1,631
Views
3
CrossRef citations to date
0
Altmetric
Original Articles

Special Hypergeometric Motives and Their L-Functions: Asai Recognition

ORCID Icon, ORCID Icon, ORCID Icon & ORCID Icon

Abstract

We recognize certain special hypergeometric motives, related to and inspired by the discoveries of Ramanujan more than a century ago, as arising from Asai L-functions of Hilbert modular forms.

1 Introduction

Motivation

The generalized hypergeometric functions are a familiar player in arithmetic and algebraic geometry. They come quite naturally as periods of certain algebraic varieties, and consequently they encode important information about the invariants of these varieties. Many authors have studied this rich interplay, including Igusa [Citation27], Dwork [Citation13], and Katz [Citation29]. More recently, authors have considered hypergeometric motives (HGMs) defined over Q, including Cohen [Citation8], Beukers–Cohen–Mellit [Citation3], and Roberts–Rodriguez-Villegas–Watkins [Citation37]. A hypergeometric motive over Q arises from a parametric family of varieties with certain periods (conjecturally) satisfying a hypergeometric differential equation; the construction of this family was made explicit by Beukers–Cohen–Mellit [Citation3] based on work of Katz [Citation29]. Following the analogy between periods and point counts (Manin’s “unity of mathematics” [Citation7]), counting points on the reduction of these varieties over finite fields is accomplished via finite field hypergeometric functions, a notion originating in work of Greene [Citation17] and Katz [Citation29]. These finite sums are analogous to truncated hypergeometric series in which Pochhammer symbols are replaced with Gauss sums, and they provide an efficient mechanism for computing the L-functions of hypergeometric motives. (Verifying the precise connection to the hypergeometric differential equation is usually a difficult task, performed only in some particular cases.)

In this paper, we illustrate some features of hypergeometric motives attached to particular arithmetically significant hypergeometric identities for 1/π and 1/π2. To motivate this study, we consider the hypergeometric function (1.1) 3F2(12, 14, 341, 1|z)=n=0(12)n(14)n(34)nn!3 zn,(1.1) where we define the Pochhammer symbol (rising factorial) by (1.2) (α)n:=Γ(α+n)Γ(α)={α(α+1)(α+n1),for n1;1,for n=0.(1.2)

Ramanujan [35, eq. (36)] more than a century ago proved the delightful identity (1.3) n=0(12)n(14)n(34)nn!3(28n+3)(148)n=16π3(1.3) involving a linear combination of the hypergeometric series (1.1) and its z-derivative (a different, but contiguous hypergeometric function). Notice the practicality of this series for computing the quantity on the right-hand side of (1.3), hence for computing 1/π and π itself.

The explanation for the identity (1.3) was already indicated by Ramanujan: the hypergeometric function can be parametrized by modular functions (see (2.4) below), and the value 1/48 arises from evaluation at a complex multiplication (CM) point! Put into the framework above, we observe that the HGM of rank 3 with parameters α={12,14,34} and β={1,1,1} corresponds to the Fermat–Dwork pencil of quartic K3 surfaces of generic Picard rank 19 defined by the equation (1.4) Xψ:x04+x14+x24+x34=4ψx0x1x2x3(1.4) whose transcendental L-function is related to the symmetric square L-function attached to a classical modular form (see Elkies–Schütt [Citation14]). At the specialization z=ψ4=1/48, the K3 surface is singular, having Picard rank 20; it arises as the Kummer surface of E×τ(E), where E is the elliptic Q-curve LMFDB label 144.1-b1 [32] defined over Q(3) attached to the CM order of discriminant –36, and τ(3)=3. The corresponding classical modular form f with LMFDB label 144.2.c.a [Citation32] has CM, and we have the identity (1.5) L(T(X),s)=L(f,s, Sym2)(1.5) where T(X) denotes the transcendental lattice of X (as a Galois representation). The rare event of CM explains the origin of the formula (1.1): for more detail, see Example 3.12 below.

Main result

With this motivation, we seek in this paper to explain similar hypergeometric Ramanujan-type formulas for 1/π2 in higher rank. Drawing a parallel between these examples, our main result is to experimentally identify that the L-function of certain specializations of hypergeometric motives (coming from these formulas) have a rare property: they arise from Asai L-functions of Hilbert modular forms of weight (2, 4) over real quadratic fields.

For example, consider the higher rank analogue (1.6) n=0(12)n(13)n(23)n(14)n(34)nn!5(252n2+63n+5)(148)n=?48π2(1.6) given by Guillera [Citation19]; the question mark above a relation indicates that it has been experimentally observed, but not proven. Here, we suggest that (1.6) is ‘explained’ by the existence of a Hilbert modular form f over Q(12) of weight (2, 4) and level (81) in the sense that we experimentally observe that (1.7) L(H(12,13,23,14,34;1,1,1,1,1 | 1/48),s)=?ζ(s2)L(f,s+1, Asai)(1.7) where notation is explained in Section 3. (By contrast, specializing the hypergeometric L-series at other values tQ generically yields a primitive L-function of degree 5.) Our main result, stated more generally, can be found in Conjecture 5.1.

In spite of a visual similarity between Ramanujan’s 3F2 formula (1.3) for 1/π and Guillera’s 5F4 formula (1.6) for 1/π2, the structure of the underlying hypergeometric motives is somewhat different. Motives attached to 3F2 hypergeometric functions are reasonably well understood (see e.g. Zudilin [40, Observation 4]), and we review them briefly in Section 2. By contrast, the 5F4 motives associated with similar formulas had not been linked explicitly to modular forms. In Conjecture 5.1, we propose that they are related to Hilbert modular forms, and we experimentally establish several other formulas analogous to (1.7).

