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

On the mixed-twist construction and monodromy of associated Picard-Fuchs systems

Andreas Malmendier Dept.​ of Mathematics, University of Connecticut, Storrs, Connecticut 06269
    Dept.​ of Mathematics & Statistics, Utah State University, Logan, UT 84322
[email protected]
 and  Michael T. Schultz Dept.​ of Mathematics, Virginia Tech, Blacksburg, VA 24060 [email protected]
Abstract.

We use the mixed-twist construction of Doran and Malmendier to obtain a multi-parameter family of K3 surfaces of Picard rank ρ16\rho\geq 16. Upon identifying a particular Jacobian elliptic fibration on its general member, we determine the lattice polarization and the Picard-Fuchs system for the family. We construct a sequence of restrictions that lead to extensions of the polarization by two-elementary lattices. We show that the Picard-Fuchs operators for the restricted families coincide with known resonant hypergeometric systems. Second, for the one-parameter mirror families of deformed Fermat hypersurfaces we show that the mixed-twist construction produces a non-resonant GKZ system for which a basis of solutions in the form of absolutely convergent Mellin-Barnes integrals exists whose monodromy we compute explicitly.

Key words and phrases:
K3 surfaces, Picard-Fuchs equations, Euler integral transform
2020 Mathematics Subject Classification:
14J27, 14J28, 14J32, 32Q25; 14D05, 33C60
A.M. acknowledges support from the Simons Foundation through grant no. 202367.

1. Introduction

In [22], Doran and Malmendier introduced the mixed-twist construction, which iteratively constructs families of Jacobian elliptic Calabi-Yau nn-folds Y(n)Y^{(n)} from a family of Jacobian elliptic Calabi-Yau (n1)(n-1)-folds Y(n1)Y^{(n-1)} for all n2n\geq 2. In fact, the new families are then fibered by the Calabi-Yau (n1)(n-1)-folds Y(n1)Y^{(n-1)} in addition to being elliptically fibered. For example, for n=2n=2 the procedure starts with a family of elliptic curves with rational total space, and the mixed-twist construction returns families of Jacobian elliptic K3 surfaces polarized by the lattice HE8(1)E8(1)2kH\oplus E_{8}(-1)\oplus E_{8}(-1)\oplus\langle-2k\rangle for certain kk\in\mathbb{N}. The central tool of the construction, which is inspired by aspects of physics related to mirror symmetry and the embedding of F theory into gauge theory, is an invariant for ramified covering maps 11\mathbb{P}^{1}\to\mathbb{P}^{1}, called the generalized functional invariant.

Central to the mixed-twist construction is the incarnation of an iterative relation between the period integrals of nn-folds Y(n)Y^{(n)} and the periods of Y(n1)Y^{(n-1)}. When applied to the the family of mirror manifolds Y(n1)Y^{(n-1)} of the family of deformed Fermat hypersurfaces X(n1)X^{(n-1)} in n\mathbb{P}^{n}

X0n+1++Xnn+1+nλX0X1Xn=0X_{0}^{n+1}+\cdots+X_{n}^{n+1}+n\lambda X_{0}X_{1}\cdots X_{n}=0

obtained by the Greene-Plesser orbifolding construction [26], Doran and Malmendier proved the existence of transcendental cycles  Σn1Hn1(Y(n1),)\Sigma_{n-1}\in H_{n-1}(Y^{(n-1)},\mathbb{Q}) such that the period integral

ωn1=Σn1η(n1)\omega_{n-1}=\int_{\Sigma_{n-1}}\,\eta^{(n-1)}

can be computed iteratively from the Hadamard product of the hypergeometric function Fn1n{}_{n}F_{n-1} and the period integral ωn2\omega_{n-2} on Y(n2)Y^{(n-2)} [22, Prop. 7.2]. Here, η(n1)\eta^{(n-1)} is a holomorphic trivializing section of the canonical bundle KY(n1)K_{Y^{(n-1)}}. We recall this result explicitly in Proposition 4.2. This result matches well known results in the literature on the periods of the mirror family Y(n1)Y^{(n-1)}, but elucidates the connection between the periods and the iterative fibration structure.

In such a situation, of particular interest are the Picard-Fuchs operators that annihilate the periods ωn1\omega_{n-1}, and the monodromy behavior of the periods as one encircles singular points in family of Calabi-Yau varieties Y(n1)Y^{(n-1)}. In the context of mirror symmetry, the Picard-Fuchs operators are often realized as resonant GKZ hypergeometric systems [34, 68] – named after the seminal work by Gel’fand, Kapranov, and Zelevinsky [24] – a vast generalization of the hypergeometric function Fn1n{}_{n}F_{n-1}. Due to resonance of these systems, the monodromy representations are reducible due to a result of Schulze and Walther [66], which makes their explicit determination much more challenging in general. In the case described above, the monodromy group of the hypergeometric Fuchsian ODE annihilating Fn1n{}_{n}F_{n-1} is known, going back to work of Levelt [45]. The mixed-twist construction offers an alternative formulation to arrive at the same monodromy group (up to conjugacy) based off the iterative period relation.

This article aims to demonstrate that the mixed twist construction is a suitable tool that allows for the computation of the monodromy group of resonant GKZ systems that arise in mirror symmetry and other contexts in algebraic geometry. We apply the mixed-twist construction in two distinct arenas, for constructing multi-parameter families of lattice polarized K3 surfaces, and the mirror family of Calabi-Yau nn-folds Y(n)Y^{(n)} described above. Our approach in each case differs in somewhat major ways.

In the former, we utilize the geometry of K3 surface constructed through the mixed-twist construction to connect to some known results in the literature, allowing us to determine the monodromy group. In particular, since the K3 surface is presented explicitly as a Jacobian elliptic fibration, the mixed-twist construction that we apply to a certain family of elliptic curves with rational total space coincides with the well known quadratic twist construction in the theory of elliptic surfaces. From the perspective of lattice polarizations, this construction is nontrivial. We prove that the new family of K3 surfaces is birationally equivalent to a family of double-sextic K3 surfaces, obtained from the minimal resolution of a double cover of 2\mathbb{P}^{2} branched along six lines (for example, studied in [52, 53, 54, 13, 49]). From here, we identify the lattice polarization LL for the family, and determine the global monodromy group, and the Picard-Fuchs system, the latter two being determined by the Aomoto-Gel’fand system E(3,6)E(3,6), as studied in [52, 53, 54]. In particular, this system is a multi-parameter resonant GKZ hypergeometric system. We naturally determine the parameter space of this family as the complement of the singular locus of the fibration. Morevover, the structure of the fibration allows us to consider natural sub-varieties of the parameter space of double-sextics where the Picard-Fuchs system restricts to known lower-rank systems of resonant hypergeometric type. In each case, the global monodromy group is determined by connecting our family to known results in K3 geometry. We then show that these restrictions lead to extensions of the lattice polarization in a chain of even, indefinite, two-elementary lattices. In this way, we are able to unify central analytical aspects for resonant generalized hypergeometric functions with geometric and lattice theoretic investigations by Hoyt [37, 38] and Hoyt and Schwarz [39].

In the second case, we look at an application of the mixed-twist construction in the context of the mirror families for the deformed Fermat pencils as outlined above. In fact, in this context the mixed-twist construction returns the mirror family of Calabi-Yau nn-folds in n+1\mathbb{P}^{n+1} fibered by mirror Calabi-Yau (n1)(n-1)-folds. In this framework, the set of periods generates a set of resonant GKZ data, which makes the analysis of the behavior of the family near regular singular points quite difficult [68]. However, we show that the mixed-twist construction also generates a second set of non-resonant GKZ data associated with the holomorphic periods, which allows us to compute the explicit monodromy matrices for the mirror families. This second part generalizes work of Chen et al. [8] where the authors constructed the monodromy group of the Picard-Fuchs differential equations associated with the one-parameter families of Calabi-Yau threefolds from Doran and Morgan  [23].

We remark that the Picard-Fuchs operators for the families of mirror Calabi-Yau nn-folds has been known since at least the work of Corti & Golyshev [15]. Our approach in this article is novel in the sense that it is inpired by the physics - in particular, by connections between effective Yang-Mills gauge theory (i.e., Seiberg-Witten theory) and string compactifications on Calabi-Yau varieties. The mixed-twist construction offers a potential to connect computations in these two realms, by geometrizing a link between families of elliptic curves and their Picard-Fuchs operators, and families of Calabi-Yau varieties and their Picard-Fuchs operators via the iterative period relation described above. In addition, the mixed-twist construction provides a mechanism by which to construct transcendental cycles on Calabi-Yau varieties. This allows for the description of the period integrals in terms of 𝒜\mathcal{A}-hypergeometric functions. This approach was utilized, for example, by Clingher, Doran, & Malmendier in [9] to obtain a description of the periods of so-called generalized Kummer surfaces in terms of Appell’s bivariate F2F_{2} hypergeometric function.

Our approach in this article is summarized as follows: in the first part we construct and analyze a family that generalizes the family of K3 surfaces whose rank-19 polarizing lattice is HD16(1)A1(1)H\oplus D_{16}(-1)\oplus A_{1}(-1) and whose Picard-Fuchs operator is the hypergeometric differential equation for F23(12,12,12;1,1|){}_{3}F_{2}(\frac{1}{2},\frac{1}{2},\frac{1}{2};1,1|\;\cdot\;). The generalization considered is a four-dimensional family of K3 surfaces whose polarizing lattice is HD10(1)D4(1)A1(1)H\oplus D_{10}(-1)\oplus D_{4}(-1)\oplus A_{1}(-1), and whose Picard-Fuchs system is the Aomoto-Gel’fand system E(3,6)E(3,6). In the second part we compute the monodromy matrices for the families of Calabi-Yau (n1)(n-1)-folds that extend the family of K3 surface whose rank-19 polarizing lattice is HE8(1)E8(1)4H\oplus E_{8}(-1)\oplus E_{8}(-1)\oplus\langle-4\rangle and whose Picard-Fuchs operator is the hypergeometric differential equation for F23(14,12,34;1,1|){}_{3}F_{2}(\frac{1}{4},\frac{1}{2},\frac{3}{4};1,1|\;\cdot\;). The generalization considered are the one-dimensional mirror families of deformed Fermat pencils whose Picard-Fuchs operator is the hypergeometric differential equation for Fn1n(1n+1,,nn+1;1,,1|){}_{n}F_{n-1}(\frac{1}{n+1},\dots,\frac{n}{n+1};1,\dots,1|\;\cdot\;). The main results of the two parts are Theorem 3.26 and Theorem 4.22, respectively.

This article is organized as follows. In §2 we review relevant background material, which includes multi-parameter Weierstrass models associated with families of Jacobian elliptic fibrations and their multivariate Picard-Fuchs operators. We also recall the fundamental definition of a generalized functional invariant and its relation to the mixed twist construction. In §3 we use the mixed-twist construction to obtain a multi-parameter family of K3 surfaces of Picard rank ρ16\rho\geq 16. Upon identifying a particular Jacobian elliptic fibration on its general member, we find the corresponding lattice polarization, the parameter space, and the Picard-Fuchs system for the family with its global monodromy group. We construct a sequence of restrictions that lead to extensions of the polarization keeping the polarizing lattice two-elementary. We show that the Picard-Fuchs operators under these restrictions coincide with well-known hypergeometric systems, the Aomoto-Gel’fand E(3,6)E(3,6) system (for ρ=17\rho=17), Appell’s F2F_{2} system (for ρ=18\rho=18), and Gauss’ hypergeometric functions of type F23{}_{3}F_{2} (for ρ=19\rho=19). This allows us to determine the global monodromy groups of each family. Finally, we will show in §4 that the mixed-twist construction produces for each mirror family a non-resonant GKZ system for which a basis of solutions in the form of absolutely convergent Mellin-Barnes integrals exists whose monodromy is then computed explicitly.

Acknowledgement

We thank the referees for their many helpful comments, suggestions, and corrections. This work was partially supported by a grant from the Simons Foundation.

2. Elliptic fibrations and the mixed-twist construction

In this section we give some well-known results on Weierstrass models and their period integrals. We also review the generalized functional invariant.

2.1. Weierstrass models and their Picard-Fuchs Operators

We begin by recalling some basic notions of elliptic fibrations and the associated Weierstrass models. Let XX and SS be normal complex algebraic varieties and π:XS\pi:X\to S an elliptic fibration, that is, π\pi is proper surjective morphism with connected fibers such that the general fiber is a nonsingular elliptic curve. Moreover, we assume that π\pi is smooth over an open subset S0SS_{0}\subset S, whose complement in SS is a divisor with at worst normal crossings. Thus, the local system H0i:=Riπ¯X|S0H^{i}_{0}:=R^{i}\pi_{*}\underline{\mathbb{Z}}_{X}{|}_{S_{0}} forms a variation of Hodge structure over S0S_{0}.

Elliptic fibrations possess the following canonical bundle formula: on SS, the fundamental line bundle denoted :=(R1π𝒪X)1\mathcal{L}:=(R^{1}\pi_{*}\mathcal{O}_{X})^{-1} and the canonical bundles 𝝎X:=topT(1,0)X\boldsymbol{\omega}_{X}:=\wedge^{\mathrm{top}}\,T^{*(1,0)}X, 𝝎S:=topT(1,0)S\boldsymbol{\omega}_{S}:=\wedge^{\mathrm{top}}\,T^{*(1,0)}S are related by

(2.1) 𝝎Xπ(𝝎S)𝒪X(D),\boldsymbol{\omega}_{X}\cong\pi^{*}(\boldsymbol{\omega}_{S}\otimes\mathcal{L})\otimes\mathcal{O}_{X}(D),

where DD is a certain effective divisor on XX depending only on divisors on SS over which π\pi has multiple fibers, and divisors on XX giving (1)(-1)-curves of π\pi. When π:XS\pi:X\to S is a Jacobian elliptic fibration, that is, when there is a section σ:SX\sigma:S\to X, the case of multiple fibers is prevented. We may avoid the presence of (1)(-1)-curves in the following way: For XX an elliptic surface, we assume that the fibration is relatively minimal, meaning that there are no (1)(-1)-curves in the fibers of π\pi. When XX is an elliptic threefold, we additionally assume that no contraction of a surface is compatible with the fibration.

Assuming these minimality constraints, we have D=0D=0, thus the canonical bundle formula (2.1) simplifies to 𝝎Xπ(𝝎S)\boldsymbol{\omega}_{X}\cong\pi^{*}(\boldsymbol{\omega}_{S}\otimes\mathcal{L}). In particular, for 𝝎S1\mathcal{L}\cong\boldsymbol{\omega}_{S}^{-1} we obtain 𝝎X𝒪X\boldsymbol{\omega}_{X}\cong\mathcal{O}_{X}. Recall that XX is a Calabi-Yau manifold if 𝝎X𝒪X\boldsymbol{\omega}_{X}\cong\mathcal{O}_{X} and hi(X,𝒪X)=0h^{i}(X,\mathcal{O}_{X})=0 for 0<i<n=dim(X)0<i<n=\dim(X). In this article we will be concerned with Jacobian elliptic fibrations on Calabi-Yau manifolds. It is well known that for XX an elliptic Calabi-Yau threefold, the base surface can have at worst log-terminal orbifold singularities. We will take the base surface SS to be a Hirzebruch surface 𝔽k\mathbb{F}_{k} (or its blowup).

It is well known that Jacobian elliptic fibrations admit Weierstrass models, i.e., given a Jacobian elliptic fibration π:XS\pi:X\to S with section σ:SX\sigma:S\to X, there is a complex algebraic variety WW together with a proper, flat, surjective morphism π^:WS\hat{\pi}:W\to S with canonical section σ^:SW\hat{\sigma}:S\to W whose fibers are irreducible cubic plane curves, together with a birational map XWX\dashrightarrow W compatible with the sections σ\sigma and σ^\hat{\sigma}; see [55]. The map from XX to WW blows down all components of the fibers that do not intersect the image σ(S)\sigma(S). If π:XS\pi:X\to S is relatively minimal, the inverse map WXW\dashrightarrow X is a resolution of the singularities of WW.

A Weierstrass model is constructed as follows: given a line bundle S\mathcal{L}\to S, and sections g2g_{2}, g3g_{3} of 4\mathcal{L}^{4}, 6\mathcal{L}^{6} such that the discriminant Δ=g2327g32\Delta=g_{2}^{3}-27g_{3}^{2} as a section of 12\mathcal{L}^{12} does not vanish, define a 2\mathbb{P}^{2}-bundle p:𝐏Sp:\mathbf{P}\to S as 𝐏:=(𝒪S23)\mathbf{P}:=\mathbb{P}\left(\mathcal{O}_{S}\oplus\mathcal{L}^{2}\oplus\mathcal{L}^{3}\right) with pp the natural projection. Moreover, let 𝒪𝐏(1)\mathcal{O}_{\mathbf{P}}(1) be the tautological line bundle. Denoting x,yx,y and zz as the sections of 𝒪𝐏(1)2\mathcal{O}_{\mathbf{P}}(1)\otimes\mathcal{L}^{2}, 𝒪P(1)3\mathcal{O}_{P}(1)\otimes\mathcal{L}^{3} and 𝒪𝐏(1)\mathcal{O}_{\mathbf{P}}(1) that correspond to the natural injections of 2\mathcal{L}^{2}, 3\mathcal{L}^{3} and 𝒪S\mathcal{O}_{S} into π𝒪𝐏(1)=𝒪S23\pi_{*}\mathcal{O}_{\mathbf{P}}(1)=\mathcal{O}_{S}\oplus\mathcal{L}^{2}\oplus\mathcal{L}^{3}, the Weierstrass model WW from above is given by the the sub-variety of 𝐏\mathbf{P} defined by the equation

(2.2) y2z=4x3g2xz2g3z3.y^{2}z=4x^{3}-g_{2}xz^{2}-g_{3}z^{3}.

The canonical section σ:SW\sigma:S\to W is given by the point [x:y:z]=[0:1:0][x:y:z]=[0:1:0] in each fiber, such that Σ:=σ(S)W\Sigma:=\sigma(S)\subset W is a Cartier divisor whose normal bundle is isomorphic to the fundamental line bundle \mathcal{L} via p𝒪𝐏(Σ)p_{*}\mathcal{O}_{\mathbf{P}}(-\Sigma)\cong\mathcal{L}. It follows that WW inherits the properties of normality and Gorenstein if SS possesses these. Thus, the canonical bundle formula  (2.1) reduces to

(2.3) 𝝎W=π(𝝎S).\boldsymbol{\omega}_{W}=\pi^{*}\left(\boldsymbol{\omega}_{S}\otimes\mathcal{L}\right).

The Jacobian elliptic fibration p:WSp:W\to S then has a Calabi-Yau total space if 𝝎S1=𝒪S(KS)\mathcal{L}\cong\boldsymbol{\omega}_{S}^{-1}=\mathcal{O}_{S}(-K_{S}) (misusing notation slightly to denote the projection map pp the as the projection from the ambient 2\mathbb{P}^{2}-bundle).

For a Jacobian elliptic fibration XX the canonical bundle 𝝎X\boldsymbol{\omega}_{X} is determined by the discriminant Δ=g2327g32\Delta=g_{2}^{3}-27g_{3}^{2}. For example, if π:XS\pi:X\to S is a Jacobian elliptic fibration for a smooth algebraic surface XX and S=1S=\mathbb{P}^{1} with homogeneous coordinates [t:s][t:s], then XX is a rational elliptic surface if the Δ\Delta is a homogeneous polynomial of degree 12 (meaning that =𝒪(1)\mathcal{L}=\mathcal{O}(1)), and XX is a K3 surface when Δ\Delta is a homogeneous polynomial of degree 24 (meaning that =𝒪(2)\mathcal{L}=\mathcal{O}(2)); these results follow readily from adjunction and Noether’s formula. The nature of the singular fibers and their effect on the canonical bundle was established by the seminal work of Kodaira  [42, 43, 41].

Of particular interest in this article are multi-parameter families of elliptic Calabi-Yau nn-folds over a base BB, a quasi-projective variety of dimension rr, denoted by π:XB\pi:~{}X\to B. Hence, each Xp=π1(p)X_{p}=\pi^{-1}(p) is a compact, complex nn-fold with trivial canonical bundle. Moreover, each XpX_{p} is elliptically fibered with section over a fixed normal variety SS. This means that we have a multi-parameter family of minimal Weierstrass models pb:WbSp_{b}:W_{b}\to S representing a family of Jacobian elliptic fibrations πb:XbS\pi_{b}:X_{b}\to S. We denote the collective family of Weierstrass models as p:WBp:W\to B.

Working within affine coordinates for BB and SS we set u=(u1,,un1)n1Su=(u_{1},\dots,u_{n-1})\in\mathbb{C}^{n-1}\subset S and b=(b1,,br)rBb=(b_{1},\dots,b_{r})\in\mathbb{C}^{r}\subset B. We then may write the Weierstrass model WbW_{b} in the form

(2.4) y2=4x3g2(u,b)xg3(u,b),y^{2}=4x^{3}-g_{2}(u,b)x-g_{3}(u,b),

where for each fiber we have chosen the affine chart of WbW_{b} given by z=1z=1 in Equation (2.2).

Part of the utility of a Weierstrass model is the explicit construction of the holomorphic nn-form on each XbX_{b}, up to fiberwise scale, allowing for the detailed study of the Picard-Fuchs operators underlying a variation of Hodge structure. In fact, consider the holomorphic sub-system HBH\to B of the local system V=Rnπ¯XBV=R^{n}\pi_{*}\underline{\mathbb{C}}_{X}\to B, whose fibers are given as the line H0(ωXb)Hn(Xb,)H^{0}(\omega_{X_{b}})\subset H^{n}(X_{b},\mathbb{C}). Here, ¯X\underline{\mathbb{C}}\to X is the constant sheaf whose stalks are \mathbb{C}. Griffiths showed [28, 27, 29, 30] that 𝒱=V𝒪B\mathcal{V}=V\otimes_{\mathbb{C}}\mathcal{O}_{B} is a vector bundle carrying a canonical flat connection \nabla, the Gauss-Manin connection. A meromorphic section of =H𝒪B𝒱\mathcal{H}=H\otimes_{\mathbb{C}}\mathcal{O}_{B}\subset\mathcal{V} is given fiberwise by the holomorphic nn-form ηbH0(𝝎Xb)Hn(Xb,)\eta_{b}\in H^{0}(\boldsymbol{\omega}_{X_{b}})\subset H^{n}(X_{b},\mathbb{C})

(2.5) ηb=du1dun1dxy,\eta_{b}=du_{1}\wedge\cdots\wedge du_{n-1}\wedge\frac{dx}{y},

where we denote the collective section as ηΓ(𝒱,B)\eta\in\Gamma(\mathcal{V},B). It is natural to consider local parallel sections of the dual bundle =H𝒪B\mathcal{H}^{*}=H^{*}\otimes_{\mathbb{C}}\mathcal{O}_{B}, where HH^{*} is the local system dual to HH; these are generated by transcendental cycles ΣbHn(Xb,)\Sigma_{b}\in H_{n}(X_{b},\mathbb{R}) that vary continuously with bBb\in B, writing the collective section as ΣΓ(𝒱,B)\Sigma\in\Gamma(\mathcal{V}^{*},B). The sections are covariantly constant since the local system V=Rnπ¯XV=R^{n}\pi_{*}\underline{\mathbb{C}}_{X} is locally topologically trivial, and thus local sections of the dual VV^{*} are as well. Utilizing the natural fiberwise de Rham pairing

Σb,ηb=Σbηb,\langle\Sigma_{b},\eta_{b}\rangle=\oint_{\Sigma_{b}}\eta_{b},

we obtain the period sheaf ΠB\Pi\to B, whose stalks are given by the local analytic function bω(b)=Σb,ηbb\mapsto\omega(b)=\langle\Sigma_{b},\eta_{b}\rangle. The function ω(b)\omega(b) is called a period integral (over Σb\Sigma_{b}) and satisfies a system of coupled linear PDEs in the variables b1,,brb_{1},\dots,b_{r} – the so called Picard-Fuchs system - whose rank is that of the period sheaf ΠB\Pi\to B, or the number of linearly independent period integrals of the family.

Given the affine local coordinates (b1,,br)rB(b_{1},\dots,b_{r})\in\mathbb{C}^{r}\subset B, fix the meromorphic vector fields j=/bj\partial_{j}=\partial/\partial b_{j} for j=1,,rj=1,\dots,r. Then each j\partial_{j} induces a covariant derivative operator j\nabla_{\partial_{j}} on 𝒱\mathcal{V}. Since \nabla is flat, the curvature tensor Ω=Ω\Omega=\Omega_{\nabla} vanishes, and hence, for all meromorphic vector fields U,VU,V on BB we have

Ω(U,V)=UVVU[U,V]=0.\Omega(U,V)=\nabla_{U}\nabla_{V}-\nabla_{V}\nabla_{U}-\nabla_{[U,V]}=0.

Substituting in the commuting coordinate vector fields i,j\partial_{i},\partial_{j}, we conclude

ij=ji.\nabla_{\partial_{i}}\nabla_{\partial_{j}}=\nabla_{\partial_{j}}\nabla_{\partial_{i}}.

