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

\KOMAoptions

DIV=10 \clearscrheadings\automark[section]subsection \cfoot[– \pagemark –]– \pagemark – \lohead\headertitle \rohead\headerauthors \RedeclareSectionCommand[font=,beforeskip=1afterskip=0.5dent=0em]section \RedeclareSectionCommands[font=,afterskip=-1em]subsection,subsubsection \RedeclareSectionCommands[font=,afterskip=-1em,indent=0pt, ]paragraph

\deffootnote

[1em]1.5em1em\thefootnotemark

Non-Holomorphic
Ramanujan-type Congruences for
Hurwitz Class Numbers

Olivia Beckwith    Martin Raum The author was partially supported by Vetenskapsrådet Grants 2015-04139 and 2019-03551.    Olav K. Richter The author was partially supported by Simons Foundation Grant #412655.

Abstract: In contrast to all other known Ramanujan-type congruences, we discover that Ramanujan-type congruences for Hurwitz class numbers can be supported on non-holomorphic generating series. We establish a divisibility result for such non-holomorphic congruences of Hurwitz class numbers. The two keys tools in our proof are the holomorphic projection of products of theta series with a Hurwitz class number generating series and a theorem by Serre, which allows us to rule out certain congruences.
Hurwitz class numbers\blacksquareRamanujan-type congruences\blacksquareholomorphic projection\blacksquareChebotarev density theorem
MSC Primary: 11E41\blacksquareMSC Secondary: 11F33, 11F37

The study of class numbers for imaginary quadratic fields and the related Hurwitz class numbers has a long and rich history. Their divisibility properties were first studied as early as the 1930s [27], but have proved highly elusive. Such divisibility properties are directly reflected in the existence of torsion elements in class groups. The Cohen-Lenstra heuristic [9] has been the guiding principle in the topic, providing conjectures of a statistical nature for the factorization of class numbers; see [15] for an experimentally supported refinement. However, essentially nothing is known about strictly regular patterns of divisibility as opposed to statistical patterns. In this light, it is natural to study Ramanujan-type congruences, i.e., divisibility properties on arithmetic progressions.

The study of congruences of modular forms originates with the Ramanujan congruences [26] for the partition function:

p(5n+4)0(mod 5),p(7n+5)0(mod 7),p(11n+6)0(mod 11).\displaystyle p(5n+4)\equiv 0\;\;(\mathrm{mod}\,5)\text{,}\quad p(7n+5)\equiv 0\;\;(\mathrm{mod}\,7)\text{,}\quad p(11n+6)\equiv 0\;\;(\mathrm{mod}\,11)\text{.}

It is now known that all weakly holomorphic modular forms, including the generating function for p(n)p(n), satisfy congruences [22, 2, 31], which arise from the theory of Galois representations associated to modular forms. Zagier [33] showed that the Hurwitz class numbers are Fourier coefficients of a weight 32\frac{3}{2} mock modular form (in today’s terminology), i.e., the holomorphic part of a harmonic Maass form [23, 24, 11]. Much less is known about congruences in this setting. However, congruences for other weight 32\frac{3}{2} mock modular forms have been studied by several authors [6, 21, 1, 3, 4].

Ramanujan-type congruences for the Hurwitz class numbers H(D)H(D) have not appeared in the literature. For this work, we have employed a computer search to discover many examples of such congruences for H(D)H(D), which can then be confirmed with our method of Section 2 in combination with Sturm bounds for modular forms modulo a prime; see also Remark (2) following Theorem A. For instance one finds that:

H(53n+25)0(mod 5),H(73n+147)0(mod 7),H(113n+242)0(mod 11).\displaystyle H(5^{3}n+25)\equiv 0\;\;(\mathrm{mod}\,5)\text{,}\quad H(7^{3}n+147)\equiv 0\;\;(\mathrm{mod}\,7)\text{,}\quad H(11^{3}n+242)\equiv 0\;\;(\mathrm{mod}\,11)\text{.}

A common theme of all three of these congruences, written in the form H(an+b)0(mod)H(an+b)\equiv 0\,\;(\mathrm{mod}\,\ell), is that b-b is a square modulo aa, which yields a generating series for H(an+b)H(an+b) that is mock modular, i.e., has a non-holomorphic modular completion. This differs from the congruences for mock theta functions of weight 12\frac{1}{2}, which are so far only known to occur when the generating function is a holomorphic modular form [3]. Our main theorem provides the following divisibility result for such non-holomorphic congruences:

Theorem A.

Fix a prime >3\ell>3, a1a\in\mathbb{Z}_{\geq 1}, and bb\in\mathbb{Z}. If b-b is a square modulo aa and

H(an+b)0(mod)\displaystyle H(an+b)\equiv 0\;\;(\mathrm{mod}\,\ell)

for all integers nn, then a\ell\mathop{\mid}a.

Remark 1.
  1. (1)

    We used the Hurwitz-Eichler relations to compute Hurwitz class numbers H(D)H(D) for D<3109D<3\cdot 10^{9}. We did not employ any computer algebra system, but implemented our software in C/C++/Julia from scratch, relying on the FLINT library [14] only for modular arithmetics. The code is available at the second author’s homepage.

  2. (2)

    All examples of Ramanujan-type congruences of Hurwitz class numbers that we discovered, including the non-holomorphic ones, can be explained by a Hecke-type factorization of Hurwitz class numbers (A). Note that all such non-holomorphic Ramanujan-type congruences for H(n)H(n) satisfy the conclusion of Theorem A. While it is not known whether this factorization implies all non-holomorphic Ramanujan-type congruences, our experimental data suggests that it does. Thus, Theorem A provides evidence for this speculation.

  3. (3)

    In the Appendix A we point out that equation  (A) also implies congruences modulo powers of \ell.

  4. (4)

    There are also examples of congruences H(an+b)0(mod)H(an+b)\equiv 0\,\;(\mathrm{mod}\,\ell) where b-b is not a square modulo aa. However, our computational data reveals that in those cases \ell does not divide aa, which is in contrast to other known Ramanujan-type congruences (see [25, 3]). The first examples for {5,7,11}\ell\in\{5,7,11\} are:

    H(33n+9)0(mod 5),H(53n+50)0(mod 7),H(29n+192)0(mod 11).\displaystyle H(3^{3}n+9)\equiv 0\;\;(\mathrm{mod}\,5)\text{,}\quad H(5^{3}n+50)\equiv 0\;\;(\mathrm{mod}\,7)\text{,}\quad H(2^{9}n+192)\equiv 0\;\;(\mathrm{mod}\,11)\text{.}
  5. (5)

    The methods in this work likely generalize to all mock modular forms of weight 32\frac{3}{2}. In particular, congruences of the Andrews spt\mathrm{spt}-function [21] should be subject to requirements similar to the ones presented in Theorem A. On the other hand, both the divisibility and the square class conditions that appear in the case of mock modular forms of weight 12\frac{1}{2}, see [3], seem to be of a different nature and originate in the principal part of mock modular forms.

The method of our proof is novel: We combine a holomorphic projection argument for products of theta series and mock modular forms, which appeared first in [16], with a theorem by Serre [29, 28] that is rooted in the Chebotarev Density Theorem. The latter was employed by Ono [22] in order to establish his celebrated results on the distribution of the partition function modulo primes. Ono applied it to establish congruences, while our proof proceeds by contradiction and uses Serre’s theorem to rule out certain congruences.

Acknowledgments

The authors thank Scott Ahlgren, Jeremy Lovejoy, and the anonymous referee for valuable suggestions.

1 Preliminaries

A recent reference which contains most of the necessary background material for this paper is [5], a more classical one on the theory of modular forms is [19].

1.1 Modular forms

Let Γ0(N)\Gamma_{0}(N), Γ1(N)\Gamma_{1}(N), and Γ(N)\Gamma(N) be the usual congruence sub-groups of SL2()\mathrm{SL}_{2}(\mathbb{Z}), Mk(Γ)\mathrm{M}_{k}(\Gamma) the space of modular forms of integral or half-integral weight kk for ΓSL2()\Gamma\subseteq\mathrm{SL}_{2}(\mathbb{Z}), and 𝕄k(Γ)\mathbb{M}_{k}(\Gamma) the corresponding space of harmonic Maass forms (satisfying the moderate growth condition at all cusps). We also consider quasi-modular forms of integral weight [32, 18]. Moreover, \mathbb{H} denotes the Poincaré upper half plane, and throughout τ\tau\in\mathbb{H}, y=Im(τ)y=\mathrm{Im}(\tau), and e(sτ):=exp(2πisτ)e(s\tau):=\exp(2\pi i\,s\tau) for ss\in\mathbb{Q}.

For a holomorphic modular form f(τ)=m0c(f,m)e(mτ)M2k(Γ(N))f(\tau)=\sum_{m\geq 0}c(f,m)e(m\tau)\in\mathrm{M}_{2-k}(\Gamma(N)) with k1k\neq 1 and N1N\in\mathbb{Z}_{\geq 1}, we recall its non-holomorphic Eichler integral

f(τ):=(2i)k1τ¯iG(w¯)¯(w+τ)k𝑑w=c(f,0)¯1ky1k(4π)k1m<0c(f,|m|)¯|m|k1Γ(1k,4π|m|y)e(mτ),\displaystyle\begin{aligned} f^{\ast}(\tau)\;&:=\;-(2i)^{k-1}\int_{-\overline{\tau}}^{i\infty}\frac{\overline{G(-\overline{w})}}{(w+\tau)^{k}}\,d\!w\\ &\hphantom{:}=\frac{\overline{c(f,0)}}{1-k}\,y^{1-k}\,-\,(4\pi)^{k-1}\sum_{m<0}\overline{c(f,|m|)\,}|m|^{k-1}\Gamma(1-k,4\pi|m|y)e(m\tau)\text{,}\end{aligned} (1.1)

where Γ\Gamma stands for the upper incomplete Gamma-function.