More generally, for a hypergeometric family, we expect interesting behavior (such as a formula involving periods) when the motivic Galois group at a specialization is smaller than the motivic Galois group at the generic point. We hope that experiments in our setting leading to this kind of explanation will lead to further interesting formulas and, perhaps, a proof.

Organization

The paper is organized as follows. After a bit of setup in Section 2, we quickly review hypergeometric motives in Section 3. In Section 4 we discuss Asai lifts of Hilbert modular forms, then in Section 5 we exhibit the conjectural hypergeometric relations. We conclude in Section 6 with some final remarks.

2 Hypergeometric functions

In this section, we begin with some basic setup. For α1,,αdQ and β1,,βd1Q>0, define the generalized hypergeometric function (2.1) dFd1(α1, α2, , αdβ1, , βd1|z):=n=0(α1)n(α2)n(αd)n(β1)n(βd1)n znn! .(2.1)

These functions possess numerous features that make them unique in the class of special functions. It is convenient to abbreviate (2.1) as (2.2) dFd1(α,β|z)=F(α,β|z),(2.2) where α={α1,,αd} and β={β1,,βd}={β1,,βd1,1} are called the parameters of the hypergeometric function: they are multisets (that is, sets with possibly repeating elements), with the additional element βd=1 introduced to reflect the appearance of n!=(1)n in the denominator in (2.1). The hypergeometric function (2.1) satisfies a linear homogeneous differential equation of order d: (2.3) D(α,β;z):(zj=1d(zddz+αj)j=1d(zddz+βj1))y=0.(2.3)

Among many arithmetic instances of the hypergeometric functions, there are those that can be parameterized by modular functions. One particular example, referenced in the introduction, is (2.4) 3F2(12, 14, 341, 1|ρ(τ))=(η(τ)16η(2τ)8+64 η(2τ)16η(τ)8)1/2,(2.4) for τC with Im(τ)>0, where (2.5) ρ(τ):=256η(τ)24η(2τ)24(η(τ)24+64η(2τ)24)2(2.5) and η(τ)=q1/24j=1(1qj) denotes the Dedekind eta function with q=e2πiτ. Taking the CM point τ=(1+31)/2, we obtain ρ(τ)=1/48 and the evaluation [26, Example 3] (2.6) 3F2(12, 14, 341, 1|148)=2π35/4(Γ(14)Γ(34))2.(2.6)

As indicated by Ramanujan [Citation35], CM evaluations of hypergeometric functions like (2.6) are accompanied by formulas for 1/π, like (1.3) given in the introduction.

Remark 2.7.

Less is known about the conjectured congruence counterpart of (2.6), (2.8) n=0p1(12)n(14)n(34)nn!3(148)n?bp(modp2)(2.8) for primes p5, where (2.9) bp:={2(x2y2)ifp1 (mod 12),p=x2+y2with3|y;(x2y2)ifp5 (mod 12),p=12(x2+y2)with3|y;0ifp3 (mod 4).(2.9)

The congruence (2.8) is in line with a general prediction of Roberts–Rodriguez-Villegas [Citation36], though stated there for z=±1 only.

Ramanujan’s and Ramanujan-type formulas for 1/π corresponding to rational values of z are tabulated in [6, –6]. Known 5F4 identities for 1/π2 are due to Guillera [Citation18–21, Citation23], also in collaboration with Almkvist [Citation1] and Zudilin [Citation25]. We list the corresponding hypergeometric data α and z for them in , we have β={1,1,1,1,1} in all these cases.

Table 1 Hypergeometric data for Guillera’s formulas for 1/π2.

Remark 2.10.

Some other entries in nicely pair up with Ramanujan’s and Ramanujan-type formulas for 1/π [Citation19]. Apart from case #9 from discussed above, we highlight another instance [19, eq. (2-4)]: (2.11) n=0(12)n(13)n(23)n(16)n(56)nn!5(5418n2+693n+29)(1803)n=?1285π2(2.11)

underlying entry #13, which shares similarities with the Ramanujan-type formula (2.12) n=0(12)n(16)n(56)nn!3(5418n+263)(1803)n=640153π.(2.12)

Remark 2.13.

The specialization points z in exhibit significant structure: writing z=a/c and 1z=b/c, so that a + b = c, we already see abc-triples of good quality! But more structure is apparent: see Remark 5.7.

3 Hypergeometric motives

In this section, we quickly introduce the theory of hypergeometric motives over Q.

Definition

Analogous to the generalized hypergeometric function (2.1), a hypergeometric motive is specified by hypergeometric data, consisting of two multisets α={α1,,αd} and β={β1,,βd} with αj,βjQ(0,1] satisfying αβ= and βd=1. Herein, we consider only those hypergeometric motives that are defined over Q, which means that the polynomials (3.1) j=1d(te2πiαj)andj=1d(te2πiβj)(3.1) have coefficients in Z—that is, they are products of cyclotomic polynomials.

Let q be a prime power that is coprime to the least common denominator of αβ, and let Fq be a finite field with q elements. Let ω:Fq×C× be a generator of the character group on Fq×, and let ψq:FqC× be a nontrivial (additive) character. For mZ, define the Gauss sum (3.2) g(m):=xFq×ω(x)mψq(x);(3.2) then g(m) is periodic in m with period q1=#Fq×.

When αj(q1),βj(q1)Z for all j, we define the finite field hypergeometric sum for tFq× by (3.3) Hq(α,β|t)=11qm=0q2ω((1)dt)m·j=1dg(m+αj(q1)) g(mβj(q1))g(αj(q1)) g(βj(q1))(3.3) by direct analogy with the generalized hypergeometric function. More generally, Beukers–Cohen–Mellit [3, Theorem 1.3] have extended this definition to include all prime powers q that are coprime to the least common denominator of αβ.