This integrability condition is crucial in obtaining a system of PDEs from the Gauss-Manin connection. Since 𝒱\mathcal{V} has rank m=dimHn(Xb,)m=\dim H^{n}(X_{b},\mathbb{C}), each sequence of parallel sections k1i1krim^η\nabla^{i_{1}}_{\partial_{k_{1}}}\cdots\nabla^{i_{\hat{m}}}_{\partial_{k_{r}}}\eta, for i1++im^=0,1,,m^i_{1}+\cdots+i_{\hat{m}}=0,1,\dots,\hat{m} and 1k1,,krr1\leq k_{1},\dots,k_{r}\leq r form the linear dependence relations

i1++im^=0m^k1,,kr=1rai1im^k1kr(b)k1i1krim^η=0\sum_{i_{1}+\cdots+i_{\hat{m}}=0}^{\hat{m}}\,\sum_{k_{1},\dots,k_{r}=1}^{r}a^{k_{1}\cdots k_{r}}_{i_{1}\cdots i_{\hat{m}}}(b)\,\nabla^{i_{1}}_{\partial_{k_{1}}}\cdots\nabla^{i_{\hat{m}}}_{\partial_{k_{r}}}\,\eta=0

for some integer 0<m^m0<\hat{m}\leq m, where ai1im^k1kr(b)a^{k_{1}\cdots k_{r}}_{i_{1}\cdots i_{\hat{m}}}(b) are meromorphic. Here, it is understood that 0=id\nabla^{0}=\mathrm{id}. As \nabla annihilates the transcendental cycle Σ\Sigma and is compatible with the pairing Σ,η\langle\Sigma,\eta\rangle, we may “differentiate under the integral sign” to obtain

bjω(b)=bjΣη=Σjη.\frac{\partial}{\partial b_{j}}\omega(b)=\frac{\partial}{\partial b_{j}}\oint_{\Sigma}\eta=\oint_{\Sigma}\nabla_{\partial_{j}}\eta.

It follows that the period integral ω(b)\omega(b) satisfies the system of linear PDEs of rank 𝗋1\mathsf{r}\geq 1, given by

(2.6) i1++im^=0m^k1,,kr=1rai1im^k1kr(b)i1++im^k1bk1krbkrω(b)=0.\sum_{i_{1}+\cdots+i_{\hat{m}}=0}^{\hat{m}}\,\sum_{k_{1},\dots,k_{r}=1}^{r}a^{k_{1}\cdots k_{r}}_{i_{1}\cdots i_{\hat{m}}}(b)\frac{\partial^{i_{1}+\dots+i_{\hat{m}}}}{\partial^{k_{1}}b_{k_{1}}\cdots\partial^{k_{r}}b_{k_{r}}}\omega(b)=0.

Equation (2.6) is the Picard-Fuchs system of the multi-parameter family π:XB\pi:X\to B of Calabi-Yau nn-folds. The resulting system is then known to be a linear Fuchsian system, i.e., the system with at worst regular singularities. This is due to analytical results of Griffiths [29] and Deligne [17] who utilized Hironaka’s resolution of singularities [32, 33] to estimate the growth of solutions of the system.

The rank 𝗋\mathsf{r} and order m^\hat{m} of the system depends on the parameter space BB and algebro-geometric data of the generic fiber XbX_{b}. For example, let π:XB\pi:X\to B be a family of Jacobian elliptic K3 surfaces which is polarized by a lattice111For the definition of lattice polarized K3 surface, see §3.2. LL of rank ρ18\rho\leq 18 such that BB defines an n=20ρn=20-\rho dimensional family of L-polarized K3 surfaces. By results due to Dolgachev [19], there is a coarse moduli space L\mathcal{M}_{L} of all lattice polarized K3 surfaces of dimension nn; in this case, we are requiring that BB be a top dimensional family of LL-polarized K3 surfaces. It then follows from the general program of Sasaki and Yoshida [65] on orbifold uniformizing differential equations that the Picard-Fuchs system (2.6) is a linear system of order m^=2\hat{m}=2 and rank 𝗋=n+2\mathsf{r}=n+2 in nn variables, the latter coming from the local coordinates in BB. Naturally, there are sub-loci of such parameter spaces BB where the lattice polarization extends to higher Picard rank and the rank of the Picard-Fuchs system drops accordingly. This behavior was studied, for example, by Doran et al. in [21], and coined the differential rank-jump property therein. In the sequel, we will analyze it by studying corresponding Weierstrass model p:WBp:W\to B. Moreover, we will see that the Picard-Fuchs system can be explicitly computed from the geometry of the elliptic fibrations and the presentation of the associated period integrals as generalized Euler integrals using GKZ systems  [24].

It is commonplace in the literature to study the Picard-Fuchs equations of one parameter families of Calabi-Yau nn-folds; in this case, the base BB is a punctured complex plane with local affine coordinate tBt\in\mathbb{C}\subset B, and an analogous construction leads to a regular Fuchsian ODE of order m\leq m with m=dimHn(Xt,)m=\dim H^{n}(X_{t},\mathbb{C}) for the general fiber XtX_{t}. In the construction of Doran and Malmendier [22], this is the central focus, with B=1{0,1,}B=\mathbb{P}^{1}-\{0,1,\infty\} and B=1{0,1,p,}B=\mathbb{P}^{1}-\{0,1,p,\infty\}. We will show that the restriction of the multi-parameter Picard-Fuchs system (2.6) above leads to the Picard-Fuchs ODE operators and families of lattice polarized K3 surfaces of Picard rank ρ=19\rho=19, for example the mirror partners of the classic deformed Fermat quartic K3.

2.2. The generalized functional invariant

We first recall the generalized functional invariant of the mixed-twist construction studied by Doran and Malmendier [22], first introduced by Doran [20]. A generalized functional invariant is a triple (i,j,α)(i,j,\alpha) with i,ji,j\in\mathbb{N} and α{12,1}\alpha\in\left\{\frac{1}{2},1\right\} such that 1i,j61\leq i,j\leq 6. To this end, the generalized functional invariant encodes a 1-parameter family of degree i+ji+j covering maps 11\mathbb{P}^{1}\to\mathbb{P}^{1}, which is totally ramified over 0, ramified to degrees ii and jj over \infty, and simply ramified over another point t~\tilde{t}. For homogeneous coordinates [v0:v1]1[v_{0}:v_{1}]\in\mathbb{P}^{1}, this family of maps (parameterized by t~1{0,1,}\tilde{t}\in\mathbb{P}^{1}-\{0,1,\infty\}) is given by

(2.7) [v0,v1][cijv1i+jt~:v0i(v0+v1)j],[v_{0},v_{1}]\mapsto[c_{ij}v_{1}^{i+j}\tilde{t}:v_{0}^{i}(v_{0}+v_{1})^{j}],

for some constant cij×c_{ij}\in\mathbb{C}^{\times}. For a family π:XB\pi:X\to B with Weierstrass models given by Equation (2.4) with complex nn-dimensional fibers and a generalized functional invariant (i,j,α)(i,j,\alpha) such that

(2.8) 0degt(g2)min(4i,4αj),0degt(g3)min(6i,6αj),0\leq\mathrm{deg}_{t}(g_{2})\leq\mathrm{min}\left(\frac{4}{i},\frac{4\alpha}{j}\right),\hskip 14.22636pt0\leq\mathrm{deg}_{t}(g_{3})\leq\mathrm{min}\left(\frac{6}{i},\frac{6\alpha}{j}\right),

Doran and Malmendier showed that a new family π~:X~B\tilde{\pi}:\tilde{X}\to B can be constructed such that the general fiber X~t~=π~1(t~)\tilde{X}_{\tilde{t}}=\tilde{\pi}^{-1}(\tilde{t}) is a compact, complex (n+1)(n+1)-manifold equipped with a Jacobian elliptic fibration over 1×S\mathbb{P}^{1}\times S. In the coordinate chart {[v0:v1],(u1,,un1)}1×S\{[v_{0}:v_{1}],(u_{1},\dots,u_{n-1})\}\in\mathbb{P}^{1}\times S the family of Weierstrass models Wt~W_{\tilde{t}} is given by

(2.9) y~2=4x~3g2(cijt~v1i+jv0i(v0+v1)j,u)v04v144α(v0+v1)4αx~g3(cijt~v1i+jv0i(v0+v1)j,u)v06v166α(v0+v1)6α\begin{split}\tilde{y}^{2}=4\tilde{x}^{3}&-g_{2}\left(\frac{c_{ij}\tilde{t}v_{1}^{i+j}}{v_{0}^{i}(v_{0}+v_{1})^{j}},u\right)v_{0}^{4}v_{1}^{4-4\alpha}(v_{0}+v_{1})^{4\alpha}\tilde{x}\\ &-g_{3}\left(\frac{c_{ij}\tilde{t}v_{1}^{i+j}}{v_{0}^{i}(v_{0}+v_{1})^{j}},u\right)v_{0}^{6}v_{1}^{6-6\alpha}(v_{0}+v_{1})^{6\alpha}\end{split}

with cij=(1)iiijj/(i+j)i+jc_{ij}=(-1)^{i}i^{i}j^{j}/(i+j)^{i+j}. The new family is called the twisted family with generalized functional invariant (i,j,α)(i,j,\alpha) of π:XB\pi:X\to B. It follows that conditions (2.8) guarantee that the twisted family is minimal and normal if the original family is. Moreover, they showed that if the Calabi-Yau condition is satisfied for the fibers of the twisted family if it is satisfied for the fibers of the original.

The twisting associated with the generalized functional invariant above is referred to as the pure twist construction; we may extend this notion to that of a mixed twist construction. This means that one combines a pure twist from above with a rational map BBB\to B, thus allowing one to change locations of the singular fibers and ramification data. This was studied in [22, Sec. 8] for linear and quadratic base changes. We may also perform a multi-parameter version of the mixed twist construction for a generalized functional invariant (i,j,α)=(1,1,1)(i,j,\alpha)=(1,1,1). For us, it will be enough to consider the two-parameter family of ramified covering maps given by

(2.10) [v0:v1][4av0(v0+v1)+(ab)v12:4v0(v0+v1)],[v_{0}:v_{1}]\mapsto[4av_{0}(v_{0}+v_{1})+(a-b)v_{1}^{2}:4v_{0}(v_{0}+v_{1})],

such that for a,b1{0,1,}a,b\in\mathbb{P}^{1}-\{0,1,\infty\} with aba\not=b the map in Equation (2.10) is totally ramified over aa and bb. We will apply the mixed twist construction to certain (families of) rational elliptic surfaces X1X\to\mathbb{P}^{1}. In [22, Sec. 5.5] the authors showed that the twisted family with generalized functional invariant (1,1,1)(1,1,1) in this case is birational to a quadratic twist family of X1X\to\mathbb{P}^{1}. We will explain the relationship in more detail and utilize it in the construction of the associated Picard-Fuchs operators in the next section.

3. A multi-parameter family of K3 surfaces

In this section, we use the mixed-twist construction to obtain a multi-parameter family of K3 surfaces of Picard rank ρ16\rho\geq 16. Upon identifying a particular Jacobian elliptic fibration on its general member, we find the corresponding lattice polarization and the Picard-Fuchs system using the results from §2.1. We construct a sequence of restrictions on the parameter space that lead to extensions of the lattice polarization, while keeping the polarizing lattice two-elementary.

Moreover, we show that the Picard-Fuchs operators under these restrictions coincide with well-known hypergeometric systems, the Aomoto-Gel’fand E(3,6)E(3,6) system (for ρ=16,17\rho=16,17), Appell’s F2F_{2} system (for ρ=18\rho=18), and Gauss’ hypergeometric functions of type F23{}_{3}F_{2} (for ρ=19\rho=19). Each such Picard-Fuchs system forms a resonsant GKZ hypergeometric system. We also determine the corresponding monodromy group for each family.

3.1. Quadratic twists and double-sextics

A two-parameter family of rational elliptic surfaces Sc,d1S_{c,d}\to\mathbb{P}^{1} is given by the affine Weierstrass model

(3.1) y2=4x3g2(t)xg3(t),{y}^{2}=4x^{3}-g_{2}(t)x-g_{3}(t),

where g2(t)g_{2}(t) and g3(t)g_{3}(t) are the following polynomials of degree four and six, respectively,

g2=43(t4(2c+d+1)t3+(c2+cd+d2+2cd+1)t2c(cd+2)t+c2),g3=427(t2(cd+2)t+2c)(t2(c+2d1)tc)(2t2(2c+d+1)t+c),\begin{split}g_{2}&=\frac{4}{3}\left(t^{4}-(2c+d+1)t^{3}+(c^{2}+cd+d^{2}+2c-d+1)t^{2}-c(c-d+2)t+c^{2}\right),\\ g_{3}&=\frac{4}{27}\left(t^{2}-(c-d+2)t+2c\right)\left(t^{2}-(c+2d-1)t-c\right)\left(2t^{2}-(2c+d+1)t+c\right),\end{split}

where tt is the affine coordinate on the base curve. Assuming general parameters c,dc,d, Equation (3.1) defines a rational elliptic surface with 6 singular fibers of Kodaira type I2I_{2} over t=0,1,,c,c+d,t=0,1,\infty,c,c+d, and c/(d1)c/(d-1). We have the following:

Lemma 3.1.

The rational elliptic surface S=Sc,dS=S_{c,d} in Equation (3.1) is birationally equivalent to the twisted Legendre pencil

(3.2) y~2=x~(x~1)(x~t)(tcdx~).\tilde{y}^{2}=\tilde{x}(\tilde{x}-1)(\tilde{x}-t)(t-c-d\tilde{x}).
Proof.

By direct computation using the transformation:

x=3t(tc)3x~+t2+(d+1c)tc,y=3y~t(tc)2(3x~+t2+(d+1c)tc)2.x=\frac{3t\left(t-c\right)}{3\tilde{x}+{t}^{2}+\left(d+1-c\right)t-c},\quad y=\frac{3\tilde{y}t\left(t-c\right)}{2\left(3\tilde{x}+{t}^{2}+\left(d+1-c\right)t-c\right)^{2}}.

A quadratic twist applied to a rational elliptic surface can be identified with Doran and Malmendier’s mixed-twist construction with generalized functional invariant (i,j,α)=(1,1,1)(i,j,\alpha)=(1,1,1). The two-parameter family of ramified covering maps in Equation (2.10) is totally ramified over a,b1{0,1,}a,b\in\mathbb{P}^{1}-\{0,1,\infty\} with aba\not=b. We apply the mixed-twist construction to the rational elliptic surface Sc,dS_{c,d}:

Proposition 3.2.

The mixed-twist construction with generalized functional invariant (i,j,α)=(1,1,1)(i,j,\alpha)=(1,1,1) applied to the rational elliptic surface in Equation (3.1) yields the family of Weierstrass models

(3.3) y^2=4x^3(ta)2(tb)2g2(t)x^(ta)3(tb)3g3(t).\hat{y}^{2}=4\hat{x}^{3}-(t-a)^{2}(t-b)^{2}g_{2}(t)\hat{x}-(t-a)^{3}(t-b)^{3}g_{3}(t).

The family is birationally equivalent to

(3.4) y2=x(x1)(xt)(ta)(tb)(tcdx).y^{2}=x(x-1)(x-t)(t-a)(t-b)(t-c-dx).

Over the four-dimensional parameter space

(3.5) ={(a,b,c,d)4|ab,(c,d)(a,0),(b,0),(0,1)},\mathcal{M}=\Big{\{}(a,b,c,d)\in\mathbb{C}^{4}\ \Big{|}\ a\not=b,\ (c,d)\not=(a,0),(b,0),(0,1)\Big{\}}\,,

Equation (3.4) defines a family of Jacobian elliptic K3 surfaces 𝐗a,b,c,d1\mathbf{X}_{a,b,c,d}\to\mathbb{P}^{1}.

Proof.

In affine base coordinates [v:1]1[v:1]\in\mathbb{P}^{1}, the map f:11f:\mathbb{P}^{1}\to\mathbb{P}^{1} from the mixed-twist construction with generalized functional invariant (i,j,α)=(1,1,1)(i,j,\alpha)=(1,1,1) in Equation (2.10) is given by

f(v)=a+ab4v(v+1).f(v)=a+\frac{a-b}{4v(v+1)}.

The pullback of the Weierstrass model for the two-parameter family of the rational elliptic surfaces in Equation (3.1) along the map t=f(v)t=f(v) is easily checked to yield the four-parameter family in Equation (3.3). Equation (3.4) follows from a direct computation, with the following transformation:

x^=3t(ta)(tb)(tc)3x+(ta)(tb)(t2+(d+1c)tc),y^=3yt(ta)(tb)(tc)2(3x+(ta)(tb)(t2+(d+1c)tc))2.\begin{split}\hat{x}=\frac{3t\left(t-a\right)\left(t-b\right)\left(t-c\right)}{3x+\left(t-a\right)\left(t-b\right)\left({t}^{2}+\left(d+1-c\right)t-c\right)},\\ \hat{y}=\frac{3yt\left(t-a\right)\left(t-b\right)\left(t-c\right)}{2\left(3x+\left(t-a\right)\left(t-b\right)\left({t}^{2}+\left(d+1-c\right)t-c\right)\right)^{2}}.\end{split}

One checks that for parameters in \mathcal{M} the minimal resolution of Equation (3.3) defines a Jacobian elliptic K3 surfaces 𝐗a,b,c,d1\mathbf{X}_{a,b,c,d}\to\mathbb{P}^{1}. In fact, Equation (3.3) is a minimal Weierstrass equation of a K3 surface if and only if aba\not=b and (c,d)(a,0),(b,0),(0,1)(c,d)\not=(a,0),(b,0),(0,1). ∎

A direct computation for the Weierstrass model yields the following:

Lemma 3.3.

Equation (3.3) defines a Jacobian elliptic fibration π:𝐗1\pi:\mathbf{X}\to~{}\mathbb{P}^{1} on a general 𝐗=𝐗a,b,c,d\mathbf{X}=\mathbf{X}_{a,b,c,d} with two singular fibers of Kodaira type I0I_{0}^{*} over t=a,bt=a,b, six singular fibers of Kodaira type I2I_{2} over t=0,1,,c,c+d,t=0,1,\infty,c,c+d, and c/(d1)c/(d-1), and the Mordell Weil group MW(𝐗,π)=(/2)2\mathrm{MW}(\mathbf{X},\pi)=(\mathbb{Z}/2\mathbb{Z})^{2}.

Equation (3.4) provides a model for the K3 surfaces 𝐗\mathbf{X} as double covers of the projective plane branched on the union of six lines. In general, we call a K3 surface 𝒳\mathcal{X} a double-sextic surface if it is the minimal resolution of a double cover of the projective plane 2\mathbb{P}^{2} branched along the union of six lines, which we denote by ={1,,6}\boldsymbol{\ell}=\{\ell_{1},\dots,\ell_{6}\}. In weighted homogeneous coordinates [t1:t2:t3:z](1,1,1,3)[t_{1}:t_{2}:t_{3}:z]\in\mathbb{P}(1,1,1,3) such a double-sextic is given by the equation

(3.6) z2=i=16(ai1t1+ai2t2+ai3t3),z^{2}=\prod_{i=1}^{6}(a_{i1}t_{1}+a_{i2}t_{2}+a_{i3}t_{3}),

where the lines i={[t1:t2:t3]|ai1t1+ai2t2+ai3t3=0}2\ell_{i}=\left\{[t_{1}:t_{2}:t_{3}]\;|\;a_{i1}t_{1}+a_{i2}t_{2}+a_{i3}t_{3}=0\right\}\subset\mathbb{P}^{2} for parameters aija_{ij}\in\mathbb{C}, i=1,,6i=1,\dots,6, j=1,2,3j=1,2,3 are assumed to be general. Let A=(aij)Mat(3,6;)A=(a_{ij})\in\mathrm{Mat}(3,6;\mathbb{C}) be the matrix whose entries are the coefficients encoding the six-line configuration \boldsymbol{\ell}. Let MM be the configuration space of six lines \boldsymbol{\ell} whose minimal resolution is a K3 surface. Then isomorphic K3 surfaces are obtained if we act on elements AMA\in M by matrices induced from automorphisms of 2\mathbb{P}^{2} on the left and overall scale changes of each line i\ell_{i}\in\boldsymbol{\ell} on the right. Thus, we are led to consider the four-dimensional quotient space

(3.7) 6=SL(3,)\M/()6,\mathcal{M}_{6}=\mathrm{SL}(3,\mathbb{C})\backslash M/(\mathbb{C}^{*})^{6},

and \mathcal{M} in Equation (3.5) can be identified with the open subspace of 6\mathcal{M}_{6}, given by elements [A]6[A]\in\mathcal{M}_{6} of the form

(11100d010111001abc)\begin{pmatrix}1&1&1&0&0&-d\\ 0&1&0&-1&-1&-1\\ 0&0&1&a&b&c\end{pmatrix}

with (a,b,c,d)(a,b,c,d)\in\mathcal{M} and t1=x,t2=t,t3=1t_{1}=x,t_{2}=-t,t_{3}=-1.

The family of double-sextics in Equation (3.6) has been studied in the literature, for example by Matsumoto [51], and Matsumoto et al. [52, 53, 54]. One takeaway from their work is that the family of double sextic K3 surfaces is, in many ways, analogous to the Legendre pencil of elliptic curves which is realized as double covers of 1\mathbb{P}^{1} branching over four points. More recently, the double-sextic family 𝒳\mathcal{X} and closely related K3 surfaces have been studied in the context of string dualities [47, 49, 13, 46, 10]. In Clingher et al. [13], the authors showed that four different elliptic fibrations on 𝒳\mathcal{X} have interpretations in F-theory/heterotic string duality. Similar constructions are relevant to anomaly cancellations [48], studied by the authors of the present article. In [10], the authors classified all Jacobian elliptic fibrations on the Shioda-Inose surface associated with 𝒳\mathcal{X}. Finally, Hosono et al.  in [35, 35] constructed compactifications of 6\mathcal{M}_{6} from GKZ data and toric geometry, suitable for the study of the Type IIA/Type IIB string duality.

3.2. Determination of the lattice polarization and monodromy

In the following we will use the following standard notations for lattices: L1L2L_{1}\oplus L_{2} is orthogonal sum of the two lattices L1L_{1} and L2L_{2}, L(λ)L(\lambda) is obtained from the lattice LL by multiplication of its form by λ\lambda\in\mathbb{Z}, R\langle R\rangle is a lattice with the matrix RR in some basis; AnA_{n}, DmD_{m}, and EkE_{k} are the positive definite root lattices for the corresponding root systems, HH is the unique even unimodular hyperbolic rank-two lattice. A lattice LL is two-elementary if its discriminant group ALA_{L} is a two-elementary abelian group, namely AL(/2)A_{L}\cong(\mathbb{Z}/2\mathbb{Z})^{\ell} with \ell being the minimal number of generators of the discriminant group ALA_{L}, also called the length of the lattice LL. Even, indefinite, two-elementary lattices LL are uniquely determined by the rank ρ\rho, the length \ell, and the parity δ\delta – which equals 11 unless the discriminant form qL(x)q_{L}(x) takes values in /2/2\mathbb{Z}/2\mathbb{Z}\subset\mathbb{Q}/2\mathbb{Z} for all xALx\in A_{L} in which case it is 0; this is a result by Nikulin [63, Thm. 4.3.2].

Let 𝐗\mathbf{X} be a smooth algebraic K3 surface over the field of complex numbers. Denote by NS(𝐗)\operatorname{NS}(\mathbf{X}) the Néron-Severi lattice of 𝐗\mathbf{X}. This is known to be an even lattice of signature (1,ρ𝐗1)(1,\rho_{\mathbf{X}}-1), where p𝐗p_{\mathbf{X}} denotes the Picard number of 𝐗\mathbf{X}, with 1ρ𝐗201\leq\rho_{\mathbf{X}}\leq 20. In this context, a lattice polarization [59, 60, 61, 62, 18] on 𝐗\mathbf{X} is, by definition, a primitive lattice embedding i:LNS(𝐗)i\colon L\hookrightarrow\operatorname{NS}(\mathbf{X}), with i(L)i(L) containing a pseudo-ample class, i.e., a numerically effective class of positive self-intersection in the Néron-Severi lattice NS(𝐗)\mathrm{NS}(\mathbf{X}). Here, LL is a choice of even lattice of signature (1,ρ)(1,\rho), with 1ρ201\leq\rho\leq 20 that admits a primitive embeddings into the K3 lattice ΛK3H3E8(1)2\Lambda_{K3}\cong H^{\oplus 3}\oplus E_{8}(-1)^{\oplus 2}. Two LL-polarized K3 surfaces (𝐗,i)(\mathbf{X},i) and (𝐗,i)(\mathbf{X}^{\prime},i^{\prime}) are said to be isomorphic222Our definition of isomorphic lattice polarizations coincides with the one used by Vinberg [72, 73, 74]. It is slightly more general than the one used in [19, Sec. 1]., if there exists an analytic isomorphism α:𝐗𝐗\alpha\colon\mathbf{X}\rightarrow\mathbf{X}^{\prime} and a lattice isometry βO(L)\beta\in O(L), such that αi=iβ\alpha^{*}\circ i^{\prime}=i\circ\beta, where α\alpha^{*} is the appropriate morphism at cohomology level. In general, LL-polarized K3 surfaces are classified, up to isomorphism, by a coarse moduli space L\mathcal{M}_{L}, which is known [19] to be a quasi-projective variety of dimension 20ρ20-\rho. A general LL-polarized K3 surface (𝐗,i)(\mathbf{X},i) satisfies i(L)=NS(𝐗)i(L)=\operatorname{NS}(\mathbf{X}).

We have the following result:

Proposition 3.4.

Over \mathcal{M} in Equation (3.5) the family

(3.8) 𝐗a,b,c,d:y2=x(x1)(xt)(ta)(tb)(tcdx).\mathbf{X}_{a,b,c,d}:\;\;y^{2}=x(x-1)(x-t)(t-a)(t-b)(t-c-dx).

is a 4-dimensional family of LL-polarized K3 surfaces where LL has rank 16 and the following isomorphic presentations:

(3.9) LHE8(1)A1(1)6HE7(1)D4(1)A1(1)3HD6(1)D4(1)2HD6(1)2A1(1)2HD10(1)A1(1)4HD8(1)D4(1)A1(1)2.\begin{split}L\ \cong\ &H\oplus E_{8}(-1)\oplus A_{1}(-1)^{\oplus 6}\ \cong\ H\oplus E_{7}(-1)\oplus D_{4}(-1)\oplus A_{1}(-1)^{\oplus 3}\\ \cong\ &H\oplus D_{6}(-1)\oplus D_{4}(-1)^{\oplus 2}\ \cong\ H\oplus D_{6}(-1)^{\oplus 2}\oplus A_{1}(-1)^{\oplus 2}\\ \ \cong\ &H\oplus D_{10}(-1)\oplus A_{1}(-1)^{\oplus 4}\cong\ H\oplus D_{8}(-1)\oplus D_{4}(-1)\oplus A_{1}(-1)^{\oplus 2}.\end{split}

In particular, LL is a primitive sub-lattice of the K3 lattice ΛK3\Lambda_{K3}.