1.2 Generating series of Hurwitz class numbers

Zagier [33] showed that the holomorphic generating series DH(D)e(Dτ)\sum_{D}H(D)e(D\tau) of Hurwitz class numbers admits the following modular completion:

E32(τ):=D=0H(D)e(Dτ)+116πθ(τ)𝕄32(Γ0(4)),\displaystyle E_{\frac{3}{2}}(\tau)\;:=\;\sum_{D=0}^{\infty}H(D)e(D\tau)\,+\,\frac{1}{16\pi}\theta^{\ast}(\tau)\;\in\;\mathbb{M}_{\frac{3}{2}}(\Gamma_{0}(4))\text{,} (1.2)

where

θ:=θ1,0M12(Γ0(4)) withθa,b(τ):=nnb(moda)e(n2τa)M12(Γ(4a)),a1,b.\displaystyle\begin{aligned} \theta&\;:=\;\theta_{1,0}\,\in\,\mathrm{M}_{\frac{1}{2}}(\Gamma_{0}(4))\text{\, with}\\ \theta_{a,b}(\tau)&\;:=\;\sum_{\begin{subarray}{c}n\in\mathbb{Z}\\ n\equiv b\,\;(\mathrm{mod}\,a)\end{subarray}}e\big{(}\tfrac{n^{2}\tau}{a}\big{)}\,\in\,\mathrm{M}_{\frac{1}{2}}(\Gamma(4a))\text{,}\quad a\in\mathbb{Z}_{\geq 1},b\in\mathbb{Z}\text{.}\end{aligned} (1.3)

For a1a\in\mathbb{Z}_{\geq 1} and bb\in\mathbb{Z}, we define the following operator on Fourier series expansions of non-holomorphic modular forms:

Ua,bnc(f;n;y)e(nτ):=nnb(moda)c(f;n;ya)e(nτa).\displaystyle\mathrm{U}_{a,b}\,\sum_{n\in\mathbb{Z}}c(f;\,n;\,y)e(n\tau)\;:=\;\sum_{\begin{subarray}{c}n\in\mathbb{Z}\\ n\equiv b\,\;(\mathrm{mod}\,a)\end{subarray}}c\big{(}f;\,n;\,\tfrac{y}{a}\big{)}e\big{(}\tfrac{n\tau}{a}\big{)}\text{.} (1.4)

The holomorphic part of Ua,bE32(τ)\mathrm{U}_{a,b}\,E_{\frac{3}{2}}(\tau) is the generating series of Hurwitz class numbers H(an+b)H(an+b) for nn\in\mathbb{Z}. A Hecke-theory-like computation (see also [8, 17]) shows that

Ua,bE32M32(Γ(4a)).\displaystyle\mathrm{U}_{a,b}\,E_{\frac{3}{2}}\,\in\,\mathrm{M}_{\frac{3}{2}}(\Gamma(4a))\text{.} (1.5)

Moreover, we have

Ua,bθ=β2b(moda)θa,βandUa,bθ=β2b(moda)aθa,β.\displaystyle\mathrm{U}_{a,b}\,\theta\;=\;\sum_{\beta^{2}\equiv b\,\;(\mathrm{mod}\,a)}\theta_{a,\beta}\quad\text{and}\quad\mathrm{U}_{a,b}\,\theta^{\ast}\;=\;\sum_{\beta^{2}\equiv-b\,\;(\mathrm{mod}\,a)}\sqrt{a}\,\theta^{\ast}_{a,\beta}\text{.} (1.6)

In particular, if b-b is not a square modulo aa, then Ua,bE3/2\mathrm{U}_{a,b}\,E_{3/2} is a holomorphic modular form.

1.3 A theorem by Serre

The following theorem by Serre and its corollary allow us to disprove that a given generating series is a quasi-modular form modulo a prime. Recall that a rational number is called \ell-integral for a prime \ell, if its denominator is not divisible by \ell.

Theorem 1.1 (Deligne-Serre [10] and Serre [29, 28]).

Fix an odd prime \ell and k,N1k,N\in\mathbb{Z}_{\geq 1}. Then there are infinitely many primes p1(modN)p\equiv 1\,\;(\mathrm{mod}\,\ell N) such that for every fMk(Γ1(N))f\in\mathrm{M}_{k}(\Gamma_{1}(N)) with \ell-integral Fourier coefficients, we have

c(f;npr)(r+1)c(f;n)(mod)\displaystyle c(f;\,np^{r})\;\equiv\;(r+1)c(f;\,n)\;\;(\mathrm{mod}\,\ell) (1.7)

for all nn\in\mathbb{Z} co-prime to pp and all r0r\in\mathbb{Z}_{\geq 0}.

Proof.

The proof of Lemma 9.6 of [10], which is stated in the special case of weight 11, extends verbatim.

Corollary 1.2.

Fix a prime >3\ell>3 and positive integers kk and NN. Then there are infinitely many primes p1(modN)p\equiv 1\,\;(\mathrm{mod}\,\ell N) such that for every quasi-modular form ff of weight kk for Γ1(N)\Gamma_{1}(N) with \ell-integral Fourier coefficients, we have the congruence (1.7).

Proof.

Recall the weight 22 quasi-modular form

E2(τ)= 124n=1σ1(n)e(nτ),σ1(n):=dnd>0d.\displaystyle E_{2}(\tau)\;=\;1-24\sum_{n=1}^{\infty}\sigma_{1}(n)e(n\tau)\text{,}\quad\sigma_{1}(n)\;:=\;\sum_{\begin{subarray}{c}d\mathop{\mid}n\\ d>0\end{subarray}}d\text{.}

Quasi-modular forms are polynomials in E2E_{2} whose coefficients are modular forms. More specifically, a quasi-modular form of weight kk for Γ1(N)\Gamma_{1}(N) can be written as

n=0dE2nfn,fnMk2n(Γ1(N)).\displaystyle\sum_{n=0}^{d}E_{2}^{n}f_{n}\text{,}\quad f_{n}\in\mathrm{M}_{k-2n}(\Gamma_{1}(N))\text{.}

Recall also that the weight 1\ell-1 and level 11 Eisenstein series E1E_{\ell-1} is a modular form (here we use that >3\ell>3) that is congruent to 11 modulo \ell, and that the weight +1\ell+1 and level 11 Eisenstein series E+1E_{\ell+1} is a modular form that is congruent to E2E_{2} modulo \ell. Thus,

n=0dE2nfnE1dn=0dE2nfnn=0d(E2E1)nE1dnfnn=0dE+1nE1dnfn(mod),\displaystyle\sum_{n=0}^{d}E_{2}^{n}f_{n}\;\equiv\;E_{\ell-1}^{d}\sum_{n=0}^{d}E_{2}^{n}f_{n}\;\equiv\;\sum_{n=0}^{d}(E_{2}E_{\ell-1})^{n}E_{\ell-1}^{d-n}f_{n}\;\equiv\;\sum_{n=0}^{d}E_{\ell+1}^{n}E_{\ell-1}^{d-n}f_{n}\;\;(\mathrm{mod}\,\ell)\text{,}

which allows us to apply Theorem 1.1 to the modular form of weight k+d(1)k+d(\ell-1) on the right hand side.

1.4 Holomorphic projection

We now revisit holomorphic projection, which allows one to map continuous functions with certain growth and modular behavior to holomorphic modular forms (for example, see [30, 13]). It is convenient for us to refer to [16] as a reference, since it provides a variant that does not project to cusp forms. Fix a weight kk\in\mathbb{Z}, k2k\geq 2 and a level N1N\in\mathbb{Z}_{\geq 1}.

Consider an NN-periodic continuous function f:f:\,\mathbb{H}\rightarrow\mathbb{C} with Fourier series expansion

f(τ)=n1Nc(f;n;y)e(τn)\displaystyle f(\tau)\;=\;\sum_{n\in\frac{1}{N}\mathbb{Z}}c(f;\,n;\,y)e(\tau n)

subject to the conditions: (i) For some a>0a>0 and all γSL2()\gamma\in\mathrm{SL}_{2}(\mathbb{Z}), there are coefficients c~(f|kγ; 0)\tilde{c}(f|_{k}\,\gamma;\,0)\in\mathbb{C}, such that (f|kγ)(τ)=c~(f|kγ; 0)+𝒪(ya)(f|_{k}\,\gamma)(\tau)=\tilde{c}(f|_{k}\,\gamma;\,0)+\mathcal{O}(y^{-a}) as yy\rightarrow\infty; (ii) For all n1N>0n\in\frac{1}{N}\mathbb{Z}_{>0}, we have c(f;n;y)=𝒪(y2k)c(f;\,n;\,y)=\mathcal{O}(y^{2-k}) as y0y\rightarrow 0. The holomorphic projection operator of weight kk is defined by

πkhol(f):=c~(f; 0)+n1N>0c(πkhol(f);n)e(nτ)withc(πkhol(f);n):=(4πn)k1Γ(k1)lims00c(f;n;y)exp(4πny)ys+k2dy.\displaystyle\begin{aligned} \pi^{\mathrm{hol}}_{k}(f)\;&:=\;\tilde{c}(f;\,0)\,+\,\sum_{n\in\frac{1}{N}\mathbb{Z}_{>0}}c\big{(}\pi^{\mathrm{hol}}_{k}(f);\,n\big{)}\,e(n\tau)\quad\text{with}\\ c\big{(}\pi^{\mathrm{hol}}_{k}(f);\,n\big{)}\;&:=\;\frac{(4\pi n)^{k-1}}{\Gamma(k-1)}\,\lim_{s\rightarrow 0}\,\int_{0}^{\infty}c(f;\,n;\,y)\exp(-4\pi ny)y^{s+k-2}\,\mathrm{d}\!y\text{.}\end{aligned} (1.8)
Remark 1.3.

