+14074q84077611q11+379612q14212509817q17+O(q20),\mathfrak{Z}_{V}(\tau)=q^{-1}+\frac{1}{2}q^{2}-\frac{701}{5}q^{5}+\frac{1407}{4}q^{8}-\frac{40776}{11}q^{11}+\frac{37961}{2}q^{14}-\frac{2125098}{17}q^{17}+O(q^{20}),

whose shadow is 4π2f4\pi^{2}f. Note that the Fourier coefficients above are indeed rational numbers, which can be shown using work of Bruinier-Ono-Rhoades [12, Theorem 1.3] or Ehlen-Li-Schwagenscheidt [20, Corollary 1.2].

4.3. Level 23

The modular curve X0(23)X_{0}(23) has genus 22 and there is one Galois orbit of newforms, generated by the form with Fourier expansion

f(τ)=qϕq2+(2ϕ1)q3+(ϕ1)q42ϕq5+O(q6),ϕ=152.f(\tau)=q-\phi q^{2}+(2\phi-1)q^{3}+(\phi-1)q^{4}-2\phi q^{5}+O(q^{6}),\quad\phi=\frac{1-\sqrt{5}}{2}.

We denote the Galois conjugate of ff by fσf^{\sigma}.

The four elements {1/19,0},{1/17,0},{1/15,0},{1/11,0}\{-1/19,0\},\{-1/17,0\},\{-1/15,0\},\{-1/11,0\} form a basis of H1(X0(23),)H_{1}(X_{0}(23),\mathbb{Z}). Consequently we find the following basis for the period lattice

c1\displaystyle c_{1} =(1.062972+2.060558i0.642714+0.710672i),c2=(1.062972+2.060558i0.642714+0.710672i),\displaystyle=\begin{pmatrix}-1.062972+2.060558i\\ 0.642714+0.710672i\end{pmatrix},\ c_{2}=\begin{pmatrix}1.062972+2.060558i\\ -0.642714+0.710672i\end{pmatrix},
c3\displaystyle c_{3} =(1.719925+0.787063i0.397219+1.860563i),c4=(1.3139062.079867).\displaystyle=\begin{pmatrix}1.719925+0.787063i\\ 0.397219+1.860563i\end{pmatrix},\ c_{4}=\begin{pmatrix}1.313906\\ 2.079867\end{pmatrix}.

Computing the intersection pairing with the help of Magma we compute the period matrices

ω\displaystyle\omega =(c1c2+c3,c2)=(0.406019+0.787063i1.062972+2.060558i1.682647+1.860563i0.642714+0.710672i),\displaystyle=(c_{1}-c_{2}+c_{3},c_{2})=\begin{pmatrix}-0.406019+0.787063i&1.062972+2.060558i\\ 1.682647+1.860563i&-0.642714+0.710672i\end{pmatrix},
ω\displaystyle\omega^{\prime} =(c2+c3c4,c2c3)=(0.6569531.273495i0.656953+1.273495i1.039933+1.149891i1.0399331.149891i),\displaystyle=(-c_{2}+c_{3}-c_{4},c_{2}-c_{3})=\begin{pmatrix}-0.656953-1.273495i&-0.656953+1.273495i\\ -1.039933+1.149891i&-1.039933-1.149891i\end{pmatrix},

so that we have

Ω=ω1ω=(0.010741690.38178940.38178940.3888885)+i(0.76664480.17307820.17307820.6607763)2.\Omega=\omega^{-1}\omega^{\prime}=\begin{pmatrix}0.01074169&-0.3817894\\ -0.3817894&0.3888885\end{pmatrix}+i\begin{pmatrix}0.7666448&-0.1730782\\ -0.1730782&0.6607763\end{pmatrix}\in\mathfrak{H}_{2}.