There exist p1,,pr,q1,,qsZ1 such that (3.4) j=1dxe2πiαjxe2πiβj=j=1r(xpj1)j=1s(xqj1),(3.4) and we define (3.5) M:=p1p1prprq1q1qsqs.(3.5)

Computing the local L-factors at good primes is completely automated in the Magma [Citation4] package of hypergeometric motives.

Motive and L-function

The finite field hypergeometric sums arose in counting points on algebraic varieties over finite fields, and they combine to give motivic L-functions following Beukers–Cohen–Mellit [Citation3], as follows. For a parameter λ, let Vλ be the pencil of varieties in weighted projective space defined by the equations (3.6) x1+x2++xr=y1++ys,λx1p1xrpr=y1q1ysqs(3.6) and subject to xi,yj0. The pencil Vλ is affine and singular [3, Section 5]; in fact, it is smooth outside of λ=1/M, where it acquires an ordinary double point.

Theorem 3.7.

Suppose that gcd(p1,,pr,q1,,qs)=1 and Mλ1. Then there exists a suitable completion Vλ¯ of Vλ such that #Vλ¯(Fq)=Prs(q)+(1)r+s1qmin(r1,s1)H(α,β|Mλ), and where Prs(q)Q(q) is explicitly given.

The completion provided in Theorem 3.7 may still be singular, and a nonsingular completion is not currently known in general; we expect that Vλ¯ has only quotient singularities, and hence behaves like a smooth manifold with respect to rational cohomology, by the nature of the toric (partial) desingularization. In any event, this theorem shows that the sums (3.3) have an explicit connection to arithmetic geometry and complex analysis.

We accordingly define hypergeometric L-functions, as follows. Let Sλ be the set of primes dividing the numerator or denominator in M together with the primes dividing the numerator or denominator of Mλ or Mλ1. A prime pSλ is called good (for α,β,λ). For a good prime p, we define the formal series (3.8) Lp(H(α,β | Mλ),T):=exp(r=1Hpr(α,β | Mλ)Trr)1+TQ[[T]].(3.8)

Corollary 3.2.

For pSλ and λFp×, we have Lp(H(α,β | Mλ),T)Q[T].

Proof.

The zeta function of Vλ¯ over Fp is a rational function by work of Dwork; the exponential series for Prs(q) is also rational, so the result follows from Theorem 3.7. □

Remark 3.10.

In fact, we expect that Lp(H(α,β | Mλ),T)1+TZ[T] is a polynomial of degree d; it should follow from the construction in Theorem 3.7 or from work of Katz [Citation28], but we could not find a published proof. We establish this property in the cases we consider, as a byproduct of our analysis.

Globalizing, we define the incomplete L-series (3.11) LS(H(α,β | Mλ),s)=pSLp(H(α,β | Mλ),ps)1(3.11) a Dirichlet series that converges in a right half-plane, but otherwise remains rather mysterious. Our goal in what follows will be to match such L-functions (coming from geometry, rapidly computable) with L-functions of modular forms in certain cases, so that the former can be completed to inherit the good properties of the latter.

Examples

We conclude this section with two examples.

Example 3.12.

We return to our motivating example, with the parameters α={12,14,34} and β={1,1,1}, we find p1=4 and q1==q4=1. Then eliminating x1 in (3.6) gives Vλ:λ(y1+y2+y3+y4)4=y1y2y3y4; and Theorem 3.7 yields #Vλ¯(Fq)=q31q1+Hq(12,14,34;1,1,1 | 44λ).

We make a change of parameters λ1=44ψ4 and consider the pencil of quartic K3 hypersurfaces with generically smooth fibers defined by (3.13) Xψ:x04+x14+x24+x34=4ψx0x1x2x3(3.13) as in (1.4), with generic Picard rank 19. The family 3.13 is known as the Fermat–Dwork family and is well studied (going back to Dwork [13, §6j, p. 73]; see e.g. Doran–Kelly–Salerno–Sperber–Voight–Whitcher [12, §1.5] for further references). In the context of mirror symmetry, one realizes Vλ¯ as the mirror of Xψ [11, §5.2] in the following way, due to Batyrev: there is an action of G=(Z/4Z)3 on Xψ, and Vλ is birational to Xψ/G. We see again that the finite field hypergeometric sum H(12,14,34;1,1,1 | ψ4) contributes nontrivially to the point counts [12, Main Theorem 1.4.1(a)].

In either model, the holomorphic periods of Vλ¯ or Xψ are given by the hypergeometric series (3.14) F(12,14,34;1,1,1 | ψ4)=n=0(12)n(14)n(34)nn!3 (44λ)n=n=0(4n)!n!4 λn.(3.14)

As mentioned in the introduction, at the specialization ψ4=44λ=1/48, the K3 surface is singular, with Picard number 20—it is this rare event that explains the formula (1.3). Computing the local L-factors, we find Lp(H(12,14,34;1,1,1 | 1/48),T)=(1χ(p)pT)(1bpT+p2T2) for p2,3, where χ(p)=(12p) is the quadratic character attached to Q(12) and bpZ defined in (2.9). Indeed, this factorization agrees with the fact that the global L-series can be completed to L(H(12,14,34;1,1,1 | 1/48),s)=L(f,s, Sym2) where f is the classical modular form with LMFDB label 144.2.c.a [Citation32]: more generally, see Elkies–Schütt [Citation14], Doran–Kelly–Salerno–Sperber–Voight–Whitcher [11, Theorem 5.1.3], or Zudilin [40, Observation 4]. Consequently, the completed hypergeometric L-series inherits analytic continuation and functional equation.

Example 3.15.