Proof.

The general member of the family in Equation (3.8) is a double-sextic whose associated K3 surface has Picard number 16. A K3 surface 𝐗\mathbf{X} obtained as the minimal resolution of the double-sextic associated with a six-line configuration \boldsymbol{\ell} in general position has the transcendental lattice T(𝐗)H(2)H(2)22\mathrm{T}(\mathbf{X})\cong H(2)\oplus H(2)\oplus\langle-2\rangle^{\oplus 2}; see [39]. Accordingly, 𝐗\mathbf{X} has a Néron-Severi lattice given by a two-elementary lattice LL of rank ρ=16\rho=16 such that AL(/2)A_{L}\cong(\mathbb{Z}/2\mathbb{Z})^{\ell} with =6\ell=6. From general lattice theory, it follows that LL is the unique two-elementary lattice with ρ=16\rho=16, =6\ell=6, δ=1\delta=1 (for ρ=16\rho=16 the two-elementary lattice must have δ=1\delta=1; see [63]), and we obtain LHE8(1)A1(1)6L\cong H\oplus E_{8}(-1)\oplus A_{1}(-1)^{\oplus 6}.

The family in Equation (3.3) is birationally equivalent to the family in Equation (3.4). In turn, Lemma 3.1 identifies the family in Equation (3.4) as a family of Jacobian elliptic K3 surfaces whose general member has the singular fibers 2I0+6I22I_{0}^{*}+6I_{2} and the Mordell-Weil group (/2)2(\mathbb{Z}/2\mathbb{Z})^{2}. We then use results in [40, Table 1] to conclude that the general member of such a K3 surface 𝐗\mathbf{X} has the Néron-Severi lattice isomorphic to HE8(1)A1(1)6H\oplus E_{8}(-1)\oplus A_{1}(-1)^{\oplus 6}. From [40, Table 1] we also read off the isomorphic presentations of LL as the Jacobian elliptic fibrations supported on 𝐗\mathbf{X} with trivial Mordell Weil group. These elliptic fibrations prove that the lattice LL has the isomorphic presentations in Equation (3.9). ∎

The Picard-Fuchs system for the family can also be determined:

Proposition 3.5.

Let ΣT(𝐗)\Sigma\in\mathrm{T}(\mathbf{X}) be a transcendental cycle on a general K3 surface 𝐗=𝐗a,b,c,d\mathbf{X}=\mathbf{X}_{a,b,c,d}, η𝐗\eta_{\mathbf{X}} the holomorphic two-form induced by dtdx/ydt\wedge dx/y in Equation (3.8), and ω=Ση𝐗\omega=\oint_{\Sigma}\eta_{\mathbf{X}} a period. The Picard-Fuchs system for 𝐗a,b,c,d\mathbf{X}_{a,b,c,d}, annihilating ω=b(bc)ω\omega^{\prime}=\sqrt{b(b-c)}\ \omega, is the rank-six Aomoto-Gel’fand system E(3,6)E(3,6) of [52, 53] and [54, §0.15] in the variables

(3.10) x1=ab,x2=acbc,x3=1b,x4=dbc.x_{1}=\frac{a}{b},\hskip 14.22636ptx_{2}=\frac{a-c}{b-c},\hskip 14.22636ptx_{3}=\frac{1}{b},\hskip 14.22636ptx_{4}=\frac{d}{b-c}.

In particular, the Picard-Fuchs system is a resonant GKZ hypergeometric system.

Proof.

In [52], a matrix A=(aij)Mat(3,6;)A=(a_{ij})\in\mathrm{Mat}(3,6;\mathbb{C}) was considered whose entries are the coefficients encoding a six-line configuration \boldsymbol{\ell}. The authors used the action of SL(3,)\mathrm{SL}(3,\mathbb{C}) and ()6(\mathbb{C}^{*})^{6} to bring AA into the standard form

(3.11) (1001110101x1x20011x3x4).\begin{pmatrix}1&0&0&1&1&1\\ 0&1&0&1&x_{1}&x_{2}\\ 0&0&1&1&x_{3}&x_{4}\end{pmatrix}.

Equivalently, the associated K3 surface 𝒳\mathcal{X} is the minimal resolution of the double-sextic

(3.12) z2=t1t2t3(t1+t2+t3)(t1+x1t2+x3t3)(t1+x2t2+x4t3).z^{2}=t_{1}t_{2}t_{3}\big{(}t_{1}+t_{2}+t_{3}\big{)}\big{(}t_{1}+x_{1}t_{2}+x_{3}t_{3}\big{)}\big{(}t_{1}+x_{2}t_{2}+x_{4}t_{3}\big{)}\,.

In [64, §4] Sasaki showed that the period integral for the non-vanishing holomorphic two-form η𝒳H0(𝝎𝒳)\eta_{\mathcal{X}}\in H^{0}(\boldsymbol{\omega}_{\mathcal{X}}) induced by dt2dt3/zdt_{2}\wedge dt_{3}/z in Equation (3.12) in the affine chart t1=1t_{1}=-1 over a transcendental cycle ΣT(𝒳)\Sigma^{\prime}\in~{}\mathrm{T}(\mathcal{X}), given by

(3.13) ω=ω(x1,x2,x3,x4)=Ση𝒳,\omega^{\prime}=\omega^{\prime}(x_{1},x_{2},x_{3},x_{4})=\oint_{\Sigma^{\prime}}\eta_{\mathcal{X}},

is a solution of the resonant rank-six Aomoto-Gel’fand system E(3,6)E(3,6) in the variables x1x_{1}, x2x_{2}, x3x_{3}, x4x_{4}. The construction of transcendental cycles Σ\Sigma^{\prime} was described in [52].

In the affine coordinate system t1=1t_{1}=-1, we consider the transformation φ:𝐗(μ)𝒳\varphi:\mathbf{X}^{(\mu)}\dashrightarrow\mathcal{X} given by

t2=x3t1x3tx1,t3=x(1x1)x3tx1,z=x3(x11)2y~(x3tx1)3,t_{2}=\frac{x_{3}t-1}{x_{3}t-x_{1}},\hskip 14.22636ptt_{3}=\frac{x(1-x_{1})}{x_{3}t-x_{1}},\hskip 14.22636ptz=\frac{x_{3}(x_{1}-1)^{2}\tilde{y}}{(x_{3}t-x_{1})^{3}},

together with the change of parameters in Equation (3.10). Here, 𝐗(μ)\mathbf{X}^{(\mu)} is the twist of the K3 surface 𝐗\mathbf{X} and given by

(3.14) μy~2=x(x1)(xt)(ta)(tb)(tcdx)\mu\tilde{y}^{2}=\,x(x-1)(x-t)(t-a)(t-b)(t-c-dx)

with μ=b(bc)\mu=b(b-c). The map φ:𝐗(μ)𝒳\varphi:\mathbf{X}^{(\mu)}\dashrightarrow\mathcal{X} extends to a birational map of K3 surfaces such that

(3.15) φη𝒳=dtdxy~.\varphi^{*}\eta_{\mathcal{X}}=dt\wedge\frac{dx}{\tilde{y}}.

It follows that periods of the two-form dtdx/y~dt\wedge dx/\tilde{y} for the family 𝐗(μ)\mathbf{X}^{(\mu)} satisfy the same Picard-Fuchs system as the periods ω\omega^{\prime} in Equation (3.13). In turn, periods ω\omega of the two-form dtdx/ydt\wedge dx/y for 𝐗\mathbf{X} in Equation (3.8) with y=μy~y=\sqrt{\mu}\tilde{y} are annihilated by the same Picard-Fuchs operator as ω/μ\omega^{\prime}/\sqrt{\mu}. ∎

We now turn our attention to the determination of the monodromy group of the period map of the family 𝐗\mathbf{X} of double sextic LL-polarized K3 surfaces. As the Picard-Fuchs system E(3,6)E(3,6) annihilating the (twisted) period integral in Proposition 3.5 is a resonant GKZ system, the monodromy representation is reducible [66], and so the determination of the monodromy group is in general more complicated. Our strategy is to connect the family 𝐗\mathbf{X} birationally to other families of K3 surfaces whose monodromy groups are known, as we have done in Proposition 3.5 with the double sextic family 𝒳\mathcal{X} studied by Matsumoto et al. [54].

We need to pay close attention to the twist factor μ=b(bc)\sqrt{\mu}=\sqrt{b(b-c)}, which causes the period map for the family 𝐗\mathbf{X} to become multi-valued; thus, the monodromy representation does not coincide with the topological monodromy of the family, i.e., the monodromy of the local system R2π¯𝐗R^{2}\pi_{*}\underline{\mathbb{Z}}_{\mathbf{X}}\to\mathcal{M}.

Let ΣT(𝐗)\Sigma\in\mathrm{T}(\mathbf{X}) be a transcendental cycle, and \nabla the Gauss-Manin connection from §2.1 associated to the system of Picard-Fuchs equations for 𝐗\mathbf{X} - the Aomoto-Gel’fand E(3,6)E(3,6) system - in Proposition 3.5. Let η𝐗\eta_{\mathbf{X}} be the holomorphic two-form on the K3 surface 𝐗\mathbf{X} induced by dtdx/ydt\wedge dx/y. As we parallel transport Σ\Sigma under \nabla around the locus b=0b=0 in \mathcal{M}, for an initial point away from c=0c=0, we obtain a new cycle Σ\Sigma^{\prime} that is related by the action of the monodromy group of the Aomoto-Gel’fand system on T(𝐗)\mathrm{T}(\mathbf{X}) and the twist μ\mu relating the families 𝒳\mathcal{X} and 𝐗\mathbf{X}; see proof of Proposition 3.5. Thus, as we switch branches of the square root of the twisting factor, we obtain the following action on a period integral:

(3.16) b(bc)Ση𝐗b(bc)Ση𝐗.\sqrt{b(b-c)}\oint_{\Sigma}\,\eta_{\mathbf{X}}\;\;\;\to\;\;\;-\sqrt{b(b-c)}\oint_{\Sigma^{\prime}}\,\eta_{\mathbf{X}}\,.

The situation can be described as follows: let Π\Pi\to\mathcal{M} be the period sheaf of the family 𝐗\mathbf{X} described in §2.1, that is the rank six complex local system whose stalks are generated by linearly independent period integrals for 𝐗\mathbf{X}. Moreover, we define a rank one integral local system 𝖲4Z(μ)\mathsf{S}\to\mathbb{C}^{4}-Z(\mu), with the monodromy group 2\mathbb{Z}_{2} around the divisor μ=0\mu=0. Here, Z(μ)Z(\mu) is the vanishing locus of μ\mu in 4\mathbb{C}^{4}. The monodromy representation of the family 𝐗\mathbf{X} acts on the tensor product 𝖲¯Π\mathsf{S}\otimes_{\underline{\mathbb{Z}}_{\mathcal{M}}}\Pi, with 2\mathbb{Z}_{2} acting nontrivially as multiplication by 𝕀-\mathbb{I}, the negative of the identity matrix, as the vanishing locus of μ\mu is encircled away from the singular locus of the family. Here, we are identifying 𝖲\mathsf{S} with its restriction to \mathcal{M}.

Let p:𝒫p:\mathcal{M}\to\mathcal{P} be the period mapping

(3.17) p:(a,b,c,d)[ω1(a,b,c,d)::ω6(a,b,c,d)],p:(a,b,c,d)\mapsto[\omega_{1}(a,b,c,d)\,:\,\cdots\,:\omega_{6}(a,b,c,d)]\,,
ωi(a,b,c,d)=Σiη𝐗,i=1,,6\omega_{i}(a,b,c,d)=\oint_{\Sigma_{i}}\,\eta_{\mathbf{X}}\,,\,\,\,i=1,\dots,6

with Σ1,,Σ6T(𝐗)\Sigma_{1},\dots,\Sigma_{6}\in\mathrm{T}(\mathbf{X}) a basis, and 𝒫5\mathcal{P}\subset\mathbb{P}^{5} the period domain of six linearly independent period integrals of the family 𝐗\mathbf{X} in Equation (3.4). Similarly, for the family 𝒳\mathcal{X} in Equation (3.12) let p~:6𝒫\tilde{p}:\mathcal{M}_{6}\to\mathcal{P} be the period map as defined by Matsumoto [54, §7]. Let AA be the Gram matrix of the lattice H(2)H(2)22H(2)\oplus H(2)\oplus\langle-2\rangle^{\oplus 2}, and let G𝒳GL(6,)G_{\mathcal{X}}\subset\mathrm{GL}(6,\mathbb{Z}) be the subgroup of the isometry group O(A,)\mathrm{O}(A,\mathbb{Z}) given by

(3.18) G𝒳={MGL(6,)|MTAM=A,M𝕀mod2}O(A,).G_{\mathcal{X}}=\left\{M\in\mathrm{GL}(6,\mathbb{Z})\;|\;M^{T}AM=A,\;M\equiv\mathbb{I}\!\mod 2\;\right\}\ \subset\ \mathrm{O}(A,\mathbb{Z})\,.

We have the following:

Proposition 3.6.

The global monodromy group G𝐗GL(6,)G_{\mathbf{X}}\subset\mathrm{GL}(6,\mathbb{Z}) of the period map p:𝒫p~{}:\mathcal{M}\to\mathcal{P} for the family 𝐗\mathbf{X} in Equation (3.4) is, up to conjugacy, the group G𝒳G_{\mathcal{X}}.

Proof.

In [54, §7], Matsumoto et al. showed that the monodromy group of the period map p~:6𝒫\tilde{p}:\mathcal{M}_{6}\to\mathcal{P} for the family 𝒳\mathcal{X} coincides with that of the monodromy group for the Aomoto-Gel’fand E(3,6)E(3,6) system, and is given by the group G𝒳O(A,)G_{\mathcal{X}}\subset\mathrm{O}(A,\mathbb{Z}) in Equation (3.18). They showed this group is the topological monodromy group of 𝒳\mathcal{X}, i.e., the monodromy group of the local system R2π¯𝒳6R^{2}\pi_{*}\underline{\mathbb{Z}}_{\mathcal{X}}\to\mathcal{M}_{6}. It then follows from Proposition 3.5 that G𝒳G𝐗G_{\mathcal{X}}\subseteq G_{\mathbf{X}}. For μ=b(bc)\mu=b(b-c), the multi-valued functions μω\sqrt{\mu}\,\omega were shown to be solutions to Aomoto-Gel’fand E(3,6)E(3,6) system. Hence, the tensor product of local systems 𝖲¯Π\mathsf{S}\otimes_{\underline{\mathbb{Z}}_{\mathcal{M}}}\Pi is the span of solutions to the Picard-Fuchs system for the family 𝐗\mathbf{X}, where 𝖲\mathsf{S} is the rank one integral local system defined above. The order-two monodromy group 2\mathbb{Z}_{2} is generated by the monodromy around the vanishing locus of μ\mu, and Π\Pi is the rank six period sheaf.

Let Λ\Lambda be subset of the parameter space corresponding to singular members of the family 𝒳\mathcal{X}. Let gγg_{\gamma} be the monodromy operator acting on the cohomology of 𝒳\mathcal{X} for any loop γ\gamma in 4\(ΛZ(μ))\mathbb{C}^{4}\backslash\big{(}\Lambda\cup Z(\mu)\big{)}. The corresponding monodromy operator hγh_{\gamma} attached to the same loop applied to the cohomology of 𝐗\mathbf{X} satisfies hγ=±gγh_{\gamma}=\pm g_{\gamma} by Equation (3.16). Since 𝕀G𝒳-\mathbb{I}\in G_{\mathcal{X}} it follows that hγG𝒳h_{\gamma}\in G_{\mathcal{X}} and G𝐗{±𝕀}=G𝒳G_{\mathbf{X}}\cdot\{\pm\mathbb{I}\}=G_{\mathcal{X}}. Since Z(μ)ΛZ(\mu)\not\subset\Lambda, it follows that 𝕀G𝐗-\mathbb{I}\in G_{\mathbf{X}}. In fact, for a loop in Z(μ)\mathcal{M}\cap Z(\mu) away from the singular locus of 𝐗\mathbf{X}, the monodromy operator acts nontrivial on the first factor of 𝖲¯Π\mathsf{S}\otimes_{\underline{\mathbb{Z}}_{\mathcal{M}}}\Pi alone. Hence, we have the equality G𝐗=G𝒳G_{\mathbf{X}}=G_{\mathcal{X}}. ∎

Remark 3.7.

The proof of Proposition 3.6 shows that the monodromy group of the family 𝐗\mathbf{X} is the same as that of 𝒳\mathcal{X} while the monodromy representations are different. Similar statements hold about the monodromy groups in Corollary 3.16, Corollary 3.20, and Corollary 3.25.

3.3. Extensions of the lattice polarization

Using the four-parameter family of K3 surfaces in Proposition 3.4, we can efficiently study certain extensions of the lattice polarization and identify the corresponding lattice polarizations, monodromy groups, and Picard-Fuchs operators.

3.3.1. Picard rank ρ=17\rho=17

We consider the extension of the lattice polarization for d=0d=0. In this case, the surface 𝐗a,b,c=𝐗a,b,c,0\mathbf{X}^{\prime}_{a,b,c}=\mathbf{X}_{a,b,c,0} becomes the twisted Legendre Pencil:

(3.19) y2=x(x1)(xt)(ta)(tb)(tc).y^{2}=x(x-1)(x-t)(t-a)(t-b)(t-c).

The minimal resolution of a general member has Picard number 17 and was studied by Hoyt [38]. We have the following:

Lemma 3.8.

Equation (3.19) defines a Jacobian elliptic fibration π:𝐗1\pi:\mathbf{X}^{\prime}\to~{}\mathbb{P}^{1} on a general 𝐗=𝐗a,b,c\mathbf{X}^{\prime}=\mathbf{X}^{\prime}_{a,b,c} with three singular fibers of Kodaira type I0I_{0}^{*} over t=a,b,ct=a,b,c, three singular fibers of Kodaira type I2I_{2}, and the Mordell Weil group MW(𝐗,π)=(/2)2\mathrm{MW}(\mathbf{X}^{\prime},\pi)=(\mathbb{Z}/2\mathbb{Z})^{2}.

Proof.

The proof is similar to the ones given in the preceding section. The statement about Picard rank and the Mordell Weil group can be found in Hoyt [38]. ∎

In particular, 𝐗\mathbf{X}^{\prime} is birational to the two-parameter quadratic twist family of the one parameter family of rational elliptic surfaces Sc,d=0S_{c,d=0} from Lemma 3.1, and hence, 𝐗\mathbf{X}^{\prime} is equivalently described by the mixed-twist construction with generalized functional invariant (i,j,α)=(1,1,1)(i,j,\alpha)=(1,1,1). We have the following:

Proposition 3.9.

Over =|d=0\mathcal{M}^{\prime}=\mathcal{M}|_{d=0} the family 𝐗a,b,c\mathbf{X}^{\prime}_{a,b,c} in Equation (3.19) is a 3-dimensional family of LL^{\prime}-polarized K3 surfaces 𝐗\mathbf{X}^{\prime} where LL^{\prime} has rank 17 and the following isomorphic presentations:

(3.20) LHE8(1)D4(1)A1(1)3HE7(1)D4(1)2HD12(1)A1(1)3HD10(1)D4(1)A1(1)HD8(1)D6(1)A1(1).\begin{split}L^{\prime}\cong\ &H\oplus E_{8}(-1)\oplus D_{4}(-1)\oplus A_{1}(-1)^{\oplus 3}\ \cong\ H\oplus E_{7}(-1)\oplus D_{4}(-1)^{\oplus 2}\\ \ \cong\ &H\oplus D_{12}(-1)\oplus A_{1}(-1)^{\oplus 3}\ \cong\ H\oplus D_{10}(-1)\oplus D_{4}(-1)\oplus A_{1}(-1)\\ \cong\ &H\oplus D_{8}(-1)\oplus D_{6}(-1)\oplus A_{1}(-1).\end{split}

In particular, LL^{\prime} is a primitive sub-lattice of the K3 lattice ΛK3\Lambda_{K3}.

Proof.

We use the same strategy as in the proof of Proposition 3.5. Using Lemma 3.8 it follows that the two-elementary lattice LL^{\prime} must have ρ=17\rho=17 and =5\ell=5. Applying Nikulin’s classification [63] it follows that there is only one such lattice admitting a primitive lattice embedding into ΛK3\Lambda_{K3}, and it must have δ=1\delta=1. We then go through the list in [67] to find the isomorphic presentations. ∎

Remark 3.10.

In [12] it was shown that the configuration of six lines \boldsymbol{\ell} associated with 𝐗\mathbf{X}^{\prime} has three lines intersecting in one point. The pencil of lines through the intersection point induces precisely the elliptic fibration of Lemma 3.8. In particular, the general K3 surface 𝐗\mathbf{X}^{\prime} is not a Jacobian Kummer surface. It is the relative Jacobian fibration of an elliptic Kummer surface associated to an abelian surface with a polarization of type (1,2)(1,2); this was proved in [12, 14].

Setting d=0d=0 in Proposition 3.5 we immediately obtain the following:

Corollary 3.11.

Let ΣT(𝐗)\Sigma\in\mathrm{T}(\mathbf{X}^{\prime}) be a transcendental cycle on a general K3 surface 𝐗=𝐗a,b,c\mathbf{X}^{\prime}=\mathbf{X}^{\prime}_{a,b,c}, η𝐗\eta_{\mathbf{X}^{\prime}} the holomorphic two-form induced by dtdx/ydt\wedge dx/y in Equation (3.19), and ω=Ση𝐗\omega=\oint_{\Sigma}\eta_{\mathbf{X}^{\prime}} a period. The Picard-Fuchs system for 𝐗a,b,c\mathbf{X}^{\prime}_{a,b,c}, annihilating ω=b(bc)ω\omega^{\prime}=\sqrt{b(b-c)}\,\omega, is the restricted rank-five Aomoto-Gel’fand system E(3,6)E(3,6) of [52, 53, 54] with x4=0x_{4}=0.

To determine the global monodromy group of the period map for the twisted Legendre pencil, we utilize the relation of 𝐗\mathbf{X}^{\prime} to the Kummer surface Kum(A)\mathrm{Kum}(A) of a principally polarized abelian surface AA. This is equivalent to determining which configurations of six lines \boldsymbol{\ell} yield total spaces that are Kummer surfaces; in particular, the lines must be mutually tangent to a common conic. In [5] the authors gave geometric characterizations of such six-line configurations. We have the following:

Proposition 3.12.

The minimal resolution of a general member in Equation (3.4) is a Jacobian Kummer surface, i.e., the Kummer surface associated with the Jacobian of a general genus-two curve, if and only if d(abb)=(ac)(bc)d(ab-b)=(a-c)(b-c).

Proof.

Using the methods of [13] we compute the square of the degree-two Dolgachev-Ortland invariant R2R^{2}. It vanishes if and only if the six lines are tangent to a common conic. It is well known that this is a necessary and sufficient criterion for the total space to be a Jacobian Kummer surface; see for example [11]. A direct computation of R2R^{2} for the six lines in Equation (3.4) yields the result. ∎

We also have the following:

Lemma 3.13.

For general parameters a,b,ca,b,c and d=(ac)(bc)/(abc)d=(a-c)(b-c)/(ab-c) Equation (3.4) defines a Jacobian elliptic fibration π:𝐗~1\pi:\widetilde{\mathbf{X}}\to~{}\mathbb{P}^{1} with the singular fibers 2I0+6I22I_{0}^{*}+6I_{2} and the Mordell Weil group MW(𝐗~,π)=(/2)21\mathrm{MW}(\widetilde{\mathbf{X}},\pi)=(\mathbb{Z}/2\mathbb{Z})^{2}\oplus\langle 1\rangle.

The connection between the parameters a,b,ca,b,c and the moduli of genus-two curves was exploited in [50, 3]. We have the following:

Proposition 3.14.

Over the subspace ~\widetilde{\mathcal{M}}, given as d=(ac)(bc)/(abc)d=(a-c)(b-c)/(ab-c) in \mathcal{M}, the family in Equation (3.1) is a three-dimensional family of L~\tilde{L}-polarized K3 surfaces 𝐗~\widetilde{\mathbf{X}} where L~\tilde{L} has the following isomorphic presentations:

(3.21) L~HD8(1)D4(1)A3(1)HD7(1)D4(1)2.\tilde{L}\ \cong\ H\oplus D_{8}(-1)\oplus D_{4}(-1)\oplus A_{3}(-1)\ \cong\ H\oplus D_{7}(-1)\oplus D_{4}(-1)^{\oplus 2}\,.

In particular, L~\tilde{L} is a primitive sub-lattice of the K3 lattice ΛK3\Lambda_{K3}.

Proof.

We established in Proposition 3.12 that the K3 surface obtained from the Weierstrass model in Equation (3.4) is a Jacobian Kummer surface if and only if the parameters a,b,c,da,b,c,d satisfy a certain relation. In [44] Kumar classified all Jacobian elliptic fibrations on a generic Kummer surface. Among them are exactly two fibrations that have a trivial Mordell Weil group, called (15) and (17). The types of reducible fibers in the two fibrations then yield isomorphic presentations for the polarizing lattice. ∎

Remark 3.15.

It was shown in [12] that the general K3 surface 𝐗~\widetilde{\mathbf{X}} in Proposition 3.14 arises as the rational double cover of a general K3 surface in Proposition 3.8. The double cover 𝐗~𝐗\widetilde{\mathbf{X}}\dasharrow\mathbf{X}^{\prime} is branched along the even eight on 𝐗\mathbf{X}^{\prime} composed of the non-central components of the two reducible fibers of type D~4\widetilde{D}_{4}.

