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

Counting modular forms by rationality field

Alex Cowan Department of Mathematics, Harvard University, Cambridge, MA 02138 USA [email protected]  and  Kimball Martin Department of Mathematics, University of Oklahoma, Norman, OK 73019 USA [email protected]
Abstract.

We investigate the distribution of degrees and rationality fields of weight 2 newforms. In particular, we give heuristic upper bounds on how often degree dd rationality fields occur for squarefree levels, and predict finiteness if d7d\geq 7. When d=2d=2, we make predictions about how frequently specific quadratic fields occur, prove lower bounds, and conjecture that (5)\mathbb{Q}(\sqrt{5}) is the most common quadratic rationality field.

AC was supported by the Simons Foundation (Collaboration Grant 550031).
KM was supported by the Simons Foundation (Collaboration Grant 512927), the Japan Society for the Promotion of Science (Invitational Fellowship L22540), and the Osaka Central Advanced Mathematical Institute (MEXT Joint Usage/Research Center on Mathematics and Theoretical Physics JPMXP0619217849).

1. Introduction

Let Sk(N)=Sk(Γ0(N))S_{k}(N)=S_{k}(\Gamma_{0}(N)) be the space of holomorphic cusp forms of weight kk and level NN with trivial nebentypus. For a newform fSk(N)f\in S_{k}(N), denote by KfK_{f} its rationality field, i.e. the number field generated by its Hecke eigenvalues. Define the (rationality) degree and discriminant of ff to be the degree and discriminant of Kf/K_{f}/\mathbb{Q} respectively. In what follows, we always assume trivial nebentypus.

Weight 2 newforms are of special interest as they correspond to modular abelian varieties, i.e., simple factors of the Jacobian J0(N)J_{0}(N) of X0(N)X_{0}(N). Namely (the Galois orbit of) a degree dd newform fS2(N)f\in S_{2}(N) corresponds to a dd-dimensional simple abelian subvariety of J0(N)J_{0}(N) which has conductor NdN^{d}. The d=1d=1 case is the celebrated bijection between rational newforms in S2(N)S_{2}(N) and isogeny classes of elliptic curves of conductor NN.

It is expected (e.g., see [35, 23, 28]) that the Galois orbit of a newform is “as large as possible” 100% of the time, so that newforms have small degree rather infrequently. On the other hand, there are a relatively large number of elliptic curves of small conductor. Watkins’ [39] refinement of the Brumer–McGuinness heuristics [7] for counting elliptic curves suggests that the number of weight 2 rational newforms with level N<XN<X grows like cX5/6cX^{5/6} for some computable constant cc. See [12, 36] for some theoretical results towards this growth rate. Note that the total number of weight 2 newforms of level N<XN<X grows roughly like X2X^{2}.

Here we consider the questions: how many weight 2 newforms of level N<XN<X are there with a given degree dd or a given rationality field KK? There is no analogue of the Brumer–McGuinness heuristics for d>1d>1, since those rely on having simple equations for elliptic curves. Moreover, as degree d>1d>1 forms are relatively rare, it is difficult to generate enough data to predict precise asymptotics based on calculations.

In fact, even for d=1d=1, it is difficult to make accurate predictions based solely on computations. E.g., as remarked in [39], the growth rate in Cremona’s database of elliptic curves is about X0.98X^{0.98}; however more recent and very extensive calculations for prime conductors in [2] align closely with the X5/6X^{5/6} heuristic.

Using a combination of heuristics and data, we predict some bounds on asymptotic orders of growth, and the relative frequency of such forms.

Conjecture 1.1.

Let ε>0\varepsilon>0. The number of degree dd weight 22 newforms of squarefree level NXN\leq X is O(X1d/6+ε)O(X^{1-d/6+\varepsilon}) as XX\to\infty. In particular, this number is finite if d7d\geq 7.

Conjecture 1.2.

Among squarefree levels NN\to\infty, 100% of degree 22 newforms in S2(N)S_{2}(N) have rationality field (5)\mathbb{Q}(\sqrt{5}).

Remark 1.3.

The heuristics for these conjectures do not require a restriction to squarefree levels, however there are special considerations for non-squarefree levels. First, one should only count quadratic twist classes for a more general analogue of 1.1. Second, CM forms (which do not occur in squarefree level with trivial nebentypus) deserve separate consideration. Third, if prNp^{r}\mid N for sufficiently large rr, then the rationality field of a newform ff of level NN must contain a certain cyclotomic subfield (e.g., if p5p\geq 5 and r3r\geq 3, then Kf(ζp)+K_{f}\supset\mathbb{Q}(\zeta_{p})^{+})—see [6, 26].

It is at least plausible that 1.2 holds for general levels, and 1.1 holds for general levels if one restricts to counting non-CM newforms up to quadratic twist. However, our data are much more limited for non-squarefree levels.

1.1 is just a conjectural upper bound, and it may not be sharp for 2d52\leq d\leq 5 (see below for more discussion). When d=1d=1, one can prove a lower bound of order X5/6X^{5/6} for elliptic curves, but we are not aware of nontrivial analogous lower bounds (or even a proof of infinitude!) for any d>1d>1. Using constructions of genus 2 curves with real multiplication, we obtain the following lower bounds for d=2d=2, without a restriction to squarefree level.

Proposition 1.4.

The number of quadratic twist classes of weight 22 newforms with rationality field (5)\mathbb{Q}(\sqrt{5}) (resp. (2)\mathbb{Q}(\sqrt{2})) and minimal level N<XN<X is X1/3\gg X^{1/3} (resp. X2/7\gg X^{2/7}).

The same result for squarefree levels would follow if one knew certain polynomials took on squarefree values sufficiently often.

Remark 1.5.

It is not even clear for which d6d\leq 6 there should exist infinitely many weight 2 newforms of squarefree level. Constructions of genus 3 curves with real multiplication suggest it may be infinite for d=3d=3—see Section 3.4. For d=4,5,6d=4,5,6, we have little theoretical evidence, but our data suggest these counts are infinite at least for each d4d\leq 4.

We will consider two approaches to predicting counts of newforms with fixed degree or rationality field. First, in Section 2, we present a heuristic using a random model for Hecke polynomials, building off of [33, 28]. In fact this random model naively suggests upper and lower bounds for counts of degree dd forms on the order of X1d/6±εX^{1-d/6\pm\varepsilon}. However, it ignores any geometric considerations for the existence of degree dd forms, so it is unclear how accurate this heuristic is. Nevertheless, comparing these predictions with data at least suggests it gives an upper bound, as asserted in 1.1.

In Section 3, we suggest an approach to predict counts of weight 2 newforms with a given rationality field KK by counting moduli points for suitable abelian varieties. In principle, this would also yield the number of counts of newforms of a fixed degree dd, and we expect this approach should give more accurate predictions than the random Hecke polynomial model. However, it requires more knowledge about the moduli spaces and the relation between heights and conductors than we currently possess. We carry out some of this analysis when d=2d=2, namely when K=(D)K=\mathbb{Q}(\sqrt{D}) for D=5,8,12,13,17D=5,8,12,13,17. This leads to 1.2, and also suggests cX3/5εcX^{3/5-\varepsilon} may be a lower bound for the total count for degree 2 newforms. However, our analysis is not definitive enough to confidently conjecture this.

A database of all prime-level forms of degree 66 or less and level 21062\cdot 10^{6} or less was computed using an algorithm of the first author [8]. In Section 4, we use this database to investigate 1.1 and 1.2, and pose some related questions.

Acknowledgements

We are especially grateful to Noam Elkies for many insights and suggestions. We have also benefited from conversations with Eran Assaf, Armand Brumer, Bjorn Poonen, Ari Shnidman, and John Voight. Computations were performed at the OU Supercomputing Center for Education & Research (OSCER) at the University of Oklahoma (OU).

2. Counts by degree

First we discuss counting newforms of fixed degree. For a newform ff, let degf=[Kf:]\deg f=[K_{f}:\mathbb{Q}] be its rationality degree. Set

𝒞d(X)=#{newforms fS2(N):N<X,N squarefree,degf=d}.\displaystyle\mathcal{C}_{d}(X)=\#\{\text{newforms }f\in S_{2}(N):\,N<X,\,N\text{ squarefree},\,\deg f=d\}.

As explained in the introduction, we restrict to squarefree NN for simplicity, though our initial discussion applies equally well to counting quadratic twist classes of non-CM weight 2 newforms.

Watkins [39], building on heuristics of Brumer and McGuinness [7], formulates heuristics that suggest

(1) 𝒞1(X)c1X56\displaystyle\mathcal{C}_{1}(X)\sim c_{1}X^{\frac{5}{6}}

for some computable constant c1c_{1}. It is known that X56𝒞1(X)X1+εX^{\frac{5}{6}}\ll\mathcal{C}_{1}(X)\ll X^{1+\varepsilon} [12]. Furthermore, Shankar–Shankar–Wang [36] show a growth rate of b1X56b_{1}X^{\frac{5}{6}} if one restricts to elliptic curves of squarefree conductor coprime to 6 with some restrictions on discriminant-conductor and discriminant-height ratios.

While the exponent 56\frac{5}{6} is not in clear agreement with databases of elliptic curves in general levels (the Cremona [10] and Stein–Watkins [37] databases), the compatibility with prime-level data is much better. Namely, Watkins’ heuristic suggests a growth rate of c1li(X56)c_{1}^{\prime}\textrm{li}(X^{\frac{5}{6}}) for prime levels, and this fits extremely well with the extensive database of elliptic curves of prime conductor in [2]. Thus there is much evidence towards (1).

For d>1d>1, the situation is much more mysterious. In [35], Serre proves a statement which strongly suggests, though does not quite imply, the bound 𝒞d(X)=o(X2)\mathcal{C}_{d}(X)=o(X^{2}). Namely, if NN\to\infty along a sequence which is coprime to a fixed prime \ell, among bases of eigenforms for S2(N)S_{2}(N), Serre proves that the number of forms of degree dd is o(dim(S2(N)))o(\dim(S_{2}(N))) as NN\to\infty. Serre’s theorem was made effective by Murty and Sinha [31], and more recently by Sarnak and Zubrilina [34].