We consider the hypergeometric data attached to Ramanujan-type formula (2.11), corresponding to #13 in and with parameters α={12,13,23,16,56} and β={1,,1}. This example is, in many aspects, runs parallel to Example 3.12 and the related mirror symmetry construction of the famous quintic threefold [Citation5]. We have Vλ:λ(y1+y2++y6)6=y1y2y6 and Theorem 3.7 implies #Vλ¯(Fq)=q51q1+H(12,13,23,16,56;1,1,1,1,1 | 66λ).

Alternatively, we consider the pencil of sextic fourfolds Xψ:x06+x16+x26+x36+x46+x56=6ψx0x1x2x3x4x5 in P5. Under the change of parameter λ1=66ψ6, we find that Vλ is birational to Xψ/G where G(Z/6Z)5. The Xψ are generically Calabi–Yau fourfolds. A computation (analogous to Candelas–de la Ossa–Greene–Parks [Citation5]) shows that the Picard–Fuchs differential operator is given by (λddλ)56λ(6λddλ+1)(6λddλ+2)(6λddλ+3)·(6λddλ+4)(6λddλ+5).

The unique (up to scalar) holomorphic solution near zero is the hypergeometric function F(α,β | 66λ)=n=0(12)n(13)n(23)n(16)n(56)nn!5 (66λ)n=n=0(6n)!n!6 λn.

Using the Magma implementation, we compute the first few (good) L-factors: (3.16) L7(T)=(1+72T)(174T2)(135T+74T2)L11(T)=(1112T)(1114T2)(130T+114T2)L13(T)=(1+132T)(164T1758·13T264·134T3+138T4)(3.16) and observe that for p2,3,5, (3.17) Lp(T)=?(1χ5(p)p2T)·(1apT+bppTχ129(p)app4T3+χ129(p)p8T4).(3.17)

Moreover, when χ129(p)=1 then bp = 0 and the quartic polynomial factors as 1apT+bppTχ129(p)app4T3+χ129(p)p8T4=?(1p4T2)(1apT+p4T2) whereas for χ129(p)=1 it is generically irreducible. This suggests again a rare event which we seek to explain using modular forms.

4 The Asai transfer of a Hilbert modular form

Having defined L-functions arising from hypergeometric motives in the previous sections, over the next two sections we follow the predictions of the Langlands philosophy and seek to identify these L-functions as coming from modular forms in the cases of interest. More precisely, we confirm experimentally a match with the Asai transfer of certain Hilbert modular forms over quadratic fields. We begin in this section by setting up the needed notation and background. As general references for Hilbert modular forms, consult Freitag [Citation15] or van der Geer [Citation16]; for a computational take, see Dembélé–Voight [Citation9].

Let F=Q(d) be a real quadratic field of discriminant d > 0 with ring of integers ZF and Galois group Gal(F | Q)=τ. By a prime of ZF we mean a nonzero prime ideal pZF. Let v1,v2:F,R be the two embeddings of F into R. For xF we write xi:=vi(x), and for γ M2(F) we write γi=vi(γ) for the coordinate-wise application of vi. An element aF× is totally positive if v1(a),v2(a)>0; we write F>0× for the group of totally positive elements. The group (4.1) GL2+(F):={γGL2(F):detγF>0×}(4.1) acts on the product H×H of upper half-planes by embedding-wise linear fractional transformations γ(z):=(γ1(z1),γ2(z2)).

Let k1,k22Z>0, write k:=(k1,k2), and let k0:=max(k1,k2) and w0:=k01. Let NZF be a nonzero ideal. Let Sk(N;ψ) denote the (finite-dimensional) C-vector space of Hilbert cusp forms of weight k, level Γ0(N), and central character ψ. Hilbert cusp forms are the analogue of classical cusp forms, but over the real quadratic field F. When the narrow class number of F is equal to 1 (i.e., every nonzero ideal of ZF is principal, generated by a totally positive element) and ψ is the trivial character, a Hilbert cusp form fSk(N) is a holomorphic function f:H×HC, vanishing at the cusps, such that (4.2) f(γz)=(c1z1+d1)k1/2(c2z2+d2)k2/2f(z)(4.2) for all γ=(abcd)GL2+(ZF) such that cN.

The space Sk(N) is equipped with an action of pairwise commuting Hecke operators Tp indexed by nonzero primes pN. A Hilbert cusp form f is a newform if f is an eigenform for all Hecke operators and f does not arise from Sk(M) with M|N a proper divisor.

Let fSk(N) be a newform. For pN, we have Tpf=apf with apC a totally real algebraic integer (the Hecke eigenvalue), and we factor 1apT+Nm(p)w0T2=(1αpT)(1βpT)C[T] where Nm(p) is the absolute norm. Then |αp|=|βp|=pw0.

For pZ prime with pNm(N), following Asai [Citation2] we define, abbreviating p=τ(p), (4.3) Lp(f,T, Asai):={(1αpαpT)(1αpβpT)(1αpβpT)(1βpβpT),if pZF=pp splits;(1αpT)(1βpT)(1ψ(p)p2w0T2),if pZF=p is inert;(1αp2T)(1βp2T)(1ψ(p)pw0T),if pZF=p2 ramifies.(4.3)

We call the factors Lp(f,T, Asai) the good L-factors of f. The partial Asai L-function of f is the Dirichlet series defined by the Euler product (4.4) LS(f,s, Asai):=pSLp(f,ps, Asai)1(4.4) where S={p:p|Nm(N)}.

The key input we need is the following theorem. For a newform fSk(N,ψ), let τ(f) be the newform of weight (k2, k1) and level τ(N) with Tpτ(f)=aτ(p)τ(f), with central character ψ°τ. Finally, for central character ψ (of the idele class group of F) let ψ0 denote its restriction (to the ideles of Q).