We now determine the monodromy group for the period map of the twisted Legendre pencil 𝐗\mathbf{X}^{\prime} in Equation (3.19). Notice that the period map for this family is the restriction pp\!\mid_{\mathcal{M}^{\prime}} to \mathcal{M}^{\prime} of the period map from Equation (3.17). We define a rank-one integral local system 𝖲3Z(μ)\mathsf{S}^{\prime}\to\mathbb{C}^{3}-Z(\mu), by restricting the local system 𝖲\mathsf{S} defined above as 𝖲=𝖲|d=0\mathsf{S}^{\prime}=\mathsf{S}|_{d=0}. The monodromy around the locus μ=0\mu=0 obtained by switching branches of the square root function and is again 2\mathbb{Z}_{2}.

In the following, for a matrix group GGL(n,)G\subseteq\mathrm{GL}(n,\mathbb{Z}), identified with its standard representation acting on n\mathbb{Z}^{n}, let 2GGL(r,)\wedge^{2}\,G\subseteq\mathrm{GL}(r,\mathbb{Z}) be the exterior square representation acting on r\mathbb{Z}^{r}, with r=(n2)r=\binom{n}{2}. In the following result, the exterior square representation of the group GG turns out to be reducible on r\mathbb{Z}^{r}, but irreducible on r1\mathbb{Z}^{r-1}. Let Γ2(2)Sp(4,)\Gamma_{2}(2)\subset~{}\mathrm{Sp}(4,\mathbb{Z}) be the Siegel congruence subgroup of level two. Hara et al. showed in [31] that the exterior square representation 2Γ2(2)GL(6,)\wedge^{2}\,\Gamma_{2}(2)\subset\mathrm{GL}(6,\mathbb{Z}) of the Siegel congruence subgroup of level two Γ2(2)Sp(4,)\Gamma_{2}(2)\subset\mathrm{Sp}(4,\mathbb{Z}) is reducible on 6\mathbb{Z}^{6}, but irreducible on 5\mathbb{Z}^{5}. Hence, we have 2Γ2(2)GL(5,)\wedge^{2}\,\Gamma_{2}(2)\subset\mathrm{GL}(5,\mathbb{Z}).

Corollary 3.16.

The global monodromy group G𝐗GL(5,)G_{\mathbf{X}^{\prime}}\subset\mathrm{GL}(5,\mathbb{Z}) of the period map pp\!\mid_{\mathcal{M}^{\prime}} is, up to conjugacy, the exterior square G𝐗=2Γ2(2)G_{\mathbf{X}^{\prime}}=\wedge^{2}\,\Gamma_{2}(2).

Proof.

The period map pp\!\mid_{\mathcal{M}^{\prime}} of the twisted Legendre pencil in Equation (3.19) was originally investigated by Hoyt in [38], where a partial analysis of its behavior for generic parameter values a,b,ca,b,c was made. There, Hoyt showed [38, §5, statements (iv), (iv′′)] that 𝐗\mathbf{X}^{\prime} was related the Kummer surface 𝐗~=Kum(Jac(C))\widetilde{\mathbf{X}}=\mathrm{Kum}(\mathrm{Jac}(C)) of a Jacobian of a general genus-two curve CC. In Braeger et al. [6, Theorem 3.12], the authors produced a dominant rational rational map ψ:𝐗~𝐗\psi:\widetilde{\mathbf{X}}\dashrightarrow\mathbf{X}^{\prime} of degree two that explicitly related the twisted Legendre parameters a,b,ca,b,c to the Rosenhain roots λ1,λ2,λ3\lambda_{1},\lambda_{2},\lambda_{3} of the genus two curve CC that pulls back the holomorphic two-form η𝐗=dtdx/y\eta_{\mathbf{X}^{\prime}}=dt\wedge dx/y to the holomorphic two-form η𝐗~\eta_{\widetilde{\mathbf{X}}} on the Kummer surface 𝐗~\widetilde{\mathbf{X}}. In particular, the induced map on homology ψ:H2(𝐗~,)H2(𝐗,)\psi_{*}:H_{2}(\widetilde{\mathbf{X}},\mathbb{C})\to H_{2}\left(\mathbf{X}^{\prime},\mathbb{C}\right) is compatible with the associated lattice polarizations LL^{\prime} on 𝐗\mathbf{X}^{\prime} and L~\tilde{L} on 𝐗~\widetilde{\mathbf{X}}. Thus, the Picard-Fuchs systems for 𝐗\mathbf{X}^{\prime} and 𝐗~\widetilde{\mathbf{X}} are equivalent. Hara et al. showed in [31] showed that the global monodromy group of the Picard-Fuchs system for 𝐗~\widetilde{\mathbf{X}} is precisely this exterior square representation 2Γ2(2)\wedge^{2}\,\Gamma_{2}(2). Hence, we have that 2Γ2(2)G𝐗\wedge^{2}\,\Gamma_{2}(2)\subseteq G_{\mathbf{X}^{\prime}}. Let Π\Pi^{\prime}\to\mathcal{M}^{\prime} is the rank five period sheaf of the family 𝐗\mathbf{X}^{\prime}, and 𝖲\mathsf{S}^{\prime} the rank one integral local system defined above. Then the argument in Proposition 3.6 applies to the tensor product 𝖲¯Π\mathsf{S}^{\prime}\otimes_{\underline{\mathbb{Z}}_{\mathcal{M}^{\prime}}}\Pi^{\prime} generated by solutions to the Picard-Fuchs equations in Corollary  3.11, and it follows that the full monodromy group is G𝐗=2Γ2(2)G_{\mathbf{X}^{\prime}}=\wedge^{2}\,\Gamma_{2}(2), as desired. ∎

3.3.2. Picard rank ρ=18\rho=18

We consider the extension of the lattice polarization for c=d=0c=d=0. In this case, the surface 𝐗a,b′′=𝐗a,b,0,0\mathbf{X}^{\prime\prime}_{a,b}=\mathbf{X}_{a,b,0,0} becomes the two-parameter twisted Legendre pencil:

(3.22) y2=x(x1)(xt)t(ta)(tb).y^{2}=x(x-1)(x-t)t(t-a)(t-b).

The minimal resolution of a general member of this family has Picard number 18. We have the following:

Lemma 3.17.

Equation (3.22) defines a Jacobian elliptic fibration π:𝐗′′1\pi:\mathbf{X}^{\prime\prime}\to~{}\mathbb{P}^{1} on a general 𝐗′′=𝐗a,b′′\mathbf{X}^{\prime\prime}=\mathbf{X}^{\prime\prime}_{a,b} with the singular fibers I2+2I0+2I2I_{2}^{*}+2I_{0}^{*}+2I_{2} and the Mordell Weil group MW(𝐗′′,π)=(/2)2\mathrm{MW}(\mathbf{X}^{\prime\prime},\pi)=(\mathbb{Z}/2\mathbb{Z})^{2}.

We then have the following:

Proposition 3.18.

Over ′′=|c=d=0\mathcal{M}^{\prime\prime}=\mathcal{M}|_{c=d=0} the family 𝐗a,b′′\mathbf{X}^{\prime\prime}_{a,b} in Equation (3.27) is a 2-dimensional family of L′′L^{\prime\prime}-polarized K3 surfaces 𝐗′′\mathbf{X}^{\prime\prime} where L′′L^{\prime\prime} has rank 18 and the following isomorphic presentations:

(3.23) L′′HE8(1)D6(1)A1(1)2HE7(1)2A1(1)2HE7(1)D8(1)A1(1)HD14(1)A1(1)2HD10(1)D6(1).\begin{split}L^{\prime\prime}\cong\ &H\oplus E_{8}(-1)\oplus D_{6}(-1)\oplus A_{1}(-1)^{\oplus 2}\ \cong\ H\oplus E_{7}(-1)^{\oplus 2}\oplus A_{1}(-1)^{\oplus 2}\\ \ \cong\ &H\oplus E_{7}(-1)\oplus D_{8}(-1)\oplus A_{1}(-1)\ \cong\ H\oplus D_{14}(-1)\oplus A_{1}(-1)^{\oplus 2}\\ \cong\ &H\oplus D_{10}(-1)\oplus D_{6}(-1).\end{split}

In particular, L′′L^{\prime\prime} is a primitive sub-lattice of the K3 lattice ΛK3\Lambda_{K3}.

Proof.

We use the same strategy as in the proof of Proposition 3.5. Using Lemma 3.17 it follows that the two-elementary lattice L′′L^{\prime\prime} must have ρ=18\rho=18 and =4\ell=4. Applying Nikulin’s classification [63] it follows that there are two such lattices admitting a primitive lattice embedding into ΛK3\Lambda_{K3}, namely the ones with δ=0,1\delta=0,1. A standard lattice computation shows that we have δ=1\delta=1. We then go through the list in [67] to find the isomorphic presentations. ∎

From [9, Corollary 2.2], the Picard-Fuchs system can now be determined explicitly:

Proposition 3.19.

Let ΣT(𝐗′′)\Sigma\in\mathrm{T}(\mathbf{X}^{\prime\prime}) be a transcendental cycle on a general K3 surface 𝐗′′\mathbf{X}^{\prime\prime}, η𝐗′′\eta_{\mathbf{X}^{\prime\prime}} the holomorphic two-form induced by dtdx/ydt\wedge dx/y in Equation (3.22), and ω=Ση𝐗′′\omega=\oint_{\Sigma}\eta_{\mathbf{X}^{\prime\prime}} a period. The Picard-Fuchs system for 𝐗a,b′′\mathbf{X}^{\prime\prime}_{a,b}, annihilating ω=ω/a\omega^{\prime}=\omega/\sqrt{a}, is the Appell’s rank four hypergeometric system F2(12,12,12;1,1|1ba,b)F_{2}(\frac{1}{2},\frac{1}{2},\frac{1}{2};1,1|1-\frac{b}{a},b).

Proof.

We consider the transformation φ:𝐗(μ)𝐗′′\varphi:\mathbf{X}^{(\mu)}\dashrightarrow\mathbf{X}^{\prime\prime} given by

t=aba+(ba)T,x=1X,y=ab(ba)Y~(a+(ba)T)2X2.t=\frac{ab}{a+(b-a)T}\;,\hskip 14.22636ptx=\frac{1}{X}\;,\hskip 14.22636pty=\frac{ab(b-a)\tilde{Y}}{(a+(b-a)T)^{2}X^{2}}\,.

Here, 𝐗(μ)\mathbf{X}^{(\mu)} is the twisted Legendre pencil

(3.24) μY~2=X(1X)T(1T)(1aTbX),\mu\,\tilde{Y}^{2}=X(1-X)T(1-T)(1-a^{\prime}T-bX)\,,

with a=1b/aa^{\prime}=1-b/a and μ=(1a)/b=1/a\mu=(1-a^{\prime})/b=1/a. The map φ:𝐗(μ)𝐗′′\varphi:\mathbf{X}^{(\mu)}\dashrightarrow\mathbf{X}^{\prime\prime} induces a birational equivalence extending to a birational map of K3 surfaces such that

(3.25) φdtdxy=dTdXY~.\varphi^{*}\frac{dt\wedge dx}{y}=\frac{dT\wedge dX}{\tilde{Y}}.

It is known that periods ω\omega^{\prime} of the two-form dTdX/YdT\wedge dX/Y for the (untwisted) family with Y=μY~Y=\sqrt{\mu}\,\tilde{Y} and

(3.26) Y2=X(1X)T(1T)(aT+bX1)Y^{2}=X(1-X)T(1-T)(a^{\prime}T+bX-1)\,

satisfy the Appell’s hypergeometric system of F2(12,12,12;1,1|a,b)F_{2}(\frac{1}{2},\frac{1}{2},\frac{1}{2};1,1|a^{\prime},b). Thus, periods ω\omega of dtdx/ydt\wedge dx/y for 𝐗′′\mathbf{X}^{\prime\prime} satisfy the same differential system as ω/μ\omega^{\prime}/\sqrt{\mu}. ∎

We now determine the monodromy group for the period map for the family 𝐗′′\mathbf{X}^{\prime\prime} in Equation (3.22). In this case, the period map coincides with the restriction of the period map p′′p\!\mid_{\mathcal{M}^{\prime\prime}} from Equation (3.17). Again, we introduce a rank-one integral local system 𝖲′′2Z(1/μ)\mathsf{S}^{\prime\prime}\to\mathbb{C}^{2}-Z(1/\mu), with μ=1/a\mu=1/a, to record the monodromy around the locus μ=0\mu=0 obtained by switching branches of the square root function.

For a matrix group GGL(n,)G\subseteq\mathrm{GL}(n,\mathbb{Z}), identified with its standard representation acting on n\mathbb{Z}^{n}, let GGGL(2n,)G\boxtimes G\subseteq\mathrm{GL}(2n,\mathbb{Z}) be the outer tensor product representation of GG acting on 2n\mathbb{Z}^{2n}. Let Γ(2)SL(2,)\Gamma(2)\subset\mathrm{SL}(2,\mathbb{Z}) the principal congruence subgroup of level two.

Corollary 3.20.

The global monodromy group G𝐗′′G_{\mathbf{X}^{\prime\prime}} of the period map p′′p\!\mid_{\mathcal{M}^{\prime\prime}} is, up to conjugacy, the outer tensor product G𝐗′′=Γ(2)Γ(2)G_{\mathbf{X}^{\prime\prime}}=\Gamma(2)\boxtimes\Gamma(2).

Proof.

In [9, Theorem 2.5], Clingher et al. showed that the period integral of the twisted Legendre pencil in Equation (3.22) of Picard rank ρ18\rho\geq 18 factorizes holomorphically into two copies of the Gauss hypergeometric function F12(12,12,1|){}_{2}F_{1}(\frac{1}{2},\frac{1}{2},1\,|\;\cdot\;). At the level of Picard-Fuchs systems, this is realized as the decoupling of the rank four Fuchsian system annihilating Appell’s F2F_{2} function from Proposition 3.19 into two copies of the rank two Fuchsian ODE annihilating F12(12,12,1|){}_{2}F_{1}(\frac{1}{2},\frac{1}{2},1\,|\;\cdot\;). The monodromy group of each ODE is known to be the principal congruence subgroup of level two Γ(2)SL(2,)\Gamma(2)\subset~{}\mathrm{SL}(2,\mathbb{Z}). It follows that Γ(2)Γ(2)G𝐗′′\Gamma(2)\boxtimes\Gamma(2)\subseteq G_{\mathbf{X}^{\prime\prime}}. Let Π′′′′\Pi^{\prime\prime}\to\mathcal{M}^{\prime\prime} be the rank four period sheaf of the family 𝐗′′\mathbf{X}^{\prime\prime}, and 𝖲′′\mathsf{S}^{\prime\prime} the rank one integral local system defined above. We apply the argument from the proof of Proposition 3.6 to the tensor product 𝖲′′¯′′Π′′\mathsf{S}^{\prime\prime}\otimes_{\underline{\mathbb{Z}}_{\mathcal{M}^{\prime\prime}}}\Pi^{\prime\prime} generated by solutions to the Picard-Fuchs equations in Proposition 3.19, and obtain the full monodromy group G𝐗′′=Γ(2)Γ(2)G_{\mathbf{X}^{\prime\prime}}=\Gamma(2)\boxtimes\Gamma(2), as desired. ∎

3.3.3. Picard rank ρ=19\rho=19

We consider the extension of the lattice polarization for c=d=0c=d=0 and b=1b=1. In this case, the surface 𝐗a′′′=𝐗a,1,0,0\mathbf{X}^{\prime\prime\prime}_{a}=\mathbf{X}_{a,1,0,0} becomes the one-parameter twisted Legendre pencil:

(3.27) y2=x(x1)(xt)t(t1)(ta).y^{2}=x(x-1)(x-t)t(t-1)(t-a).

This family was studied in detail by Hoyt [36]; the general member has Picard number ρ=19\rho=19. We have the following:

Lemma 3.21.

Equation (3.27) defines a Jacobian elliptic fibration π:𝐗′′′1\pi:\mathbf{X}^{\prime\prime\prime}\to~{}\mathbb{P}^{1} on a general 𝐗′′′=𝐗a′′′\mathbf{X}^{\prime\prime\prime}=\mathbf{X}^{\prime\prime\prime}_{a} with the singular fibers 2I2+I0+2I22I_{2}^{*}+I_{0}^{*}+2I_{2} and the Mordell Weil group MW(𝐗′′′,π)=(/2)2\mathrm{MW}(\mathbf{X}^{\prime\prime\prime},\pi)=(\mathbb{Z}/2\mathbb{Z})^{2}.

We then have the following:

Proposition 3.22.

Over ′′′=|b=1,c=d=0\mathcal{M}^{\prime\prime\prime}=\mathcal{M}|_{b=1,c=d=0} the family 𝐗a′′′\mathbf{X}^{\prime\prime\prime}_{a} in Equation (3.27) is a 1-dimensional family of L′′′L^{\prime\prime\prime}-polarized K3 surfaces 𝐗′′′\mathbf{X}^{\prime\prime\prime} where L′′′L^{\prime\prime\prime} has rank 19 and the following isomorphic presentations:

(3.28) L′′′HE8(1)E7(1)A1(1)2HE7(1)D10(1)HE8(1)D8(1)A1(1)HD16(1)A1(1).\begin{split}L^{\prime\prime\prime}\ \cong\ &H\oplus E_{8}(-1)\oplus E_{7}(-1)\oplus A_{1}(-1)^{\oplus 2}\ \cong\ H\oplus E_{7}(-1)\oplus D_{10}(-1)\\ \cong\ &H\oplus E_{8}(-1)\oplus D_{8}(-1)\oplus A_{1}(-1)\ \cong\ H\oplus D_{16}(-1)\oplus A_{1}(-1).\end{split}

In particular, L′′′L^{\prime\prime\prime} is a primitive sub-lattice of the K3 lattice ΛK3\Lambda_{K3}.

Proof.

We use the same strategy as in the proof of Proposition 3.5. Using Lemma 3.21 it follows that the two-elementary lattice L′′′L^{\prime\prime\prime} must have ρ=19\rho=19 and =3\ell=3. Applying Nikulin’s classification [63] it follows that there is only one such lattice admitting a primitive lattice embedding into ΛK3\Lambda_{K3}, and it must have δ=1\delta=1. We then go through the list in [67] to find the isomorphic presentations. ∎

We have the following:

Proposition 3.23.

Let ΣT(𝐗′′′)\Sigma\in\mathrm{T}(\mathbf{X}^{\prime\prime\prime}) be a transcendental cycle on a general K3 surface 𝐗′′′=𝐗a′′′\mathbf{X}^{\prime\prime\prime}=\mathbf{X}^{\prime\prime\prime}_{a}, η𝐗′′′\eta_{\mathbf{X}^{\prime\prime\prime}} the holomorphic two-form induced by dtdx/ydt\wedge dx/y in Equation (3.27), and ω=Ση𝐗′′′\omega=\oint_{\Sigma}\eta_{\mathbf{X}^{\prime\prime\prime}} a period. The Picard-Fuchs operator for 𝐗a′′′\mathbf{X}^{\prime\prime\prime}_{a}, annihilating ω=ω/a\omega^{\prime}=\omega/\sqrt{a}, is the rank three ordinary differential operator annihilating the generalized hypergeometric function F23(12,12,12;1,1|11a){}_{3}F_{2}(\frac{1}{2},\frac{1}{2},\frac{1}{2};1,1|1-\frac{1}{a}).

Remark 3.24.

The results in Propositions 3.23 and 3.19 are in agreement with [71, Thm. 2.1] and [9] where it was shown that the two restrictions

(3.29) F23(α,β1, 1+αγ2γ1, 1+αγ2+β2|z1)andF2(α;β1,β2γ1,γ2|z1,1){}_{3}F_{2}\!\left(\left.{\genfrac{}{}{0.0pt}{}{\alpha,\,\beta_{1},\,1+\alpha-\gamma_{2}}{\gamma_{1},\,1+\alpha-\gamma_{2}+\beta_{2}}}\right|z_{1}\right)\quad\text{and}\quad F_{2}\!\left(\left.{\genfrac{}{}{0.0pt}{}{\alpha;\;\beta_{1},\beta_{2}}{\gamma_{1},\gamma_{2}}}\right|z_{1},1\right)

satisfy the same ordinary differential equation.

Proof.

We consider the transformation φ:𝐗(μ)𝐗′′′\varphi:\mathbf{X}^{(\mu)}\dashrightarrow\mathbf{X}^{\prime\prime\prime} given by

t=aa+(1a)T,x=a(1X)(a+(1a)T)X,y=(1a)a2Y~(a+(1a)T)3X2,t=\frac{a}{a+(1-a)T}\;,\hskip 14.22636ptx=-\frac{a(1-X)}{(a+(1-a)T)X}\;,\hskip 14.22636pty=-\frac{(1-a)a^{2}\tilde{Y}}{(a+(1-a)T)^{3}X^{2}}\,,

Here, 𝐗(μ)\mathbf{X}^{(\mu)} is the twisted Legendre pencil

(3.30) μY~2=X(1X)T(1T)(1aTX),\mu\,\tilde{Y}^{2}=X(1-X)T(1-T)(1-a^{\prime}TX)\,,

with a=11/aa^{\prime}=1-1/a and μ=1a\mu=1-a^{\prime}. The map φ:𝐗(μ)𝐗′′′\varphi:\mathbf{X}^{(\mu)}\dashrightarrow\mathbf{X}^{\prime\prime\prime} induces a birational equivalence extending to a birational map of K3 surfaces such that

(3.31) φdtdxy=dTdXY~.\varphi^{*}\frac{dt\wedge dx}{y}=\frac{dT\wedge dX}{\tilde{Y}}.

It is known that periods ω\omega^{\prime} of the two-form dTdX/YdT\wedge dX/Y for the (untwisted) family with Y=μY~Y=\sqrt{\mu}\,\tilde{Y} and

(3.32) Y2=X(1X)T(1T)(1aTX)Y^{2}=X(1-X)T(1-T)(1-a^{\prime}TX)\,

satisfy the differential equation of F23(12,12,12;1,1|a){}_{3}F_{2}(\frac{1}{2},\frac{1}{2},\frac{1}{2};1,1|a^{\prime}). Thus, periods ω\omega of dtdx/ydt\wedge dx/y for 𝐗′′′\mathbf{X}^{\prime\prime\prime} satisfy the same differential equation as ω/μ\omega^{\prime}/\sqrt{\mu}. ∎

We determine the monodromy group for the period map for the family 𝐗′′′\mathbf{X}^{\prime\prime\prime} in Equation (3.27). The period map coincides with the restriction of the period map p′′′p\!\mid_{\mathcal{M}^{\prime\prime\prime}} from Equation (3.17). We again define here a rank-one integral local system 𝖲′′′Z(1/μ)\mathsf{S}^{\prime\prime\prime}\to\mathbb{C}-Z(1/\mu) by restricting the local system 𝖲′′\mathsf{S}^{\prime\prime} in Corollary 3.20 and the preceding discussion there as 𝖲′′′=𝖲′′|b=1\mathsf{S}^{\prime\prime\prime}=\mathsf{S}^{\prime\prime}|_{b=1}, as to record the monodromy around the locus 1/μ=01/\mu=0 obtained by switching branches of the square root μ\sqrt{\mu} with μ=1/a\mu=1/a.

In the following, for a matrix group GGL(n,)G\subseteq\mathrm{GL}(n,\mathbb{Q}), identified with its standard representation acting on n\mathbb{Q}^{n}, let GGGL(r,)G\odot G\subseteq\mathrm{GL}(r,\mathbb{Q}) be the symmetric square representation acting on r\mathbb{Z}^{r}, with r=n(n+1)/2r=n(n+1)/2. We also denote by Γ(2):=Γ(2),w\Gamma(2)^{*}:=\langle\Gamma(2),w\rangle with w=(01220)w=\big{(}\begin{smallmatrix}0&-\frac{1}{2}\\ 2&0\end{smallmatrix}\big{)} the Fricke involution.

Corollary 3.25.

The global monodromy group GX′′′GL(3,)G_{X^{\prime\prime\prime}}\subset\mathrm{GL}(3,\mathbb{Z}) of the period map p′′′p\!\mid_{\mathcal{M}^{\prime\prime\prime}} is, up to conjugacy, the direct product G𝐗′′′=Γ(2)Γ(2)G_{\mathbf{X}^{\prime\prime\prime}}=\Gamma(2)^{*}\odot\Gamma(2)^{*}.

Proof.