Choosing the characteristic α=(1/2,0)tr\alpha=(1/2,0)^{tr}, β=(1/2,1/2)tr\beta=(1/2,1/2)^{tr}, we find that θ(0,Ω)=0\theta(0,\Omega)=0, so we directly obtain from Theorem 1.1 that the components of the function

14π2X0(23)ωIm(Ω)ω¯tr\frac{1}{4\pi^{2}}\mathfrak{Z}_{X_{0}(23)}\omega\operatorname{Im}(\Omega)\overline{\omega}^{tr}

yield preimages of the newforms ff and fσf^{\sigma} resp. under ξ0\xi_{0}, up to the addition of a meromorphic modular form. Their meromorphic parts are given by

0.259008q1+1.000942+4.868978q+18.294037q2+68.247223q3+252.912538q4+938.377980q5+3477.898343q6+12892.503560q7+47787.961740q8+O(q9)0.259008q^{-1}+1.000942+4.868978q+18.294037q^{2}+68.247223q^{3}+252.912538q^{4}\\ +938.377980q^{5}+3477.898343q^{6}+12892.503560q^{7}+47787.961740q^{8}+O(q^{9})

and

0.505669q11.9541676.9786217q26.191387q297.573609q3\displaystyle-0.505669q^{-1}-1.954167-6.9786217q-26.191387q^{2}-97.573609q^{3}
361.535343q41341.254086q54971.053026q618427.581035q7\displaystyle-361.535343q^{4}-1341.254086q^{5}-4971.053026q^{6}-18427.581035q^{7}
68304.578170q8+O(q9).\displaystyle-68304.578170q^{8}+O(q^{9}).

We consider the sum of the components of the vector-valued function. This will be a scalar-valued polar mock modular form 𝔷V(τ)\mathfrak{z}_{V}(\tau) whose shadow is some linear combination of the newforms ff and fσf^{\sigma}. Note that the coefficient of q43q^{43} and that of q109q^{109} in ff both vanish, so it follows from the work of Bruinier-Ono-Rhoades [12], that the coefficients of q43q^{43} and q109q^{109} of a good preimage of ff under ξ0\xi_{0} should be algebraic numbers. Even though 𝔷V\mathfrak{z}_{V} is not guaranteed to be a good preimage in the sense of [12], we still find that

𝔷V(τ)=q1+3.864515+0.142266q+0.319448q2+0.193313q3+0.304709q4+0.055558q5+0.059060q6+0.080332q7+0.572492q80.190607q9+O(q10)\mathfrak{z}_{V}(\tau)=q^{-1}+3.864515+0.142266q+0.319448q^{2}+0.193313q^{3}+0.304709q^{4}\\ +0.055558q^{5}+0.059060q^{6}+0.080332q^{7}+0.572492q^{8}-0.190607q^{9}+O(q^{10})

and the coefficient of q43q^{43} is (within computational precision) 27/4327/43 and that of q109q^{109} is 942/109942/109.

We conclude this example by mentioning a few numerical observations.

  1. (1)

    In this particular case, we see that the matrix P=12i(ω¯ωtrω¯ωtr)P=\frac{1}{2i}\left(\overline{\omega}\omega^{\prime tr}-\overline{\omega^{\prime}}\omega^{tr}\right) from Theorem 1.1 is diagonal, in fact we have, up to computational precision

    P=(3.741508005.347829)=4π2(f,f00fσ,fσ),P=\begin{pmatrix}3.741508&0\\ 0&5.347829\end{pmatrix}=4\pi^{2}\begin{pmatrix}\langle f,f\rangle&0\\ 0&\langle f^{\sigma},f^{\sigma}\rangle\end{pmatrix},

    where ,\langle\cdot,\cdot\rangle denotes the Petersson inner product. Possibly, this is a consequence of Haberland’s formula for subgroups (see e.g. [16, Theorem 5.2]).

  2. (2)

    By the Petersson coefficient formula we can write the (conditionally convergent) cuspidal Poincaré series 𝒫(2,1,23;τ)=γΓ\Γ0(23)e2πiτ|2γ\mathcal{P}(2,1,23;\tau)=\sum_{\gamma\in\Gamma_{\infty}\backslash\Gamma_{0}(23)}e^{2\pi i\tau}|_{2}\gamma as

    𝒫(2,1,23;τ)=14πf,ff(τ)+14πfσ,fσfσ(τ).\mathcal{P}(2,1,23;\tau)=\frac{1}{4\pi\langle f,f\rangle}f(\tau)+\frac{1}{4\pi\langle f^{\sigma},f^{\sigma}\rangle}f^{\sigma}(\tau).