Theorem 4.5

(Krishnamurty [Citation31], Ramakrishnan [Citation34]). Let fSk(N,ψ) be a Hilbert newform, and suppose that τ(f) is not a twist of f. Then the partial L-function LS(f,s, Asai) can be completed to a Q-automorphic L-function Λ(f,s, Asai)=Ns/2ΓC(s)2L(f,s, Asai) of degree 4, conductor NZ>0, with central character ψ02.

More precisely, there exists a cuspidal automorphic representation Π=Π(pΠp) of GL4(AQ) such that Lp(f,ps, Asai)=L(s,Πp)1 for all pNm(N). In particular, L(f,s, Asai) is entire and satisfies a functional equation Λ(s)=εΛ¯(8s) with |ε|=1.

The automorphic representation Π in Theorem 4.5 goes by the name Asai transfer, Asai lift, or tensor induction of the automorphic representation π attached to f, and we write Π= Asai(π).

Proof.

We may identify GL2(AF)ResF|QGL2(AQ), with L-group (4.6) L(ResF|QGL2)GL2(C)×GL2(C)GalQ,(4.6) where GalQ:=Gal(Qal | Q). We define the 4-dimensional representation (4.7) r:GL2(C)×GL2(C)GalQGL(C2C2)GL4(C)r(g1,g2,σ)={g1g2, if σ|F=id;g2g1, if σ|F=τ.(4.7)

For a place v of Q, let rv be the restriction of r to GL2(C)×GL2(C)GalQv.

Let π=π(pπp) be the cuspidal automorphic representation of GL2(AF) attached to f. Then πp is an admissible representation of GL2(FQp) corresponding to an L-parameter (4.8) ϕp:WpGL2(C)×GL2(C)GalQp,(4.8) where Wp is the Weil–Deligne group of Qp. We define  Asai(πp) to be the irreducible admissible representation of GL4(Qp) attached to rp°ϕp by the local Langlands correspondence, and we combine these to (4.9)  Asai(π):= Asai(π)p Asai(πp).(4.9)

By a theorem of Ramakrishnan [34, Theorem D] or Krishnamurty [31, Theorem 6.7], Asai(π) is an automorphic representation of GL4(AQ) whose L-function is defined by (4.10) L(s,π, Asai):=L(s,π,r°ϕ)pL(s,πp,rp°ϕp)(4.10) whose good L-factors agree with (4.3) [31, §4]. Under the hypothesis that τ(f) is not a twist of f, we conclude that Asai(π) is cuspidal [34, Theorem D(b)]. Consequently, we may take Π= Asai(π) in the theorem. □

Remark 4.11.

Some authors also define the representation Asai(π), which is the quadratic twist of Asai(π) by the quadratic character attached to F.

In addition to the direct construction (4.4) and the automorphic realization in Theorem 4.5, one can also realize the Asai L-function via Galois representations. By Taylor [38, Theorem 1.2], attached to f is a Galois representation ρ:GalFGL(V)GL2(Qlal)

such that for each prime pN, we have det(1ρ(Frobp)T)=1apT+Nm(p)w0T2.

Then there is a natural extension of ρ to GalQ, a special case of multiplicative induction (or tensor induction) [33, §7] defined as follows: for a lift of τ to GalQ which by abuse is also denoted τ, we define [30, p. 1363] (taking a left action) (4.12)  Asai(ρ):GalQGL(VV)GL4(Qlal) Asai(ρ)(σ)(xy)={ρ(σ)(x)ρ(τ1στ)(y),if σ|F=id;ρ(στ)(y)ρ(τ1σ)(x),if σ|F=τ|F.(4.12)

Up to isomorphism, this representation does not depend on the choice of lift τ. A direct computation [30, Lemma 3.3.1] then verifies that det(1 Asai(ρ)(Frobp)T)=Lp(f,T, Asai) as defined in (4.4).

The bad L-factors Lp(f,T, Asai) and conductor N of L(f,s, Asai) are uniquely determined by the good L-factors, but they are not always straightforward to compute.

5 Matching the hypergeometric and Asai L-functions

We now turn to the main conjecture of this paper.

Main conjecture

We propose the following conjecture.

Conjecture 5.1. Let α be a set of parameters from and let β={1,1,1,1,1}. Then there exist quadratic Dirichlet characters χ,ε and a Hilbert cusp form f over a real quadratic field F of weight (2, 4) such that for all good primes p we have Lp(H(α,β | z),T)=?(1χ(p)p2T)Lp(f,T/p, Asai,ε).

In particular, we have the identity L(H(α,β | z),s)=?L(s2,χ)L(f,s+1, Asai,ε).

We can be more precise in Conjecture 5.1 for some of the rows, as follows. Let #n be a row in with n7,13,14,15,16. Then we conjecture that the central character ψ of f is a quadratic character of the class group of F induced from a Dirichlet character; and the conductors of χ,ε,ψ, the discriminant dF of F, and the level N of f are indicated in .

Table 2 Hilbert modular form data.

Evidence

We verified Conjecture 5.1 for the complete rows indicated in using Magma [Citation4]; the algorithms for hypergeometric motives were implemented by Watkins, algorithms for L-functions implemented by Tim Dokchitser, and algorithms for Hilbert modular forms by Dembélé, Donnelly, Kirschmer, and Voight. The code is available online [Citation10].

Moreover, using the L-factor data in , we have confirmed the functional equation for L(H(α,β | z),s) up to 20 decimal digits for all but #13. When the discriminant dF and the level N are coprime, we observe that the conductor of L(f,s, Asai) is N=dFNm(N).

Table 3 L-factor data for Lp(H(α,β | z),T)=?Lp(χ,p2T)Lp(f,T/p, Asai,ε).

Remark 5.2.