The formulation in [16] is missing the limit s0s\rightarrow 0, which is required to ensure convergence in all situations. Nevertheless, the arguments in [16] are still valid without any restrictions after adding that missing regularization.

Proposition 4 and Theorem 5 of [16] provide some key properties of the holomorphic projections operator (for vector-valued modular forms): If ff is holomorphic, then πkhol(f)=f\pi^{\mathrm{hol}}_{k}(f)=f. Furthermore, if ff transforms like a modular form of weight 22 for the group Γ1(N)\Gamma_{1}(N), then π2hol(f)\pi^{\mathrm{hol}}_{2}(f) is a quasi-modular form of weight 22 for Γ1(N)\Gamma_{1}(N).

2 Proof of Theorem A

We will apply holomorphic projection to products of holomorphic modular forms of weight 12\frac{1}{2} and harmonic Maass forms of weight 32\frac{3}{2}. These products when inserted into (1.8) lead to the integrals evaluated in the next two lemmas.

Lemma 2.1.

Given n>0n\in\mathbb{Q}_{>0}, we have

π2hol(y12e(nτ))= 2πne(nτ).\displaystyle\pi^{\mathrm{hol}}_{2}\Big{(}y^{-\frac{1}{2}}\,e(n\tau)\Big{)}\;=\;2\pi\sqrt{n}\,e(n\tau)\text{.}

Proof.

We suppress the limit s0s\rightarrow 0 from our calculation, since the integral is convergent at s=0s=0. We then have to evaluate

4πnΓ(1)0y12exp(4πny)dy=4πnΓ(1)(4πn)12Γ(12)=2πn.\displaystyle\frac{4\pi n}{\Gamma(1)}\int_{0}^{\infty}y^{-\frac{1}{2}}\,\exp(-4\pi ny)\,\mathrm{d}\!y=\frac{4\pi n}{\Gamma(1)}(4\pi n)^{-\frac{1}{2}}\Gamma(\tfrac{1}{2})=2\pi\sqrt{n}\text{.}

Lemma 2.2.

Given rational numbers n,n~n,\tilde{n}\in\mathbb{Q} satisfying n~<0\tilde{n}<0 and n+n~>0n+\tilde{n}>0, we have

π2hol(Γ(12, 4π|n~|y)e((n+n~)τ))=2π(n+n~)|n~|(n+|n~|)e((n+n~)τ).\displaystyle\pi^{\mathrm{hol}}_{2}\Big{(}\Gamma\big{(}-\tfrac{1}{2},\,4\pi|\tilde{n}|y\big{)}e\big{(}(n+\tilde{n})\tau\big{)}\Big{)}=\frac{2\sqrt{\pi}(n+\tilde{n})}{\sqrt{|\tilde{n}|}\,(\sqrt{n}+\sqrt{|\tilde{n}|})}\,e\big{(}(n+\tilde{n})\tau\big{)}\text{.} (2.1)

Proof.

Again, we suppress the limit s0s\rightarrow 0 from our calculation, since the integral is convergent at s=0s=0. We have to evaluate

0Γ(12, 4π|n~|y)exp(4π(n+n~)y)dy.\displaystyle\int_{0}^{\infty}\Gamma\big{(}-\tfrac{1}{2},\,4\pi|\tilde{n}|y\big{)}\exp(-4\pi(n+\tilde{n})y)\,\mathrm{d}\!y\text{.}

We apply case 1 of (6.455) on p. 657 of [12] with α4π|n~|\alpha\leadsto 4\pi|\tilde{n}|, β4π(n+n~)\beta\leadsto 4\pi(n+\tilde{n}), μ1\mu\leadsto 1, and ν12\nu\leadsto-\frac{1}{2}. The assumptions Re(α+β)=4π((n+n~)+|n~|)>0\mathrm{Re}(\alpha+\beta)=4\pi((n+\tilde{n})+|\tilde{n}|)>0 (since n+n~>0n+\tilde{n}>0), Re(μ)=1>0\mathrm{Re}(\mu)=1>0, and Re(μ+ν)=12>0\mathrm{Re}(\mu+\nu)=\frac{1}{2}>0 are satisfied. As a result, we obtain

(4π|n~|)12Γ(12)1(4π((n+n~)+|n~|))12F12(1,12; 2;n+n~n+n~+|n~|)=14πn|n~|F12(1,12; 2;n+n~n).\displaystyle\frac{(4\pi|\tilde{n}|)^{-\frac{1}{2}}\,\Gamma(\frac{1}{2})}{1\cdot(4\pi((n+\tilde{n})+|\tilde{n}|))^{\frac{1}{2}}}\,{}_{2}\mathrm{F}_{1}\Big{(}1,\tfrac{1}{2};\,2;\,\frac{n+\tilde{n}}{n+\tilde{n}+|\tilde{n}|}\Big{)}=\frac{1}{4\sqrt{\pi n|\tilde{n}|}}\,{}_{2}\mathrm{F}_{1}\Big{(}1,\tfrac{1}{2};\,2;\,\frac{n+\tilde{n}}{n}\Big{)}\text{.}

To evaluate the hypergeometric function, we employ 15.4.17 of [20] with a12a\leadsto\frac{1}{2}. It allows us to simplify the previous expression to

14πn|n~|(12+121n+n~n)1=12π|n~|(n+|n~|).\displaystyle\frac{1}{4\sqrt{\pi n|\tilde{n}|}}\,\Big{(}\frac{1}{2}+\frac{1}{2}\sqrt{1-\frac{n+\tilde{n}}{n}}\Big{)}^{-1}=\frac{1}{2\sqrt{\pi|\tilde{n}|}\,(\sqrt{n}+\sqrt{|\tilde{n}|})}\text{.}

We combine this with the leading factor 4π(n+n~)4\pi(n+\tilde{n}) in the defining equation (1.8) of the holomorphic projection to finish the proof.

The next three results compute the nn-th coefficient of π2hol((Ua,bE32)(θa,β+θa,β))\pi_{2}^{\mathrm{hol}}\big{(}(\mathrm{U}_{a,b}E_{\frac{3}{2}})\cdot(\theta_{a,\beta}+\theta_{a,-\beta})\big{)} for certain nn. The first elementary lemma establishes the existence of a subprogression of a+b={an+b:n}a\mathbb{Z}+b=\{an+b:n\in\mathbb{Z}\} satisfying arithmetic conditions that will be useful in the proof of Theorem A. To state it, we let ordp(a)\operatorname{ord}_{p}(a) be the maximal exponent for powers of a prime pp dividing an nonzero integer aa.

Lemma 2.3.

Let a~1\tilde{a}\in\mathbb{Z}_{\geq 1} and b~\tilde{b}\in\mathbb{Z} be such that b~-\tilde{b} is a square modulo a~\tilde{a}. Denote by PP the set of prime divisors of a~\tilde{a}. Then there exist a1a\in\mathbb{Z}_{\geq 1} and bb\in\mathbb{Z} such that

  1. (i)

    We have a~a\tilde{a}\mathop{\mid}a, and if pp is a prime divisor of aa then pPp\in P.

  2. (ii)

    The integer b-b is a square modulo aa.

  3. (iii)

    We have

    h(p)>{2k(p),if p2;2k(p)+4,if p=2;\displaystyle h(p)\;>\;\begin{cases}2k(p)\text{,}&\text{if }p\neq 2\text{;}\\ 2k(p)+4\text{,}&\text{if }p=2\text{;}\end{cases}

    where h(p)=ordp(a)h(p)=\operatorname{ord}_{p}(a) and k(p)=ordp(b)k(p)=\operatorname{ord}_{p}(b).

  4. (iv)

    For pPp\in P and for all disjoint sets P1,P2P_{1},P_{2} with P1P2=P\{p}P_{1}\cup P_{2}=P\backslash\{p\}, we have

    4b{pk(p)qP1qk(q)qP2q2h(q)(modph(p)),if p2,pk(p)+2qP1qk(q)qP2qh(q)(modph(p)),if p=2.\displaystyle-4b\;\not\equiv\;\left\{\begin{aligned} &p^{k(p)}\prod_{q\in P_{1}}q^{k(q)}\prod_{q\in P_{2}}q^{2h(q)}&&\;\;(\mathrm{mod}\,p^{h(p)})\text{,}\qquad&&\text{if }p\neq 2\text{,}\\ &p^{k(p)+2}\prod_{q\in P_{1}}q^{k(q)}\prod_{q\in P_{2}}q^{h(q)}&&\;\;(\mathrm{mod}\,p^{h(p)})\text{,}\qquad&&\text{if }p=2\text{.}\end{aligned}\right. (2.2)

Proof.

In order to produce integers aa and bb that satisfy Conditions (i)(iv), we initially set a=a~a=\tilde{a} and b=b~b=\tilde{b} and then repeatedly replace aa and bb by aaaa^{\prime} and b+abb+ab^{\prime} for integers a,ba^{\prime},b^{\prime} in such a way that successively more of these conditions are met. In accordance with Condition (i), the prime divisors of aa^{\prime} must be elements of PP.

Recall that h(p)=ordp(a)h(p)=\operatorname{ord}_{p}(a) and k(p)=ordp(b)k(p)=\operatorname{ord}_{p}(b). For each prime pPp\in P, we define ap:=a/ph(p)a_{p}:=a/p^{h(p)} and similarly define bp:=b/pk(p)b_{p}:=b/p^{k(p)}. We start by making the substitutions aaaa\leadsto aa^{\prime} and bb+abb\leadsto b+ab^{\prime} several times, in such a way that (b+ab)-(b+ab^{\prime}) is still a square modulo aaaa^{\prime} (hence Condition (ii) remains true) and that Condition (iii) is satisfied.