Since we do not know a good way to predict precise asymptotics for 𝒞d(X)\mathcal{C}_{d}(X), we aim to predict weaker estimates of the form

(2) Xαd𝒞d(X)Xβd\displaystyle X^{\alpha_{d}}\ll\mathcal{C}_{d}(X)\ll X^{\beta_{d}}

which are nontrivial, i.e., αd>0\alpha_{d}>0 or βd<2\beta_{d}<2. Computations of modular forms, as well as heuristics in [28], suggest βd\beta_{d} is decreasing in dd, and thus we should at least be able to take βd56\beta_{d}\leq\frac{5}{6} for each d1d\geq 1. In [28, Question 3.1], it was also suggested that one may have βd=0\beta_{d}=0 for d0d\gg 0.

To our knowledge, 1.1 is the first prediction of more precise upper bounds (for either squarefree or general levels). In particular, it predicts that one can take βd=0\beta_{d}=0 for d7d\geq 7, and βd\beta_{d} arbitrarily small for d=6d=6. However, we do not have insight into whether the upper bounds in 1.1 should be sharp for 2d52\leq d\leq 5.

Note that 1.1 implies that αd=0\alpha_{d}=0 is optimal among lower bounds of the form (2) for d6d\geq 6. In addition, Proposition 1.4 suggests that one may take α213\alpha_{2}\geq\frac{1}{3} for d=2d=2. (Note that Proposition 1.4 does not prove a lower bound for squarefree levels, only for general levels.) This lower bound is almost certainly not sharp. We do not have any predictions for lower bounds when 3d53\leq d\leq 5.

2.1. Random Hecke polynomial model

Here we present a random model to estimate the distribution of degree dd newforms that will lead us to 1.1. This is based on ideas for heuristics suggested in [33] and [28].

Consider a newspace S2knew(N)S_{2k}^{\mathrm{new}}(N). One can further decompose this space into 2ω(N)2^{\omega(N)} joint eigenspaces of the Atkin–Lehner operators WpW_{p} for pNp\mid N, which we call the Atkin–Lehner eigenspaces. Each Atkin–Lehner eigenspace is Galois invariant. For non-squarefree levels, one can further decompose each Atkin–Lehner eigenspace into smaller Galois invariant subspaces according to local inertia types of non-CM forms (see [11]) and the subspace of CM forms.

For simplicity, assume NN is squarefree. Then there are no CM forms of trivial nebentypus and there is only one local inertial type. Let SS be an Atkin–Lehner eigenspace in S2knew(N)S_{2k}^{\mathrm{new}}(N). For a newform fSf\in S, the single Fourier coefficient ap(f)a_{p}(f) generates KfK_{f} for 100% of pp [22], and it is conjectured that this is true for all but finitely many pp if [Kf:]>4[K_{f}:\mathbb{Q}]>4 [32]. Hence, for fixed pNp\nmid N, the factorization type of the characteristic polynomial cTp(x)[x]c_{T_{p}}(x)\in\mathbb{Z}[x] of the Hecke operator TpT_{p} will usually tell us the degrees of the newforms in SS. In fact, it will always give us lower bounds.

Let n=dimSn=\dim S. As in [33] and [28], we can model cTp(x)c_{T_{p}}(x) as a random polynomial in the set Hn=Hn(k,p)H_{n}=H_{n}(k,p) of degree nn monic integral polynomials whose roots α\alpha satisfy |α|2pk1/2|\alpha|\leq 2p^{k-1/2}. Alternatively, one can consider the set of Weil qq-polynomials of degree 2n2n where q=pkq=p^{k}, or the isogeny classes of nn-dimensional abelian varieties over 𝔽pk\mathbb{F}_{p^{k}}.

Set h(n)=#Hnh(n)=\#H_{n}. As discussed in [28, §2.1], the number of polynomials in HnH_{n} with a degree d<n2d<\tfrac{n}{2} factor is approximately h(d)h(nd)h(d)h(n-d). Thus, if we select polynomials in HnH_{n} uniformly at random, then

(3) Prob(pHn has a degree d factor)h(d)h(nd)h(n).\displaystyle\mathrm{Prob}\!\left(p\in H_{n}\text{ has a degree $d$ factor}\right)\approx\frac{h(d)h(n-d)}{h(n)}.

In this section, by approximately (\approx), we mean that for fixed dd both sides have the same growth rate in nn as nn\to\infty.

For fixed qq, no good asymptotics are known for h(n)h(n) to directly estimate this probability. There is an asymptotic for h(n)h(n) when nn is fixed and qq varies. Instead, [33] and [28, §2.1] analyzed how this probability behaves if one uses a known asymptotic for #Hn(k,p)\#H_{n}(k,p) in pkp^{k} when nn is fixed. (The result is certainly too small, as it would predict only finitely many degree 1 forms.)

To circumvent this lack of precise asymptotics for h(n)h(n) as nn\to\infty, we rewrite the right hand side of (3) as

h(d)h(nd)h(n)\displaystyle\frac{h(d)h(n-d)}{h(n)} =h(d)h(nd)h(nd+1)h(nd+1)h(nd+2)h(n1)h(n).\displaystyle=h(d)\frac{h(n-d)}{h(n-d+1)}\frac{h(n-d+1)}{h(n-d+2)}\cdots\frac{h(n-1)}{h(n)}.

One should have that h(n2)h(n1)h(n1)h(n)\frac{h(n-2)}{h(n-1)}\approx\frac{h(n-1)}{h(n)}, so applying this a small fixed number of times for a given dd yields

(4) h(d)h(nd)h(n)(h(n1)h(n))d.\frac{h(d)h(n-d)}{h(n)}\approx\left(\frac{h(n-1)}{h(n)}\right)^{d}.

Combining (3) and (4) suggests that the probability of a degree dd factor of cTpc_{T_{p}} should approximately be the dd-th power of the probability of a degree 11 factor of cTpc_{T_{p}}. The latter typically corresponds to a degree 1 form, and so we can model it using well-known expectations about counts of elliptic curves.

We will also use the following lemma.

Lemma 2.1.

Let ν2k(X)\nu_{2k}(X) be the number of Atkin–Lehner eigenspaces in NS2knew(N)\bigcup_{N}S_{2k}^{\mathrm{new}}(N), where NN ranges over squarefree levels, having dimension less than XX. Then Xν2k(X)X1+εX\ll\nu_{2k}(X)\ll X^{1+\varepsilon}, for any ε>0\varepsilon>0.

Proof.

Since dimS2knew(N)N\dim S_{2k}^{\mathrm{new}}(N)\ll N, the lower bound is obvious.

Let us show the upper bound. First, it follows from the dimension formulas for Atkin–Lehner eigenspaces from [27] that any Atkin–Lehner eigenspace in S2knew(N)S_{2k}^{\mathrm{new}}(N) has dimension (k1)ϕ(N)122ω(N)+O(N1/2+ε)\frac{(k-1)\phi(N)}{12\cdot 2^{\omega(N)}}+O(N^{1/2+\varepsilon}). (The necessary argument, though not the statement, is given in the proof of [29, Proposition 3.10].) This dimension is N1ε\gg N^{1-\varepsilon}. Hence if an Atkin–Lehner eigenspace has dimension less than XX, it occurs in a level NX1+εN\ll X^{1+\varepsilon}. Now the number of Atkin–Lehner eigenspaces in levels less than tt is bounded by Nt2ω(N)\sum_{N\leq t}2^{\omega(N)}. It is known that this latter sum is 6π2tlogt+O(t)\frac{6}{\pi^{2}}t\log t+O(t). ∎

Lemma 2.1 combines with (3) and (4) to produce the following heuristic.

Heuristic 2.2.

Suppose the number of rational newforms of weight 2k2k and squarefree level N<XN<X is O(X1α)O(X^{1-\alpha}) for some α<1\alpha<1. Then, for any ε>0\varepsilon>0, the number of degree dd weight 2k2k newforms of squarefree level NXN\leq X is O(X1αd+ε)O(X^{1-\alpha d+\varepsilon}) as XX\to\infty.

Our reasoning for this heuristic is as follows. Under the hypothetical bound O(X1α)O(X^{1-\alpha}), the lemma indicates that the probability of an Atkin–Lehner space of dimension nn having a size 1 Galois orbit is approximately nαn^{-\alpha}. Assuming uniform distribution of Hecke polynomials in H(n)H(n), the probability of a size dd Galois orbit is approximately the probability of a degree dd factor of cTpc_{T_{p}}, which by (3) and (4) is approximately ndαn^{-d\alpha}. Applying the lemma again leads to the stated heuristic.

Combining the Brumer–McGuinness and Watkins heuristics for d=1d=1 with 2.2 now suggests 1.1 from the introduction.

2.2. Assessment of the model

The random model for Hecke polynomials in Section 2.1 uses the counting measure on HnH_{n}, i.e., all polynomials in HnH_{n} are equally likely. If this were the case, the heuristic reasoning above would suggest both upper and lower bounds: X1d6ε𝒞d(X)X1d6+εX^{1-\frac{d}{6}-\varepsilon}\ll\mathcal{C}_{d}(X)\ll X^{1-\frac{d}{6}+\varepsilon}, so the upper bound in 1.1 would be essentially optimal.

However, there are other factors controlling the distribution of Hecke polynomials in HnH_{n}. For instance, trace formulas place arithmetic conditions on the roots of Hecke polynomials. Moreover, there are vertical and horizontal equidistribution results about convergence of the roots to Plancherel and Sato–Tate measures. For this reason, we view our random model as a first approximation to counting degree dd forms.

In Section 4.1, we will present data which suggests this heuristic does give an upper bound, but possibly not an optimal one. It is really the data that lends credence to 1.1.