In a recent arithmetic study of his formulas for 1/π2, Guillera [Citation24] comes up with an explicit recipe to cook up the two quadratic characters for each such formula. He calls them χ0 and ε0 and records them in [24, ]. Quite surprisingly, they coincide with our χ and ε in .

Example 5.3.

Consider row #1. In the space S(2,4)(4) of Hilbert cusp forms over F=Q(5) of weight (2, 4) and level (4) with trivial central character, we find a unique newform f with first few Hecke eigenvalues a(2)=0,a(3)=30,a(5)=10,a(7)=70, and ap,aτ(p)=12±85, giving for example L3(f,T, Asai)=(136T2)(1+10·3T+36T2);

we then match L3(H(12,,12;1,,1 | 1/22),T)=(132T)L3(f,T/3, Asai).

We matched L-factors for all good primes p such that a prime p of F lying over p has Nm(p)200.

Example 5.4.

For row #9, the space of Hilbert cusp forms over F=Q(12) of weight (2, 4) and level N=(81) has dimension 2186 with a newspace of dimension 972. We find a form f with Hecke eigenvalues a(5)=140,a(7)=98, …; accordingly, we find (5.5) L5(H(α,β | 1/48),T)=(152T)L5(f,T/5, Asai)=(152T)2(1+52T)(1+28T+625T2),(5.5) and so on. We again matched Hecke eigenvalues up to prime norm 200.

Remark 5.6.

To match row #16 in with a candidate Hilbert modular form, we would need to extend the implementation of hypergeometric motives to apply for specialization at points zQ; we expect this extension to be straightforward, given the current implementation of finite field hypergeometric sums.

By contrast, to match the final rows #17 and #13–#15, we run into difficulty with computing spaces of Hilbert modular forms: we looked for forms in low level, but the dimensions grow too quickly with the level. We also currently lack the ability to efficiently compute with arbitrary nontrivial central character. We plan to return to these examples with a new approach to computing systems of Hecke eigenvalues for Hilbert modular forms in future work.

Remark 5.7. Returning to Remark 2.13, we observe structure in the specialization points z from : beyond patterns in the factorization of z and 1z, we also note that for these points the completed L-function typically has unusually small conductor N, as in . (Perhaps a twist of #15 has smaller conductor?) Some general observations that may explain this conductor drop:

  • Factor N=N1N2 where N1 consists of the product of primes p|N that divide the least common denominator of α or the numerator or denominator of z. Then N2 should be the squarefree part of the numerator of 1z; this numerator is divisible by a nontrivial square in ten of the fifteen cases.

  • The power of p dividing the numerator or denominator of z is itself a multiple of p for most primes p dividing a denominator in α.

  • For a prime p, define sp(α)=0 if α is coprime to p and otherwise let sp(α)=ordp(α)+1/(p1). If ordp(z) is a multiple of j=15sp(αj), then ordp(N) tends to be especially small.

These last two phenomena were first observed by Rodriguez-Villegas; we thank the referee for these observations.

While not making any assertions about completeness, these observations give some indication of why our is so short: the specialization points z like those listed are quite rare, and they seem to depend on a pleasing but remarkable arithmetic confluence. It would be certainly valuable to be able to predict more generally and precisely the conductor of hypergeometric L-functions.

Method

We now discuss the recipe by which we found a match. For simplicity, we exclude the case #8 and suppose that the central character ψ is trivial. In a nutshell, our method uses good split ordinary primes to recover the Hecke eigenvalues up to sign.

We start with the hypergeometric motive and compute Lp(H,T):=Lp(H(α,β | z),T) for many good primes p. We first guess χ and dF by factoring Lp(H,T)=(1χ(p)p2T)Qp(T): for primes p that are split in F, we usually have Qp(T) irreducible whereas and for inert primes we find (1p4T2)|Qp(T). We observe in many cases that dF is (up to squares) the numerator of 1z. Combining this information gives us a good guess for χ and dF.

We now try to guess the Hecke eigenvalues of a candidate Hilbert newform f of weight (2, 4). Let p=pτ(p) be a good split prime, and suppose that p is ordinary for f, i.e., the normalized valuations ordp(ap), ordp(aτ(p))=0,1 are as small as possible, or equivalently, factoring (5.8) Lp(f,T)=1apT+p3T2=(1αpT)(1βpT)Lτ(p)(f,T)=1aτ(p)T+p3T2=(1ατ(p)T)(1βτ(p)T)(5.8) we may choose p so that αp,ατ(p)/p are p-adic units. We expect that such primes will be abundant, though that seems difficult to prove. Then Lp(f,T, Asai) has Hodge–Tate weights (i.e., reciprocal roots with valuations) (0,3)(1,2)=(1,2,4,5) (adding pairwise) so the Tate twist Lp(f,T/p, Asai) has Hodge–Tate weights (0, 1, 3, 4) and coefficients with valuations 0, 0, 1, 3, 4, 8, matching that of the hypergeometric motive.

So we factor Qp(T) over the p-adic numbers, identifying ordinary p when the roots δ0,δ1,δ3,δ4 have corresponding valuations 0, 1, 3, 4. Then we have the equations (5.9) pδ0=αpατ(p)pδ1=αpβτ(p)(5.9) and two similar equations for δ3,δ4. Therefore (5.10) p2δ0δ1=αp2ατ(p)βτ(p)=αp2p3(5.10) so (5.11) αp=±δ0δ1p;(5.11) and this determines the Hecke eigenvalue (5.12) ap=αp+βp=αp+p3/αp(5.12) up to sign.

We then go hunting in Magma by slowly increasing the level and looking for newforms whose Hecke eigenvalues match the value ap in (5.12) up to sign. With a candidate in hand, we then compute all good L-factors using (4.3) to identify a precise match. The bottleneck in this approach is the computation of systems of Hecke eigenvalues for Hilbert modular forms.