    It is well-known that the preimage of a cuspidal Poincaré series under the ξ\xi-operator is given by the so-called Maass-Poincaré series of dual weight, denoted by 𝒬(2,1,23;τ)\mathcal{Q}(2,-1,23;\tau) (see e.g. [8, Theorem 6.10]). Computing the Fourier expansion of this Poincaré series numerically (see [8, Theorem 6.10] for a description of the coefficients) strongly suggests that indeed

    𝔷^V(τ)=𝒬(2,1,23;τ)+C\widehat{\mathfrak{z}}_{V}(\tau)=\mathcal{Q}(2,-1,23;\tau)+C

    for some constant CC, which would imply that 𝔷^V\widehat{\mathfrak{z}}_{V} indeed has no poles within the upper half-plane. Since their shadows are equal, we know that the difference 𝔷^V(τ)𝒬(2,1,23;τ)\widehat{\mathfrak{z}}_{V}(\tau)-\mathcal{Q}(2,-1,23;\tau) is a meromorphic modular function. By analyzing the behavior at cusps (see Theorem 1.5), we see that this function must have all its poles in the upper half-plane.

  3. (3)

    Since f|W23=ff|W_{23}=-f, it follows by construction that V(τ)+V|W23(τ)\mathfrak{Z}_{V}(\tau)+\mathfrak{Z}_{V}|W_{23}(\tau) should be a meromorphic modular function for Γ0(23)\Gamma_{0}(23) or rather the group Γ0(23)+\Gamma_{0}(23)^{+}. Indeed we find, within computational accuracy, that

    V(τ)+V|W23(τ)\displaystyle\mathfrak{Z}_{V}(\tau)+\mathfrak{Z}_{V}|W_{23}(\tau)
    =(C(q1+α+4q+7q2+13q3+19q4+33q5+47q6+O(q7))C(q1+α+4q+7q2+13q3+19q4+33q5+47q6+O(q7)))\displaystyle=\begin{pmatrix}C(q^{-1}+\alpha+4q+7q^{2}+13q^{3}+19q^{4}+33q^{5}+47q^{6}+O(q^{7}))\\ C^{\prime}(q^{-1}+\alpha^{\prime}+4q+7q^{2}+13q^{3}+19q^{4}+33q^{5}+47q^{6}+O(q^{7}))\end{pmatrix}

    for constants C=2.732921,C=3.7329211,α=0.019847,α=2.543165C=-2.732921...,C^{\prime}=3.7329211...,\alpha=-0.019847,\alpha^{\prime}=2.543165.... Note that the coefficients given above agree, apart from the constant term, with those of the Hauptmodul for the group Γ0(23)+\Gamma_{0}(23)^{+} given by

    T23+(τ)=t(τ)+4t(τ)(t(τ)1),t(τ)=η(τ)η(23τ)η(2τ)η(46τ)T_{23+}(\tau)=t(\tau)+4\frac{t(\tau)}{(t(\tau)-1)},\quad t(\tau)=\frac{\eta(\tau)\eta(23\tau)}{\eta(2\tau)\eta(46\tau)}

    (see e.g. [17, Tables 3 and 4a], correcting an error in loc.cit.).

4.4. Level 256

We consider the newform

f(τ)=q+22q+5q922q11+6q1762q19+O(q21)S2(256),f(\tau)=q+2\sqrt{2}q+5q^{9}-2\sqrt{2}q^{11}+6q^{17}-6\sqrt{2}q^{19}+O(q^{21})\in S_{2}(256),

which has CM by (2)\mathbb{Q}(\sqrt{-2}). This is the smallest level for which there exists a CM newform with non-rational coefficients. As in the case of level 2727 in Section 4.2 the coefficients of the Kleinian mock modular form are algebraic. As before we denote the Galois conjugate of ff by fσf^{\sigma}.