An alternative perspective, which we will explore in Section 3, is that it is more natural to model the distribution of degree dd forms by modeling dd-dimensional modular abelian varieties. The analysis we do there for d=2d=2 is compatible with the notion that the random Hecke polynomial heuristic gives a valid upper bound which might not be optimal.

2.3. Finiteness questions

Related to the question of asymptotics are several questions about finiteness. We do not investigate them here, but suggest them for future consideration.

  1. (1)

    We can ask: for what dd are there infinitely many weight 2 newforms of squarefree level? 1.1 asserts such dd must be at most 6, but also suggests the answer could be negative for d=6d=6. We know a positive answer for d=1d=1, and expect a positive answer for d=2d=2. Section 3.4 and our data suggest the answers may be positive for d=3,4d=3,4 also.

  2. (2)

    More generally, one can ask the same question in weight 2k2k. Roberts’ conjecture [33] implies that there are only finitely many quadratic twist classes of non-CM rational newforms in weight 2k62k\geq 6. So 2.2 suggests that there are only finitely many newforms of squarefree level of fixed weight 2k62k\geq 6 and any fixed degree d1d\geq 1. One might similarly expect to have finitely many quadratic twist classes of non-CM newforms of fixed degree dd and weight 2k62k\geq 6.

  3. (3)

    One can also ask whether there should be a uniform version of the finiteness part of 1.1, i.e., whether for sufficiently large squarefree NN and some fixed d0d_{0} (possibly d0=6d_{0}=6), each Atkin–Lehner eigenspace has a unique Galois orbit of size dd0d\geq d_{0}. This seems plausible based on 2.2. See 4.4 for a more precise question in prime level.

3. Hecke fields

In Section 2 we considered the question of how often degree dd newforms occur and presented a random Hecke polynomial model, which, at least for prime levels, appears to give asymptotic upper bounds. Here we consider the refined question of how often a specific degree dd rationality field KK should occur, and relate this question to rational points on Hilbert modular varieties. We discuss possible lower bounds for fixed quadratic fields, prove some lower bounds, and predict that (5)\mathbb{Q}(\sqrt{5}) is the most common quadratic rationality field.

3.1. Modular varieties

First recall the connection between weight 2 modular forms and abelian varieties.

Let N1N\geq 1. To a newform fS2(N)f\in S_{2}(N) with [Kf:]=d[K_{f}:\mathbb{Q}]=d, Shimura constructed a dd-dimensional simple abelian variety Af/A_{f}/\mathbb{Q} satisfying the following properties. First, AfA_{f} is a quotient of J0(N)J_{0}(N). Moreover AfA_{f} is isogenous to AgA_{g} if and only if ff and gg are Galois conjugates. The endomorphism algebra End0(Af)End(Af)Kf\mathrm{End}^{0}(A_{f})\coloneqq\mathrm{End}(A_{f})\otimes\mathbb{Q}\simeq K_{f}. The conductor of AfA_{f} is NdN^{d}. Finally, L(s,Af)=σL(s,fσ)L(s,A_{f})=\prod_{\sigma}L(s,f^{\sigma}), where fσf^{\sigma} ranges over the Galois conjugates of ff.

In general, the center of the endomorphism algebra of a dd-dimensional abelian variety AA has degree d\leq d. If End0(A)\mathrm{End}^{0}(A) contains a totally real field KK of degree d=dimAd=\dim A, then we say AA has maximal real multiplication (RM). Any AfA_{f} as above has maximal RM, and conversely if A/A/\mathbb{Q} is a simple abelian variety with maximal RM, then it is isogenous to some AfA_{f} [26, Lemma 3.1]. Hence the correspondence fAff\mapsto A_{f} yields a bijection between degree dd newforms ff of weight 22 and isogeny classes of dd-dimensional simple abelian varieties A/A/\mathbb{Q} with maximal RM.

We propose a heuristic approach to predicting coarse asymptotic counts of such objects. Let KK be a totally real number field of degree dd, and 𝔞\mathfrak{a} be an ideal in 𝒪K\mathcal{O}_{K}. The quotient d/SL(𝒪K𝔞)\mathfrak{H}^{d}/\mathrm{SL}(\mathcal{O}_{K}\oplus\mathfrak{a}) parametrizes dd-dimensional complex abelian varieties with RM by 𝒪K\mathcal{O}_{K} together with a polarization structure corresponding to 𝔞\mathfrak{a} (see [15] for a precise statement). Compactifying this quotient and desingularizing gives a Hilbert modular variety Y(𝒪K𝔞)Y(\mathcal{O}_{K}\oplus\mathfrak{a}).

Now consider a newform fS2(N)f\in S_{2}(N) with Kf=KK_{f}=K. The abelian variety AfA_{f} has endomorphism ring an order in 𝒪K\mathcal{O}_{K}. Typically we expect it is all of 𝒪K\mathcal{O}_{K}, but if not, one can replace AfA_{f} by an isogenous variety with RM by 𝒪K\mathcal{O}_{K}. Thus ff corresponds to a rational point yy on Y(𝒪K𝔞)Y(\mathcal{O}_{K}\oplus\mathfrak{a}) for some 𝔞\mathfrak{a}, which we can take to be in a given set of representatives for Cl+(K)\mathrm{Cl}^{+}(K).

This correspondence is far from one-to-one. First, replacing AfA_{f} by an isogenous variety, or modifying the polarization structure, may give a different point yy on Y(𝒪K𝔞)Y(\mathcal{O}_{K}\oplus\mathfrak{a}). Second, if gg is another weight 2 newform and AgA_{g} is \mathbb{C}-isogenous to AfA_{f}, then both ff and gg correspond to the same rational points. Third, this is not a fine moduli space, so not all rational points on Y(𝒪K𝔞)Y(\mathcal{O}_{K}\oplus\mathfrak{a}) will correspond to abelian varieties defined over \mathbb{Q}, and of those that do, some will correspond to non-simple abelian varieties.

That said, it seems reasonable to expect that, generically, quadratic twist classes of Galois orbits of weight 2 newforms correspond to finite sets of rational points on Y=aCl+(K)Y(𝒪K𝔞)Y=\bigcup_{a\in\mathrm{Cl}^{+}(K)}Y(\mathcal{O}_{K}\oplus\mathfrak{a}). Thus one can attempt estimate the number of quadratic twist classes by estimating counts of rational points on YY. A priori, it is not clear how different orderings of classes of newforms (e.g., by minimal level) will correlate with different orderings of sets of rational points (e.g., by minimal height, for some choice of height function), and we will speculate more on this for d=2d=2 anon.

Note that typically (each component of) YY will be of general type, and one might expect that it has finitely many (and often no) rational points. Hence, for a given dd, to estimate counts of degree dd newforms, it should in principle suffice to consider finitely many YY. Moreover, this philosophy suggests that some totally real degree dd rationality fields will be more common than others, roughly according to whether the moduli spaces YY have many or few rational points. Of course this is not the only consideration, due to various complications of the correspondence between newforms and rational points mentioned above.

This philosophy is in line with Coleman’s conjecture (e.g., see [5]), which predicts there are only finitely many isomorphism classes of endomorphism algebras for dd-dimensional abelian varieties over \mathbb{Q}. Hence Coleman’s conjecture implies that, for a fixed dd, only finitely many degree dd rationality fields KfK_{f} occur as ff varies over weight 2 newforms.

3.2. Rational points on Hilbert modular surfaces

Now we estimate point counts on certain Hilbert modular surfaces, and pursue the ideas of the previous section for d=2d=2.

Let D>0D>0 be a fundamental discriminant and 𝒪D\mathcal{O}_{D} be the ring of integers of (D)\mathbb{Q}(\sqrt{D}). Let Y(D)Y_{-}(D) be the Hilbert modular surface constructed from the quotient 2/SL(𝒪DD𝒪D)\mathfrak{H}^{2}/\mathrm{SL}(\mathcal{O}_{D}\oplus\sqrt{D}\mathcal{O}_{D}). This parametrizes principally polarized abelian surfaces with RM by 𝒪D\mathcal{O}_{D} (together with a polarization structure). See [38], [15] for details. For brevity, we will write RM DD for RM by 𝒪D\mathcal{O}_{D}.

For our heuristic point counts, we will use explicit models for Hilbert modular surfaces. For D<100D<100, Elkies and Kumar [14] computed models for Y(D)Y_{-}(D). By work of Hirzebruch and Zagier [19], Y(D)Y_{-}(D) is rational (i.e., birational to 2\mathbb{P}^{2}) if and only if D{5,8,12,13,17}D\in\{5,8,12,13,17\}.

We expect that 100% of degree 2 weight 2 newforms correspond to rational points on Hilbert modular surfaces with the most rational points, i.e., the rational surfaces. While the Hilbert modular surfaces parametrizing non-principally polarizable surfaces with RM DD are rational over \mathbb{C} for D=12,21,24,28,33,60D=12,21,24,28,33,60 (see [38, Theorem VII.3.3]), we at least expect that the 5 rational Y(D)Y_{-}(D)’s should account for a positive proportion of degree 2 weight 2 newforms, and this is supported by data.

To be more precise, the polarization classes of abelian surfaces with RM DD are in bijection with the narrow ideal classes Cl+((D))\mathrm{Cl}^{+}(\mathbb{Q}(\sqrt{D})). In particular, if D=5,8,13,17D=5,8,13,17, the abelian surface is automatically principally polarizable. For D=12D=12, the narrow class number is 2, and the moduli spaces for each polarization type are rational (at least over \mathbb{C}), so it is not clear whether a positive proportion of abelian surfaces with RM 1212 should be principally polarizable. We cannot yet analyze counts for the non-prinicipal polarization types as we do not know models for the corresponding moduli spaces together with appropriate invariants. However, our data suggest that, at least for prime level, most degree 2 weight 2 newforms have rationality field (D)\mathbb{Q}(\sqrt{D}) with D=5D=5 or 88, and thus correspond to points on Y(5)Y_{-}(5) and Y(8)Y_{-}(8).