First choose a=1a^{\prime}=1, and pick a suitable bb^{\prime} such that h(p)k(p)h(p)\geq k(p) holds for each pPp\in P after replacing bb by b+abb+ab^{\prime}. Next let P0:={pP:h(p)=k(p)}P_{0}:=\{p\in P\,:\,h(p)=k(p)\}. For each pP0p\in P_{0}, let e(p)=1e(p)=1 if h(p)h(p) is even, and let e(p)=2e(p)=2 if h(p)h(p) is odd. Consider a=pP0pe(p)a^{\prime}=\prod_{p\in P_{0}}p^{e(p)}. Let bb^{\prime} be any integer satisfying gcd(b,pP\P0p)=1\gcd(b^{\prime},\prod_{p\in P\backslash P_{0}}p)=1 and

b{(1+bp)ap1(modp),if pP0 and 2h1(p);(p+bp)ap1(modp2),if pP0 and 2h1(p).\displaystyle b^{\prime}\;\equiv\;\left\{\begin{aligned} &-(1+b_{p})a_{p}^{-1}&&\;\;(\mathrm{mod}\,p)\text{,}\qquad&&\text{if $p\in P_{0}$ and $2\mathop{\mid}h_{1}(p)$;}\\ &-(p+b_{p})a_{p}^{-1}&&\;\;(\mathrm{mod}\,p^{2})\text{,}&&\text{if $p\in P_{0}$ and $2\mathop{\nmid}h_{1}(p)$.}\end{aligned}\right.

Note that the minus signs of the terms (1+bp)-(1+b_{p}) and (p+bp)-(p+b_{p}) are required to ensure that Condition (ii) holds after replacing aa and bb by aaaa^{\prime} and b+abb+ab^{\prime}. After this substitution we may assume that h(p)>k(p)h(p)>k(p) for all pPp\in P.

We take this idea a step further when p=2p=2. If h(2)k(2)=1h(2)-k(2)=1, then we choose a=2a^{\prime}=2, and bb^{\prime} is 0 or 11 depending on whether bpb_{p} is 33 or 11 modulo 44. We find that we may assume h(2)k(2)2h(2)-k(2)\geq 2. Similarly, if we assume that h(2)k(2)=2h(2)-k(2)=2, then we can choose a=2a^{\prime}=2, and bb^{\prime} is 0 or 11 depending on whether bpb_{p} is 77 or 33 modulo 88. Thus, we can assume that h(2)k(2)3h(2)-k(2)\geq 3.

After making the above assumptions on h(p)h(p) and k(p)k(p), we conclude that b-b is also a square modulo aaaa^{\prime} for any aa^{\prime} satisfying Condition (i). In particular, after choosing an appropriate aa^{\prime}, we may assume that h(p)h(p) is as large as we wish. Thus, we can assume that Condition (iii) holds.

It remains to make one more substitution aaaa\leadsto aa^{\prime} and bb+abb\leadsto b+ab^{\prime} in order to ensure Condition (iv). Observe that the right hand side of (2.2) can take at most 2|P|12^{|P|-1} different values, corresponding to the 2|P|12^{|P|-1} different choices of P1P_{1} and P2P_{2}. Given aa^{\prime}, we let h(p):=ordp(a)h^{\prime}(p):=\operatorname{ord}_{p}(a^{\prime}) for pPp\in P. Any choice of bb^{\prime} preserves Conditions (i), (ii), and (iii), and this yields ph(p)p^{h^{\prime}(p)} different values for the left hand side of (2.2) after replacing bb by b+abb+ab^{\prime}. Finally, choose aa^{\prime} in such a way that ph(p)>2|P|1p^{h^{\prime}(p)}>2^{|P|-1} for every pPp\in P, and then pick a suitable bb^{\prime} to validate Condition (iv).

The next result will be useful in the computation of the Fourier series coefficients of π2hol((Ua,bE32)(θa,β+θa,β))\pi_{2}^{\mathrm{hol}}\big{(}(\mathrm{U}_{a,b}E_{\frac{3}{2}})\cdot(\theta_{a,\beta}+\theta_{a,-\beta})\big{)}. We continue using PP to denote the set of prime divisors of aa, and we assume the definition of h(p)h(p) and k(p)k(p) from Lemma 2.3. We also let β\beta be such that β2b(moda)\beta^{2}\equiv-b\;(\mathrm{mod}\,a). Thus, we have a=paph(p)a=\prod_{p\mathop{\mid}a}p^{h(p)}, and β=βpapk(p)/2\beta=\beta^{\prime}\prod_{p\mathop{\mid}a}p^{k(p)/2}, where β\beta^{\prime} is co-prime to aa.

Lemma 2.4.

Assume that a1a\in\mathbb{Z}_{\geq 1} and bb\in\mathbb{Z} satisfy Conditions (i)(iv) in Lemma 2.3, and let \ell be a prime that does not divide aa.

For each prime pPp\in P, let q(p)q(p) be a prime that is congruent 11 modulo ph(p)p^{h(p)}, congruent to 2β/pk(p)/22\beta/p^{k(p)/2} modulo ph(p)k(p)/2p^{h(p)-k(p)/2}, and congruent modulo \ell to a unit such that

ph(p)+q(p)pk(p)\displaystyle p^{h(p)}+q(p)p^{k(p)}  1(mod),\displaystyle\;\equiv\;1\;\;(\mathrm{mod}\,\ell)\text{,}\quad if p2p\neq 2, or
ph(p)+q(p)pk(p)+1\displaystyle p^{h(p)}+q(p)p^{k(p)+1}  1(mod),\displaystyle\;\equiv\;1\;\;(\mathrm{mod}\,\ell)\text{,}\quad if p=2p=2.

Moreover, let

na:={pPq(p)pk(p),if 2a;8pPq(p)pk(p),if 2a.\displaystyle n_{a}\;:=\;\left\{\begin{aligned} &\hphantom{8{}}\prod_{p\in P}q(p)p^{k(p)}\text{,}\quad&&\text{if\/ $2\mathop{\nmid}a$;}\\ &8\prod_{p\in P}q(p)p^{k(p)}\text{,}\quad&&\text{if\/ $2\mathop{\mid}a$.}\end{aligned}\right.

Then we have

β~2b(moda)ϵ=±1ana=d1d2d1,d2>0d1ϵβ+β~(moda)d2ϵββ~(moda)(d1+d2) 2(mod).\displaystyle\sum_{\begin{subarray}{c}\tilde{\beta}^{2}\equiv-b\,\;(\mathrm{mod}\,a)\\ \epsilon=\pm 1\end{subarray}}\sum_{\begin{subarray}{c}an_{a}=d_{1}d_{2}\\ d_{1},d_{2}>0\\ d_{1}\equiv\epsilon\beta+\tilde{\beta}\,\;(\mathrm{mod}\,a)\\ d_{2}\equiv\epsilon\beta-\tilde{\beta}\,\;(\mathrm{mod}\,a)\end{subarray}}\big{(}d_{1}+d_{2}\big{)}\;\equiv\;2\;\;(\mathrm{mod}\,\ell)\text{.}

Proof.

Because the symmetry β~β~\tilde{\beta}\mapsto-\tilde{\beta} swaps d1d_{1} and d2d_{2}, we have

β~2b(moda)ϵ=±1ana=d1d2d1,d2>0d1ϵβ+β~(moda)d2ϵββ~(moda)(d1+d2)= 2β~2b(moda)ϵ=±1ana=d1d2d1,d2>0d1ϵβ+β~(moda)d2ϵββ~(moda)d1.\displaystyle\sum_{\begin{subarray}{c}\tilde{\beta}^{2}\equiv-b\,\;(\mathrm{mod}\,a)\\ \epsilon=\pm 1\end{subarray}}\sum_{\begin{subarray}{c}an_{a}=d_{1}d_{2}\\ d_{1},d_{2}>0\\ d_{1}\equiv\epsilon\beta+\tilde{\beta}\,\;(\mathrm{mod}\,a)\\ d_{2}\equiv\epsilon\beta-\tilde{\beta}\,\;(\mathrm{mod}\,a)\end{subarray}}\big{(}d_{1}+d_{2}\big{)}\;=\;2\cdot\sum_{\begin{subarray}{c}\tilde{\beta}^{2}\equiv-b\,\;(\mathrm{mod}\,a)\\ \epsilon=\pm 1\end{subarray}}\sum_{\begin{subarray}{c}an_{a}=d_{1}d_{2}\\ d_{1},d_{2}>0\\ d_{1}\equiv\epsilon\beta+\tilde{\beta}\,\;(\mathrm{mod}\,a)\\ d_{2}\equiv\epsilon\beta-\tilde{\beta}\,\;(\mathrm{mod}\,a)\end{subarray}}d_{1}\text{.} (2.3)

If p2p\neq 2, the square roots β~\tilde{\beta} of b(modph(p))-b\,\;(\mathrm{mod}\,p^{h(p)}) are of the form ±β+ph(p)k(p)β~\pm\beta+p^{h(p)-k(p)}\tilde{\beta}^{\prime} for some β~/pk(p)\tilde{\beta}^{\prime}\in\mathbb{Z}/p^{k(p)}\mathbb{Z}. In the sum, we therefore have the conditions

d1ϵβ±β+ph(p)k(p)β~(modph(p)),d2ϵββph(p)k(p)β~(modph(p)).\displaystyle d_{1}\;\equiv\;\epsilon\beta\pm\beta+p^{h(p)-k(p)}\tilde{\beta}^{\prime}\;\;(\mathrm{mod}\,p^{h(p)})\text{,}\quad d_{2}\;\equiv\;\epsilon\beta\mp\beta-p^{h(p)-k(p)}\tilde{\beta}^{\prime}\;\;(\mathrm{mod}\,p^{h(p)})\text{.}