6 Conclusion

The 1/π story brings many more puzzles into investigation, as formulas discussed in this note do not exhaust the full set of mysteries. Some of them are associated with the special 4F3 evaluations of 1/π, like the intermediate one in the trio (6.1) n=0(12)n(14)n(34)nn!3(40n+3)174n=4939π,(6.1) (6.2) n=0(18)n(38)n(58)n(78)nn!3(32)n(1920n2+1072n+55)174n=19673π,(6.2) (6.3) n=0(12)n(18)n(38)n(58)n(78)nn!5(1920n2+304n+15)174n=?567π2.(6.3)

Here the first equation is from Ramanujan’s list [35, eq. (42)], the second one is recently established by Guillera [22, Equationeq. (1.6)], while the third one corresponds to Entry #15 in and is given in [19, eq. (2-5)]. There is also one formula for 1/π3, due to B. Gourevich (2002), (6.4) n=0(12)n7n!7(168n3+76n2+14n+1)126n=?32π3,(6.4) which shares similarities with Ramanujan's [35, eq. (29)] (6.5) n=0(12)n3n!3(42n+5)126n=16π(6.5)

(observe that 168 = 42 × 4). And the pattern extends even further with the support of the experimental findings (6.6) n=0(12)n7(14)n(34)nn!9·(43680n4+20632n3+4340n2+466n+21)1212n=?2048π4,(6.6) due to J. Cullen (December 2010), and (6.7) n=0(12)n5(13)n(23)n(14)n(34)nn!9·(4528n4+3180n3+972n2+147n+9)(27256)n=?768π4,(6.7) due to Yue Zhao [Citation39] (September 2017). On the top of these examples there are ‘divergent’ hypergeometric formulas for 1/π3 and 1/π4 coming from ‘reversing’ Zhao’s experimental formulas for π4 and ζ(5) in [Citation39], and corresponding to the hypergeometric data 7F6(12, 12, 12, 12, 12, 13, 231, , 1|3322)and9F8(12, 12, 12, 12, 12, 15, 25, 35, 451, , 1|55210), respectively. We hope to address the arithmetic-geometric origins of the underlying motives in the near future.

Simons Foundation10.13039/100000893;

It is our pleasure to thank Frits Beukers, Henri Cohen, Vasily Golyshev, Jesús Guillera, Günter Harder, Yuri Manin, Anton Mellit, David Roberts, Alexander Varchenko, Fernando Rodriguez-Villegas, Mark Watkins and Don Zagier for valuable feedback, stimulating discussions, and crucial observations. The authors would also like to thank the anonymous referee for their stimulating report, insightful comments, and helpful suggestions.

Acknowledgments

This project commenced during Zudilin’s visit in the Fourier Institute in Grenoble in June 2017, followed by the joint visit of Dembélé, Panchishkin, and Zudilin to the Max Planck Institute for Mathematics in Bonn in July 2017, followed by collaboration between Voight and Zudilin during the trimester on Periods in Number Theory, Algebraic Geometry and Physics at the Hausdorff Research Institute for Mathematics in Bonn in March–April 2018. We thank the staff of these institutes for providing such excellent conditions for research.

Declaration of interest

No potential conflict of interest was reported by the author(s).

Correction Statement

This article has been republished with minor changes. These changes do not impact the academic content of the article.

Additional information

Funding

Dembélé and Voight were supported by a Simons Collaboration grant (550029, to Voight).