Equation (3.31) proves that the monodromy group of the ODE annihilating F23(12,12,12;1,1|){}_{3}F_{2}(\frac{1}{2},\frac{1}{2},\frac{1}{2};1,1|\;\cdot\;) is the symmetric square representation in GL(3,)\mathrm{GL}(3,\mathbb{Z}) of the monodromy group for the ODE annihilating F12(12,12;1|){}_{2}F_{1}(\frac{1}{2},\frac{1}{2};1|\;\cdot\;), after adjoining the involution that is generated by the monodromy operator for loops around the singular fiber at t=at=a or, equivalently, t=0t=0. One checks that in terms of the modular parameter the action is conjugate to the action of the Fricke involution ww. Hence, we have Γ(2)Γ(2)G𝐗′′′\Gamma(2)^{*}\odot\Gamma(2)^{*}\subseteq~{}G_{\mathbf{X}^{\prime\prime\prime}}. Let Π′′′′′′\Pi^{\prime\prime\prime}\to\mathcal{M}^{\prime\prime\prime} be the rank three period sheaf of the family 𝐗′′′\mathbf{X}^{\prime\prime\prime}, and 𝖲′′′\mathsf{S}^{\prime\prime\prime} the rank one integral local system defined above. Applying the argument from the proof of Proposition 3.6 to the tensor product 𝖲′′′¯′′′Π′′′\mathsf{S}^{\prime\prime\prime}\otimes_{\underline{\mathbb{Z}}_{\mathcal{M}^{\prime\prime\prime}}}\Pi^{\prime\prime\prime} generated by solutions to the Picard-Fuchs equations in Proposition 3.23, we obtain the full monodromy group G𝐗′′′=Γ(2)Γ(2)G_{\mathbf{X}^{\prime\prime\prime}}=\Gamma(2)^{*}\odot\Gamma(2)^{*}, as  desired. ∎

In general, if LLΛK3L\leqslant L^{\prime}\leqslant\Lambda_{K3} are lattices primitively embedded in the K3 lattice, then there is a map LL\mathcal{M}_{L^{\prime}}\to\mathcal{M}_{L} of moduli spaces which depends on the particular choice of the lattice embeddings. In particular, the map may have degree greater than one. We have constructed a family of K3 surfaces 𝐗a,b,c,d\mathbf{X}_{a,b,c,d} such that the period map (from the base of the family) to the coarse moduli space L\mathcal{M}_{L} of LL-polarized K3 surfaces is birational. We then showed that the restriction of the Weierstrass model for 𝐗a,b,c,d\mathbf{X}_{a,b,c,d} to a suitable subspace \mathcal{M}^{\prime}\subset\mathcal{M} with dim=dimL\dim\mathcal{M}^{\prime}=\dim\mathcal{M}_{L^{\prime}} determines an extension of the lattice polarization L=HKL^{\prime}=H\oplus K^{\prime} of L=HKL=H\oplus K as extension of the associated root lattices Kroot(K)rootK^{\text{root}}\hookrightarrow(K^{\prime})^{\text{root}} in the Weierstrass model. We have the following main result:

Theorem 3.26.

Over the subspaces, obtained by restriction and given by

(3.33) =|d=0′′=|c=d=0′′′=|b=1,c=d=0,\mathcal{M}\ \supset\ \mathcal{M}^{\prime}=\mathcal{M}\Big{|}_{d=0}\ \supset\ \mathcal{M}^{\prime\prime}=\mathcal{M}\Big{|}_{c=d=0}\ \supset\ \mathcal{M}^{\prime\prime\prime}=\mathcal{M}\Big{|}_{b=1,c=d=0}\,,

the polarization of the family 𝐗a,b,c,d\mathbf{X}_{a,b,c,d} extends in a chain of even, indefinite, two-elementary lattices, given by

(3.34) LLL′′L′′′,L\ \leqslant\ L^{\prime}\ \leqslant\ L^{\prime\prime}\ \leqslant\ L^{\prime\prime\prime}\,,

where the lattices are uniquely determined by (rank, length, parity) with (ρ,,δ)=(16+k,6k,1)(\rho,\ell,\delta)=(16+k,6-k,1) for k=0,1,2,3k=0,1,2,3 such that dim(k)=dimL(k)=4k\dim\mathcal{M}^{(k)}=\dim\mathcal{M}_{L^{(k)}}=4-k. Their Picard-Fuchs systems are determined in Proposition 3.5, Corollary 3.11, and Propositions 3.19, 3.23, and the global monodromy groups in Proposition 3.6, and Corollaries 3.16, 3.20, 3.25.

Proof.

Restricting (i) d=0d=0, (ii) c=d=0c=d=0, (iii) b=1,c=d=0b=1,c=d=0 in the family of K3 surfaces in Equation (3.4), the theorem collect statements from Propositions 3.4, 3.9, 3.18, 3.22 and their respective proofs, as well as from Proposition 3.5, Corollary 3.11, Propositions 3.19, 3.23 and Proposition 3.6, Corollaries 3.16, 3.20, 3.25. ∎

4. GKZ Description of the Univariate Mirror Families

In this section we will show that the generalized functional invariant of the mixed-twist construction captures all key features of the one-parameter mirror families for the Fermat pencils. In particular, we will show that the mixed-twist construction allows us to obtain a non-resonant GKZ system for which a basis of solutions in the form of absolutely convergent Mellin-Barnes integrals exists whose monodromy is computed explicitly.

4.1. The Mirror Families

Let us briefly review the construction of the mirror family for the deformed Fermat hypersurface. Let n(n+1)\mathbb{P}^{n}(n+1) be the general family of hypersurfaces of degree (n+1)(n+1) in n\mathbb{P}^{n}. The general member of n(n+1)\mathbb{P}^{n}(n+1) is a smooth hypersurface Calabi-Yau (n1)(n-1)-fold. Let [X0::Xn][X_{0}:\cdots:X_{n}] be the homogeneous coordinates on n\mathbb{P}^{n}. The following family

(4.1) X0n+1++Xnn+1+nλX0X1Xn=0X_{0}^{n+1}+\cdots+X_{n}^{n+1}+n\lambda X_{0}X_{1}\cdots X_{n}=0

determines a one-parameter single-monomial deformation Xλ(n1)X^{(n-1)}_{\lambda} of the classical Fermat hypersurface in n(n+1)\mathbb{P}^{n}(n+1). Cox and Katz determined [16] what deformations of Calabi-Yau hypersurfaces remain Calabi-Yau. For example, for n=5n=5 there are 101 parameters for the complex structure, which determine the coefficients of additional terms in the quintic polynomials. Starting with a Fermat-type hypersurfaces VV in n\mathbb{P}^{n}, Yui [76, 75, 70] and Goto [25] classified all discrete symmetries GG such that the quotients V/GV/G are singular Calabi-Yau varieties with at worst Abelian quotient singularities. A theorem by Greene, Roan, and Yau [26] guarantees that there are crepant resolutions of V/GV/G. This is known as the Greene-Plesser orbifolding construction.

For the family (4.1), the discrete group of symmetries needed for the Greene-Plesser orbifolding is readily constructed: it is generated by the action (X0,Xj)(ζn+1nX0,ζn+1Xj)(X_{0},X_{j})\mapsto(\zeta_{n+1}^{n}X_{0},\zeta_{n+1}X_{j}) for 1jn1\leq j\leq n and the root of unity ζn+1=exp(2πin+1)\zeta_{n+1}=\exp{(\frac{2\pi i}{n+1})}. Since the product of all generators multiplies the homogeneous coordinates by a common phase, the symmetry group is Gn1=(/(n+1))n1G_{n-1}=(\mathbb{Z}/(n+1)\,\mathbb{Z})^{n-1}. One checks that the affine variables

t=(1)n+1λn+1,x1=X1n(n+1)X0X2Xnλ,,xn=X2n(n+1)X0X1Xn1λ,t=\frac{(-1)^{n+1}}{\lambda^{n+1}},\,x_{1}=\frac{X_{1}^{n}}{(n+1)\,X_{0}\cdot X_{2}\cdots X_{n}\,\lambda},\,\dots,\,x_{n}=\frac{X_{2}^{n}}{(n+1)\,X_{0}\cdot X_{1}\cdots X_{n-1}\,\lambda},

are invariant under the action of Gn1G_{n-1}, hence coordinates on the quotient Xλ(n1)/Gn1X^{(n-1)}_{\lambda}/G_{n-1}. A family of special hypersurfaces Yt(n1)Y^{(n-1)}_{t} is then defined by the remaining relation between x1,,xnx_{1},\dots,x_{n}, namely the equation

(4.2) fn(x1,,xn,t)=x1xn(x1++xn+1)+(1)n+1t(n+1)n+1=0.f_{n}(x_{1},\dots,x_{n},t)=x_{1}\cdots x_{n}\,\Big{(}x_{1}+\dots+x_{n}+1\Big{)}+\frac{(-1)^{n+1}\,t}{(n+1)^{n+1}}=0\;.

Moreover, it was proved by Batyrev and Borisov in  [1] that the family of special Calabi-Yau hypersurfaces Yt(n1)Y^{(n-1)}_{t} of degree (n+1)(n+1) in n\mathbb{P}^{n} given by Equation (4.2) is mirror to a general hypersurface n(n+1)\mathbb{P}^{n}(n+1) of degree (n+1)(n+1) and co-dimension one in n\mathbb{P}^{n}, in the sense that the Hodge diamonds are mirror images, hi,j(Xλn1)=hj,i(Ytn1)h^{i,j}(X^{n-1}_{\lambda})=h^{j,i}(Y^{n-1}_{t}) for all n2n\geq 2 and appropriate λ,t\lambda,t. For n=2,3,4n=2,3,4 the mirror family is a family of elliptic curves, K3 surfaces, and Calabi-Yau threefolds, respectively.

Each mirror family can be realized as a fibration of Calabi-Yau (n2)(n-2)-folds associated with a generalized functional invariant. The following was proved by Doran and Malmendier:

Proposition 4.1.

For n2n\geq 2 the family of hypersurfaces Yt(n1)Y^{(n-1)}_{t} in Equation (4.2) is a fibration over 1\mathbb{P}^{1} by hypersurfaces Yt~(n2)Y^{(n-2)}_{\tilde{t}} constructed as mixed-twist with the generalized functional invariant (1,n,1)(1,n,1).

Proof.

For each xn0,1x_{n}\not=0,-1 substituting x~i=xi/(xn+1)\tilde{x}_{i}=x_{i}/(x_{n}+1) for 1in11\leq i\leq n-1 and t~=nnt/((n+1)n+1xn(xn+1)n)\tilde{t}=-n^{n}\,t/((n+1)^{n+1}x_{n}\,(x_{n}+1)^{n}) defines a fibration of the hypersurface (4.2) by fn1(x~1,,x~n1t~)=0f_{n-1}(\tilde{x}_{1},\dots,\tilde{x}_{n-1}\tilde{t})=0 since

(4.3) fn(x1,,xn,t)=xn(xn+1)nfn1(x~1,,x~n1,t~)=0.f_{n}(x_{1},\dots,x_{n},t)=x_{n}\,(x_{n}+1)^{n}\,f_{n-1}(\tilde{x}_{1},\dots,\tilde{x}_{n-1},\tilde{t}\,)=0\;.

This is the mixed-twist construction with generalized functional invariant (1,n,1)(1,n,1). ∎

4.2. GKZ data of the mirror family

In the GKZ formalism, the construction of the family Yt(n1)Y^{(n-1)}_{t} is described as follows: from the homogeneous degrees of the defining Equation (4.1) and the coordinates of the ambient projective space for the family Xλ(n1)X^{(n-1)}_{\lambda} we obtain the lattice 𝕃=((n+1),1,1,,1)n+2\mathbb{L}^{\prime}=\mathbb{Z}(-(n+1),1,1,\dots,1)\subset\mathbb{Z}^{n+2}. We define a matrix 𝖠Mat(n+1,n+2;)\mathsf{A}^{\prime}\in\operatorname{Mat}(n+1,n+2;\mathbb{Z}) as a matrix row equivalent to the (n+1)×(n+2)(n+1)\times(n+2) matrix with columns of the (n+1)×(n+1)(n+1)\times(n+1) identity matrix as the first (n+1)(n+1) columns, followed by the generator of 𝕃\mathbb{L}^{\prime}:

(4.4) (100(n+1)01010100011)𝖠=(111101010100011),\scalebox{0.9}{$\left(\begin{array}[]{ccccc}1&0&0&\dots&(n+1)\\ 0&1&0&\dots&-1\\ 0&\ddots&\ddots&\ddots&-1\\ \vdots&&&&\vdots\\ 0&0&\dots&0\qquad 1&-1\end{array}\right)\quad\sim\quad\mathsf{A}^{\prime}=\left(\begin{array}[]{ccccr}1&1&1&\dots&1\\ 0&1&0&\dots&-1\\ 0&\ddots&\ddots&\ddots&-1\\ \vdots&&&&\vdots\\ 0&0&\dots&0\qquad 1&-1\end{array}\right)$}\;,

and let 𝒜={a1,,an+2}\mathcal{A}^{\prime}=\{\vec{a}^{\prime}_{1},\dots,\vec{a}^{\prime}_{n+2}\} denote the columns of the right-handed matrix obtained by a basis transformation in n+1\mathbb{Z}^{n+1} from the matrix on the left hand side. The finite subset 𝒜n+1\mathcal{A}^{\prime}\subset\mathbb{Z}^{n+1} generates n+1\mathbb{Z}^{n+1} as an abelian group and can be equipped with a group homomorphism h:n+1h^{\prime}:\mathbb{Z}^{n+1}\to\mathbb{Z}, in this case the projection onto the first coordinate, such that h(𝒜)=1h^{\prime}(\mathcal{A}^{\prime})=1. This means that 𝒜\mathcal{A}^{\prime} lies in an affine hyperplane in n+1\mathbb{Z}^{n+1}. The lattice of linear relations between the vectors in 𝒜\mathcal{A}^{\prime} is easily checked to be precisely 𝕃=((n+1),1,1,,1)n+2\mathbb{L}^{\prime}=\mathbb{Z}(-(n+1),1,1,\dots,1)\subset\mathbb{Z}^{n+2}. From 𝖠\mathsf{A}^{\prime} we form the Laurent polynomial

P𝖠(z1,,zn+1)=a𝒜caz1a1z2a2zn+1an+1=c1z1+c2z1z2+c3z1z3++cn+2z1z21zn+11,\begin{split}P_{\mathsf{A}^{\prime}}(z_{1},\dots,z_{n+1})&=\sum_{\vec{a}^{\prime}\in\mathcal{A}^{\prime}}c_{\vec{a}}\,z_{1}^{a_{1}}\cdot z_{2}^{a_{2}}\cdots z_{n+1}^{a_{n+1}}\\ &=c_{1}\,z_{1}+c_{2}\,z_{1}\,z_{2}+c_{3}\,z_{1}\,z_{3}+\dots+c_{n+2}z_{1}\,z_{2}^{-1}\cdots z_{n+1}^{-1}\;,\end{split}

and observe that the dehomogenized Laurent polynomial yields

x1xnc1P𝖠(1,c1x1c2,c1x2c3,,c1xncn+1)=fn(x1,,xn,t=(1)n+1(n+1)n+1c2cn+2c1n+1).\frac{x_{1}\cdots x_{n}}{c_{1}}\;P_{\mathsf{A}^{\prime}}\left(1,\frac{c_{1}x_{1}}{c_{2}},\frac{c_{1}x_{2}}{c_{3}},\dots,\frac{c_{1}x_{n}}{c_{n+1}}\right)\\ =f_{n}\left(x_{1},\dots,x_{n},t=(-1)^{n+1}\frac{(n+1)^{n+1}\,c_{2}\cdots c_{n+2}}{c_{1}^{n+1}}\right)\;.

In the context of toric geometry, this is interpreted as follows: a secondary fan is constructed from the data (𝒜,𝕃)(\mathcal{A}^{\prime},\mathbb{L}^{\prime}). This secondary fan is a complete fan of strongly convex polyhedral cones in 𝕃=Hom(𝕃,)\mathbb{L}^{\prime\vee}_{\mathbb{R}}=\operatorname{Hom}(\mathbb{L}^{\prime},\mathbb{R}) which are generated by vectors in the lattice 𝕃=Hom(𝕃,)\mathbb{L}^{\prime\vee}_{\mathbb{Z}}=\operatorname{Hom}(\mathbb{L}^{\prime},\mathbb{Z}). As the coefficients c1,,cn+2c_{1},\dots,c_{n+2} – or effectively tt – vary, the zero locus of P𝒜P_{\mathcal{A}^{\prime}} sweeps out the family of hypersurfaces Yt(n1)Y^{(n-1)}_{t} in ()n+1/=()n(\mathbb{C}^{*})^{n+1}/\mathbb{C}^{*}=(\mathbb{C}^{*})^{n}. Both ()n(\mathbb{C}^{*})^{n} and the hypersurfaces can then be compactified. The members of the family Yt(n1)Y^{(n-1)}_{t} are Calabi-Yau varieties since the original Calabi-Yau varieties had codimension one in the ambient space; see Batyrev and van Straten [2].

4.3. Recurrence relation between holomorphic periods

We now describe the construction of the period integrals. A result of Doran and Malmendier – referenced below as Lemma 4.2 – shows that the fibration on Yt(n1)1Y^{(n-1)}_{t}\to\mathbb{P}^{1} by Calabi-Yau hypersurfaces Yt~(n2)Y^{(n-2)}_{\tilde{t}} allows for a recursive construction of the period integrals for Yt(n1)Y^{(n-1)}_{t} by integrating a twisted period integral over a transcendental homology cycle. It turns out that the result can be obtained explicitly as the Hadamard product of certain generalized hypergeometric functions. Recall that the Hadamard of two analytic functions f(t)=k0fktkf(t)=\sum_{k\geq 0}f_{k}t^{k}, g(t)=k0gktkg(t)=\sum_{k\geq 0}g_{k}t^{k} is the analytic function fgf\star g given by

(fg)(t)=k=0fkgktk.(f\star g)(t)=\sum_{k=0}^{\infty}f_{k}g_{k}t^{k}.

The unique holomorphic (n1)(n-1)-form on Yt(n1)Y^{(n-1)}_{t} is given by

(4.5) ηt(n1)=dx2dx3dxnx1fn(x1,,xn,t).\eta_{t}^{(n-1)}=\dfrac{dx_{2}\wedge dx_{3}\wedge\dots\wedge dx_{n}}{\partial_{x_{1}}f_{n}(x_{1},\dots,x_{n},t)}\;.

The formula is obtained from the Griffiths-Dwork technique (see, for example, Morrison [56]). One then defines an (n1)(n-1)-cycle Σn1\Sigma_{n-1} on Yt(n1)Y^{(n-1)}_{t} by requiring that the period integral of ηt(n1)\eta^{(n-1)}_{t} over Σn1\Sigma_{n-1} corresponds by a residue computation in x1x_{1} to the integral over the middle dimensional torus cycle Tn1(𝐫):=Sr11××Srn11Hn1(Ytn1,)T_{n-1}(\vec{\mathbf{r}}):=S^{1}_{r_{1}}\times\dots\times S^{1}_{r_{n-1}}\in H_{n-1}(Y^{n-1}_{t},\mathbb{Q}) with Srj1={|x|=rj}S^{1}_{r_{j}}~{}=\{|x|=r_{j}\}\subset\mathbb{C} and 𝐫n1=(r1,,rn1)+n1\vec{\mathbf{r}}_{n-1}=(r_{1},\dots,r_{n-1})\in\mathbb{R}^{n-1}_{+}, i.e.,

(4.6) Σn1dx2dxnx1fn(x1,,xn,t)=c12πiTn1(r)P𝒜(1,c1x1c2,c1x2c3,,c1xncn+1)1dx2x2dxnxn.\underbrace{\int\dots\int}_{\Sigma_{n-1}}\dfrac{dx_{2}\wedge\dots\wedge dx_{n}}{\partial_{x_{1}}f_{n}(x_{1},\dots,x_{n},t)}\\ =\frac{c_{1}}{2\pi i}\underbrace{\int\dots\int}_{T_{n-1}(r)}P_{\mathcal{A}}\left(1,\frac{c_{1}x_{1}}{c_{2}},\frac{c_{1}x_{2}}{c_{3}},\dots,\frac{c_{1}x_{n}}{c_{n+1}}\right)^{-1}\frac{dx_{2}}{x_{2}}\wedge\dots\wedge\frac{dx_{n}}{x_{n}}.

The right hand side of Equation (4.6) is a resonant 𝒜\mathcal{A}-hypergeometric integral in the sense of [24, Thm. 2.7] derived from the data (𝒜,𝕃)(\mathcal{A}^{\prime},\mathbb{L}^{\prime}) and

(4.7) α=α1,β11,,βn1t=1,0,,0t=i=1n+2γiai\vec{\alpha}^{\prime}=\langle\alpha^{\prime}_{1},-\beta^{\prime}_{1}-1,\dots,-\beta^{\prime}_{n}-1\rangle^{t}=\langle-1,0,\dots,0\rangle^{t}=\sum_{i=1}^{n+2}\gamma^{\prime}_{i}\,\vec{a}^{\prime}_{i}

with 𝜸0=(γ1,,γn+2)=(1,0,,0)\boldsymbol{\gamma}^{\prime}_{0}=(\gamma^{\prime}_{1},\dots,\gamma^{\prime}_{n+2})=(-1,0,\dots,0). We will denote the period integral by ωn1(t)=Σn1ηt(n1)\omega_{n-1}(t)=\oint_{\;\Sigma_{n-1}}\eta_{t}^{(n-1)}.

We recall the following result, which connects the GKZ data above to the iterative twist construction of Doran and Malmendier:

Proposition 4.2.

[22, Prop. 7.2] For n1n\geq 1 and |t|1|t|\leq 1, there is a family of transcendental (n1)(n-1)-cycles Σn1\Sigma_{n-1} on Yt(n1)Y^{(n-1)}_{t} such that

(4.8) ωn1(t)=Σn1ηt(n1)=(2πi)n1Fn1n(1n+1nn+11 1|t).\omega_{n-1}(t)=\oint_{\Sigma_{n-1}}\eta_{t}^{(n-1)}=(2\pi i)^{n-1}\,{}_{n}F_{n-1}\!\left(\left.{\genfrac{}{}{0.0pt}{}{\frac{1}{n+1}\quad\dots\quad\frac{n}{n+1}}{1\;\dots\;1}}\right|t\right)\;.

The iterative structure in Proposition 4.1 induces the iterative period relation

(4.9) ωn1(t)=(2πi)Fn1n(1n+1nn+11nn1n|t)ωn2(t)for n2.\begin{split}\omega_{n-1}(t)&=(2\pi i)\,{}_{n}F_{n-1}\!\left(\left.{\genfrac{}{}{0.0pt}{}{\frac{1}{n+1}\quad\dots\quad\frac{n}{n+1}}{\frac{1}{n}\;\dots\;\frac{n-1}{n}}}\right|t\right)\star\omega_{n-2}(t)\quad\text{for $n\geq 2$}.\end{split}

Here, the symbol \star denotes the Hadamard product. The cycles Σn1\Sigma_{n-1} are determined by T~n1(𝐫n1):=nn+1(Tn2(𝐫n2)×Srn21)\tilde{T}_{n-1}(\vec{\mathbf{r}}_{n-1}):=\frac{n}{n+1}\cdot\left(T_{n-2}(\vec{\mathbf{r}}_{n-2})\times S^{1}_{r_{n-2}}\right) as in  (4.6), with rj=1jj+1r_{j}~{}=~{}1-\frac{j}{j+1}, and nn+1(Tn2(𝐫n2)×Srn11)\frac{n}{n+1}~{}\cdot\left(T_{n-2}(\vec{\mathbf{r}}_{n-2})\times S^{1}_{r_{n-1}}\right) indicates that coordinates are scaled by a factor of nn+1\frac{n}{n+1}.

Hence, the iterative structure in Proposition 4.1, namely, the generalized functional invariant (1,n,1)(1,n,1), determines the iterative period relations of the mirror family and the corresponding 𝒜\mathcal{A}-hypergeometric data (𝒜,𝕃,𝜸0)(\mathcal{A}^{\prime},\mathbb{L}^{\prime},\boldsymbol{\gamma}_{0}^{\prime}) in the GKZ formalism.

4.3.1. The mirror family of K3 surfaces

Narumiya and Shiga [58] showed that the mirror family of K3 surfaces in Equation (4.2) with n=3n=3 is birationally equivalent to a family of Weierstrass model. In fact, if we set

(4.10) x1=(4u2λ2+3Xλ2+u3+u)(4u2λ2+3Xλ2+u32u)6λ2u(16u3λ23iYλ2+12Xuλ2+4u4+4u2),x2=16u3λ23iYλ2+12Xuλ2+4u4+4u28u(4u2λ2+3Xλ2+u32u),x3=u2(4u2λ2+3Xλ2+u32u)2λ2(16u3λ23iYλ2+12Xuλ2+4u4+4u2),\begin{split}x_{1}&=-{\frac{\left(4\,{u}^{2}{\lambda}^{2}+3\,X{\lambda}^{2}+{u}^{3}+u\right)\left(4\,{u}^{2}{\lambda}^{2}+3\,X{\lambda}^{2}+{u}^{3}-2\,u\right)}{{6\lambda}^{2}u\left(16\,{u}^{3}{\lambda}^{2}-3\,iY{\lambda}^{2}+12\,Xu{\lambda}^{2}+4\,{u}^{4}+4\,{u}^{2}\right)}}\,,\\ x_{2}&=-\,{\frac{16\,{u}^{3}{\lambda}^{2}-3\,iY{\lambda}^{2}+12\,Xu{\lambda}^{2}+4\,{u}^{4}+4\,{u}^{2}}{8u\left(4\,{u}^{2}{\lambda}^{2}+3\,X{\lambda}^{2}+{u}^{3}-2\,u\right)}}\,,\\ x_{3}&={\frac{{u}^{2}\left(4\,{u}^{2}{\lambda}^{2}+3\,X{\lambda}^{2}+{u}^{3}-2\,u\right)}{{2\lambda}^{2}\left(16\,{u}^{3}{\lambda}^{2}-3\,iY{\lambda}^{2}+12\,Xu{\lambda}^{2}+4\,{u}^{4}+4\,{u}^{2}\right)}}\,,\end{split}

in Equation (4.2), we obtain the Weierstrass equation

(4.11) Y2=4X3g2(u)Xg3(u),Y^{2}=4X^{3}-g_{2}(u)\,X-g_{3}(u)\,,

with coefficients

(4.12) g2=43λ4u2(u4+8λ2u3+(4λ21)(4λ2+1)u2+8λ2u+1),g3=427λ6u3(u2+4λ2u+1)(2u4+16λ2u3+(32λ45)u2+16λ2u+2).\begin{split}g_{2}&=\frac{4}{3\,\lambda^{4}}\,{u}^{2}\,\left(u^{4}+8\lambda^{2}u^{3}+(4\lambda^{2}-1)(4\lambda^{2}+1)u^{2}+8\lambda^{2}u+1\right)\,,\\ g_{3}&=\frac{4}{27\,\lambda^{6}}\,{u}^{3}\,\left(u^{2}+4{\lambda}^{2}u+1\right)\left(2u^{4}+16\lambda^{2}u^{3}+(32\lambda^{4}-5)u^{2}+16\lambda^{2}u+2\right)\,.\end{split}

For generic parameter λ\lambda, Equation (4.11) defines a Jacobian elliptic fibration with the singular fibers 2I4+4I12I_{4}^{*}+4I_{1} and the Mordell-Weil group /21\mathbb{Z}/2\mathbb{Z}\oplus\langle 1\rangle, generated by a two-torsion section and an infinite-order section of height pairing one; see [58, 6]. Using the Jacobian elliptic fibration one has the following:

Proposition 4.3 ([58]).