For the remainder of the section, assume D{5,8,12,13,17}D\in\{5,8,12,13,17\}. Then Y(D)Y_{-}(D) is birational to m,n2\mathbb{P}^{2}_{m,n}. Let 𝒜2\mathcal{A}_{2} be the moduli space for principally polarized abelian surfaces. Forgetting the RM action yields a map Y(D)𝒜2Y_{-}(D)\to\mathcal{A}_{2}.

Let 2\mathcal{M}_{2} be the moduli space of genus 2 curves. To a genus 2 curve C:y2=h(x)C:y^{2}=h(x), one associates Igusa–Clebsch invariants I2j(C)I_{2j}(C) for j=1,2,3,5j=1,2,3,5. Here I2j(C)I_{2j}(C) can be regarded as a degree 2j2j-polynomial in the coefficients of h(x)h(x), and I10(C)I_{10}(C) is the discriminant of hh. One can realize 𝒜2\mathcal{A}_{2} as weighted projective space 1,2,3,53\mathbb{P}^{3}_{1,2,3,5} with coordinates (I2:I4:I6:I10)(I_{2}:I_{4}:I_{6}:I_{10}). The Torelli map 2𝒜2\mathcal{M}_{2}\to\mathcal{A}_{2} sends the moduli of CC to (I2(C):I4(C):I6(C):I10(C))(I_{2}(C):I_{4}(C):I_{6}(C):I_{10}(C)), and the image is the complement of the hyperplane I10=0I_{10}=0. We note that Igusa–Clebsch invariants are only isomorphism invariants of CC up to weighted projective scaling.

Elkies and Kumar [14] gave a birational model for Y(D)Y_{-}(D). In particular, for generic affine coordinates (m,n)𝔸2(m,n)\in\mathbb{A}^{2}, one has an associated point on Y(D)Y_{-}(D) and thus weighted projective coordinates (I2(m,n):I4(m,n):I6(m,n):I10(m,n))𝒜2(I_{2}(m,n):I_{4}(m,n):I_{6}(m,n):I_{10}(m,n))\in\mathcal{A}_{2}, where the I2j(m,n)I_{2j}(m,n)’s are explicit rational functions in m,nm,n.

Now we will attempt to estimate the number of rational points (m,n)(m,n) with bounded Igusa–Clebsch invariants. First we want to scale Igusa–Clebsch invariants (in 1,2,3,53\mathbb{P}^{3}_{1,2,3,5}) to be integral, as will be the case for the I2j(C)I_{2j}(C)’s given a curve CC over \mathbb{Z}. Let us write (m,n)=(a/c,b/c)(m,n)=(a/c,b/c) for a,b,ca,b,c\in\mathbb{Z} with gcd(a,b,c)=1\gcd(a,b,c)=1. Regarding a,b,ca,b,c as variables, we scale the I2j(m,n)I_{2j}(m,n)’s to get polynomials I2j(a,b,c)[a,b,c]I_{2j}(a,b,c)\in\mathbb{Z}[a,b,c]’s which are minimal integral over [a,b,c]\mathbb{Z}[a,b,c]. That is, we scale out denominators, and also any factors of the numerators π\pi so that πjI2j(a,b,c)\pi^{j}\mid I_{2j}(a,b,c) for all j{1,2,3,5}j\in\{1,2,3,5\} implies π\pi is a unit in [a,b,c]\mathbb{Z}[a,b,c]. The resulting I10(a,b,c)I_{10}(a,b,c)’s (which are uniquely determined up to ±1\pm 1) are given in 3.1.

DD I10(a,b,c)I_{10}(a,b,c)
55 8(a510a3b2+25ab4+5a4c50a2b2c+125b4c8(a^{5}-10a^{3}b^{2}+25ab^{4}+5a^{4}c-50a^{2}b^{2}c+125b^{4}c
5a3c2+25ab2c245a2c3+225b2c3+108c5)2-5a^{3}c^{2}+25ab^{2}c^{2}-45a^{2}c^{3}+225b^{2}c^{3}+108c^{5})^{2}
88 8c3(ac)3(a+c)6(16a2b2+32b4+a3c56ab2c+9a2c272b2c2+27ac3+27c4)28c^{3}(a-c)^{3}(a+c)^{6}(-16a^{2}b^{2}+32b^{4}+a^{3}c-56ab^{2}c+9a^{2}c^{2}-72b^{2}c^{2}+27ac^{3}+27c^{4})^{2}
1212 (a+c)3(ac)9(27a2+b2+27c2)2(a2b+9a2c8c3)3(a+c)^{3}(a-c)^{9}(-27a^{2}+b^{2}+27c^{2})^{2}(a^{2}b+9a^{2}c-8c^{3})^{3}
1313 23311(267a3+72a2bab23552a2c+1440abc128b2c+768ac2)22^{3}\cdot 3^{11}\cdot(-267a^{3}+72a^{2}b-ab^{2}-3552a^{2}c+1440abc-128b^{2}c+768ac^{2})^{2}
(12a3+3a2c+b2c)4(a3150a2c+6abc264ac2+120bc2+64c3)4\cdot\,(-12a^{3}+3a^{2}c+b^{2}c)^{4}(-a^{3}-150a^{2}c+6abc-264ac^{2}+120bc^{2}+64c^{3})^{4}
1717 215311(132a+b+3c)3(256a31200a2c+18abc6006ac2+99bc2+41c3)52^{15}\cdot 3^{11}\cdot(-132a+b+3c)^{3}(-256a^{3}-1200a^{2}c+18abc-6006ac^{2}+99bc^{2}+41c^{3})^{5}
(456a2+ab+723ac8bc+24c2)3(4608a31728a2c+b2c+216ac29c3)2\cdot\,(456a^{2}+ab+723ac-8bc+24c^{2})^{3}(4608a^{3}-1728a^{2}c+b^{2}c+216ac^{2}-9c^{3})^{2}
Table 3.1.

I10I_{10} polynomials for Y(D)Y_{-}(D)

Specializing a,b,ca,b,c to integers, we denote by I2jmin(a,b,c)I_{2j}^{\min}(a,b,c)\in\mathbb{Z} scalings which are minimal integral over \mathbb{Z}. Note that I2jmin(a,b,c)I2j(a,b,c)I_{2j}^{\min}(a,b,c)\mid I_{2j}(a,b,c) but they are often not equal. E.g., when D=5D=5, then the invariants I2j(1,3,2)I_{2j}(1,3,2)’s are (2453,2854,2155599,22138)(-2^{4}\cdot 5^{3},2^{8}\cdot 5^{4},-2^{15}\cdot 5\cdot 599,2^{21}\cdot 3^{8}), whereas the \mathbb{Z}-minimal invariants I2jmin(a,b,c)I_{2j}^{\min}(a,b,c) are obtained by scaling out a factor of 242^{4}, i.e., they are (53,54,235599,238)(-5^{3},5^{4},-2^{3}\cdot 5\cdot 599,2\cdot 3^{8}).

First we want to estimate, in terms of a real parameter TT, the growth of the cardinality of

ZD(T){(a,b,c)3UD:gcd(a,b,c)=1 and |I2jmin(a,b,c)|<T2j for j1,2,3,5},Z_{D}(T)\coloneqq\{(a,b,c)\in\mathbb{Z}^{3}-U_{D}:\gcd(a,b,c)=1\text{ and }|I^{\min}_{2j}(a,b,c)|<T^{2j}\text{ for }j\in 1,2,3,5\},

where UDU_{D} consists of (a,b,c)(a,b,c) such that the map (a/c,b/c)𝒜2(a/c,b/c)\to\mathcal{A}_{2} is either undefined (e.g., c=0c=0) or is not finite-to-one (e.g., for D=8D=8, all points with m=a/c=1m=a/c=-1 map to (1:0:0:0)𝒜2(1:0:0:0)\in\mathcal{A}_{2}). Really our interest is just in bounding I10minI^{\min}_{10}, but we impose bounds on the other I2jminI_{2j}^{\min}’s to guarantee finiteness of ZD(T)Z_{D}(T).

Precise estimates are difficult, so we make two simplifications which are sufficient to get lower bounds: (1) We impose the stronger bound |I2jmin(a,b,c)||I2j(a,b,c)|<T2j|I^{\min}_{2j}(a,b,c)|\leq|I_{2j}(a,b,c)|<T^{2j}. (2) We will suppose each monomial in I2j(a,b,c)I_{2j}(a,b,c) is bounded by T2jT^{2j}. Note that (1) and (2) can respectively be thought of as non-archimedean and archimedean simplifications to monomials.

Proposition 3.2.

We have #ZD(T)TrD\#Z_{D}(T)\gg T^{r_{D}} as TT\to\infty, where respectively rD=3,32,2,1,1r_{D}=3,\tfrac{3}{2},2,1,1 for D=5,8,12,13,17D=5,8,12,13,17.

Proof.

Each I2j(a,b,c)I_{2j}(a,b,c) is a homogeneous polynomial in a,b,ca,b,c, say of degree djd_{j}. Taking a,b,ca,b,c independently up to size T2j/djT^{2j/d_{j}} shows there are T6j/dj\gg T^{6j/d_{j}} tuples (a,b,c)(a,b,c) with |I2j(a,b,c)|<T2j|I_{2j}(a,b,c)|<T^{2j}. Moreover, one checks the ratio j/djj/d_{j} is independent of the choice of jj. Since the conditions gcd(a,b,c)=1\gcd(a,b,c)=1 and (a,b,c)UD(a,b,c)\not\in U_{D} are satisfied for a positive proportion of (a,b,c)(a,b,c), we get the asymptotic lower bound #ZD(T)T30/d\#Z_{D}(T)\gg T^{30/d}, where d=degI10(a,b,c)d=\deg I_{10}(a,b,c). We respectively have d=10,20,25,30,30d=10,20,25,30,30 for D=5,8,12,13,17D=5,8,12,13,17, which gives the asserted lower bounds for D=5,8,13,17D=5,8,13,17.