In this case one may check (e.g. by going through a list of generators of Γ0(256)\Gamma_{0}(256)) that the associated period lattice in 2\mathbb{C}^{2} is generated by

c1\displaystyle c_{1} =2πi49/512f(t)dt=(2.7675052.767505i1.146338+1.146338i),\displaystyle=-2\pi i\int_{49/512}^{\infty}\vec{f}(t)\mathrm{d}t=\begin{pmatrix}2.767505-2.767505i\\ 1.146338+1.146338i\end{pmatrix},
c2\displaystyle c_{2} =2πi55/512f(t)dt=(1.9569211.956921i0.8105830.810583i),\displaystyle=-2\pi i\int_{55/512}^{\infty}\vec{f}(t)\mathrm{d}t=\begin{pmatrix}1.956921-1.956921i\\ -0.810583-0.810583i\end{pmatrix},
c3\displaystyle c_{3} =2πi29/256f(t)dt=(0.810583+1.956921i1.956921+0.810583i),\displaystyle=-2\pi i\int_{29/256}^{\infty}\vec{f}(t)\mathrm{d}t=\begin{pmatrix}0.810583+1.956921i\\ 1.956921+0.810583i\end{pmatrix},
c4\displaystyle c_{4} =2πi39/256f(t)dt=(1.956921+1.956921i0.810583+0.810583i).\displaystyle=-2\pi i\int_{39/256}^{\infty}\vec{f}(t)\mathrm{d}t=\begin{pmatrix}1.956921+1.956921i\\ -0.810583+0.810583i\end{pmatrix}.

A suitable basis as described in Section 4.1 is given as follows

ω\displaystyle\omega =(c1,c2c4)=(2.7675052.767505i3.913843i1.146338+1.146338i1.621166i),\displaystyle=(c_{1},c_{2}-c_{4})=\begin{pmatrix}2.767505-2.767505i&-3.913843i\\ 1.146338+1.146338i&-1.621166i\end{pmatrix},
ω\displaystyle\omega^{\prime} =(c2+c3+2c4,c2+c3+c4)\displaystyle=(c_{2}+c_{3}+2c_{4},c_{2}+c_{3}+c_{4})
=(6.681348+3.913843i4.724426+1.956921i0.474828+1.621166i0.335754+0.810583i),\displaystyle=\begin{pmatrix}6.681348+3.913843i&4.724426+1.956921i\\ -0.474828+1.621166i&0.335754+0.810583i\end{pmatrix},

yielding

Ω=ω1ω=i(1.4142130.7071060.7071060.707106)=i22(2111)2.\Omega=\omega^{-1}\omega^{\prime}=i\begin{pmatrix}1.414213&0.707106\\ 0.707106&0.707106\end{pmatrix}=i\frac{\sqrt{2}}{2}\begin{pmatrix}2&1\\ 1&1\end{pmatrix}\in\mathfrak{H}_{2}.

The latter equality is to be understood within computational precision (100 digits), the authors are not aware of a rigorous proof for this observation.

The sum of the components 𝔷V\mathfrak{z}_{V} of V\mathfrak{Z}_{V} is then given by

𝔷V(τ)\displaystyle\mathfrak{z}_{V}(\tau) =q1+(1.828427+0.585786i)q+(0.8284273.313708i)q3\displaystyle=q^{-1}+(1.828427+0.585786i)q+(-0.828427-3.313708i)q^{3}
+(5.2548337.484271i)q5+(7.487074+45.254833i)q7\displaystyle\qquad\qquad\qquad+(-5.254833-7.484271i)q^{5}+(-7.487074+45.254833i)q^{7}
+(27.456710+44.496608i)q9+O(q11)\displaystyle\qquad\qquad\qquad\qquad\qquad+(27.456710+44.496608i)q^{9}+O(q^{11})
=q1+[(1+22)+i(22)]q+[(222)+i(882)]q3\displaystyle=q^{-1}+[(-1+2\sqrt{2})+i(2-\sqrt{2})]q+[(2-2\sqrt{2})+i(8-8\sqrt{2})]q^{3}
+15[(2001602)+i(104+1002)]q5+17[(144110562)+i2242]q7\displaystyle\quad+\frac{1}{5}[(200-160\sqrt{2})+i(104+100\sqrt{2})]q^{5}+\frac{1}{7}[(1441-1056\sqrt{2})+i224\sqrt{2}]q^{7}
+19[(521135102)+i(5238+39872)]q9+O(q11).\displaystyle\quad\quad+\frac{1}{9}[(5211-3510\sqrt{2})+i(-5238+3987\sqrt{2})]q^{9}+O(q^{11}).