References

  • Almkvist, G., Guillera, J. (2012). Ramanujan-like series for 1/π2 and string theory. Exp. Math. 21(3): 223–234. doi:10.1080/10586458.2012.656059
  • Asai, T. (1977). On certain Dirichlet series associated with Hilbert modular forms and Rankin’s method. Math. Ann. 226(1): 81–94. doi:10.1007/BF01391220
  • Beukers, F., Cohen, H., Mellit, A. (2015). Finite hypergeometric functions. Pure Appl. Math. Q. 11(4): 559–589. doi:10.4310/PAMQ.2015.v11.n4.a2
  • Bosma, W., Cannon, J., Playoust, C. (1997). The Magma algebra system. I. The user language. J. Symbolic Comput. 24(3–4): 235–265. doi:10.1006/jsco.1996.0125
  • Candelas, P., de la Ossa, X., Green, P. S., Parkes, L. (1991). A pair of Calabi–Yau manifolds as an exactly soluble superconformal theory. Nuclear Phys. B. 359(1): 21–74. doi:10.1016/0550-3213(91)90292-6
  • Chan, H. H., Cooper, S. (2012). Rational analogues of Ramanujan’s series for 1/π. Math. Proc. Camb. Phil. Soc. 153(2): 361–383. doi:10.1017/S0305004112000254
  • Clemens, C. H. (2003). A scrapbook of complex curve theory. 2nd ed., Vol. 55. Providence, RI: Graduate Studies in Mathematics, American Mathematical Society.
  • Cohen, H. (2015). Computing L-functions: A survey. J. Théor. Nombres Bordeaux. 27(3): 699–726. doi:10.5802/jtnb.920
  • Dembélé, L., Voight, J. (2013). Explicit methods for Hilbert modular forms. In: Darmon, H., Diamond, F., Dieulefait, L. V., Edixhoven, B., Rotger, V., eds. Elliptic Curves, Hilbert Modular Forms and Galois Deformations. Basel: Advanced Courses in Mathematics CRM Barcelona. Birkhäuser/Springer, pp. 135–198. doi:10.1007/978-3-0348-0618-3_4
  • Dembélé, L., Voight, J. (2019). Code for Asai recognition. Available at: http://www.math.dartmouth.edu/∼jvoight/HGM_Asai.m
  • Doran, C. F., Kelly, T. L., Salerno, A., Sperber, S., Voight, J., Whitcher, U. (2018). Zeta functions of alternate mirror Calabi–Yau families. Isr. J. Math. 228(2): 665–705. doi:10.1007/s11856-018-1783-0
  • Doran, C. F., Kelly, T. L., Salerno, A., Sperber, S., Voight, J., Whitcher, U. (2020). Hypergeometric decomposition of symmetric K3 quartic pencils. Res. Math. Sci. 7(2): Art. 7. doi:10.1007/s40687-020-0203-3
  • Dwork, B. (1969). p-adic cycles. Inst. Hautes Études Sci. Publ. Math. 37(1): 27–115. doi:10.1007/BF02684886
  • Elkies, N. D., Schütt, M. (2008). K3 families of high Picard rank. Available at: http://www2.iag.uni-hannover.de/∼schuett/K3-fam.pdf
  • Freitag, E. (1990). Hilbert Modular Forms. Berlin: Springer-Verlag.
  • van der Geer, G. (1988), Hilbert Modular Surfaces. Berlin: Springer-Verlag.
  • Greene, J. (1987). Hypergeometric functions over finite fields. Trans. Amer. Math. Soc. 301(1): 77–101. doi:10.1090/S0002-9947-1987-0879564-8
  • Guillera, J. (2002). Some binomial series obtained by the WZ-method. Adv. Appl. Math. 29(4): 599–603. doi:10.1016/S0196-8858(02)00034-9
  • Guillera, J. (2003). About a new kind of Ramanujan-type series. Exp. Math. 12(4): 507–510. doi:10.1080/10586458.2003.10504518
  • Guillera, J. (2006). Generators of some Ramanujan formulas. Ramanujan J. 11(1): 41–48. doi:10.1007/s11139-006-5306-y
  • Guillera, J. (2011). A new Ramanujan-like series for 1/π2. Ramanujan J. 26(3): 369–374. doi:10.1007/s11139-010-9259-9
  • Guillera, J. (2017). More Ramanujan–Orr formulas for 1/π. New Zealand J. Math. 47: 151–160.
  • Guillera, J. (2019). Bilateral sums related to Ramanujan-like series. Preprint arXiv:1610.04839v2 [math.NT], 13 pp.
  • Guillera, J. (2019). Bilateral Ramanujan-like series for 1/πk and their congruences. Preprint arXiv:1908.05123 [math.NT].
  • Guillera, J., Zudilin, W. (2012). “Divergent” Ramanujan-type supercongruences. Proc. Amer. Math. Soc. 140(3): 765–777. 9939-2011-10950-X doi:10.1090/S0002-
  • Guillera, J., Zudilin, W. (2013). Ramanujan-type formulae for1/π: the art of translation. In: The Legacy of Srinivasa Ramanujan, B. C. Berndt, D. Prasad, eds., Ramanujan Math. Soc. Lecture Notes Ser. Vol. 20, pp. 181–195, Ramanujan Math. Soc., Mysore.
  • Igusa, J. (1958). Class number of a definite quaternion with prime discriminant. Proc. Nat. Acad. Sci. USA. 44(4): 312–314. doi:10.1073/pnas.44.4.312
  • Katz, N. M. (1976). p-Adic interpolation of real analytic Eisenstein series. Ann. of Math. 104(3): 459–571. doi:10.2307/1970966
  • Katz, N. M. (1990). Exponential sums and differential equations. In: Annals of Mathematics Studies. Vol. 124. Princeton, NJ: Princeton University Press.
  • Krishnamurthy, M. (2012). Determination of cusp forms on GL(2) by coefficients restricted to quadratic subfields (with an appendix by Dipendra Prasad and Dinakar Ramakrishnan). J. Number Theory. 132(6): 1359–1384. doi:10.1016/j.jnt.2011.09.014
  • Krishnamurthy, M. (2003). The Asai transfer to GL4 via the Langlands-Shahidi method. Int. Math. Res. Not. 2003(41): 2221–2254. doi:10.1155/S1073792803130528
  • The LMFDB Collaboration. (2019). The L-functions and modular forms database. Available at: http://www.lmfdb.org.
  • Prasad, D. (1992). Invariant forms for representations of GL2 over a local field. Amer. J. Math. 114(6): 1317–1363. doi:10.2307/2374764
  • Ramakrishnan, D. (2002). Modularity of solvable Artin representations of GO(4)-type. Int. Math. Res. Not. 2002(1): 1–54. doi:10.1155/S1073792802000016
  • Ramanujan, S. (1914). Modular equations and approximations to π. Quart. J. Math. Oxford Ser. 45(2): 350–372. Reprinted in: Hardy, G. H., Sechu Aiyar, P. V., Wilson, B. M. (eds.), Collected papers of Srinivasa Ramanujan. Cambridge University Press, Cambridge (1927) & Chelsea Publ., New York (1962), 23–39.
  • Roberts, D. P., Rodriguez-Villegas, F. (2019). Hypergeometric supercongruences. In: Wood, D. R., de Gier, J., Praeger, C. E., Tao, T. (eds), 2017 MATRIX Annals, MATRIX Book Ser. 2, Springer, Cham, pp. 435–439.
  • Roberts, D., Rodriguez-Villegas, F., Watkins, M., Hypergeometric motives, in preparation.
  • Taylor, R. (1989). On Galois representations associated to Hilbert modular forms. Invent. Math. 98(2): 265–280. doi:10.1007/BF01388853
  • Zhao, Y. (2017). A mysterious connection between Ramanujan-type formulas for 1/πk and hypergeometric motives, September. Available at: http://mathoverflow.net/questions/281009/
  • Zudilin, W. (2018). A hypergeometric version of the modularity of rigid Calabi–Yau manifolds. SIGMA. 14: 086. doi:10.3842/SIGMA.2018.086