For D=12D=12, one can get better lower bounds using a 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} parametrization for (m,n)(m,n). Namely, write (m,n)=(r/s,t/u)(m,n)=(r/s,t/u) for r,s,t,ur,s,t,u\in\mathbb{Z}. Let I2j(r,s,t,u)I_{2j}(r,s,t,u) be the minimal invariants over [r,s,t,u]\mathbb{Z}[r,s,t,u]. These have degrees 6,12,18,306,12,18,30 for j=1,2,3,5j=1,2,3,5. Using same argument as above gives a lower bound of #Z12(T)T4/3\#Z_{12}(T)\gg T^{4/3}. However, if we regard I2j(r,s,t,u)I_{2j}(r,s,t,u) as polynomials only in tt and uu, the respective degrees are 2,4,6,102,4,6,10 for j=1,2,3,5j=1,2,3,5. Thus by taking r,sr,s uniformly bounded and |t|,|u|T|t|,|u|\ll T yields #Z12(T)T2\#Z_{12}(T)\gg T^{2} as claimed. ∎

The lower bounds in the proposition are the optimal ones we could find using the 2\mathbb{P}^{2} or 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} parametrizations for (m,n)(m,n) by allowing either each of a,b,ca,b,c or r,s,t,ur,s,t,u to vary independently up to some power of TT (not necessarily the same power for each variable). We remark that the optimal exponents for lower bounds using the 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} parametrization for D=5,8,13,17D=5,8,13,17 are 2,4/3,1,12,4/3,1,1, respectively. To get these exponents, for D=5D=5 one can take each of |r|,|s|,|t|,|u|T|r|,|s|,|t|,|u|\ll T. For D=8D=8, one takes |r|,|s|T2/3|r|,|s|\ll T^{2/3} and |t|,|u|1|t|,|u|\ll 1. For both D=13D=13 and D=17D=17, one takes |r|,|s|1|r|,|s|\ll 1 and |t|,|u|T1/2|t|,|u|\ll T^{1/2}.

Question 3.3.

For D{5,8,12,13,17}D\in\{5,8,12,13,17\}, is #ZD(T)TrD+ε\#Z_{D}(T)\ll T^{r_{D}+\varepsilon} for any ε>0\varepsilon>0?

It is not clear if the simplifications to monomials affect the exponents in our estimates, but if the I2jI_{2j} polynomials are sufficiently general type, one might expect they only account for a multiplicative factor of size O(1+Tε)O(1+T^{\varepsilon}). Note that in the Brumer–McGuinness heuristics, it is believed the analogous archimedean simplification (2) only affects counts by an O(1)O(1) factor.

A more serious reason to doubt the exponents in these lower bounds are optimal is that there may exist (i) other rational parametrizations of Y(D)Y_{-}(D) where the I2jI_{2j} degrees are smaller, or (ii) special curves on Y(D)Y_{-}(D) which intersect ZDZ_{D} in an especially large number of rational points. Indeed, the proof makes clear that different parametrizations may yield different counts for a given surface.

Now we explain how these estimates for #ZD(T)\#Z_{D}(T) are related to counting quadratic twist classes of weight 2 newforms ff with rationality field (D)\mathbb{Q}(\sqrt{D}). As explained above, a positive proportion of ff (at least if D12D\neq 12) should correspond to rational points on Y(D)Y_{-}(D). Conversely, a rational point on Y(D)Y_{-}(D) may not come from a simple abelian surface with RM defined over \mathbb{Q}—one needs that the Mestre conic has a point, the RM is defined over \mathbb{Q}, and the Jacobian is nonsplit. However, we expect that these obstructions will only contribute logarithmic factors to asymptotics. (See [9] for details about when the Mestre obstruction vanishes and when the RM is defined over \mathbb{Q}.)

Consider a point (a,b,c)ZD(T)(a,b,c)\in Z_{D}(T) which corresponds to a ¯\bar{\mathbb{Q}}-isomorphism class of simple abelian surfaces A/A/\mathbb{Q}. Generically this ¯\bar{\mathbb{Q}}-isomorphism class should be the family of Jacobians of quadratic twists of a genus 2 curve C/C/\mathbb{Q} with RM DD. Suppose this, and assume CC is a minimal quadratic twist of conductor NCN_{C}. Then (a,b,c)(a,b,c) corresponds to the quadratic twist class of some minimal newform ff of level NfN_{f} where NC=Nf2N_{C}=N_{f}^{2}. One can write down a minimal integral model for CC, and the polynomially-defined Igusa–Clebsch invariants I2j(C)I_{2j}(C) are necessarily divisible by I2jmin(a,b,c)I_{2j}^{\min}(a,b,c). Thus I10min(a,b,c)I_{10}^{\min}(a,b,c) divides the minimal discriminant ΔC=212I10(C)\Delta_{C}=2^{-12}I_{10}(C) of CC. One also knows that the conductor NCΔCN_{C}\mid\Delta_{C}.

Now we would like to understand how NCN_{C} relates to I10min(a,b,c)I_{10}^{\min}(a,b,c) or I10(a,b,c)I_{10}(a,b,c). There are no general upper or lower bounds, and in fact there are competing issues in opposite directions. One is that ΔC\Delta_{C} may be much larger than either I10min(a,b,c)I_{10}^{\min}(a,b,c) or I10(a,b,c)I_{10}(a,b,c), and numerically this is quite typical. E.g., if pmI10min(a,b,c)p^{m}\parallel I_{10}^{\min}(a,b,c) it often happens that pm+10ΔCp^{m+10}\mid\Delta_{C} (see [24] for local results). On the other hand, the prime powers occurring in NCN_{C} are often smaller than those in ΔC\Delta_{C}. E.g., if p3ΔCp^{3}\parallel\Delta_{C} then necessarily p2NCp^{2}\parallel N_{C} or pNCp\nmid N_{C}. Some preliminary investigations suggest that the latter issue has more impact, and asymptotic counts of curves by I10minI^{\min}_{10} may essentially be lower bounds (up to logarithmic factors) of counts by conductor. This leads us to ask:

Question 3.4.

Let C2tw(D;X)C^{\mathrm{tw}}_{2}(D;X) be the number of quadratic twist classes of non-CM weight 22 newforms of minimal level N<XN<X with rationality field (D)\mathbb{Q}(\sqrt{D}). Is C2tw(D;X)XαεC^{\mathrm{tw}}_{2}(D;X)\gg X^{\alpha-\varepsilon} for α\alpha such that #ZD(T)Tα/5\#Z_{D}(T)\gg T^{\alpha/5}? Note that one can take α=35,310,25,15,15\alpha=\tfrac{3}{5},\tfrac{3}{10},\tfrac{2}{5},\tfrac{1}{5},\tfrac{1}{5} for D=5,8,12,13,17D=5,8,12,13,17.

Observe that all of these exponents are less than the exponent of 2/3 from the d=2d=2 case of 1.1. We will compare these speculative lower bounds with our prime level data below.

Even if these lower bounds hold, there are several reasons why they may not be sharp. For one, there is the issue of 3.3. Perhaps most serious is the issue of how prime powers in NCN_{C} relate to prime powers in ΔC\Delta_{C} mentioned above.

Another potential issue comes from the way we defined ZD(T)Z_{D}(T): for the comparison with conductors, we are only interested in bounds on I10minI_{10}^{\min}, and there may be many points with I10minI^{\min}_{10} small relative to I2min,I4min,I6minI_{2}^{\min},I_{4}^{\min},I_{6}^{\min}. In particular, for D=12D=12, if one views I2j(a,b,c)I_{2j}(a,b,c) as a polynomial in bb with a=0a=0 and cc fixed, then the degrees are 2,2,4,42,2,4,4, so one gets at least T5/2T^{5/2} points with I10minT10I_{10}^{\min}\ll T^{10}, which is a better than the lower bound T2\gg T^{2} in Proposition 3.2. These rational points correspond to the curve m=0m=0 on Y(12)Y_{-}(12), and numerical investigations suggests this is a Shimura curve parametrizing abelian surfaces with geometric endomorphism algebra the quaternion algebra of discriminant 6. Consequently, one might be able to take α=12\alpha=\tfrac{1}{2} in 3.4 when D=12D=12. However, it is not clear that there is a family of genus 2 curves with RM 12 over \mathbb{Q} that would achieve α=12\alpha=\tfrac{1}{2}.

3.3. Lower bounds for quadratic fields

There are several known families of genus 2 curves with RM 5 and RM 8. These can be used to give lower bounds on counting such curves with bounded discriminant, and therefore conductor. For a genus 2 curve C/C/\mathbb{Q}, let ΔC\Delta_{C} denote the minimal integral discriminant.

Proposition 3.5.

The number of ¯\bar{\mathbb{Q}}-isomorphism classes of genus 22 curves C/C/\mathbb{Q} with RM 55 (resp. RM 88) with |ΔC|<X|\Delta_{C}|<X is X1/6\gg X^{1/6} (resp. X1/7\gg X^{1/7}).

Proof.

First consider RM 5. Brumer exhibited a 3-parameter family of curves Cb,c,dC_{b,c,d} with RM 5 over \mathbb{Q} (see [6] for an announcement and [18] for a proof), however it is not clear when two such curves are isomorphic, either over \mathbb{Q} or ¯\bar{\mathbb{Q}}. We consider the 1-parameter subfamily CdC_{d} with b=c=0b=c=0, which is given by

Cd:y2+(x3+x+1)y=dx3+x2+x.C_{d}:y^{2}+(x^{3}+x+1)y=-dx^{3}+x^{2}+x.