Again, the last equality is to be understood within computational accuracy.

References

  • [1] Claudia Alfes, Michael Griffin, Ken Ono, and Larry Rolen. Weierstrass mock modular forms and elliptic curves. Res. Number Theory, 1:Paper No. 24, 31, 2015.
  • [2] Claudia Alfes-Neumann, Jens Funke, Michael Mertens, and Eugenia Rosu. On Jacobi–Weierstrass mock modular forms, 2023. preprint, available at https://arxiv.org/abs/2303.01445.
  • [3] George E. Andrews, Robert C. Rhoades, and Sander P. Zwegers. Modularity of the concave composition generating function. Algebra Number Theory, 7(9):2103–2139, 2013.
  • [4] Henry F. Baker. An Introduction to the Theory of Multiply Periodic Functions. Cambridge University Press, 1907.
  • [5] Rolf Berndt and Ralf Schmidt. Elements of the representation theory of the Jacobi group. Modern Birkhäuser Classics. Birkhäuser/Springer Basel AG, Basel, 1998. [2011 reprint of the 1998 original] [MR1634977].
  • [6] Christina Birkenhake and Herbert Lange. Complex abelian varieties, volume 302 of Grundlehren der mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, second edition, 2004.
  • [7] Wieb Bosma, John Cannon, and Catherine Playoust. The Magma algebra system. I. The user language. J. Symbolic Comput., 24(3-4):235–265, 1997. Computational algebra and number theory (London, 1993).
  • [8] Kathrin Bringmann, Amanda Folsom, Ken Ono, and Larry Rolen. Harmonic Maass forms and mock modular forms: theory and applications, volume 64 of American Mathematical Society Colloquium Publications. American Mathematical Society, Providence, RI, 2017.
  • [9] Kathrin Bringmann and Ken Ono. Lifting cusp forms to Maass forms with an application to partitions. Proc. Natl. Acad. Sci. USA, 104(10):3725–3731, 2007.
  • [10] Jan Bruinier and Ken Ono. Heegner divisors, LL-functions and harmonic weak Maass forms. Ann. of Math. (2), 172(3):2135–2181, 2010.
  • [11] Jan H. Bruinier. Borcherds products on O(2, ll) and Chern classes of Heegner divisors, volume 1780 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 2002.
  • [12] Jan H. Bruinier, Ken Ono, and Robert C. Rhoades. Differential operators for harmonic weak Maass forms and the vanishing of Hecke eigenvalues. Math. Ann., 342(3):673–693, 2008.
  • [13] Jan Hendrik Bruinier and Jens Funke. On two geometric theta lifts. Duke Math. J., 125(1):45–90, 2004.
  • [14] Victor M. Buchstaber, Victor Z. Enolskiĭ, and Dmitri V. Leĭkin. Hyperelliptic Kleinian functions and applications. In Solitons, geometry, and topology: on the crossroad, volume 179 of Amer. Math. Soc. Transl. Ser. 2, pages 1–33. Amer. Math. Soc., Providence, RI, 1997.
  • [15] Ching-Li Chai. The period matrices and theta functions of Riemann. In The legacy of Bernhard Riemann after one hundred and fifty years. Vol. I, volume 35 of Adv. Lect. Math. (ALM), pages 79–106. Int. Press, Somerville, MA, 2016.
  • [16] Henri Cohen. Haberland’s formula and numerical computation of Petersson scalar products. In ANTS X—Proceedings of the Tenth Algorithmic Number Theory Symposium, volume 1 of Open Book Ser., pages 249–270. Math. Sci. Publ., Berkeley, CA, 2013.