Using h(p)>2k(p)h(p)>2k(p), asserted by Condition (iii) of Lemma 2.3, and (2.2), we find that the only possibility that ana=d1d2an_{a}=d_{1}d_{2} satisfies these congruences is if we have ϵ=+1\epsilon=+1, β~0(modph(p))\tilde{\beta}^{\prime}\equiv 0\,\;(\mathrm{mod}\,p^{h(p)}), and

ph(p)d1,q(p)pk(p)d2orq(p)pk(p)d1,ph(p)d2.\displaystyle p^{h(p)}\mathop{\mid}d_{1}\text{,}\;q(p)p^{k(p)}\mathop{\mid}d_{2}\quad\text{or}\quad q(p)p^{k(p)}\mathop{\mid}d_{1}\text{,}\;p^{h(p)}\mathop{\mid}d_{2}\text{.} (2.4)

The case of p=2p=2 allows for more square roots of b(mod 2h(p))-b\,\;(\mathrm{mod}\,2^{h(p)}). Specifically, β~={1,c3,c5,c7}β+ph(p)k(p)β~\tilde{\beta}=\{1,c_{3},c_{5},c_{7}\}\beta+p^{h(p)-k(p)}\tilde{\beta}^{\prime}, where by the set notation we indicate one of the factors occurs and c33(mod 8)c_{3}\equiv 3\,\;(\mathrm{mod}\,8), c55(mod 8)c_{5}\equiv 5\,\;(\mathrm{mod}\,8), c77(mod 8)c_{7}\equiv 7\,\;(\mathrm{mod}\,8) are roots of 11 modulo 2h(2)2^{h(2)}. We now have the conditions

d1\displaystyle d_{1} {2,1+c3,1+c5,1+c7}β+2h(2)k(2)β~(mod 2h(2)),\displaystyle\;\equiv\;\{2,1+c_{3},1+c_{5},1+c_{7}\}\beta+2^{h(2)-k(2)}\tilde{\beta}^{\prime}\;\;(\mathrm{mod}\,2^{h(2)})\text{,}
d2\displaystyle d_{2} {0,1c3,1c5,1c7}β2h(2)k(2)β~(mod 2h(2)).\displaystyle\;\equiv\;\{0,1-c_{3},1-c_{5},1-c_{7}\}\beta-2^{h(2)-k(2)}\tilde{\beta}^{\prime}\;\;(\mathrm{mod}\,2^{h(2)})\text{.}

We use that k(p)+3<h(p)k(p)k(p)+3<h(p)-k(p), asserted by Condition (iii) of Lemma 2.3, and (2.2) in order to see that the only possibilities are ϵ=+1\epsilon=+1, β~0(mod 2h(2))\tilde{\beta}^{\prime}\equiv 0\,\;(\mathrm{mod}\,2^{h(2)}), and the divisibilities in (2.4) with p=2p=2.

Observe that modulo \ell, the sum actually factors for our choice of nan_{a}. Specifically, we have

β~2b(moda)ana=d1d2d1,d2>0d1β+β~(moda)d2ββ~(moda)d1=pap2(ph(p)+q(p)pk(p)){1,if 2a;2h(2)+q(2)2k(2)+1,if 2a.\sum_{\tilde{\beta}^{2}\equiv-b\,\;(\mathrm{mod}\,a)}\sum_{\begin{subarray}{c}an_{a}=d_{1}d_{2}\\ d_{1},d_{2}>0\\ d_{1}\equiv\beta+\tilde{\beta}\,\;(\mathrm{mod}\,a)\\ d_{2}\equiv\beta-\tilde{\beta}\,\;(\mathrm{mod}\,a)\end{subarray}}d_{1}\\ \;=\;\prod_{\begin{subarray}{c}p\mathop{\mid}a\\ p\neq 2\end{subarray}}\big{(}p^{h(p)}+q(p)p^{k(p)}\big{)}\,\cdot\,\left\{\begin{aligned} &1\text{,}\quad&&\text{if $2\mathop{\nmid}a$;}\\ &2^{h(2)}+q(2)2^{k(2)+1}\text{,}\quad&&\text{if $2\mathop{\mid}a$.}\end{aligned}\right. (2.5)

Our choice of q(p)q(p) ensures that this is congruent to 11 modulo \ell. The additional factor 22 in (2.3) yields the desired result.

Next we compute certain coefficients of π2hol((Ua,bE32)(θa,β+θa,β))\pi^{\mathrm{hol}}_{2}\big{(}(\mathrm{U}_{a,b}\,E_{\frac{3}{2}})\cdot(\theta_{a,\beta}+\theta_{a,-\beta})\big{)}. We assume the notation above for nan_{a}, k(p)k(p), h(p)h(p), and β\beta.

Proposition 2.5.

Assume that H(an+b)0(mod)H(an+b)\equiv 0\;(\mathrm{mod}\,\ell) for all nn. Furthermore, assume that β2b(moda)\beta^{2}\equiv-b\,\;(\mathrm{mod}\,a), and that aa and bb satisfy Conditions (i)(iv) of Lemma 2.3.

Then π2hol((Ua,bE32)(θa,β+θa,β))\pi^{\mathrm{hol}}_{2}\big{(}(\mathrm{U}_{a,b}\,E_{\frac{3}{2}})\cdot(\theta_{a,\beta}+\theta_{a,-\beta})\big{)} is a quasi-modular form for Γ1(4a)\Gamma_{1}(4a) and

π2hol((Ua,bE32)(θa,β+θa,β))=n=0c(n)e(nτ),\displaystyle\pi^{\mathrm{hol}}_{2}\big{(}(\mathrm{U}_{a,b}\,E_{\frac{3}{2}})\cdot(\theta_{a,\beta}+\theta_{a,-\beta})\big{)}\;=\;\sum_{n=0}^{\infty}c(n)e(n\tau)\text{,}

where

c(napq)2(1+q)(mod)andc(nap2q)2(1+p+q)(mod)\displaystyle c(n_{a}pq)\;\equiv\;-2(1+q)\;\;(\mathrm{mod}\,\ell)\quad\text{and}\quad c(n_{a}p^{2}q)\;\equiv\;-2(1+p+q)\;\;(\mathrm{mod}\,\ell)

for any primes p,q1(mod)p,q\equiv 1\,\;(\mathrm{mod}\,\ell) such that

p2>(ana)q>(ana)2p>(ana)3.\displaystyle p^{2}>(an_{a})q>(an_{a})^{2}p>(an_{a})^{3}\text{.} (2.6)

Remark 2.6.

In light of Lemma 2.3, the assumptions on aa and bb are no restriction. For any a,ba,b with b-b a square modulo aa, we can find a subprogression of a+ba\mathbb{Z}+b satisfying these conditions. Furthermore, for any sufficiently large p1(moda)p\equiv 1\,\;(\mathrm{mod}\,a\ell), there is a prime q1(mod)q\equiv 1\,\;(\mathrm{mod}\,\ell) satisfying the Conditions in (2.6)—this follows from the Prime Number Theorem for arithmetic progressions.

Proof.

As we assume that H(an+b)0(mod)H(an+b)\equiv 0\,\;(\mathrm{mod}\,\ell) for all nn\in\mathbb{Z}, we have

π2hol((Ua,bE32)(θa,β+θa,β))\displaystyle\pi^{\mathrm{hol}}_{2}\big{(}(\mathrm{U}_{a,b}\,E_{\frac{3}{2}})\cdot(\theta_{a,\beta}+\theta_{a,-\beta})\big{)}
=\displaystyle\;=\; (Ua,bD=0H(D)e(Dτ))(θa,β+θa,β)+116ππ2hol((Ua,bθ)(θa,β+θa,β))\displaystyle\big{(}\mathrm{U}_{a,b}\sum_{D=0}^{\infty}H(D)e(D\tau)\big{)}\cdot(\theta_{a,\beta}+\theta_{a,-\beta})\,+\,\frac{1}{16\pi}\pi^{\mathrm{hol}}_{2}\big{(}(\mathrm{U}_{a,b}\,\theta^{\ast})\cdot(\theta_{a,\beta}+\theta_{a,-\beta})\big{)}
\displaystyle\;\equiv\; 116ππ2hol((Ua,bθ)(θa,β+θa,β))(mod).\displaystyle\frac{1}{16\pi}\pi^{\mathrm{hol}}_{2}\big{(}(\mathrm{U}_{a,b}\,\theta^{\ast})\cdot(\theta_{a,\beta}+\theta_{a,-\beta})\big{)}\;\;(\mathrm{mod}\,\ell)\text{.} (2.7)

Now (1.2) and (1.6) lead us to the study of aθa,β\sqrt{a}\,\theta^{\ast}_{a,\beta}, which arise from Ua,bθ\mathrm{U}_{a,b}\,\theta^{\ast}. Write δ\delta_{\bullet} for the Kronecker δ\delta-function. For general β,β~\beta,\tilde{\beta}\in\mathbb{Z}, we use (1.1) to compute that

aθa,β~(τ)θa,β(τ)=\displaystyle\sqrt{a}\,\theta^{\ast}_{a,\tilde{\beta}}(\tau)\cdot\theta_{a,\beta}(\tau)\;=\; 2aδβ~0(moda)mβ(moda)y12e(m2τa)\displaystyle-2\sqrt{a}\delta_{\tilde{\beta}\equiv 0\,\;(\mathrm{mod}\,a)}\,\sum_{m\equiv\beta\,\;(\mathrm{mod}\,a)}y^{-\frac{1}{2}}e\big{(}\frac{m^{2}\tau}{a}\big{)}
2πmβ(moda)m~β~(moda)m~0|m~|Γ(12, 4πm~2ay)e((m2m~2)τa).\displaystyle-2\sqrt{\pi}\sum_{\begin{subarray}{c}m\equiv\beta\,\;(\mathrm{mod}\,a)\\ \tilde{m}\equiv\tilde{\beta}\,\;(\mathrm{mod}\,a)\\ \tilde{m}\neq 0\end{subarray}}|\tilde{m}|\Gamma\big{(}-\tfrac{1}{2},\,4\pi\frac{\tilde{m}^{2}}{a}y\big{)}e\big{(}\frac{(m^{2}-\tilde{m}^{2})\tau}{a}\big{)}\text{.}