This defines a genus 2 curve with RM 5 for all dd\in\mathbb{Z} (I10I_{10} is never 0 for dd\in\mathbb{Z}), and the discriminant of this model, (27d381d234d103)2(27d^{3}-81d^{2}-34d-103)^{2}, is degree 66 in dd. Thus to complete the RM 5 case of the proposition, it suffices to show that the number of CdC_{d^{\prime}} isomorphic to a given CdC_{d} over ¯\bar{\mathbb{Q}} is finite and uniformly bounded.

Let I2jI_{2j} (resp. I2jI_{2j}^{\prime}) denote the polynomial Igusa–Clebsch invariants for CdC_{d} (resp. CdC_{d^{\prime}}) for j=1,2,3,5j=1,2,3,5. Then I4/I22I_{4}/I_{2}^{2} and is an absolute invariant for CdC_{d}. Thus if CdC_{d} and CdC_{d^{\prime}} are ¯\bar{\mathbb{Q}}-isomorphic, one must have F:=(I2)2I4I22I4=0F:=(I_{2}^{\prime})^{2}I_{4}-I_{2}^{2}I_{4}^{\prime}=0. Now F=0F=0 defines a union of 3 curves in the (d,d)(d,d^{\prime})-plane, none of which are of the form d=d0d=d_{0}. So the number of CdC_{d^{\prime}} which are ¯\bar{\mathbb{Q}}-isomorphic to a fixed CdC_{d} is bounded by degF\deg F.

Now consider RM 8. Here we use Mestre’s 2-parameter family Ca,bC^{\prime}_{a,b} of genus 2 RM 8 curves over \mathbb{Q} from [30]. Consider the subfamily Cb=C2,bC^{\prime}_{b}=C^{\prime}_{2,b}, which is given by

Cb:y2=7500x5+(75b+3400)x4+(34b+2283)x3+(3b+1111)x2+177x+9.C^{\prime}_{b}:y^{2}=7500x^{5}+(-75b+3400)x^{4}+(-34b+2283)x^{3}+(-3b+1111)x^{2}+177x+9.

This is a genus 2 curve with RM 8 for b{88,112}b\in\mathbb{Z}-\{-88,112\}, and the discriminant has degree 7 in bb. One can complete the argument just as in the RM 5 case. ∎

Corollary 3.6.

The number of quadratic twist classes of weight 22 newforms with rationality field (5)\mathbb{Q}(\sqrt{5}) (resp. (2)\mathbb{Q}(\sqrt{2})) and minimal level N<XN<X is X1/3\gg X^{1/3} (resp. X2/7\gg X^{2/7}).

Proof.

If CC is a genus 2 curve with RM DD over \mathbb{Q} with nonsplit Jacobian, then modularity tells us that CC corresponds to a weight 2 newform ff of level N|DC|N\leq\sqrt{|D_{C}|} and rationality field (D)\mathbb{Q}(\sqrt{D}), so it suffices to show that a positive proportion of the curves in the proofs have nonsplit Jacobian. For the RM 5 CdC^{\prime}_{d} family above, one can check that the above model has discriminant coprime to 55 when d1mod5d\equiv 1\bmod 5. Moreover computing the LL-polynomial shows the mod 5 Jacobian is nonsplit, whence the Jacobian of CdC_{d} over \mathbb{Q} is nonsplit. Similarly, for the RM 8 family, the curve CbC^{\prime}_{b} has nonsplit Jacobian when b1mod7b\equiv 1\bmod 7. ∎

Note that these lower bounds are significantly smaller than those in 3.4.

Remark 3.7.
  1. (1)

    Under the Bateman–Horn conjecture [1], 27d381d234d10327d^{3}-81d^{2}-34d-103 is prime for X/logX\gg X/\log X integers d1mod5d\equiv 1\bmod 5. For such dd, CdC_{d} has prime-squared discriminant and the associated newform ff has prime conductor. Consequently, subject to this conjecture, the above argument shows there are X1/3/logX\gg X^{1/3}/\log X weight 2 newforms ff with rationality field (5)\mathbb{Q}(\sqrt{5}) and prime level <X<X. The analogous argument does not work for RM 8 as the discriminant of CbC^{\prime}_{b} splits into linear factors over \mathbb{Z}.

  2. (2)

    Elkies [13] recently gave similar lower bounds for genus 2 curves with RM 5 which satisfy an Eisenstein congruence, but with the exponent 13\tfrac{1}{3} replaced by 14\tfrac{1}{4}.

  3. (3)

    One should similarly be able to give explicit lower bounds for D=12,13,17D=12,13,17 (and some other DD). Namely, [14] gives infinite families of RM DD genus 2 curves for various DD, though without explicit models. One can use Mestre’s algorithm to construct models, and then follow the above argument. However, this typically yields families with discriminants of very large degree, and so one would need to do more work to get decent lower bounds.

3.4. Remarks for higher degree

Much less is known about Hilbert modular nn-folds for n>2n>2 than for n=2n=2. Grundman and Lippincott ([16, 17]) have done work towards classifying such spaces by arithmetic genus for n=3,4n=3,4. Then Borisov and Gunnells [4] studied the geometry of a Hilbert modular 3-fold attached to (ζ7+)\mathbb{Q}(\zeta_{7}^{+}). However, to our knowledge, not much is known about the explicit geometry of higher-dimensional Hilbert modular varieties beyond these works and their references.

One case where we do know of geometric constructions leading to weight 2 modular forms of degree d>2d>2 is the following. In [30], Mestre constructs (among other things) families of genus d=p12d=\frac{p-1}{2} hyperelliptic curves with potential RM by [ζp+]\mathbb{Z}[\zeta_{p}^{+}]. When p=5,7p=5,7, the RM is actually defined over \mathbb{Q}, and generically the curves should correspond to degree 22 and 33 modular forms with rationality fields (5)\mathbb{Q}(\sqrt{5}) and (ζ7+)\mathbb{Q}(\zeta_{7}^{+}).

Mestre’s constructions have been extended by various authors. For instance, [21] and [20] construct genus 3 curves CC with RM by an order in (ζ7+)\mathbb{Q}(\zeta_{7}^{+}) such that the RM is generically defined over the base field. This at least suggests there may be infinitely many weight 2 newforms of squarefree level with rationality field (ζ7+)\mathbb{Q}(\zeta_{7}^{+}).

4. Data

The LMFDB [25] currently contains all weight 2 newforms of level N104N\leq 10^{4} [3]. However, this range is not nearly sufficient to study asymptotic behavior of the distribution of newform degrees in squarefree level. For instance, one is still forced to have many small degree forms of level close to 10000 as the size of Atkin–Lehner spaces can still be small. For instance, for N=9870=235747N=9870=2\cdot 3\cdot 5\cdot 7\cdot 47, there are 183 newforms in S2(N)S_{2}(N) divided among 32 Atkin–Lehner eigenspaces, and each Atkin–Lehner eigenspace has dimension between 33 and 88.

One can mitigate this effect by restricting to levels with at most 2 or 3 prime factors, or ordering counts by dimensions of Atkin–Lehner eigenspaces rather than level. However, even with such considerations, the range is still not large enough to say much about asymptotic counts.

Instead, we analyze data we computed in prime level using the algorithms from [8]. Namely, we computed all weight 2 newforms of prime level less than 21062\cdot 10^{6} and degree at most 66, as well as the degrees of all newforms of prime level less than 10610^{6}. Our data is available on the first author’s personal webpage, and is currently in the process of being added to the LMFDB.

4.1. Data for counts by degree

In 4.1, we plot counts 𝒞d(X)\mathcal{C}_{d}^{\prime}(X) of degree dd Galois orbits in weight 2 and prime level at most XX for 1d41\leq d\leq 4. The plot uses a log-log scale, and because the number of primes up to XX is approximately li(X)\text{li}(X), we do a least-squares fit of functions of the form y=log(ali(exp(x)b)))y=\log(a\text{li}(\exp(x)^{b}))) to the data

{(logX,log𝒞d(X)):X<2106,X prime,𝒞d(X)1}.\left\{(\log X,\log\mathcal{C}_{d}^{\prime}(X))\,:\,X<2\cdot 10^{6},\,X\text{ prime},\,\mathcal{C}_{d}^{\prime}(X)\geq 1\right\}.
[Uncaptioned image]
Figure 4.1.

Number of forms with prime level less than XX by degree, with least-squares fits to the log-log data.

For degree 11, our best fit values a0.97a\approx 0.97 and b0.832b\approx 0.832 are in agreement with the elliptic curve database [2], which, for XX up to 21092\cdot 10^{9}, finds that a0.97a\approx 0.97 and b0.833b\approx 0.833. This agrees very well with the Brumer–McGuinness–Watkins heuristic (1). For d=2,3,4d=2,3,4, the best fit exponents are approximately 0.6290.629, 0.3360.336, 0.1080.108.

A least-squares fit of y=ali(xb)y=a\text{li}(x^{b}) to the data (X,𝒞d(X))(X,\mathcal{C}_{d}^{\prime}(X)) yields similar values; the exponents are 0.8410.841, 0.6220.622, 0.3300.330, 0.1070.107 for d=1,2,3,4d=1,2,3,4, respectively.

When 1d41\leq d\leq 4, 1.1 is consistent with our prime level data. In 4.1, the growth rate O(X1d/6)O(X^{1-d/6}) appears to be an upper bound. In this range the best fit has a notably lower exponent for d=3,4d=3,4, but there may exist logarithmic factors in the main asymptotics for d2d\geq 2. For example, for d=2d=2, the Mestre obstruction to rationality of genus 2 curves (see Section 3.2) may introduce a logarithmic factor in the denominator. Note that in this range, logX>X16\log X>X^{\frac{1}{6}}, so it is difficult to distinguish between logarithmic factors and small powers of XX.

4.2 shows the number of newform orbits of degree at most 66 with prime level between 11 and 10410^{4}, between 10410^{4} and 10610^{6}, and between 10610^{6} and 21062\cdot 10^{6}. Note that the first data column counts prime level forms which were already in the LMFDB.