  • [17] John H. Conway and Simon P. Norton. Monstrous moonshine. Bull. London Math. Soc., 11(3):308–339, 1979.
  • [18] David A. Cox. Primes of the form x2+ny2x^{2}+ny^{2}. Pure and Applied Mathematics (Hoboken). John Wiley & Sons, Inc., Hoboken, NJ, second edition, 2013. Fermat, class field theory, and complex multiplication.
  • [19] Atish Dabholkar, Sameer Murty, and Don Zagier. Quantum black holes, wall crossing, and mock modular forms. to appear in Cambridge Monographs in Mathematical Physics.
  • [20] Stephan Ehlen, Markus Schwagenscheidt, and Yingkun Li. Harmonic Maass forms associated with CM newforms. preprint, available at https://arxiv.org/abs/2210.07341, 2022.
  • [21] Matthew England and Chris Athorne. Generalised elliptic functions. Cent. Eur. J. Math., 10(5):1655–1672, 2012.
  • [22] Eberhard Freitag. Complex analysis. 2. Universitext. Springer, Heidelberg, 2011. Riemann surfaces, several complex variables, abelian functions, higher modular functions.
  • [23] Pavel Guerzhoy. A mixed mock modular solution of the Kaneko-Zagier equation. Ramanujan J., 36(1-2):149–164, 2015.
  • [24] Haruzo Hida. Congruence of cusp forms and special values of their zeta functions. Invent. Math., 63(2):225–261, 1981.
  • [25] Felix Klein. Ueber hyperelliptische Sigmafunctionen. Math. Ann., 32(3):351–380, 1888.
  • [26] Loïc Merel. Intersections sur des courbes modulaires. Manuscripta Math., 80(3):283–289, 1993.
  • [27] Michael H. Mertens, Ken Ono, and Larry Rolen. Mock modular Eisenstein series with Nebentypus. Int. J. Number Theory, 17(3):683–697, 2021.
  • [28] David Mumford. Tata lectures on theta. I, volume 28 of Progress in Mathematics. Birkhäuser Boston, Inc., Boston, MA, 1983. With the assistance of C. Musili, M. Nori, E. Previato and M. Stillman.
  • [29] Atsushi Nakayashiki. On algebraic expressions of sigma functions for (n,s)(n,s) curves. Asian J. Math., 14(2):175–211, 2010.
  • [30] Andrew P. Ogg. Hyperelliptic modular curves. Bull. Soc. Math. France, 102:449–462, 1974.
  • [31] Ken Ono. Unearthing the visions of a master: harmonic Maass forms and number theory. In Current developments in mathematics, 2008, pages 347–454. Int. Press, Somerville, MA, 2009.
  • [32] Larry Rolen. A new construction of Eisenstein’s completion of the Weierstrass zeta function. Proc. Amer. Math. Soc., 144(4):1453–1456, 2016.
  • [33] William Stein. Modular forms, a computational approach, volume 79 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2007. With an appendix by Paul E. Gunnells.
  • [34] W. A. Stein et al. Sage Mathematics Software (Version 9.8). The Sage Development Team, 2023. http://www.sagemath.org.
  • [35] The PARI Group, Univ. Bordeaux. PARI/GP version 2.13.4, 2022. available from http://pari.math.u-bordeaux.fr/.
  • [36] Xiang Dong Wang. 22-dimensional simple factors of J0(N)J_{0}(N). Manuscripta Math., 87(2):179–197, 1995.
  • [37] André Weil. Elliptic functions according to Eisenstein and Kronecker. Classics in Mathematics. Springer-Verlag, Berlin, 1999. Reprint of the 1976 original.
  • [38] Jiong Yang and LinSheng Yin. Differential operators for Siegel-Jacobi forms. Sci. China Math., 59(6):1029–1050, 2016.