We apply Lemmas (2.1) and (2.2) to find that π2hol(aθa,β~(τ)θa,β(τ))\pi^{\mathrm{hol}}_{2}\big{(}\sqrt{a}\,\theta^{\ast}_{a,\tilde{\beta}}(\tau)\cdot\theta_{a,\beta}(\tau)\big{)} is equal to

4π(δβ~0(moda)mβ(moda)m0|m|e(m2τa)+mβ(moda)m~β~(moda)m~0m2m~2|m|+|m~|e((m2m~2)τa)).\displaystyle-4\pi\Big{(}\delta_{\tilde{\beta}\equiv 0\,\;(\mathrm{mod}\,a)}\,\sum_{\begin{subarray}{c}m\equiv\beta\,\;(\mathrm{mod}\,a)\\ m\neq 0\end{subarray}}|m|e\big{(}\frac{m^{2}\tau}{a}\big{)}\,+\,\sum_{\begin{subarray}{c}m\equiv\beta\,\;(\mathrm{mod}\,a)\\ \tilde{m}\equiv\tilde{\beta}\,\;(\mathrm{mod}\,a)\\ \tilde{m}\neq 0\end{subarray}}\frac{m^{2}-\tilde{m}^{2}}{|m|+|\tilde{m}|}e\big{(}\frac{(m^{2}-\tilde{m}^{2})\tau}{a}\big{)}\Big{)}\text{.}

Summarizing, we find that

c(π2hol(aθa,β~(τ)θa,β(τ));n)=4π(δβ~0(moda)mβ(moda)m0an=m2|m|+anmβ(moda)m~β~(moda)m~0an=m2m~21|m|+|m~|).c\Big{(}\pi^{\mathrm{hol}}_{2}\big{(}\sqrt{a}\,\theta^{\ast}_{a,\tilde{\beta}}(\tau)\cdot\theta_{a,\beta}(\tau)\big{)};\,n\Big{)}\\ =\;-4\pi\Big{(}\delta_{\tilde{\beta}\equiv 0\,\;(\mathrm{mod}\,a)}\,\sum_{\begin{subarray}{c}m\equiv\beta\,\;(\mathrm{mod}\,a)\\ m\neq 0\\ an=m^{2}\end{subarray}}|m|\,+\,an\sum_{\begin{subarray}{c}m\equiv\beta\,\;(\mathrm{mod}\,a)\\ \tilde{m}\equiv\tilde{\beta}\,\;(\mathrm{mod}\,a)\\ \tilde{m}\neq 0\\ an=m^{2}-\tilde{m}^{2}\end{subarray}}\frac{1}{|m|+|\tilde{m}|}\Big{)}\text{.} (2.8)

We can drop the first contribution, since b0(moda)-b\not\equiv 0\,\;(\mathrm{mod}\,a) and hence β~0(moda)\tilde{\beta}\not\equiv 0\,\;(\mathrm{mod}\,a). It remains to analyze the second term on the right hand side.

When writing an=m2m~2=(m+m~)(mm~)an=m^{2}-\tilde{m}^{2}=(m+\tilde{m})(m-\tilde{m}) we recognize that the summation runs over factorizations of anan. Assume that an=d1d2an=d_{1}d_{2} is a factorization corresponding to (m+m~)(mm~)(m+\tilde{m})(m-\tilde{m}), then m=(d1+d2)/2m=(d_{1}+d_{2})/2 and m~=(d1d2)/2\tilde{m}=(d_{1}-d_{2})/2. Since an>0an>0, we conclude that d1d_{1} and d2d_{2} have the same sign. We treat only the positive case; the negative case yields the same sum and hence contributes an additional factor of 22 in (2.9).

Since d1,d2>0d_{1},d_{2}>0, we have m>0m>0. If we assume that anan is not a square, then m~0\tilde{m}\neq 0 and the sign of m~\tilde{m} is positive if d1>d2d_{1}>d_{2} and negative if d2>d1d_{2}>d_{1}. As a result we find that |m|+|m~||m|+|\tilde{m}| equals the larger factor in d1d2d_{1}d_{2}. Summarizing, we have

anmβ(moda)m~β~(moda)m~0an=m2m~21|m|+|m~|\displaystyle an\sum_{\begin{subarray}{c}m\equiv\beta\,\;(\mathrm{mod}\,a)\\ \tilde{m}\equiv\tilde{\beta}\,\;(\mathrm{mod}\,a)\\ \tilde{m}\neq 0\\ an=m^{2}-\tilde{m}^{2}\end{subarray}}\frac{1}{|m|+|\tilde{m}|} = 2an=d1d2d1,d2>0d1β+β~(moda)d2ββ~(moda)d1d2d1δd1>d2+d2δd2>d1\displaystyle\;=\;2\sum_{\begin{subarray}{c}an=d_{1}d_{2}\\ d_{1},d_{2}>0\\ d_{1}\equiv\beta+\tilde{\beta}\,\;(\mathrm{mod}\,a)\\ d_{2}\equiv\beta-\tilde{\beta}\,\;(\mathrm{mod}\,a)\end{subarray}}\frac{d_{1}d_{2}}{d_{1}\delta_{d_{1}>d_{2}}+d_{2}\delta_{d_{2}>d_{1}}}
= 2an=d1d2d1,d2>0d1β+β~(moda)d2ββ~(moda)(d1δd1<d2+d2δd2<d1).\displaystyle\;=\;2\sum_{\begin{subarray}{c}an=d_{1}d_{2}\\ d_{1},d_{2}>0\\ d_{1}\equiv\beta+\tilde{\beta}\,\;(\mathrm{mod}\,a)\\ d_{2}\equiv\beta-\tilde{\beta}\,\;(\mathrm{mod}\,a)\end{subarray}}\big{(}d_{1}\delta_{d_{1}<d_{2}}+d_{2}\delta_{d_{2}<d_{1}}\big{)}\text{.} (2.9)

We next want to separate the archimedean and nonarchimedean conditions on the right hand side.

We write n=nann=n_{a}n^{\prime}, where nn^{\prime} is either pqpq or p2qp^{2}q and where pp and qq are as in the statement of the proposition. Then for any factorization n=d1d2n^{\prime}=d^{\prime}_{1}d^{\prime}_{2}, we have d1/d2>anad^{\prime}_{1}/d^{\prime}_{2}>an_{a} or d2/d1>anad^{\prime}_{2}/d^{\prime}_{1}>an_{a}. This assumption ensures that in the resulting factorization of anan only the archimedean condition associated with the factorizaton n=d1d2n^{\prime}=d^{\prime}_{1}d^{\prime}_{2} plays a role. Summarizing, we have

an=d1d2d1,d2>0d1β+β~(moda)d2ββ~(moda)(d1δd1<d2+d2δd2<d1)=anan=d1d2d1,d2>0d1β+β~(moda)d2ββ~(moda)(d1δd1<d2+d2δd1<d2),\displaystyle\sum_{\begin{subarray}{c}an=d_{1}d_{2}\\ d_{1},d_{2}>0\\ d_{1}\equiv\beta+\tilde{\beta}\,\;(\mathrm{mod}\,a)\\ d_{2}\equiv\beta-\tilde{\beta}\,\;(\mathrm{mod}\,a)\end{subarray}}\big{(}d_{1}\delta_{d_{1}<d_{2}}+d_{2}\delta_{d_{2}<d_{1}}\big{)}\;=\;\sum_{\begin{subarray}{c}an_{a}n^{\prime}=d_{1}d_{2}\\ d_{1},d_{2}>0\\ d_{1}\equiv\beta+\tilde{\beta}\,\;(\mathrm{mod}\,a)\\ d_{2}\equiv\beta-\tilde{\beta}\,\;(\mathrm{mod}\,a)\end{subarray}}\big{(}d_{1}\delta_{d^{\prime}_{1}<d^{\prime}_{2}}+d_{2}\delta_{d^{\prime}_{1}<d^{\prime}_{2}}\big{)}\text{,}

where d1=gcd(d1,n)d^{\prime}_{1}=\gcd(d_{1},n^{\prime}) and d2=gcd(d2,n)d^{\prime}_{2}=\gcd(d_{2},n^{\prime}). Since gcd(ana,n)=1\gcd(an_{a},n^{\prime})=1, we can sum over two factorizations ana=da,1da,2an_{a}=d_{a,1}d_{a,2} and n=d1d2n^{\prime}=d^{\prime}_{1}d^{\prime}_{2}. Every factor of nn^{\prime} is congruent to 11 modulo aa, so the congruence condition applies only to the factors of anaan_{a}. We have