Degree Level range Total
1 – 1041\text{ -- }10^{4} 104 – 10610^{4}\text{ -- }10^{6} 106 – 210610^{6}\text{ -- }2\cdot 10^{6}
11 329329 88438843 64066406 1557815578
22 212212 22002200 10961096 35083508
33 7676 142142 3535 253253
44 2828 1010 11 3939
55 2020 22 2222
66 1111 11 1212
Table 4.2.

Number of prime-level newform orbits by degree and level. Blank entries are 0.

Many of the forms of degree 55 and 66 in this dataset have very small levels. For instance, 1212 of the 2222 degree 55 forms are the largest-degree forms in their Atkin–Lehner eigenspaces, and similarly for 77 of the 1212 degree 66 forms. Only 55 of the degree 55 forms and 33 of the degree 66 forms have levels greater than 10001000.

Given this paucity of data, we refrain from any quantitative analysis of these forms, but remark that the counts for d=5,6d=5,6 appear to be consistent with 1.1.

4.3 gives the decomposition type, i.e., sizes of Galois orbits, of all prime levels up to 1 million which have two or more newform orbits of degree at least 77 in the same Atkin–Lehner eigenspace. Here S2±(p)S_{2}^{\pm}(p) denotes the subspace of S2(p)S_{2}(p) with Atkin–Lehner eigenvalue ±1\pm 1 at pp.

Level S2+(p)S_{2}^{+}(p) S2(p)S_{2}^{-}(p)
607607 5+7+75+7+7 3131
911911 9+149+14 5353
12231223 3434 9+599+59
12491249 7+377+37 5959
47514751 153153 18+22518+225
Table 4.3.

Atkin–Lehner eigenspaces with multiple orbits of size at least 77

4.3 shows that there are only 66 newforms orbits of prime level less than 10610^{6} with degree 77 or more that are not the unique largest in their Atkin–Lehner eigenspaces (22 of which are tied for the largest), with degrees 77, 77, 77, 99, 99, and 1818. This data for d7d\geq 7 appears to be consistent with 1.1.

Moreover, 4.3 shows that, for each prime level between 47514751 and 10610^{6}, the newforms in each Atkin–Lehner eigenspace consist of a single large Galois orbit together with orbits of size 6\leq 6. This prompts us to ask the following:

Question 4.4.

Is there a prime p>4751p>4751 and a sign ±\pm such that S2±(p)S_{2}^{\pm}(p) contains two or more newform orbits each of degree 77 or more?

This question can be viewed as a uniform version of 1.1 in prime level. It seems plausible to us that in fact 4.3 is a complete list of all “mid-sized forms” of prime level.

4.2. Data counts by quadratic field

In this section we investigate 1.2 and 3.4 empirically. 4.5 and 4.6 present the relevant contents of our dataset.

Disc Level range Total
1 – 1041\text{ -- }10^{4} 104 – 10610^{4}\text{ -- }10^{6} 106 – 210610^{6}\text{ -- }2\cdot 10^{6}
55 158158 19001900 986986 30443044
88 3737 242242 100100 379379
1212 11 1414 33 1818
1313 1313 4040 66 5959
1717 11 11
Table 4.5.

Number of prime-level degree 22 newform orbits by discriminant and level, for discriminants DD such that Y(D)Y_{-}(D) is rational. Blank entries are 0.

[Uncaptioned image]
Figure 4.6.

Counts of degree 2 forms by discriminant DD for rational surfaces Y(D)Y_{-}(D). The plot on the right excludes discriminants 55 and 88.

The growth rate of the counts plotted in 4.6 appears to be the largest for D=5D=5, and D=8D=8 appears to be the next largest. This is consistent with 1.2.

Let 𝒞d,D(X)\mathcal{C}_{d,D}^{\prime}(X) denote the number of newform orbits of degree dd, discriminant DD, and prime level at most XX. For each of D=5D=5, 88, 1212, and 1313, we compute least-squares fits to the following four sets of data points:

  1. (1)

    {(logX,log𝒞d,D(X)): 1<X<2106,X prime,𝒞d,D(X)1}\left\{(\log X,\log\mathcal{C}_{d,D}^{\prime}(X))\,:\,1<X<2\cdot 10^{6},X\text{ prime},\mathcal{C}_{d,D}^{\prime}(X)\geq 1\right\}

  2. (2)

    {(logX,log𝒞d,D(X)): 104<X<2106,X prime,𝒞d,D(X)1}\left\{(\log X,\log\mathcal{C}_{d,D}^{\prime}(X))\,:\,10^{4}<X<2\cdot 10^{6},X\text{ prime},\mathcal{C}_{d,D}^{\prime}(X)\geq 1\right\}

  3. (3)

    {(X,𝒞d,D(X)): 1<X<2106,X prime,𝒞d,D(X)1}\left\{(X,\mathcal{C}_{d,D}^{\prime}(X))\,:\,1<X<2\cdot 10^{6},X\text{ prime},\mathcal{C}_{d,D}^{\prime}(X)\geq 1\right\}

  4. (4)

    {(X,𝒞d,D(X)): 104<X<2106,X prime,𝒞d,D(X)1}\left\{(X,\mathcal{C}_{d,D}^{\prime}(X))\,:\,10^{4}<X<2\cdot 10^{6},X\text{ prime},\mathcal{C}_{d,D}^{\prime}(X)\geq 1\right\}

In the first two cases we fit functions of the form y=log(ali(exp(x)b)))y=\log(a\mathrm{li}(\exp(x)^{b}))), and in the last two y=ali(xb)y=a\mathrm{li}(x^{b}). The best-fit exponents bb we obtain vary depending on our choice of model and range of XX values. We present these best-fit values of bb in 4.7.

Data XX range Best-fit exponents
55 88 1212 1313
(logX,log𝒞d,D(X))(\log X,\log\mathcal{C}_{d,D}^{\prime}(X)) 1121062\cdot 10^{6} 0.730.73 0.610.61 0.390.39 0.420.42
(logX,log𝒞d,D(X))(\log X,\log\mathcal{C}_{d,D}^{\prime}(X)) 10410^{4}21062\cdot 10^{6} 0.640.64 0.540.54 0.390.39 0.380.38
(X,𝒞d,D(X))(X,\mathcal{C}_{d,D}^{\prime}(X)) 1121062\cdot 10^{6} 0.670.67 0.560.56 0.370.37 0.370.37
(X,𝒞d,D(X))(X,\mathcal{C}_{d,D}^{\prime}(X)) 10410^{4}21062\cdot 10^{6} 0.650.65 0.530.53 0.370.37 0.350.35
Table 4.7.

Best-fit values of bb when fitting functions of the form y=log(ali(exp(x)b)))y=\log(a\mathrm{li}(\exp(x)^{b}))) or y=ali(xb)y=a\mathrm{li}(x^{b}) as appropriate to data of counts of degree 22 newform orbits with prime level and prescribed discriminant

The best-fit exponents we obtain are all higher than the lower bounds proposed in 3.4, in many cases substantially, except for discriminant 1212, where the value is slightly lower than the 0.40.4 appearing in 3.4. There is only one form of discriminant 1717 and prime level less than 21062\cdot 10^{6}, at level 7565375653.

The data presented in 4.5, 4.6, and 4.7, as well as the heuristics from 3.4, support 1.2. Namely, the suggested lower bounds for counts by quadratic rationality field are largest for (5)\mathbb{Q}(\sqrt{5}). Since the heuristics do not rely on a restriction to prime level, one is led to ask:

Question 4.8.

Do 100% of quadratic twist classes (ordered by minimal level) of weight 22 degree 22 non-CM newforms have rationality field (5)\mathbb{Q}(\sqrt{5})?

There is an arithmetic reason to expect a relative scarcity of certain quadratic fields in prime level compared to the lower bounds for arbitrary levels suggested in 3.4. Namely, if CC is a genus 2 curve with RM DD, then we typically expect odd primes dividing I10minI_{10}^{\min} to divide the conductor NCN_{C}. Based on the factorizations of I10I_{10}’s in 3.1, we expect I10minI_{10}^{\min} to be a 2-power times a prime power very infrequently for D=12,13,17D=12,13,17. Indeed, the LMFDB [25] lists 54855485, 39483948 21892189, 12301230, and 16431643 forms in all levels N10000N\leq 10000 for D=5D=5, 88, 1212, 1313, and 1717, respectively. Restricting to squarefree level, these numbers are 18201820, 11241124, 445445, 319319, and 461461.

4.3. Data counts for cubic fields

While we have not attempted to carry out the approach outlined in Section 3.1 to estimate counts of degree 3 forms with a given cubic rationality field KK, the data, though more limited, behaves similarly as in the degree 2 case. 4.9 and 4.10 present counts of all prime level degree 33 newform orbits by their Hecke field discriminant. These discriminants are sufficient to specify the Hecke field, in the sense that if ff and gg are degree 33 forms of prime level less than 21062\cdot 10^{6} and Disc(Kf)=Disc(Kg)\mathrm{Disc}(K_{f})=\mathrm{Disc}(K_{g}), then Kf=KgK_{f}=K_{g}.

Disc Level range Total
1 – 1041\text{ -- }10^{4} 104 – 10610^{4}\text{ -- }10^{6} 106 – 210610^{6}\text{ -- }2\cdot 10^{6}
4949 3434 9090 3030 154154
8181 33 1313 1616
148148 1212 66 1818
169169 22 66 33 1111
229229 88 2020 11 2929
257257 99 66 11 1616
321321 22 11 33
404404 22 22
469469 11 11
473473 22 22
621621 11 11
Table 4.9.

Number of prime-level degree 33 newform orbits by discriminant and level. Blank entries are 0.

[Uncaptioned image]
Figure 4.10.

Counts of degree 3 forms by discriminant. The plot on the right excludes discriminant 4949 and is on a log\log scale. Not shown are the 99 forms with discriminant 321321, 404404, 469469, 473473, or 621621.