The family in Equation (4.11) is a family of M2M_{2}-polarized K3 surfaces with M2HE8(1)E8(1)4M_{2}\cong H\oplus E_{8}(-1)\oplus E_{8}(-1)\oplus\langle-4\rangle such that the image of the period map is birational with M2\mathcal{M}_{M_{2}}.

Proposition 4.3 shows why the family (4.11) can be called the mirror family of K3 surfaces. Dolgachev’s mirror symmetry for K3 surfaces identifies marked deformations of K3 surfaces with given Picard lattice NN with a complexified Kähler cone K(M)={x+iy:y,y>0,x,yM}K(M)=\{x+iy:\,\langle y,y\rangle>0,\;x,y\in M_{\mathbb{R}}\} for some mirror lattice MM; see [19]. In the case of the rank-one lattice Nk=2kN_{k}=\langle 2k\rangle, one can construct the mirror lattice explicitly by taking a copy of HH out of the orthogonal complement NkN_{k}^{\perp} in the K3 lattice ΛK3\Lambda_{K3}. It turns out that the mirror lattice MkHE8(1)E8(1)2kM_{k}\cong H\oplus E_{8}(-1)\oplus E_{8}(-1)\oplus\langle-2k\rangle is unique if kk has no square divisor. In our situation, the general quartic hypersurfaces in Equation (4.1) with n=3n=3 have a Néron-Severi lattice isomorphic to N2=4N_{2}=\langle 4\rangle, and the mirror family in Equation (4.11) is polarized by the lattice M2M_{2} such that N2HM2N_{2}^{\perp}\cong H\oplus M_{2}.

It turns out that the holomorphic solution of the Picard-Fuchs equation governing the family of K3 surfaces in Equation (4.11) equals

(4.13) ω2=(F12(18,381|1λ4))2=F23(14,12,341,1|t).\omega_{2}=\left({}_{2}F_{1}\!\left(\left.{\genfrac{}{}{0.0pt}{}{\frac{1}{8},\,\frac{3}{8}}{1}}\right|\frac{1}{\lambda^{4}}\right)\right)^{2}=\ {}_{3}F_{2}\!\left(\left.{\genfrac{}{}{0.0pt}{}{\frac{1}{4},\frac{1}{2},\frac{3}{4}}{1,1}}\right|t\right)\,.

The first equality was proved by Narumiya and Shiga, and the second equality is Clausen’s formula, found by Thomas Clausen, expressing the square of a Gaussian hypergeometric series as a generalized hypergeometric series.

4.4. Monodromy of the mirror family

We will now show how the monodromy representations for the mirror families for general nn are computed using the iterative period relations. The results of this section are consistent with the original work of Levelt [45] up to conjugacy.

The Picard-Fuchs operators of the periods given in Proposition 4.2 are the associated rank nn-hypergeometric differential operators annihilating Fn1n{}_{n}F_{n-1}. But yet more is afforded by pursuing the GKZ description of the period integrals. In fact, the Euler-integral formula for the hypergeometric functions Fn1n{}_{n}F_{n-1} generates a second set of non-resonant GKZ data (𝒜,𝕃,𝜸0)(\mathcal{A},\mathbb{L},\boldsymbol{\gamma}_{0}) from the resonant GKZ data (𝒜,𝕃,𝜸0)(\mathcal{A}^{\prime},\mathbb{L}^{\prime},\boldsymbol{\gamma}^{\prime}_{0}) by integration. The GKZ data (𝒜,𝕃,𝜸0)(\mathcal{A},\mathbb{L},\boldsymbol{\gamma}_{0}) determines local Frobenius bases of solutions around t=0t=0 and t=t=\infty. Their Mellin-Barnes integral representation determines the transition matrix between them by analytic continuation.

We will always assume that we have nn rational parameters, namely ρ1,,ρn(0,1)\rho_{1},\dots,\rho_{n}\in(0,1)\cap\mathbb{Q}, and consider the generalized hypergeometric function

Fn1n(ρ1ρn1 1|t),{}_{n}F_{n-1}\!\left(\left.{\genfrac{}{}{0.0pt}{}{\rho_{1}\;\dots\;\rho_{n}}{1\;\dots\;1}}\right|t\right)\;,

which include all periods from Propositions 4.2 and 3.23. The Euler-integral formula then specializes to the identity

(4.14) [i=1n1Γ(ρi)Γ(1ρi)]Fn1n(ρ1ρn1 1|t)=[i=1n101dzizi1ρi(1zi)ρi](1tz1zn1)ρn.\left[\prod_{i=1}^{n-1}\Gamma(\rho_{i})\,\Gamma(1-\rho_{i})\right]\,{}_{n}F_{n-1}\!\left(\left.{\genfrac{}{}{0.0pt}{}{\rho_{1}\;\dots\;\rho_{n}}{1\;\dots\;1}}\right|t\right)\\ =\left[\prod_{i=1}^{n-1}\int_{0}^{1}\!\frac{dz_{i}}{z_{i}^{1-\rho_{i}}(1-z_{i})^{\rho_{i}}}\right](1-t\,z_{1}\cdots z_{n-1})^{-\rho_{n}}.

The rank-nn hypergeometric differential equation satisfied by Fn1n{}_{n}F_{n-1} is given by

(4.15) [θnt(θ+ρ1)(θ+ρn)]F(t)=0\Big{[}\theta^{n}-t\,(\theta+\rho_{1})\cdots(\theta+\rho_{n})\Big{]}\,F(t)=0

with θ=tddt\theta=t\frac{d}{dt}, and it has the Riemann symbol

(4.16) 𝒫(0100ρ101ρ20n2ρn10n1j=1nρjρn|t).\mathcal{P}\left.\left(\begin{array}[]{ccc}0&1&\infty\\ \hline\cr 0&0&\rho_{1}\\ 0&1&\rho_{2}\\ \vdots&\vdots&\vdots\\ 0&n-2&\rho_{n-1}\\ 0&n-1-\sum_{j=1}^{n}\rho_{j}&\rho_{n}\end{array}\right|t\right)\;.

In particular, we read from the Riemann symbol that for each n1n\geq 1, the periods from Proposition 4.2 have a point of maximally unipotent monodromy at t=0t=0. This is well known to be consistent with basic considerations for mirror symmetry [57].

From the Euler-integral (4.14), using the GKZ formalism, we immediately read off the left hand side matrix, and convert to the A-matrix 𝖠Mat(2n1,2n;)\mathsf{A}\in\mathrm{Mat}(2n-1,2n;\mathbb{Z}) given by

(4.17) (110000001100000011010001000101000001)𝖠=(10001000010001000010001000010001000110000001010000010010),\scalebox{0.9}{$\left(\begin{array}[]{ccccccc}1&1&0&0&\dots&0&0\\ 0&0&1&1&\dots&0&0\\ \vdots&&&&\ddots&&\vdots\\ 0&0&0&0&\;\ddots&1&1\\ \hline\cr 0&1&0&0&\dots&0&1\\ 0&0&0&1&\dots&0&1\\ \vdots&&&&\ddots&&\vdots\\ 0&0&0&0&\;\ddots&0&1\end{array}\right)\quad\sim\quad\mathsf{A}=\left(\begin{array}[]{cccc|c|cccc|c}1&0&\dots&0&0&1&0&\dots&0&0\\ 0&1&&0&0&0&1&&0&0\\ \vdots&&\ddots&\vdots&\vdots&\vdots&&\ddots&\vdots&\vdots\\ 0&0&\dots&1&0&0&0&\dots&1&0\\ \hline\cr 0&0&\dots&0&1&0&0&\dots&0&1\\ \hline\cr 0&0&\dots&0&1&1&0&\dots&0&0\\ 0&0&&0&1&0&1&&0&0\\ \vdots&&\ddots&\vdots&\vdots&\vdots&&\ddots&\vdots&\vdots\\ 0&0&\dots&0&1&0&0&\dots&1&0\\ \end{array}\right)$}\;,

using elementary row operations, as in §4.2. Let 𝒜={a1,,a2n}\mathcal{A}=\{\vec{a}_{1},\dots,\vec{a}_{2n}\} denote the columns of the matrix 𝖠\mathsf{A}. The entries for the matrix on the left hand side of (4.17) are determined as follows: the first nn entries in each column label which of the nn terms (1zi)ρi(1-z_{i})^{\rho_{i}} or (1tz1zn1)ρn(1-t\,z_{1}\cdots z_{n-1})^{-\rho_{n}} in the integrand of the Euler-integral (4.14) is specified. For each term, two column vectors are needed and the entries in rows n+1,,2n1n+1,\dots,2n-1 label the exponents of variables ziz_{i} appearing. For example, the last two columns determine the term (1tz1zn1)ρn(1-t\,z_{1}\cdots z_{n-1})^{-\rho_{n}}. The finite subset 𝒜2n1\mathcal{A}\subset\mathbb{Z}^{2n-1} generates 2n1\mathbb{Z}^{2n-1} as an abelian group and is equipped with a group homomorphism h:2n1h:\mathbb{Z}^{2n-1}\to\mathbb{Z}, in this case the sum of the first nn coordinates such that h(𝒜)=1h(\mathcal{A})=1. The lattice of linear relations between the vectors in 𝒜\mathcal{A} is easily checked to be 𝕃=(1,,1,1,,1)2n\mathbb{L}=\mathbb{Z}(1,\dots,1,-1,\dots,-1)\subset\mathbb{Z}^{2n}. The toric data (𝖠,𝕃)(\mathsf{A},\mathbb{L}) has an associated GKZ system of differential equations which is equivalent to the differential equation (4.15). Equivalently, the right hand side of Equation (4.14) is the 𝒜\mathcal{A}-hypergeometric integral in the sense of [24, Thm. 2.7] derived from the data (𝒜,𝕃)(\mathcal{A},\mathbb{L}) and the additional vector

α=α1,,αn1,β11,,βn1t=ρ1,,ρn,ρ1,,ρn1t=i=12nγiai,\vec{\alpha}\ =\ \langle\alpha_{1},\dots,\alpha_{n-1},-\beta_{1}-1,\dots,-\beta_{n}-1\rangle^{t}\\ =\ \langle-\rho_{1},\dots,-\rho_{n},-\rho_{1},\dots,-\rho_{n-1}\rangle^{t}\ =\ \sum_{i=1}^{2n}\gamma_{i}\,\vec{a}_{i},

where we have set 𝜸0=(γ1,,γ2n)=(0,,0,ρ1,,ρn)2n\boldsymbol{\gamma}_{0}=(\gamma_{1},\dots,\gamma_{2n})=(0,\dots,0,-\rho_{1},\dots,-\rho_{n})\subset\mathbb{Z}^{2n}. We always have the freedom to shift 𝜸0\boldsymbol{\gamma}_{0} by elements in 𝕃\mathbb{L}\otimes\mathbb{R} while leaving α\vec{\alpha} and any 𝒜\mathcal{A}-hypergeometric integral unchanged. Thus we have the following:

Proposition 4.4.

The GKZ data (𝒜,𝕃,𝛄0)(\mathcal{A},\mathbb{L},\boldsymbol{\gamma}_{0}) is non-resonant.

Proof.

We observe that αi,βj\alpha_{i},\beta_{j}\not\in\mathbb{Z} for i=1,,n1i=1,\dots,n-1 and j=1,,nj=1,\dots,n and iαi+jβjρnmod1\sum_{i}\alpha_{i}+\sum_{j}\beta_{j}\!\equiv-\rho_{n}\mod{1}\not\in\mathbb{Z}. It was proved in [24, Ex. 2.17] that this is equivalent to the non-resonance of the GKZ system. ∎

4.4.1. Construction of convergent period integrals

In this section, we show how from the toric data of the GKZ system convergent period integrals can be constructed. We are following the standard notation for GKZ systems; see, for example, Beukers [4].

Let us define the B-matrix of the lattice relations 𝕃\mathbb{L} for 𝒜\mathcal{A} as the matrix containing its integral generating set as the rows. Since the rank of 𝕃\mathbb{L} is 1, we simply have 𝖡=(1,,1,1,,1)Mat(1,2n;)Hom(2n,)\mathsf{B}=(1,\dots,1,-1,\dots,-1)\in\mathrm{Mat}(1,2n;\mathbb{Z})\cong\mathrm{Hom}_{\mathbb{Z}}(\mathbb{Z}^{2n},\mathbb{Z}). Of course, the B-matrix then satisfies 𝖠𝖡t=0\mathsf{A}\cdot\mathsf{B}^{t}=0, as this is the defining property of the lattice 𝕃\mathbb{L}. The space 𝕃2n\mathbb{L}\otimes\mathbb{R}\subset\mathbb{R}^{2n} is clearly a line, and is parameterized by the tuple (s,,s,s,,s)2n(s,\dots,s,-s,\dots,-s)\in\mathbb{R}^{2n} with ss\in\mathbb{R}. To be used later in this subsection, the polytope Δ𝒜\Delta_{\mathcal{A}} defined as convex hull of the vectors contained in 𝒜\mathcal{A} is the primary polytope associated with 𝒜\mathcal{A}. Also for later, we may also write 𝖡=bie^i\mathsf{B}=\sum b_{i}\hat{e}_{i} in terms of the standard basis {e^i}i=12n2n\{\hat{e}_{i}\}_{i=1}^{2n}\subset\mathbb{Z}^{2n}.

We can obtain a short exact sequence

0𝕃2n2n100\longrightarrow\mathbb{L}\longrightarrow\mathbb{Z}^{2n}\longrightarrow\mathbb{Z}^{2n-1}\to 0

by mapping each vector =lie^i2n\boldsymbol{\ell}=\sum l_{i}\hat{e}_{i}\in\mathbb{Z}^{2n} to the vector liai2n1\sum l_{i}\,\vec{a}_{i}\in\mathbb{Z}^{2n-1}. As the linear relations between vectors in 𝒜\mathcal{A} are given by the lattice 𝕃\mathbb{L}, this sequence is exact. The corresponding dual short exact sequence (over \mathbb{R}) is given by

02n12n𝜋𝕃0,0\longrightarrow\mathbb{R}^{2n-1}\longrightarrow\mathbb{R}^{2n}\overset{\pi}{\longrightarrow}\mathbb{L}^{\vee}_{\mathbb{R}}\cong\mathbb{R}\longrightarrow 0,

with π(u1,,u2n)=u1++unun+1u2n\pi(u_{1},\dots,u_{2n})=u_{1}+\dots+u_{n}-u_{n+1}-\dots-u_{2n}. Restricting π\pi to the positive orthant in 2n\mathbb{R}^{2n} and calling it π^\hat{\pi}, we observe that for each ss\in\mathbb{R} the set π^1(s)\hat{\pi}^{-1}(s) is a convex polyhedron. For s𝕃s\in\mathbb{L}^{\vee}_{\mathbb{R}}, there are two maximal cones 𝒞+\mathcal{C}_{+} and 𝒞\mathcal{C}_{-} in the secondary fan of 𝒜\mathcal{A} for positive and negative real value ss, respectively. The lists of vanishing components for the vertex vectors in each π^1(s)\hat{\pi}^{-1}(s) are given by

T𝒞+=k=1n{{1,,k^,,n,n+1,2n}=:Ik},T𝒞=k=1n{{1,,n,n+1,k+n^,2n}=:Ik+n}.\begin{split}T_{\mathcal{C}_{+}}&=\bigcup_{k=1}^{n}\Big{\{}\underbrace{\{1,\dots,\widehat{k},\dots,n,n+1,\dots 2n\}}_{=:I_{k}}\Big{\}},\\ T_{\mathcal{C}_{-}}&=\bigcup_{k=1}^{n}\Big{\{}\underbrace{\{1,\dots,n,n+1,\widehat{k+n},\dots\dots 2n\}}_{=:I_{k+n}}\Big{\}}.\end{split}

The symbol k^\widehat{k} indicates that the entry kk has been suppressed. For each member II of T𝒞±T_{\mathcal{C}_{\pm}}, we define 𝜸I=𝜸0μI𝖡\boldsymbol{\gamma}^{I}=\boldsymbol{\gamma}_{0}-\mu^{I}\mathsf{B} such that 𝜸iI=0\boldsymbol{\gamma}^{I}_{i}=0 for iIi\not\in I. We then have

𝜸I={𝜸0forIT𝒞+,μI=0,(ρk,,ρk,ρkρ1,,0,,ρkρn)forI=In+kT𝒞,μIn+k=ρk.\boldsymbol{\gamma}^{I}=\left\{\begin{array}[]{lll}\boldsymbol{\gamma}_{0}&\text{for}\,I\in T_{\mathcal{C}_{+}},&\mu^{I}=0,\\ (-\rho_{k},\dots,-\rho_{k},\rho_{k}-\rho_{1},\dots,0,\dots,\rho_{k}-\rho_{n})&\text{for}\,I=I_{n+k}\in T_{\mathcal{C}_{-}},&\mu^{I_{n+k}}=\rho_{k}.\end{array}\right.

Then for IkT𝒞±I_{k}\in T_{\mathcal{C}_{\pm}} we denote the convergence direction by

(4.18) 𝝂Ik=(ν1,,ν2p)=(δik)i=12p𝕃,\boldsymbol{\nu}^{I_{k}}=(\nu_{1},\dots,\nu_{2p})=(\delta_{i}^{k})_{i=1}^{2p}\in\mathbb{L}\otimes\mathbb{R},

where δik\delta^{k}_{i} is the Kronecker delta, such that π^(𝝂Ik)=±1\hat{\pi}(\boldsymbol{\nu}^{I_{k}})=\pm 1.

Using the B-matrix, one defines the zonotope

𝖹𝖡={14i=12nμibi|μi(1,1)}=(n2,n2)𝕃.\mathsf{Z}_{\mathsf{B}}=\left.\left\{\frac{1}{4}\sum_{i=1}^{2n}\mu_{i}\,b_{i}\right|\mu_{i}\in(-1,1)\right\}=\left(-\frac{n}{2},\frac{n}{2}\right)\subset\mathbb{L}^{\vee}_{\mathbb{R}}\cong\mathbb{R}.

The zonotope contains crucial data about the nature and form of the solutions to the GKZ system above. A crucial result of Beukers [4, Cor. 4.2] can then be phrased as follows:

Proposition 4.5.

[4, Cor. 4.2] Let 𝐮,𝛕\boldsymbol{u},\boldsymbol{\tau} be the vector with 𝐮=(u1,,u2n)\boldsymbol{u}=(u_{1},\dots,u_{2n}), uj=|uj|exp(2πiτj)u_{j}=|u_{j}|\,\exp{(2\pi i\tau_{j})}, and 𝛕=(τ1,,τ2n)\boldsymbol{\tau}=(\tau_{1},\dots,\tau_{2n}). For any 𝐮\boldsymbol{u} with 𝛕\boldsymbol{\tau} such that biτi𝖹\sum b_{i}\tau_{i}\in\mathsf{Z}_{\mathcal{B}} and any 𝛄\boldsymbol{\gamma} equivalent to 𝛄0\boldsymbol{\gamma}_{0} up to elements in 𝕃\mathbb{L}\otimes\mathbb{R} with γn+i<σ<γi\gamma_{n+i}<\sigma<-\gamma_{i} for all i=1,,ni=1,\dots,n, the Mellin-Barnes integral given by

(4.19) 𝖬𝝉(u1,,u2n)=σ+i[i=12nΓ(γibis)uiγi+bis]𝑑s,\begin{split}\mathsf{M}_{\boldsymbol{\tau}}(u_{1},\dots,u_{2n})=&\int_{\sigma+i\,\mathbb{R}}\left[\prod_{i=1}^{2n}\Gamma(-\gamma_{i}-b_{i}s)\;u_{i}^{\gamma_{i}+b_{i}s}\right]\,ds\;,\end{split}

is absolutely convergent and satisfies the GKZ differential system for (𝖠,𝕃)(\mathsf{A},\mathbb{L}).

A toric variety 𝒱𝒜\mathcal{V}_{\mathcal{A}} can be associated with the secondary fan by gluing together certain affine schemes, one scheme for every maximal cone in the secondary fan. Details can be found in [69]. In the situation of the hypergeometric differential equation (4.15), the secondary fan has two maximal cones 𝒞+\mathcal{C}_{+} and 𝒞\mathcal{C}_{-}, and one can easily see that the toric variety 𝒱𝒜\mathcal{V}_{\mathcal{A}} is the projective line 𝒱𝒜=1\mathcal{V}_{\mathcal{A}}=\mathbb{P}^{1} which is the the domain of definition for the variable tt in Equation (4.14). Each member in the list for a maximal cone contains 2n12n-1 integers and define a subdivision of the primary polytope Δ𝒜\Delta_{\mathcal{A}} by polytopes generated by the subdivision, called regular triangulations. In our case, these regular triangulations are unimodular, i.e.,

for allIkT𝒞±:|det(ai)iIk|=|bk|=1.\text{for all}\,I_{k}\in T_{\mathcal{C}_{\pm}}:\quad\Big{|}\det\left(\vec{a}_{i}\right)_{i\in I_{k}}\Big{|}=\Big{|}\,b_{k}\Big{|}=1\;.

Given 𝒜\mathcal{A} and its secondary fan, we define a ring 𝒜\mathcal{R}_{\mathcal{A}} by dividing the free polynomial ring in 2n2n variables by the ideal 𝒜\mathcal{I}_{\mathcal{A}} generated by the linear relations of 𝒜\mathcal{A} and the ideal 𝒞±\mathcal{I}_{\mathcal{C}_{\pm}} generated by the regular triangulations. In our situation, we obtain 𝒜\mathcal{R}_{\mathcal{A}} from the list of generators given by

ϵ=(ϵ1,,ϵ2n)=ϵ(1,,1,1,,1)𝒜\boldsymbol{\epsilon}=(\epsilon_{1},\dots,\epsilon_{2n})=\epsilon\,(1,\dots,1,-1,\dots,-1)\in\mathcal{R}_{\mathcal{A}}

with relation ϵn=0\epsilon^{n}=0, i.e., 𝒜=[ϵ]/(ϵn)\mathcal{R}_{\mathcal{A}}=\mathbb{Z}[\epsilon]/(\epsilon^{n}) is a free \mathbb{Z}-module of rank nn. Thus, we have the following:

Corollary 4.6.

A solution for the hypergeometric differential equation (4.15) is given by restricting to u2==u2n=1u_{2}=\dots=u_{2n}=1 and u1=(1)ntu_{1}=(-1)^{n}t in Equation (4.19).

Remark 4.7.

In the case of the hypergeometric differential equation (4.15), it follows crucially from Beukers [4, Prop. 4.6] that there is a basis of Mellin-Barnes integrals since the zonotope 𝖹\mathsf{Z}_{\mathcal{B}} contains nn distinct points {n12+k}k=0n1\{-\frac{n-1}{2}+k\}_{k=0}^{n-1} whose coordinates differ by integers.

4.4.2. A basis of solutions around zero

Using the toric data, we may now derive a local basis of solutions of the differential equation (4.15) around the point t=0t=0 [69]. For the convergence direction 𝝂I1\boldsymbol{\nu}^{I_{1}} in T𝒞+T_{\mathcal{C}_{+}}, the Γ\Gamma-series is a series solutions of the GKZ system for (𝕃,𝜸0)(\mathbb{L},\boldsymbol{\gamma}_{0}) and given by

(4.20) Φ𝕃,𝜸0(u1,,u2n)=𝕃u1γ1+1u2nγ2n+2nΓ(γ1+1+1)Γ(γ2n+2n+1).\Phi_{\mathbb{L},\boldsymbol{\gamma}_{0}}(u_{1},\dots,u_{2n})=\sum_{\boldsymbol{\ell}\in\mathbb{L}}\frac{u_{1}^{\gamma_{1}+\ell_{1}}\cdots u_{2n}^{\gamma_{2n}+\ell_{2n}}}{\Gamma(\gamma_{1}+\ell_{1}+1)\cdots\Gamma(\gamma_{2n}+\ell_{2n}+1)}\;.

We have the following:

Lemma 4.8.

For the convergence direction 𝛎I1\boldsymbol{\nu}^{I_{1}} in T𝒞+T_{\mathcal{C}_{+}}, the Γ\Gamma-series for (𝕃,𝛄0)(\mathbb{L},\boldsymbol{\gamma}_{0}) equals

(4.21) Φ𝕃,𝜸0(u1,,u2n)=[i=1n1Γ(1ρi)un+iρi]Fn1n(ρ1ρn1 1|t)\Phi_{\mathbb{L},\boldsymbol{\gamma}_{0}}(u_{1},\dots,u_{2n})=\left[\prod_{i=1}^{n}\frac{1}{\Gamma(1-\rho_{i})\,u_{n+i}^{\rho_{i}}}\right]\;{}_{n}F_{n-1}\!\left(\left.{\genfrac{}{}{0.0pt}{}{\rho_{1}\;\dots\;\rho_{n}}{1\;\dots\;1}}\right|t\right)

for t=(1)nu1un/(un+1u2n)>0t=(-1)^{n}u_{1}\cdots u_{n}/(u_{n+1}\cdots u_{2n})>0. Moreover, convergence of Equation (4.21) in the convergence direction 𝛎I1=(ν1,,ν2p)\boldsymbol{\nu}^{I_{1}}=(\nu_{1},\dots,\nu_{2p}) is guaranteed for all u1,,u2nu_{1},\dots,u_{2n} with |ui|=tνi|u_{i}|=t^{\nu_{i}} and 0t<10\leq t<1.

Proof.