anan=d1d2d1,d2>0d1β+β~(moda)d2ββ~(moda)(d1δgcd(d1,n)<gcd(d2,n)+d2δgcd(d1,n)<gcd(d2,n))\displaystyle\hphantom{{}\;=\;{}}\sum_{\begin{subarray}{c}an_{a}n^{\prime}=d_{1}d_{2}\\ d_{1},d_{2}>0\\ d_{1}\equiv\beta+\tilde{\beta}\,\;(\mathrm{mod}\,a)\\ d_{2}\equiv\beta-\tilde{\beta}\,\;(\mathrm{mod}\,a)\end{subarray}}\big{(}d_{1}\delta_{\gcd(d_{1},n^{\prime})<\gcd(d_{2},n^{\prime})}+d_{2}\delta_{\gcd(d_{1},n^{\prime})<\gcd(d_{2},n^{\prime})}\big{)}
=ana=da,1da,2n=d1d2da,1,da,2,d1,d2>0da,1β+β~(moda)da,2ββ~(moda)(da,1d1δd1<d2+da,2d2δd2<d1)\displaystyle\;=\;\sum_{\begin{subarray}{c}an_{a}=d_{a,1}d_{a,2}\\ n^{\prime}=d^{\prime}_{1}d^{\prime}_{2}\\ d_{a,1},d_{a,2},d^{\prime}_{1},d^{\prime}_{2}>0\\ d_{a,1}\equiv\beta+\tilde{\beta}\,\;(\mathrm{mod}\,a)\\ d_{a,2}\equiv\beta-\tilde{\beta}\,\;(\mathrm{mod}\,a)\end{subarray}}\big{(}d_{a,1}d^{\prime}_{1}\delta_{d^{\prime}_{1}<d^{\prime}_{2}}+d_{a,2}d^{\prime}_{2}\delta_{d^{\prime}_{2}<d^{\prime}_{1}}\big{)}
=(ana=d1d2d1,d2>0d1β+β~(moda)d2ββ~(moda)d1+d2)(n=d1d2d1,d2>0d1δd1<d2+d2δd2<d1).\displaystyle\;=\;\Big{(}\sum_{\begin{subarray}{c}an_{a}=d_{1}d_{2}\\ d_{1},d_{2}>0\\ d_{1}\equiv\beta+\tilde{\beta}\,\;(\mathrm{mod}\,a)\\ d_{2}\equiv\beta-\tilde{\beta}\,\;(\mathrm{mod}\,a)\end{subarray}}d_{1}+d_{2}\Big{)}\,\cdot\,\Big{(}\sum_{\begin{subarray}{c}n^{\prime}=d_{1}d_{2}\\ d_{1},d_{2}>0\end{subarray}}d_{1}\delta_{d_{1}<d_{2}}+d_{2}\delta_{d_{2}<d_{1}}\Big{)}\text{.} (2.10)

By Lemma 2.4, the first factor in (2.10) is congruent to 22 modulo \ell.

We next inspect the second factor in (2.10). We have

qp=d1d2d1,d2>0(d1δd1<d2+d2δd2<d1)=(1+1)(1+q)=2(1+q)qp2=d1d2d1,d2>0(d1δd1<d2+d2δd2<d1)=(1+1)(1+q)+(p+p)=2(1+q+p).\displaystyle\begin{aligned} \sum_{\begin{subarray}{c}qp=d_{1}d_{2}\\ d_{1},d_{2}>0\end{subarray}}\big{(}d_{1}\delta_{d_{1}<d_{2}}+d_{2}\delta_{d_{2}<d_{1}}\big{)}&\;=\;(1+1)(1+q)=2(1+q)\\ \sum_{\begin{subarray}{c}qp^{2}=d_{1}d_{2}\\ d_{1},d_{2}>0\end{subarray}}\big{(}d_{1}\delta_{d_{1}<d_{2}}+d_{2}\delta_{d_{2}<d_{1}}\big{)}&\;=\;(1+1)(1+q)+(p+p)=2(1+q+p)\text{.}\end{aligned} (2.11)

We now combine the archimedean and nonarchimedean factors in (2.10) to determine the Fourier coefficients of (2.7). Our final expression in (2.12) receives several contributions: 1/16π1/16\pi from (2.7); 4π-4\pi from (2.8); 22 from (2.9); 22 from Lemma 2.4, computing the first factor in (2.10); and 2(1+p)2(1+p) or 2(1+q+p)2(1+q+p) from (2.11), computing the second factor in (2.10). This yields

c(π2hol(Ua,bE32(θa,β+θa,β));naqp)2(1+q)(mod),c(π2hol(Ua,bE32(θa,β+θa,β));naqp2)2(1+q+p)(mod).\displaystyle\begin{aligned} c\big{(}\pi^{\mathrm{hol}}_{2}\big{(}\mathrm{U}_{a,b}\,E_{\frac{3}{2}}\cdot(\theta_{a,\beta}+\theta_{a,-\beta})\big{)};\,&n_{a}qp\big{)}&&\;\equiv\;-2(1+q)&&\quad\;(\mathrm{mod}\,\ell)\text{,}\\ c\big{(}\pi^{\mathrm{hol}}_{2}\big{(}\mathrm{U}_{a,b}\,E_{\frac{3}{2}}\cdot(\theta_{a,\beta}+\theta_{a,-\beta})\big{)};\,&n_{a}qp^{2}\big{)}&&\;\equiv\;-2(1+q+p)&&\quad\;(\mathrm{mod}\,\ell)\text{.}\end{aligned} (2.12)

We are now in a position to apply Theorem 1.1 and its Corollary 1.2.

Proof of Theorem A.

We are now in position to apply Theorem 1.1 to deduce a contradiction. Replacing aa and bb with aaaa^{\prime} and babb-ab^{\prime}, where aa^{\prime} and bb^{\prime} are as in Lemma 2.3, we may assume that aa and bb satisfy the conditions of Proposition 2.5.

Theorem 1.1 asserts that there are infinitely many primes p1(moda)p\equiv 1\,\;(\mathrm{mod}\,a\ell) such that c(f;npr)(r+1)c(f;n)(mod)c(f;\,np^{r})\equiv(r+1)c(f;\,n)\,\;(\mathrm{mod}\,\ell) for all nn\in\mathbb{Z} co-prime to pp, r0r\in\mathbb{Z}_{\geq 0}, and fM2(Γ1(4a))f\in\mathrm{M}_{2}(\Gamma_{1}(4a)). For sufficiently large pp there is a prime qq with q1(mod)q\not\equiv 1\,\;(\mathrm{mod}\,\ell) and q1(moda)q\equiv 1\,\;(\mathrm{mod}\,a) that satisfies (2.6) (see Remark 2.6).

If the qq-th Fourier coefficient of π2hol(Ua,bE32(θa,β+θa,β))\pi^{\mathrm{hol}}_{2}\big{(}\mathrm{U}_{a,b}\,E_{\frac{3}{2}}\cdot(\theta_{a,\beta}+\theta_{a,-\beta})\big{)} is divisible by \ell, then the two congruences in (2.12) yield the contradiction 1+q1+q+p0(mod)1+q\equiv 1+q+p\equiv 0\,\;(\mathrm{mod}\,\ell). Otherwise, they incur the relation

3(1+q)2(1+q+p)(mod),\displaystyle 3(1+q)\equiv 2(1+q+p)\;\;(\mathrm{mod}\,\ell)\text{,}

which is equivalent to q1(mod)q\equiv 1\,\;(\mathrm{mod}\,\ell), a contradiction.

Appendix A Hecke-type congruences

The fact that DH(D)e(Dτ)\sum_{D}H(D)e(D\tau) is a Hecke eigenform is conveniently captured by the Hurwitz class number formula. For fundamental discriminants D-D and positive integers ff, the formulas for class numbers of imaginary quadratic fields and H(D)H(D) (see for example, pages 228 and 230 of [7]) imply the following:

H(Df2)=H(D)w(Df2)w(D)dfdpd(1\mfrac1p(\mfracDp))\displaystyle H(Df^{2})\;=\;H(D)\frac{w(-Df^{2})}{w(-D)}\,\sum_{d\mathop{\mid}f}d\prod_{p\mathop{\mid}d}\Big{(}1-\mfrac{1}{p}\left(\mfrac{-D}{p}\right)\Big{)}

where the product is over primes pp dividing dd and where w(D)6w(-D)\mathop{\mid}6 is the number of roots of unity in the quadratic order of discriminant D-D. Throughout the paper we follow Zagier [33], who defines H(D)H(D) for nonnegative arguments, while Cohen [7] uses the opposite sign convention. We restrict to congruences modulo powers of primes >3\ell>3 and hence may ignore the factor w(Df2)/w(D)w(-Df^{2})/w(-D). We assume no further knowledge of H(D)H(D), and we only employ the sum over divisors dd of ff in (A) to obtain congruences for H(an+b)H(an+b). Note that this sum is multiplicative in ff, which later allows us to restrict to the case of prime powers aa.

From the introduction recall the congruences

H(53n+25)0(mod 5),H(73n+147)0(mod 7),H(113n+242)0(mod 11).\displaystyle H(5^{3}n+25)\equiv 0\;\;(\mathrm{mod}\,5)\text{,}\quad H(7^{3}n+147)\equiv 0\;\;(\mathrm{mod}\,7)\text{,}\quad H(11^{3}n+242)\equiv 0\;\;(\mathrm{mod}\,11)\text{.}

Observe that for 53n+25=(5n+1)52=Df25^{3}n+25=(5n+1)5^{2}=Df^{2} with a fundamental discriminant D-D we have D1(mod 5)D\equiv 1\,\;(\mathrm{mod}\,5) and 5f5\mathop{\mid}f. Thus, writing f5ff_{5}\mathop{\mid}f for the highest power of 55 that divides ff, we have f51f_{5}\neq 1 and the right hand side of (A) has the factor

df5d5d(1\mfrac15(\mfracD5))1(\mfracD5)0(mod 5),\displaystyle\sum_{d\mathop{\mid}f_{5}}d\prod_{5\mathop{\mid}d}\Big{(}1-\mfrac{1}{5}\left(\mfrac{-D}{5}\right)\Big{)}\equiv 1\,-\,\left(\mfrac{-D}{5}\right)\equiv 0\;\;(\mathrm{mod}\,5)\text{,}

which shows that H(53n+25)0(mod 5)H(5^{3}n+25)\equiv 0\;\;(\mathrm{mod}\,5). The congruences modulo 77 and 1111, and the congruences in the remark after Theorem A follow similarly.