Like in Section 4.2, we fit functions of the form either y=log(ali(exp(x)b)))y=\log(a\mathrm{li}(\exp(x)^{b}))) or y=ali(xb)y=a\mathrm{li}(x^{b}) as appropriate to data points of the form either (logX,log𝒞d,D(X))(\log X,\log\mathcal{C}_{d,D}^{\prime}(X)) or (X,𝒞d,D(X))(X,\mathcal{C}_{d,D}^{\prime}(X)), for prime XX between either 11 and 21062\cdot 10^{6} or 10410^{4} and 21062\cdot 10^{6}. The best-fit exponents we obtain in each of these four cases, for D=49,81,148,169,229D=49,81,148,169,229, and 257257, are shown in 4.11.

Data XX range Best-fit exponents by DD
4949 8181 148148 169169 229229 257257
(logX,log𝒞d,D(X))(\log X,\log\mathcal{C}_{d,D}^{\prime}(X)) 1121062\cdot 10^{6} 0.420.42 0.340.34 0.120.12 0.390.39 0.240.24 0.180.18
(logX,log𝒞d,D(X))(\log X,\log\mathcal{C}_{d,D}^{\prime}(X)) 10410^{4}21062\cdot 10^{6} 0.420.42 0.320.32 0.110.11 0.380.38 0.200.20 0.190.19
(X,𝒞d,D(X))(X,\mathcal{C}_{d,D}^{\prime}(X)) 1121062\cdot 10^{6} 0.430.43 0.230.23 0.110.11 0.370.37 0.190.19 0.190.19
(X,𝒞d,D(X))(X,\mathcal{C}_{d,D}^{\prime}(X)) 10410^{4}21062\cdot 10^{6} 0.430.43 0.220.22 0.110.11 0.370.37 0.180.18 0.190.19
Table 4.11.

Best-fit values of bb when fitting functions of the form y=log(ali(exp(x)b)))y=\log(a\mathrm{li}(\exp(x)^{b}))) or y=ali(xb)y=a\mathrm{li}(x^{b}) as appropriate to data of counts of degree 33 newform orbits with prime level and prescribed discriminant

Analogous to the degree 22 case, it is natural to ask:

Question 4.12.

Among squarefree levels NN\to\infty, do 100% of degree 33 newforms in S2(N)S_{2}(N) have rationality field (ζ7)+\mathbb{Q}(\zeta_{7})^{+}?

References

  • [1] Paul T. Bateman and Roger A. Horn. A heuristic asymptotic formula concerning the distribution of prime numbers. Math. Comp., 16:363–367, 1962.
  • [2] Michael A. Bennett, Adela Gherga, and Andrew Rechnitzer. Computing elliptic curves over \mathbb{Q}. Math. Comp., 88(317):1341–1390, 2019.
  • [3] Alex J. Best, Jonathan Bober, Andrew R. Booker, Edgar Costa, John E. Cremona, Maarten Derickx, Min Lee, David Lowry-Duda, David Roe, Andrew V. Sutherland, and John Voight. Computing classical modular forms. In Arithmetic geometry, number theory, and computation, Simons Symp., pages 131–213. Springer, Cham, [2021] ©2021.
  • [4] Lev A. Borisov and Paul E. Gunnells. On Hilbert modular threefolds of discriminant 49. Selecta Math. (N.S.), 19(4):923–947, 2013.
  • [5] Nils Bruin, E. Victor Flynn, Josep González, and Victor Rotger. On finiteness conjectures for endomorphism algebras of abelian surfaces. Math. Proc. Cambridge Philos. Soc., 141(3):383–408, 2006.
  • [6] Armand Brumer. The rank of J0(N)J_{0}(N). Number 228, pages 3, 41–68. 1995. Columbia University Number Theory Seminar (New York, 1992).
  • [7] Armand Brumer and Oisín McGuinness. The behavior of the Mordell-Weil group of elliptic curves. Bull. Amer. Math. Soc. (N.S.), 23(2):375–382, 1990.
  • [8] Alex Cowan. Computing newforms using supersingular isogeny graphs. Res. Number Theory, 8(4):Paper No. 96, 2022.
  • [9] Alex Cowan and Kimball Martin. Moduli for rational genus 2 curves with real multiplication for discriminant 5. J. Théor. Nombres Bordeaux, to appear.
  • [10] J. Cremona. ecdata. https://github.com/JohnCremona/ecdata, January 2023.
  • [11] Luis Victor Dieulefait, Ariel Pacetti, and Panagiotis Tsaknias. On the number of Galois orbits of newforms. J. Eur. Math. Soc. (JEMS), 23(8):2833–2860, 2021.
  • [12] W. Duke and E. Kowalski. A problem of Linnik for elliptic curves and mean-value estimates for automorphic representations. Invent. Math., 139(1):1–39, 2000. With an appendix by Dinakar Ramakrishnan.
  • [13] Noam Elkies. Families of genus-2 curves with 5-torsion, Preprint.
  • [14] Noam Elkies and Abhinav Kumar. K3 surfaces and equations for Hilbert modular surfaces. Algebra Number Theory, 8(10):2297–2411, 2014.
  • [15] Eyal Z. Goren. Lectures on Hilbert modular varieties and modular forms, volume 14 of CRM Monograph Series. American Mathematical Society, Providence, RI, 2002. With the assistance of Marc-Hubert Nicole.
  • [16] H. G. Grundman and L. E. Lippincott. Hilbert modular threefolds of arithmetic genus one. J. Number Theory, 95(1):72–76, 2002.
  • [17] H. G. Grundman and L. E. Lippincott. Computing the arithmetic genus of Hilbert modular fourfolds. Math. Comp., 75(255):1553–1560, 2006.
  • [18] Ki-ichiro Hashimoto. On Brumer’s family of RM-curves of genus two. Tohoku Math. J. (2), 52(4):475–488, 2000.
  • [19] F. Hirzebruch and D. Zagier. Classification of Hilbert modular surfaces. In Complex analysis and algebraic geometry, pages 43–77. Iwanami Shoten, Tokyo, 1977.
  • [20] J. William Hoffman, Zhibin Liang, Yukiko Sakai, and Haohao Wang. Genus 3 curves whose Jacobians have endomorphisms by (ζ7+ζ¯7)\mathbb{Q}(\zeta_{7}+\overline{\zeta}_{7}). J. Symbolic Comput., 74:561–577, 2016.
  • [21] J. William Hoffman and Haohao Wang. 7-gons and genus three hyperelliptic curves. Rev. R. Acad. Cienc. Exactas Fís. Nat. Ser. A Mat. RACSAM, 107(1):35–52, 2013.
  • [22] Koopa Tak-Lun Koo, William Stein, and Gabor Wiese. On the generation of the coefficient field of a newform by a single Hecke eigenvalue. J. Théor. Nombres Bordeaux, 20(2):373–384, 2008.
  • [23] Michael Lipnowski and George J. Schaeffer. Detecting large simple rational Hecke modules for Γ0(N){\Gamma}_{0}(N) via congruences. Int. Math. Res. Not. IMRN, (19):6149–6168, 2020.
  • [24] Qing Liu. Conducteur et discriminant minimal de courbes de genre 22. Compositio Math., 94(1):51–79, 1994.
  • [25] The LMFDB Collaboration. The L-functions and modular forms database. http://www.lmfdb.org, 2023. [Online; accessed 3 January 2023].
  • [26] Kimball Martin. Local conductors bounds for modular abelian varieties. Acta Arith., to appear.
  • [27] Kimball Martin. Refined dimensions of cusp forms, and equidistribution and bias of signs. J. Number Theory, 188:1–17, 2018.
  • [28] Kimball Martin. An on-average Maeda-type conjecture in the level aspect. Proc. Amer. Math. Soc., 149(4):1373–1386, 2021.
  • [29] Kimball Martin and Jordan Wiebe. Zeroes of quaternionic modular forms and central LL-values. J. Number Theory, 217:460–494, 2020.
  • [30] J.-F. Mestre. Familles de courbes hyperelliptiques à multiplications réelles. In Arithmetic algebraic geometry (Texel, 1989), volume 89 of Progr. Math., pages 193–208. Birkhäuser Boston, Boston, MA, 1991.
  • [31] M. Ram Murty and Kaneenika Sinha. Effective equidistribution of eigenvalues of Hecke operators. J. Number Theory, 129(3):681–714, 2009.
  • [32] V. Kumar Murty. Frobenius distributions and Galois representations. In Automorphic forms, automorphic representations, and arithmetic (Fort Worth, TX, 1996), volume 66 of Proc. Sympos. Pure Math., pages 193–211. Amer. Math. Soc., Providence, RI, 1999.
  • [33] David P. Roberts. Newforms with rational coefficients. Ramanujan J., 46(3):835–862, 2018.
  • [34] Peter Sarnak and Nina Zubrilina. Convergence to the Plancherel measure of Hecke eigenvalues, 2022. arXiv 2201.03523.
  • [35] Jean-Pierre Serre. Répartition asymptotique des valeurs propres de l’opérateur de Hecke TpT_{p}. J. Amer. Math. Soc., 10(1):75–102, 1997.
  • [36] Ananth N. Shankar, Arul Shankar, and Xiaoheng Wang. Large families of elliptic curves ordered by conductor. Compos. Math., 157(7):1538–1583, 2021.
  • [37] William A. Stein and Mark Watkins. A database of elliptic curves—first report. In Algorithmic number theory (Sydney, 2002), volume 2369 of Lecture Notes in Comput. Sci., pages 267–275. Springer, Berlin, 2002.
  • [38] Gerard van der Geer. Hilbert modular surfaces, volume 16 of Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)]. Springer-Verlag, Berlin, 1988.
  • [39] Mark Watkins. Some heuristics about elliptic curves. Experiment. Math., 17(1):105–125, 2008.