We observe that

(4.22) Φ𝕃,𝜸0(u1,,u2n)k0u1kunkun+1ρ1ku2nρnk(k!)nΓ(ρ1k+1)Γ(ρnk+1)=[i=1n1Γ(1ρi)un+iρi]k0(ρ1)k(ρn)k(k!)ntk.\begin{split}&\Phi_{\mathbb{L},\boldsymbol{\gamma}_{0}}(u_{1},\dots,u_{2n})\sum_{k\geq 0}\frac{u_{1}^{k}\cdots u_{n}^{k}\cdot u_{n+1}^{-\rho_{1}-k}\cdots u_{2n}^{-\rho_{n}-k}}{(k!)^{n}\,\Gamma(-\rho_{1}-k+1)\cdots\Gamma(-\rho_{n}-k+1)}\\ =&\left[\prod_{i=1}^{n}\frac{1}{\Gamma(1-\rho_{i})\,u_{n+i}^{\rho_{i}}}\right]\sum_{k\geq 0}\frac{(\rho_{1})_{k}\cdots(\rho_{n})_{k}}{(k!)^{n}}\,t^{k}\;.\end{split}

The summation over 𝕃\mathbb{L} reduces to non-negative integers as the other terms vanish when 1/Γ(k+1)=01/\Gamma(k+1)=0 for k<0k<0. Using the identities

(4.23) (ρ)k=(1)kΓ(1ρ)Γ(1kρ),Γ(z)Γ(1z)=πsin(πz),(\rho)_{k}=(-1)^{k}\frac{\Gamma(1-\rho)}{\Gamma(1-k-\rho)},\quad\Gamma(z)\,\Gamma(1-z)=\frac{\pi}{\sin{(\pi z)}}\;,

we obtain Equation (4.21). Equation (4.20) shows that restricting the variables u2==u2n=1u_{2}=\dots=u_{2n}=1 to a base point, the convergence of the Γ\Gamma-series Φ𝕃,𝜸0((1)nt,1,1)\Phi_{\mathbb{L},\boldsymbol{\gamma}_{0}}((-1)^{n}t,1\dots,1) is guaranteed for |u1|=t|u_{1}|=t with tt sufficiently small. ∎

Remark 4.9.

We obtain the same Γ\Gamma-series for all convergence directions 𝛎Ir\boldsymbol{\nu}^{I_{r}} with 1rn1\leq r\leq n in T𝒞+T_{\mathcal{C}_{+}}. This is due to the fact that in the Riemann symbol (4.16) at t=0t=0 the critical exponent 0 has multiplicity nn.

However, from the maximal cone 𝒞+\mathcal{C}_{+} of the secondary fan of 𝒜\mathcal{A}, we can still construct a local basis of solutions of the GKZ system around t=0t=0 by expanding the twisted power series Φ𝕃,𝜸0+ϵ(u1,,u2n)\Phi_{\mathbb{L},\boldsymbol{\gamma}_{0}+\boldsymbol{\epsilon}}(u_{1},\dots,u_{2n}) over 𝒜\mathcal{R}_{\mathcal{A}}; see [69]. Similarly, a twisted hypergeometric series can be introduced, for example, by defining the following renormalized generating function:

(4.24) f(ϵ,t)=tϵFn1(ϵ)n(ρ1ρn1 1|t)=k0(ρ1+ϵ)k(ρn+ϵ)k(1+ϵ)kntk+ϵ.f(\epsilon,t)=t^{\epsilon}{}_{n}F^{(\epsilon)}_{n-1}\!\left(\left.{\genfrac{}{}{0.0pt}{}{\rho_{1}\;\dots\;\rho_{n}}{1\;\dots\;1}}\right|t\right)=\sum_{k\geq 0}\frac{(\rho_{1}+\epsilon)_{k}\cdots(\rho_{n}+\epsilon)_{k}}{(1+\epsilon)_{k}^{n}}\,t^{k+\epsilon}.

We have the following:

Lemma 4.10.

For |t|<1|t|<1, choosing the principal branch of tϵ=exp(ϵlnt)t^{\epsilon}=\exp{(\epsilon\ln{t})} the twisted power series over 𝒜\mathcal{R}_{\mathcal{A}} is given by

(4.25) Φ𝕃,𝜸0+ϵ(u1,,u2n)=e2πiϵΓ(1+ϵ)n[i=1n1Γ(1ρiϵ)un+iρi]tϵFn1(ϵ)n(ρ1ρn1 1|t).\begin{split}&\Phi_{\mathbb{L},\boldsymbol{\gamma}_{0}+\boldsymbol{\epsilon}}(u_{1},\dots,u_{2n})=\frac{e^{2\pi i\epsilon}}{\Gamma(1+\epsilon)^{n}}\left[\prod_{i=1}^{n}\frac{1}{\Gamma(1-\rho_{i}-\epsilon)\,u_{n+i}^{\rho_{i}}}\right]\;\;t^{\epsilon}\,{}_{n}F^{(\epsilon)}_{n-1}\!\left(\left.{\genfrac{}{}{0.0pt}{}{\rho_{1}\;\dots\;\rho_{n}}{1\;\dots\;1}}\right|t\right).\end{split}
Proof.

The proof uses 1/(1+ϵ)kn=O(ϵn)=01/(1+\epsilon)_{k}^{n}=O(\epsilon^{n})=0 for k<0k<0, where (a)k(a)_{k} is the Pochammer symbol, because for kk\in\mathbb{Z} we have

1(1+ϵ)k=Γ(1+ϵ)Γ(k+1+ϵ)={ϵ(ϵ1)(ϵ+k+1)if k<0,1if k=0,1(1+ϵ)(2+ϵ)(m+ϵ)if k>0.\frac{1}{(1+\epsilon)_{k}}=\frac{\Gamma(1+\epsilon)}{\Gamma(k+1+\epsilon)}=\left\{\begin{array}[]{lcl}\epsilon(\epsilon-1)\cdots(\epsilon+k+1)&&\text{if $k<0$,}\\ 1&&\text{if $k=0$,}\\[-5.0pt] \dfrac{1}{(1+\epsilon)(2+\epsilon)\cdots(m+\epsilon)}&&\text{if $k>0$.}\end{array}\right.

For r=0,,n1r=0,\dots,n-1, we also introduce the functions

yr(t)=1r!rϵr|ϵ=0Fn1(ϵ)n(ρ1ρn1 1|t),y0(t)=f(0,t)=Fn1n(ρ1ρn1 1|t).y_{r}(t)=\frac{1}{r!}\left.\frac{\partial^{r}}{\partial\epsilon^{r}}\right|_{\epsilon=0}\!{}_{n}F^{(\epsilon)}_{n-1}\!\left(\left.{\genfrac{}{}{0.0pt}{}{\rho_{1}\;\dots\;\rho_{n}}{1\;\dots\;1}}\right|t\right),\;y_{0}(t)=f(0,t)={}_{n}F_{n-1}\!\left(\left.{\genfrac{}{}{0.0pt}{}{\rho_{1}\;\dots\;\rho_{n}}{1\;\dots\;1}}\right|t\right).

We have the following:

Lemma 4.11.

For |t|<1|t|<1, the following identity holds

(4.26) f(ϵ,t)=m=0n1(2πiϵ)mfm(t)=m=0n1(2πiϵ)mr=0m1r!(lnt2πi)rymr(t)(2πi)mr,f(\epsilon,t)=\sum_{m=0}^{n-1}\Big{(}2\pi i\epsilon\Big{)}^{m}\,f_{m}(t)=\sum_{m=0}^{n-1}\Big{(}2\pi i\epsilon\Big{)}^{m}\;\sum_{r=0}^{m}\frac{1}{r!}\,\left(\frac{\ln{t}}{2\pi i}\right)^{r}\,\frac{y_{m-r}(t)}{(2\pi i)^{m-r}},

where fm(t)=1(2πi)mm!mϵm|ϵ=0f(ϵ,t)f_{m}(t)=\frac{1}{(2\pi i)^{m}m!}\frac{\partial^{m}}{\partial\epsilon^{m}}|_{\epsilon=0}f(\epsilon,t) for m=0,,n1m=0,\dots,n-1.

As proved in [69], the functions {fr}r=0n1\{f_{r}\}_{r=0}^{n-1} form a local basis of solutions around t=0t=0, and the functions yr(t)y_{r}(t) with r=0,n1r=0,\dots n-1 are holomorphic in a neighborhood of t=0t=0. The local monodromy group is generated by the cycle (u1,,u2n)=(R1exp(iφ),R2,,R2n)(u_{1},\dots,u_{2n})=(R_{1}\exp{(i\varphi)},R_{2},\dots,R_{2n}) based at the point (R1,,R2n)(R_{1},\dots,R_{2n}) for φ[0,2π]\varphi\in[0,2\pi]. Equivalently, we consider the local monodromy of the hypergeometric differential equation generated by t=t0exp(iφ)t=t_{0}\exp{(i\varphi)} for 0<t0<10<t_{0}<1 and φ[0,2π]\varphi\in[0,2\pi] (by setting |u2|==|u2n|=1|u_{2}|=\dots=|u_{2n}|=1 and |u1|=t|u_{1}|=t). The monodromy of the functions {fr}r=0n1\{f_{r}\}_{r=0}^{n-1} can be read off Equation (4.26) immediately. We have the following:

Proposition 4.12.

The local monodromy of the basis 𝐟t=fn1,,f0t\boldsymbol{f}^{t}=\langle f_{n-1},\dots,f_{0}\rangle^{t} of solutions to the differential equation (4.15) at t=0t=0 is given by

(4.27) 𝗆0=(11121(n2)!0111(n3)!1001).\mathsf{m}_{0}=\left(\begin{array}[]{ccccc}1&1&\frac{1}{2}&\dots&\frac{1}{(n-2)!}\\ 0&1&1&\dots&\frac{1}{(n-3)!}\\ \vdots&\ddots&\ddots&&\vdots\\ \vdots&&\ddots&\ddots&1\\ 0&\dots&\dots&0&1\end{array}\right)\;.
Proof.

Lemma 4.11 proves that

fm(t)=r=0m1r!(lnt2πi)rymr(t)(2πi)mr.f_{m}(t)=\sum_{r=0}^{m}\frac{1}{r!}\,\left(\frac{\ln{t}}{2\pi i}\right)^{r}\,\frac{y_{m-r}(t)}{(2\pi i)^{m-r}}\,.

The functions yk(t)y_{k}(t) are invariant for t=t0exp(iφ)t=t_{0}\exp{(i\varphi)} for 0<t0<10<t_{0}<1 and φ2π\varphi\to 2\pi. The result then follows. ∎

Corollary 4.13.

The monodromy matrix 𝗆0\mathsf{m}_{0} is maximally unipotent.

4.4.3. A basis of solutions around infinity

We assume 0<ρ1<<ρn<10<\rho_{1}<\dots<\rho_{n}<1. Using the toric data we can derive a local basis of solutions of the differential equation (4.15) around the point t=t=\infty. For the convergence direction 𝝂In+r\boldsymbol{\nu}^{I_{n+r}} in T𝒞T_{\mathcal{C}_{-}}, the Γ\Gamma-series is a series solutions of the GKZ system for (𝕃,𝜸In+r)(\mathbb{L},\boldsymbol{\gamma}^{I_{n+r}}) and given by

(4.28) Φ𝕃,𝜸In+r(u1,,u2n)=𝕃u1γ1μIr+n+1u2nγ2n+μIr+n+2nΓ(γ1μIr+n+1+1)Γ(γ2n+μIr+n+2n+1).\begin{split}&\Phi_{\mathbb{L},\boldsymbol{\gamma}^{I_{n+r}}}(u_{1},\dots,u_{2n})=\sum_{\boldsymbol{\ell}\in\mathbb{L}}\frac{u_{1}^{\gamma_{1}-\mu^{I_{r+n}}+\ell_{1}}\cdots u_{2n}^{\gamma_{2n}+\mu^{I_{r+n}}+\ell_{2n}}}{\Gamma(\gamma_{1}-\mu^{I_{r+n}}+\ell_{1}+1)\cdots\Gamma(\gamma_{2n}+\mu^{I_{r+n}}+\ell_{2n}+1)}\;.\end{split}

We have the following:

Lemma 4.14.

For the convergence direction 𝛎In+r\boldsymbol{\nu}^{I_{n+r}} in T𝒞T_{\mathcal{C}_{-}} Equation (4.28) is a series solution for (𝕃,𝛄In+r)(\mathbb{L},\boldsymbol{\gamma}^{I_{n+r}}). The following identity holds

(4.29) Φ𝕃,𝜸In+r(u1,,u2n)=eπinρrΓ(1ρr)n[i=1n1Γ(1+ρrρi)un+iρi]×tρrFn1n(ρrρr1+ρrρ11^ 1+ρrρn|1t)\begin{split}&\Phi_{\mathbb{L},\boldsymbol{\gamma}^{I_{n+r}}}(u_{1},\dots,u_{2n})=\frac{e^{\pi in\rho_{r}}}{\Gamma(1-\rho_{r})^{n}}\left[\prod_{i=1}^{n}\frac{1}{\Gamma(1+\rho_{r}-\rho_{i})\,u_{n+i}^{\rho_{i}}}\right]\\ &\qquad\times\,t^{-\rho_{r}}\,{}_{n}F_{n-1}\!\left(\left.{\genfrac{}{}{0.0pt}{}{\rho_{r}\qquad\dots\qquad\dots\qquad\rho_{r}}{1+\rho_{r}-\rho_{1}\;\dots\widehat{1}\dots\;1+\rho_{r}-\rho_{n}}}\right|\frac{1}{t}\right)\end{split}

for t=(1)nu1un/(un+1u2n)>0t=(-1)^{n}u_{1}\cdots u_{n}/(u_{n+1}\cdots u_{2n})>0. The symbol 1^\widehat{1} indicates that the entry 1+ρrρi1+\rho_{r}-\rho_{i} for i=ri=r has been suppressed. In particular, restricting variables u1==un+r^==u2n=1u_{1}=\dots=\widehat{u_{n+r}}=\dots=u_{2n}=1 the convergence of the Γ\Gamma-series Φ𝕃,𝛄In+r(1,,(1)n/t,,1)\Phi_{\mathbb{L},\boldsymbol{\gamma}^{I_{n+r}}}(1,\dots,(-1)^{n}/t,\dots,1) is guaranteed for t>1t>1.

Proof.

A direct computation shows that the Γ\Gamma-series satisfies

Φ𝕃,𝜸In+r(u1,,u2n)=eπinρrΓ(1ρr)n[i=1n1Γ(1+ρrρi)un+iρi](un+1u2n(1)nu1un)ρr×k0(ρr)kn(1+ρrρ1)k(1+ρrρn+1)k(un+1u2n(1)nu1un)k.\begin{split}\Phi_{\mathbb{L},\boldsymbol{\gamma}^{I_{n+r}}}(u_{1},\dots,u_{2n})=&\,\frac{e^{\pi in\rho_{r}}}{\Gamma(1-\rho_{r})^{n}}\left[\prod_{i=1}^{n}\frac{1}{\Gamma(1+\rho_{r}-\rho_{i})\,u_{n+i}^{\rho_{i}}}\right]\left(\frac{u_{n+1}\cdots u_{2n}}{(-1)^{n}u_{1}\cdots u_{n}}\right)^{\rho_{r}}\\ \times&\sum_{k\geq 0}\frac{(\rho_{r})^{n}_{k}}{(1+\rho_{r}-\rho_{1})_{k}\cdots(1+\rho_{r}-\rho_{n}+1)_{k}}\left(\frac{u_{n+1}\cdots u_{2n}}{(-1)^{n}u_{1}\cdots u_{n}}\right)^{k}.\end{split}

The result follows. ∎

Based on the assumption that 0<ρ1<<ρn<10<\rho_{1}<\dots<\rho_{n}<1, we have the following:

Lemma 4.15.

There are nn different Γ\Gamma-series for the convergence directions 𝛎In+r\boldsymbol{\nu}^{I_{n+r}} with 1rn1\leq r\leq n in T𝒞T_{\mathcal{C}_{-}}.

The local monodromy group is generated by the cycle based at (R1,,R2n)(R_{1},\dots,R_{2n}) given by (u1,,un+r,,u2n)=(R1,,Rn+rexp(iφ),,R2n)(u_{1},\dots,u_{n+r},\dots,u_{2n})=(R_{1},\dots,R_{n+r}\exp{(-i\varphi)},\dots,R_{2n}) for φ[0,2π]\varphi\in[0,2\pi] Equivalently, we consider the local monodromy generated by t=t0exp(iφ)t=t_{0}\exp{(i\varphi)} for t01t_{0}\gg 1 and φ[0,2π]\varphi\in[0,2\pi] (by setting |u1|==|u2n|=1|u_{1}|=\dots=|u_{2n}|=1 and |un+r|=1/t|u_{n+r}|=1/t). We have the following:

Proposition 4.16.

The local monodromy of the basis 𝐅t=Fn,,F1t\boldsymbol{F}^{t}=\langle F_{n},\dots,F_{1}\rangle^{t} of solutions to the differential equation (4.15) at t=t=\infty is given by

(4.30) 𝖬=(e2πiρne2πiρ1).\mathsf{M}_{\infty}=\left(\begin{array}[]{ccccc}e^{-2\pi i\rho_{n}}&&\\ &\ddots&\\ &&e^{-2\pi i\rho_{1}}\end{array}\right)\;.
Proof.

From the Riemann symbol (4.16), we observe that the functions

(4.31) Fr(t)=ArtρrFn1n(ρrρr1+ρrρ11^ 1+ρrρn|1t)F_{r}(t)=A_{r}\;t^{-\rho_{r}}\,{}_{n}F_{n-1}\!\left(\left.{\genfrac{}{}{0.0pt}{}{\rho_{r}\qquad\dots\qquad\dots\qquad\rho_{r}}{1+\rho_{r}-\rho_{1}\;\dots\widehat{1}\dots\;1+\rho_{r}-\rho_{n}}}\right|\frac{1}{t}\right)

for r=1,,nr=1,\dots,n and any non-zero constants ArA_{r}, form a Frobenius basis of solutions to the differential equation (4.15) at t=t=\infty. The claim follows. ∎

4.4.4. The transition matrix

The solution (4.24) has an integral representation of Mellin-Barnes type [4] given by

(4.32) f(ϵ,t)=tϵ2πiΓ(1+ϵ)nΓ(ρ1+ϵ)Γ(ρn+ϵ)σ+i𝑑sΓ(s+ρ1+ϵ)Γ(s+ρn+ϵ)Γ(s+1+ϵ)nπ(t)ssin(πs),f(\epsilon,t)=\frac{t^{\epsilon}}{2\pi i}\,\frac{\Gamma(1+\epsilon)^{n}}{\Gamma(\rho_{1}+\epsilon)\cdots\Gamma(\rho_{n}+\epsilon)}\int_{\sigma+i\mathbb{R}}\!\!\!\!\!\!\!ds\,\frac{\Gamma(s+\rho_{1}+\epsilon)\cdots\Gamma(s+\rho_{n}+\epsilon)}{\Gamma(s+1+\epsilon)^{n}}\cdot\frac{\pi\,(-t)^{s}}{\sin{(\pi s)}}\;,

where σ(ρ1,0)\sigma\in(-\rho_{1},0). For |t|<1|t|<1 the contour integral can be closed to the right. We have the following:

Lemma 4.17.

For |t|<1|t|<1, Equation (4.32) coincides with Equation (4.24).

Proof.

For |t|<1|t|<1 the contour integral can be closed to the right, and the Γ\Gamma-series in Equation (4.24) is recovered as a sum over the enclosed residua at r0r\in\mathbb{N}_{0} where we have used

forallr0:Ress=r(π(t)ssin(πs))=tr.\mathrm{for\;all}\;r\in\mathbb{N}_{0}:\;\operatorname{Res}_{s=r}\left(\frac{\pi\,(-t)^{s}}{\sin{(\pi s)}}\right)=t^{r}.

For |t|>1|t|>1 the contour integral must be closed to the left. The relation to the local basis of solutions at t=t=\infty can be explicitly computed:

Proposition 4.18.

For |t|>1|t|>1, we obtain for f(ϵ,t)f(\epsilon,t) in Equation (4.32)

(4.33) f(ϵ,t)=r=1nBr(ϵ)Fr(t)f(\epsilon,t)=\sum_{r=1}^{n}B_{r}(\epsilon)\,F_{r}(t)

where Fr(t)F_{r}(t) is given by

(4.34) Fr(t)=ArtρrFn1n(ρrρr1+ρrρ11^ 1+ρrρn|1t)F_{r}(t)=A_{r}\;t^{-\rho_{r}}\,{}_{n}F_{n-1}\!\left(\left.{\genfrac{}{}{0.0pt}{}{\rho_{r}\qquad\dots\qquad\dots\qquad\rho_{r}}{1+\rho_{r}-\rho_{1}\;\dots\widehat{1}\dots\;1+\rho_{r}-\rho_{n}}}\right|\frac{1}{t}\right)

and

(4.35) Ar=eπiρri=1irnΓ(ρiρr)Γ(ρi)Γ(1ρr),Br(ϵ)=eπiϵ[i=1nΓ(ρi)Γ(1+ϵ)Γ(ρi+ϵ)]sin(πρr)sin(πρr+πϵ),A_{r}=-e^{-\pi i\rho_{r}}\prod_{\begin{subarray}{c}i=1\\ i\not=r\end{subarray}}^{n}\frac{\Gamma(\rho_{i}-\rho_{r})}{\Gamma(\rho_{i})\,\Gamma(1-\rho_{r})},\;B_{r}(\epsilon)=e^{-\pi i\epsilon}\left[\prod_{i=1}^{n}\frac{\Gamma(\rho_{i})\,\Gamma(1+\epsilon)}{\Gamma(\rho_{i}+\epsilon)}\right]\frac{\sin{(\pi\rho_{r})}}{\sin{(\pi\rho_{r}+\pi\epsilon)}},

such that Br(0)=1B_{r}(0)=1 for r=1,,nr=1,\dots,n.

Proof.

For |t|>1|t|>1 the contour integral in Equation (4.32) must be closed to the left. Using 1/(1+ϵ)kn=O(ϵn)=01/(1+\epsilon)_{k}^{n}=O(\epsilon^{n})=0 for k<0k<0, we observe that the poles are located at s=ϵρiks=-\epsilon-\rho_{i}-k for i=1,,ni=1,\dots,n and k0k\in\mathbb{N}_{0}. Using

r0:Ress=r(Γ(s)(t)s)=trr!.\forall r\in\mathbb{N}_{0}:\;\operatorname{Res}_{s=-r}\Big{(}\Gamma(s)\,(-t)^{s}\Big{)}=\frac{t^{-r}}{r!}.

and Equations (4.23) the result follows. ∎

Equation (4.33) allows to compute the transition matrix between the Frobenius basis fn1,,f0t\langle f_{n-1},\dots,f_{0}\rangle^{t} of solutions for the differential equation (4.15) at t=0t=0 with local monodromy given by the matrix (4.27) and the Frobenius basis Fn,,F1t\langle F_{n},\dots,F_{1}\rangle^{t} of solutions at t=t=\infty with local monodromy given by the matrix (4.30). We obtain:

Corollary 4.19.

The transition matrix 𝖯\mathsf{P} between the analytic continuations of the bases 𝐟t=fn1,,f0t\boldsymbol{f}^{t}=\langle f_{n-1},\dots,f_{0}\rangle^{t} at t=0t=0 and 𝐅t=Fn,,F1t\boldsymbol{F}^{t}=\langle F_{n},\dots,F_{1}\rangle^{t} at t=t=\infty is given by

(4.36) (fn1f1f0)=(Bn(n1)(0)(2πi)n1(n1)!B1(n1)(0)(2πi)n1(n1)!Bn(0)2πiB1(0)2πi11)(FnF2F1)\left(\begin{array}[]{c}f_{n-1}\\ \vdots\\ f_{1}\\ f_{0}\end{array}\right)=\left(\begin{array}[]{ccc}\frac{B_{n}^{(n-1)}(0)}{(2\pi i)^{n-1}(n-1)!}&\dots&\frac{B_{1}^{(n-1)}(0)}{(2\pi i)^{n-1}(n-1)!}\\ \vdots&\ddots&\vdots\\ \frac{B_{n}^{{}^{\prime}}(0)}{2\pi i}&\dots&\frac{B_{1}^{{}^{\prime}}(0)}{2\pi i}\\[1.99997pt] 1&\dots&1\end{array}\right)\cdot\left(\begin{array}[]{c}F_{n}\\ \vdots\\ F_{2}\\ F_{1}\end{array}\right)

with Br(ϵ)B_{r}(\epsilon) given in Equation (4.35).

Proof.

The transition matrix 𝖯\mathsf{P} between the analytically continued Frobenius basis of solutions 𝒇t=fn1,,f0t\boldsymbol{f}^{t}=\langle f_{n-1},\dots,f_{0}\rangle^{t} at t=0t=0 and the analytic continuation of the Frobenius basis 𝑭t=Fn,,F1t\boldsymbol{F}^{t}=\langle F_{n},\dots,F_{1}\rangle^{t} at t=t=\infty is obtained by first comparing the expression of f(ϵ,t)f(\epsilon,t) from Equation (4.24) as a linear combination of the solutions 𝑭\boldsymbol{F} at t=t=\infty from Equation (4.33), and subsequently applying Lemma 4.11 to find the explicit linear relations between 𝒇\boldsymbol{f} and 𝑭\boldsymbol{F}. By differentiation of the functions Br(ϵ)B_{r}(\epsilon) in Equation (4.35) and evaluating at ϵ=0\epsilon=0, we recover the matrix (4.36). ∎

We can now compute the monodromy of the analytic continuation of 𝒇\boldsymbol{f} around any singular point:

Corollary 4.20.