The above reasoning extends to powers m\ell^{m} for arbitrary primes \ell as follows: Assume that a=ea=\ell^{e} and b=cub=\ell^{c}u for an integer uu with gcd(,u)=1\gcd(\ell,u)=1. For simplicity, we further suppose that e>c2e>c\geq 2. Set c=c/21c^{\prime}=\lfloor c/2\rfloor\geq 1 and c′′=min{c,m}c^{\prime\prime}=\min\{c^{\prime},m\}. We have the factorization

an+b=en+cu=2cc2c(ec+u)=f2D,\displaystyle an+b=\ell^{e}n+\ell^{c}u=\ell^{2c^{\prime}}\,\ell^{c-2c^{\prime}}\big{(}\ell^{e-c}+u\big{)}=f^{2}D\text{,}

which yields cf\ell^{c^{\prime}}\mathop{\mid}f and D0(mod)D\equiv 0\,\;(\mathrm{mod}\,\ell) if cc is odd and Du(mod)D\equiv u\,\;(\mathrm{mod}\,\ell) if cc is even. Thus, the right hand side of (A) has the factor

dcdd(1\mfrac1(\mfracD))=1+n=1cn(1\mfrac1(\mfracD))1+n=1c′′n1((\mfracD))=σ1(c′′1)(1(\mfracD))+c′′(modm).\sum_{d\mathop{\mid}\ell^{c^{\prime}}}d\prod_{\ell\mathop{\mid}d}\Big{(}1-\mfrac{1}{\ell}\left(\mfrac{-D}{\ell}\right)\Big{)}=1\,+\,\sum_{n=1}^{c^{\prime}}\ell^{n}\Big{(}1-\mfrac{1}{\ell}\left(\mfrac{-D}{\ell}\right)\Big{)}\equiv 1+\sum_{n=1}^{c^{\prime\prime}}\ell^{n-1}\Big{(}\ell-\left(\mfrac{-D}{\ell}\right)\Big{)}\\ =\sigma_{1}\big{(}\ell^{c^{\prime\prime}-1}\big{)}\Big{(}1-\left(\mfrac{-D}{\ell}\right)\Big{)}+\ell^{c^{\prime\prime}}\;\;(\mathrm{mod}\,\ell^{m})\text{.}

As a result, we find congruences modulo m\ell^{m} if c2mc\geq 2m (and hence c′′mc^{\prime\prime}\geq m) is even and u-u is a square modulo \ell. In particular, we obtain non-holomorphic Ramanujan-type congruences for all primes >3\ell>3.

References

  • [1] Scott Ahlgren, Kathrin Bringmann and Jeremy Lovejoy \ell-adic properties of smallest parts functions” In Adv. Math. 228.1, 2011, pp. 629–645
  • [2] Scott Ahlgren and Ken Ono “Congruence properties for the partition function” In Proc. Natl. Acad. Sci. USA 98.23, 2001, pp. 12882–12884
  • [3] Nickolas Andersen “Classification of congruences for mock theta functions and weakly holomorphic modular forms” In Q. J. Math. 65.3, 2014, pp. 781–805
  • [4] George E. Andrews, Donny Passary, James A. Sellers and Ae Ja Yee “Congruences related to the Ramanujan/Watson mock theta functions ω(q)\omega(q) and ν(q)\nu(q) In Ramanujan J. 43.2, 2017, pp. 347–357
  • [5] Kathrin Bringmann, Amanda Folsom, Ken Ono and Larry Rolen “Harmonic Maass forms and mock modular forms: theory and applications” 64, American Mathematical Society Colloquium Publications American Mathematical Society, Providence, RI, 2017, pp. xv+391
  • [6] Kathrin Bringmann and Jeremy Lovejoy “Overpartitions and class numbers of binary quadratic forms” In Proc. Natl. Acad. Sci. USA 106.14, 2009, pp. 5513–5516
  • [7] Henri Cohen “A course in computational algebraic number theory” 138, Graduate Texts in Mathematics Springer-Verlag, Berlin, 1993, pp. xii+534
  • [8] Henri Cohen “Sums involving the values at negative integers of LL-functions of quadratic characters” In Math. Ann. 217.3, 1975, pp. 271–285
  • [9] Henri Cohen and Hendrik W. Lenstra “Heuristics on class groups of number fields” In Number theory, Noordwijkerhout 1983 (Noordwijkerhout, 1983) 1068, Lecture Notes in Math. Springer, Berlin, 1984, pp. 33–62
  • [10] Pierre Deligne and Jean-Pierre Serre “Formes modulaires de poids 11 In Ann. Sci. École Norm. Sup. (4) 7, 1974, pp. 507–530 (1975)
  • [11] W. Duke “Almost a century of answering the question: what is a mock theta function?” In Notices Amer. Math. Soc. 61.11, 2014, pp. 1314–1320
  • [12] Israil S. Gradshteyn and Iosif M. Ryzhik “Table of integrals, series, and products” Elsevier/Academic Press, Amsterdam, 2007, pp. xlviii+1171
  • [13] Benedict H. Gross and Don B. Zagier “Heegner points and derivatives of LL-series” In Invent. Math. 84.2, 1986, pp. 225–320
  • [14] William Hart, Fredrik Johansson and Sebastian Pancratz “FLINT: Fast Library for Number Theory” http://flintlib.org, 2015
  • [15] Samuel Holmin, Nathan Conrad Jones, Pär Kurlberg, Cameron McLeman and Kathleen L. Petersen “Missing class groups and class number statistics for imaginary quadratic fields” In Exp. Math. 28.2, 2019, pp. 233–254
  • [16] Özlem Imamoğlu, Martin Raum and Olav K. Richter “Holomorphic projections and Ramanujan’s mock theta functions” In Proc. Natl. Acad. Sci. USA 111.11, 2014, pp. 3961–3967
  • [17] Naomi Jochnowitz “Congruences between modular forms of half integral weights and implications for class numbers and elliptic curves”, preprint
  • [18] Masanobu Kaneko and Don Zagier “A generalized Jacobi theta function and quasimodular forms” In The moduli space of curves (Texel Island, 1994) 129, Progr. Math. Birkhäuser Boston, Boston, MA, 1995, pp. 165–172
  • [19] Serge Lang “Introduction to modular forms” With appendixes by D. Zagier and Walter Feit, Corrected reprint of the 1976 original 222, Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences] Springer-Verlag, Berlin, 1995, pp. x+261
  • [20] “NIST Digital Library of Mathematical Functions” Version 1.0.27, http://dlmf.nist.gov/, 2020
  • [21] Ken Ono “Congruences for the Andrews spt function” In Proc. Natl. Acad. Sci. USA 108.2, 2011, pp. 473–476
  • [22] Ken Ono “Distribution of the partition function modulo mm In Ann. of Math. (2) 151.1, 2000, pp. 293–307
  • [23] Ken Ono “The web of modularity: arithmetic of the coefficients of modular forms and qq-series” 102, CBMS Regional Conference Series in Mathematics Published for the Conference Board of the Mathematical Sciences, Washington, DC; by the American Mathematical Society, Providence, RI, 2004, pp. viii+216
  • [24] Ken Ono “Unearthing the visions of a master: harmonic Maass forms and number theory” In Current developments in mathematics, 2008 Int. Press, Somerville, MA, 2009, pp. 347–454
  • [25] Cristian-Silviu Radu “Proof of a conjecture by Ahlgren and Ono on the non-existence of certain partition congruences” In Trans. Amer. Math. Soc. 365.9, 2013, pp. 4881–4894
  • [26] Srinivasa Ramanujan “Congruence properties of partitions.” In Proc. Lond. Math. Soc. (2) 18, 1920, pp. xix–xx
  • [27] Arnold Scholz “Über die Beziehung der Klassenzahlen quadratischer Körper zueinander” In J. Reine Angew. Math. 166, 1932, pp. 201–203
  • [28] Jean-Pierre Serre “Divisibilité de certaines fonctions arithmétiques” In Enseign. Math. (2) 22.3-4, 1976, pp. 227–260
  • [29] Jean-Pierre Serre “Divisibilité des coefficients des formes modulaires de poids entier” In C. R. Acad. Sci. Paris Sér. A 279, 1974, pp. 679–682
  • [30] Jacob Sturm “Projections of CC^{\infty} automorphic forms” In Bull. Amer. Math. Soc. (N.S.) 2.3, 1980, pp. 435–439
  • [31] Stephanie Treneer “Congruences for the coefficients of weakly holomorphic modular forms” In Proc. London Math. Soc. (3) 93.2, 2006, pp. 304–324
  • [32] Don Zagier “Modular forms and differential operators” K. G. Ramanathan memorial issue In Proc. Indian Acad. Sci. Math. Sci. 104.1, 1994, pp. 57–75
  • [33] Don B. Zagier “Nombres de classes et formes modulaires de poids 3/23/2 In C. R. Acad. Sci. Paris Sér. A-B 281.21, 1975, pp. Ai\bibrangessepA883–A886

 

Olivia Beckwith Department of Mathematics, University of Illinois at Urbana-Champaign, Urbana, IL 61801, USA E-mail: [email protected] Homepage: https://sites.google.com/view/olivia-beckwith-homepage/home

Martin Raum Chalmers tekniska högskola och Göteborgs Universitet, Institutionen för Matematiska vetenskaper, SE-412 96 Göteborg, Sweden E-mail: [email protected] Homepage: http://raum-brothers.eu/martin

Olav K. Richter Department of Mathematics, University of North Texas, Denton, TX 76203, USA E-mail: [email protected] Homepage: http://www.math.unt.edu/~richter/