The monodromy of the analytic continuation of 𝐟\boldsymbol{f} around t=0t=0, t=t=\infty, and t=1t=1 is given by 𝗆0\mathsf{m}_{0} in (4.27), 𝗆=𝖯𝖬𝖯1\mathsf{m}_{\infty}=\mathsf{P}\cdot\mathsf{M}_{\infty}\cdot\mathsf{P}^{-1} for 𝖬\mathsf{M}_{\infty} in (4.30), and 𝗆1=𝗆𝗆01\mathsf{m}_{1}=\mathsf{m}_{\infty}\cdot\mathsf{m}_{0}^{-1}, respectively.

4.4.5. Monodromy after rescaling

For C>0C>0 the rescaled hypergeometric differential equation satisfied by F~(t)=Fn1n(Ct)\tilde{F}(t)={}_{n}F_{n-1}(Ct) is given by

(4.37) [θnCt(θ+ρ1)(θ+ρn)]F~(t)=0.\Big{[}\theta^{n}-C\,t\,(\theta+\rho_{1})\cdots(\theta+\rho_{n})\Big{]}\,\tilde{F}(t)=0\;.

For |t|<1/C|t|<1/C we introduce f~(ϵ,t)=Cϵf(ϵ,Ct)\tilde{f}(\epsilon,t)=C^{-\epsilon}f(\epsilon,Ct) such that

(4.38) f~(ϵ,t)=m=0n1(2πiϵ)mf~m(t)withf~m(t)=1(2πi)mm!mϵm|ϵ=0f(ϵ,Ct)\tilde{f}(\epsilon,t)=\sum_{m=0}^{n-1}\Big{(}2\pi i\epsilon\Big{)}^{m}\,\tilde{f}_{m}(t)\quad\text{with}\quad\tilde{f}_{m}(t)=\frac{1}{(2\pi i)^{m}m!}\left.\frac{\partial^{m}}{\partial\epsilon^{m}}\right|_{\epsilon=0}\!\!f(\epsilon,Ct)

for j=0,,n1j=0,\dots,n-1. The local monodromy around t=0t=0 with respect to the Frobenius basis f~n1,,f~0t\langle\tilde{f}_{n-1},\dots,\tilde{f}_{0}\rangle^{t} is still given by the matrix 𝗆0\mathsf{m}_{0} in (4.27). Similarly, for |t|>1/C|t|>1/C we introduce F~k(t)=Fk(Ct)\tilde{F}_{k}(t)=F_{k}(Ct) for k=1,,nk=1,\dots,n. The local monodromy (around t=t=\infty) with respect to the Frobenius basis F~n,,F~1t\langle\tilde{F}_{n},\dots,\tilde{F}_{1}\rangle^{t} is given by the matrix 𝖬\mathsf{M}_{\infty} in (4.30). We obtain:

Proposition 4.21.

The transition matrix 𝖯~\tilde{\mathsf{P}} between the analytic continuation of 𝐟~\boldsymbol{\tilde{f}} and 𝐅~\boldsymbol{\tilde{F}} such that 𝐟~=𝖯~𝐅~\boldsymbol{\tilde{f}}=\tilde{\mathsf{P}}\cdot\boldsymbol{\tilde{F}} is given by

(4.39) 𝖯~=(𝖯~nj,n+1k)j=0,k=1n1,nwith𝖯~nj,n+1k=1(2πi)jj!jϵj|ϵ=0[CϵBk(ϵ)].\tilde{\mathsf{P}}=\Big{(}\tilde{\mathsf{P}}_{n-j,n+1-k}\Big{)}_{j=0,k=1}^{n-1,n}\quad\text{with}\quad\tilde{\mathsf{P}}_{n-j,n+1-k}=\frac{1}{(2\pi i)^{j}j!}\left.\frac{\partial^{j}}{\partial\epsilon^{j}}\right|_{\epsilon=0}\!\!\Big{[}C^{-\epsilon}B_{k}(\epsilon)\Big{]}\;.

The monodromy of the analytic continuation of 𝐟~\boldsymbol{\tilde{f}} around t=t=\infty and t=1/Ct=1/C is given by 𝗆=𝖯~𝖬𝖯~1\mathsf{m}_{\infty}=\tilde{\mathsf{P}}\cdot\mathsf{M}_{\infty}\cdot\tilde{\mathsf{P}}^{-1} and 𝗆1/C=𝗆𝗆01\mathsf{m}_{1/C}=\mathsf{m}_{\infty}\cdot\mathsf{m}_{0}^{-1}, respectively.

Proof.

One emulates the proof of Corollaries 4.19 and 4.20 directly with new analytic continuations 𝒇~\boldsymbol{\tilde{f}} and 𝑭~\boldsymbol{\tilde{F}} around t=0t=0 and t=t=\infty, respectively. In this case, one finds that the functions Br(ϵ)B_{r}(\epsilon) appearing in Equation (4.33) acquire a factor of CϵC^{-\epsilon}. The result then follows suit as claimed. ∎

In summary, we obtained the monodromy matrices 𝗆0\mathsf{m}_{0} in (4.27), 𝗆=𝖯~𝖬𝖯~1\mathsf{m}_{\infty}=\tilde{\mathsf{P}}\cdot\mathsf{M}_{\infty}\cdot\tilde{\mathsf{P}}^{-1} for 𝖬\mathsf{M}_{\infty} in (4.30) and 𝖯~\tilde{\mathsf{P}} in Equation (4.39), and 𝗆1/C=𝗆𝗆01\mathsf{m}_{1/C}=\mathsf{m}_{\infty}\cdot\mathsf{m}_{0}^{-1} for the hypergeometric differential equation (4.37)(\ref{hpg_ode_b}). Thus, we have the following main result:

Theorem 4.22.

For the family of hypersurfaces Yt(n1)Y^{(n-1)}_{t} in Equation (4.2) with n2n\geq 2 the mixed-twist construction defines a non-resonant GKZ system. Then a basis of solutions exists given as absolutely convergent Mellin-Barnes integrals whose monodromy around t=0,1/C,t=0,1/C,\infty is, up to conjugation, 𝗆0,𝗆1/C,𝗆\mathsf{m}_{0},\mathsf{m}_{1/C},\mathsf{m}_{\infty}, respectively, for ρk=k/(n+1)\rho_{k}=k/(n+1) with k=1,,nk=1,\dots,n and C=(n+1)n+1C=(n+1)^{n+1}.

Proof.

The theorem combines the statements of Propositions 4.4, 4.5, 4.12, 4.16, 4.21 that were proven above. ∎

We have the following:

Corollary 4.23.

Set κ4=200ζ(3)(2πi)3\kappa_{4}=-200\frac{\zeta(3)}{(2\pi i)^{3}}, and κ5=420ζ(3)(2πi)3\kappa_{5}=420\frac{\zeta(3)}{(2\pi i)^{3}}. The monodromy matrices of Theorem 4.22 for 2n52\leq n\leq 5 are given by Table 1.

Proof.

We obtain from the multiplication formula for the Γ\Gamma-function, i.e.,

k=0m1Γ(z+km)=(2π)12(m1)m12mzΓ(mz),\prod_{k=0}^{m-1}\Gamma\left(z+\frac{k}{m}\right)=\left(2\pi\right)^{\frac{1}{2}(m-1)}m^{\frac{1}{2}-mz}\,\Gamma(mz),

the identity

CϵBk(ϵ)=Γ(1+ϵ)n+1Γ(1+(n+1)ϵ)sin(πρk)sin(πρk+πϵ)eπiϵ.C^{-\epsilon}B_{k}(\epsilon)=\frac{\Gamma(1+\epsilon)^{n+1}}{\Gamma\big{(}1+(n+1)\epsilon\big{)}}\frac{\sin{(\pi\rho_{k})}}{\sin{(\pi\rho_{k}+\pi\epsilon)}}e^{-\pi i\epsilon}\;.

We then compute the monodromy of the analytic continuation of 𝒇~\boldsymbol{\tilde{f}} around t=0,1/C,t=0,1/C,\infty where we have set κ4=200ζ(3)(2πi)3\kappa_{4}=-200\frac{\zeta(3)}{(2\pi i)^{3}} and κ5=420ζ(3)(2πi)3\kappa_{5}=420\frac{\zeta(3)}{(2\pi i)^{3}}. We obtain the results listed in Table 1. ∎

The case n=4n=4, reproduces up to conjugation the monodromy matrices for the quintic threefold case by Candelas et al. [7] and [8].In particular, our results are consistent with the original work of Levelt [45] up to conjugacy, for any n>2n>2.

nn Yt(n1)Y^{(n-1)}_{t} 𝗆0\mathsf{m}_{0} 𝗆1/C\mathsf{m}_{1/C} 𝗆\mathsf{m}_{\infty}
22 EC (1101)\left(\begin{array}[]{cc}1&1\\ 0&1\end{array}\right) (1031)\left(\begin{array}[]{rc}1&0\\ -3&1\end{array}\right) (1132)\left(\begin{array}[]{cc}1&1\\ -3&-2\end{array}\right)
33 K3 (1112011001)\left(\begin{array}[]{ccc}1&1&\frac{1}{2}\\ 0&1&1\\ 0&0&1\end{array}\right) (0014010400)\left(\begin{array}[]{rcr}0&0&-\frac{1}{4}\\ 0&1&0\\ -4&0&0\end{array}\right) (0014011442)\left(\begin{array}[]{rrr}0&0&-\frac{1}{4}\\ 0&1&1\\ -4&-4&-2\end{array}\right)
44 CY3 (1112160111200110001)\left(\begin{array}[]{cccc}1&1&\frac{1}{2}&\frac{1}{6}\\ 0&1&1&\frac{1}{2}\\ 0&0&1&1\\ 0&0&0&1\end{array}\right) (1+κ405κ412κ425251211251445κ41200105025121κ4)\left(\begin{array}[]{rrrr}1+\kappa_{4}&0&\frac{5\kappa_{4}}{12}&\frac{\kappa_{4}^{2}}{5}\\ -\frac{25}{12}&1&-\frac{125}{144}&-\frac{5\kappa_{4}}{12}\\ 0&0&1&0\\ -5&0&-\frac{25}{12}&1-\kappa_{4}\end{array}\right) (1+κ41+κ412+11κ41216+7κ412+κ425251213121311441031445κ41200115555122312κ4)\left(\begin{array}[]{rrrr}1+\kappa_{4}&1+\kappa_{4}&\frac{1}{2}+\frac{11\kappa_{4}}{12}&\frac{1}{6}+\frac{7\kappa_{4}}{12}+\frac{\kappa_{4}^{2}}{5}\\ -\frac{25}{12}&-\frac{13}{12}&-\frac{131}{144}&-\frac{103}{144}-\frac{5\kappa_{4}}{12}\\ 0&0&1&1\\ -5&-5&-\frac{55}{12}&-\frac{23}{12}-\kappa_{4}\end{array}\right)
55 CY4 (11121612401112160011120001100001)\left(\begin{array}[]{ccccc}1&1&\frac{1}{2}&\frac{1}{6}&\frac{1}{24}\\ 0&1&1&\frac{1}{2}&\frac{1}{6}\\ 0&0&1&1&\frac{1}{2}\\ 0&0&0&1&1\\ 0&0&0&0&1\end{array}\right) (756405551211κ538412124576κ515κ58κ52611κ5384154043325κ58555120001060154κ57564)\left(\begin{array}[]{rrrrr}\frac{75}{64}&0&\frac{55}{512}&-\frac{11\kappa_{5}}{384}&-\frac{121}{24576}\\ -\kappa_{5}&1&-\frac{5\kappa_{5}}{8}&\frac{\kappa_{5}^{2}}{6}&\frac{11\kappa_{5}}{384}\\ -\frac{15}{4}&0&-\frac{43}{32}&\frac{5\kappa_{5}}{8}&\frac{55}{512}\\ 0&0&0&1&0\\ -6&0&-\frac{15}{4}&\kappa_{5}&\frac{75}{64}\end{array}\right) (7564756435551211κ5384+15551211κ5384+239924576κ5κ5+19κ58+1(4κ53)(κ54)24κ526125κ5384+16154154103325κ5863325κ583695120001166274κ5194κ56164)\left(\begin{array}[]{rrrrr}\frac{75}{64}&\frac{75}{64}&\frac{355}{512}&-\frac{11\kappa_{5}}{384}+\frac{155}{512}&-\frac{11\kappa_{5}}{384}+\frac{2399}{24576}\\ -\kappa_{5}&-\kappa_{5}+1&-\frac{9\kappa_{5}}{8}+1&\frac{(4\kappa_{5}-3)(\kappa_{5}-4)}{24}&\frac{\kappa_{5}^{2}}{6}-\frac{125\kappa_{5}}{384}+\frac{1}{6}\\ -\frac{15}{4}&-\frac{15}{4}&-\frac{103}{32}&\frac{5\kappa_{5}}{8}-\frac{63}{32}&\frac{5\kappa_{5}}{8}-\frac{369}{512}\\ 0&0&0&1&1\\ -6&-6&-\frac{27}{4}&\kappa_{5}-\frac{19}{4}&\kappa_{5}-\frac{61}{64}\end{array}\right)
Table 1. Monodromy matrices for the mirror families with 2n52\leq n\leq 5

References

  • [1] V. V. Batyrev and L. A. Borisov. Dual cones and mirror symmetry for generalized Calabi-Yau manifolds. In Mirror symmetry, II, volume 1 of AMS/IP Stud. Adv. Math., pages 71–86. Amer. Math. Soc., Providence, RI, 1997.
  • [2] V. V. Batyrev and D. van Straten. Generalized hypergeometric functions and rational curves on Calabi-Yau complete intersections in toric varieties. Comm. Math. Phys., 168(3):493–533, 1995.
  • [3] L. Beshaj, R. Hidalgo, S. Kruk, A. Malmendier, S. Quispe, and T. Shaska. Rational points in the moduli space of genus two. In Higher genus curves in mathematical physics and arithmetic geometry, volume 703 of Contemp. Math., pages 83–115. Amer. Math. Soc., [Providence], RI, 2018.
  • [4] F. Beukers. Monodromy of AA-hypergeometric functions. J. Reine Angew. Math., 718:183–206, 2016.
  • [5] C. Birkenhake and H. Wilhelm. Humbert surfaces and the Kummer plane. Trans. Amer. Math. Soc., 355(5):1819–1841, 2003.
  • [6] N. Braeger, A. Malmendier, and Y. Sung. Kummer sandwiches and Greene-Plesser construction. J. Geom. Phys., 154:103718, 18, 2020.
  • [7] P. Candelas, X. C. de la Ossa, P. S. Green, and L. Parkes. An exactly soluble superconformal theory from a mirror pair of Calabi-Yau manifolds. In Particles, strings and cosmology (Boston, MA, 1991), pages 667–680. World Sci. Publ., River Edge, NJ, 1992.
  • [8] Y.-H. Chen, Y. Yang, and N. Yui. Monodromy of Picard-Fuchs differential equations for Calabi-Yau threefolds. J. Reine Angew. Math., 616:167–203, 2008. With an appendix by Cord Erdenberger.
  • [9] A. Clingher, C. F. Doran, and A. Malmendier. Special function identities from superelliptic Kummer varieties. Asian J. Math., 21(5):909–951, 2017.
  • [10] A. Clingher, T. Hill, and A. Malmendier. The duality between F-theory and the heterotic string in D=8D=8 with two Wilson lines. Lett. Math. Phys., 110(11):3081–3104, 2020.
  • [11] A. Clingher and A. Malmendier. Normal forms for Kummer surfaces. London Mathematical Society Lecture Note Series, 2(459):107–162, 2019.
  • [12] A. Clingher and A. Malmendier. Kummer surfaces associated with (1,2)(1,2)-polarized abelian surfaces, arXiv:1704.04884[math.AG].
  • [13] A. Clingher, A. Malmendier, and T. Shaska. Six line configurations and string dualities. Comm. Math. Phys., 371(1):159–196, 2019.
  • [14] A. Clingher, A. Malmendier, and T. Shaska. On isogenies among certain abelian surfaces. Michigan Math. J., pages 1–43, 2021.
  • [15] A. Corti and V. Golyshev. Hypergeometric equations and weighted projective spaces. Sci. China Math., 54(8):1577–1590, 2011.
  • [16] D. A. Cox and S. Katz. Mirror symmetry and algebraic geometry, volume 68 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 1999.
  • [17] P. Deligne. Équations différentielles à points singuliers réguliers. Lecture Notes in Mathematics, Vol. 163. Springer-Verlag, Berlin-New York, 1970.
  • [18] I. Dolgachev. Integral quadratic forms: applications to algebraic geometry (after V. Nikulin). In Bourbaki seminar, Vol. 1982/83, volume 105 of Astérisque, pages 251–278. Soc. Math. France, Paris, 1983.
  • [19] I. V. Dolgachev. Mirror symmetry for lattice polarized K3K3 surfaces. volume 81, pages 2599–2630. 1996. Algebraic geometry, 4.
  • [20] C. F. Doran. Algebro-geometric isomonodromic deformations linking Hauptmoduls: variation of the mirror map. In Proceedings on Moonshine and related topics (Montréal, QC, 1999), volume 30 of CRM Proc. Lecture Notes, pages 27–35. Amer. Math. Soc., Providence, RI, 2001.
  • [21] C. F. Doran, A. Harder, H. Movasati, and U. Whitcher. Humbert surfaces and the moduli of lattice polarized K3 surfaces. In String-Math 2014, volume 93 of Proc. Sympos. Pure Math., pages 109–140. Amer. Math. Soc., Providence, RI, 2016.
  • [22] C. F. Doran and A. Malmendier. Calabi-Yau manifolds realizing symplectically rigid monodromy tuples. Adv. Theor. Math. Phys., 23(5):1271–1359, 2019.
  • [23] C. F. Doran and J. W. Morgan. Mirror symmetry and integral variations of Hodge structure underlying one-parameter families of Calabi-Yau threefolds. In Mirror symmetry. V, volume 38 of AMS/IP Stud. Adv. Math., pages 517–537. Amer. Math. Soc., Providence, RI, 2006.
  • [24] I. M. Gel’fand, M. M. Kapranov, and A. V. Zelevinsky. Generalized Euler integrals and AA-hypergeometric functions. Adv. Math., 84(2):255–271, 1990.
  • [25] Y. Goto. Arithmetic of weighted diagonal surfaces over finite fields. J. Number Theory, 59(1):37–81, 1996.
  • [26] B. R. Greene, S.-S. Roan, and S.-T. Yau. Geometric singularities and spectra of Landau-Ginzburg models. Comm. Math. Phys., 142(2):245–259, 1991.
  • [27] P. A. Griffiths. Periods of integrals on algebraic manifolds. I. Construction and properties of the modular varieties. Amer. J. Math., 90:568–626, 1968.
  • [28] P. A. Griffiths. Periods of integrals on algebraic manifolds. II. Local study of the period mapping. Amer. J. Math., 90:805–865, 1968.
  • [29] P. A. Griffiths. Periods of integrals on algebraic manifolds. III. Some global differential-geometric properties of the period mapping. Inst. Hautes Études Sci. Publ. Math., (38):125–180, 1970.
  • [30] P. A. Griffiths. Periods of integrals on algebraic manifolds: Summary of main results and discussion of open problems. Bull. Amer. Math. Soc., 76:228–296, 1970.
  • [31] M. Hara, T. Sasaki, and M. Yoshida. Tensor products of linear differential equations—a study of exterior products of hypergeometric equations. Funkcial. Ekvac., 32(3):453–477, 1989.
  • [32] H. Hironaka. Resolution of Singularities of an Algebraic Variety Over a Field of Characteristic Zero: I. Annals of Mathematics, 79(1):109–203, 1964. Publisher: Annals of Mathematics.
  • [33] H. Hironaka. Resolution of Singularities of an Algebraic Variety Over a Field of Characteristic Zero: II. Annals of Mathematics, 79(2):205–326, 1964. Publisher: Annals of Mathematics.
  • [34] S. Hosono, A. Klemm, S. Theisen, and S.-T. Yau. Mirror symmetry, mirror map and applications to complete intersection Calabi-Yau spaces. Nuclear Phys. B, 433(3):501–552, 1995.
  • [35] S. Hosono, B. H. Lian, H. Takagi, and S.-T. Yau. K3 surfaces from configurations of six lines in 2\mathbb{P}^{2} and mirror symmetry I. Commun. Number Theory Phys., 14(4):739–783, 2020.
  • [36] W. L. Hoyt. Notes on elliptic K3K3 surfaces. In Number theory (New York, 1984–1985), volume 1240 of Lecture Notes in Math., pages 196–213. Springer, Berlin, 1987.
  • [37] W. L. Hoyt. Elliptic fiberings of Kummer surfaces. In Number theory (New York, 1985/1988), volume 1383 of Lecture Notes in Math., pages 89–110. Springer, Berlin, 1989.
  • [38] W. L. Hoyt. On twisted Legendre equations and Kummer surfaces. In Theta functions—Bowdoin 1987, Part 1 (Brunswick, ME, 1987), volume 49 of Proc. Sympos. Pure Math., pages 695–707. Amer. Math. Soc., Providence, RI, 1989.
  • [39] W. L. Hoyt and C. F. Schwartz. Yoshida surfaces with Picard number ρ17\rho\geq 17. In Proceedings on Moonshine and related topics (Montréal, QC, 1999), volume 30 of CRM Proc. Lecture Notes, pages 71–78. Amer. Math. Soc., Providence, RI, 2001.
  • [40] R. Kloosterman. Classification of all Jacobian elliptic fibrations on certain K3K3 surfaces. J. Math. Soc. Japan, 58(3):665–680, 2006.
  • [41] K. Kodaira. On compact analytic surfaces. II, III. Ann. of Math. (2) 77 (1963), 563–626; ibid., 78:1–40, 1963.
  • [42] K. Kodaira. On the structure of compact complex analytic surfaces. II. Amer. J. Math., 88:682–721, 1966.
  • [43] K. Kodaira. On the structure of compact complex analytic surfaces. III. Amer. J. Math., 90:55–83, 1968.
  • [44] A. Kumar. Elliptic fibrations on a generic Jacobian Kummer surface. J. Algebraic Geom., 23(4):599–667, 2014.
  • [45] A. H. M. Levelt. Hypergeometric functions. Drukkerij Holland N. V., Amsterdam, 1961. Doctoral thesis, University of Amsterdam.
  • [46] A. Malmendier. Kummer surfaces associated with Seiberg-Witten curves. J. Geom. Phys., 62(1):107–123, 2012.
  • [47] A. Malmendier and D. R. Morrison. K3 surfaces, modular forms, and non-geometric heterotic compactifications. Lett. Math. Phys., 105(8):1085–1118, 2015.
  • [48] A. Malmendier and M. T. Schultz. From the signature theorem to anomaly cancellation. Rocky Mountain J. Math., 50(1):181–212, 2020.
  • [49] A. Malmendier and T. Shaska. The Satake sextic in F-theory. J. Geom. Phys., 120:290–305, 2017.
  • [50] A. Malmendier and T. Shaska. A universal genus-two curve from Siegel modular forms. SIGMA Symmetry Integrability Geom. Methods Appl., 13:Paper No. 089, 17, 2017.
  • [51] K. Matsumoto. Theta functions on the classical bounded symmetric domain of type I2,2{\rm I}_{2,2}. Proc. Japan Acad. Ser. A Math. Sci., 67(1):1–5, 1991.
  • [52] K. Matsumoto, T. Sasaki, and M. Yoshida. The period map of a 44-parameter family of K3K3 surfaces and the Aomoto-Gel’fand hypergeometric function of type (3,6)(3,6). Proc. Japan Acad. Ser. A Math. Sci., 64(8):307–310, 1988.
  • [53] K. Matsumoto, T. Sasaki, and M. Yoshida. The Aomoto-Gel’fand hypergeometric function and period mappings. Sūgaku, 41(3):258–263, 1989.
  • [54] K. Matsumoto, T. Sasaki, and M. Yoshida. The monodromy of the period map of a 44-parameter family of K3K3 surfaces and the hypergeometric function of type (3,6)(3,6). Internat. J. Math., 3(1):164, 1992.
  • [55] R. Miranda. Smooth models for elliptic threefolds. In The birational geometry of degenerations (Cambridge, Mass., 1981), volume 29 of Progr. Math., pages 85–133. Birkhäuser, Boston, Mass., 1983.
  • [56] D. R. Morrison. Picard-Fuchs equations and mirror maps for hypersurfaces. In Essays on mirror manifolds, pages 241–264. Int. Press, Hong Kong, 1992.
  • [57] D. R. Morrison. Compactifications of moduli spaces inspired by mirror symmetry. Number 218, pages 243–271. 1993. Journées de Géométrie Algébrique d’Orsay (Orsay, 1992).
  • [58] N. Narumiya and H. Shiga. The mirror map for a family of K3K3 surfaces induced from the simplest 3-dimensional reflexive polytope. In Proceedings on Moonshine and related topics (Montréal, QC, 1999), volume 30 of CRM Proc. Lecture Notes, pages 139–161. Amer. Math. Soc., Providence, RI, 2001.
  • [59] V. V. Nikulin. An analogue of the Torelli theorem for Kummer surfaces of Jacobians. Izv. Akad. Nauk SSSR Ser. Mat., 38:22–41, 1974.
  • [60] V. V. Nikulin. Kummer surfaces. Izv. Akad. Nauk SSSR Ser. Mat., 39(2):278–293, 471, 1975.
  • [61] V. V. Nikulin. Finite groups of automorphisms of Kählerian K3K3 surfaces. Trudy Moskov. Mat. Obshch., 38:75–137, 1979.
  • [62] V. V. Nikulin. Integer symmetric bilinear forms and some of their geometric applications. Izv. Akad. Nauk SSSR Ser. Mat., 43(1):111–177, 238, 1979.
  • [63] V. V. Nikulin. Quotient-groups of groups of automorphisms of hyperbolic forms by subgroups generated by 22-reflections. Algebro-geometric applications. In Current problems in mathematics, Vol. 18, pages 3–114. Akad. Nauk SSSR, Vsesoyuz. Inst. Nauchn. i Tekhn. Informatsii, Moscow, 1981.
  • [64] T. Sasaki. Contiguity relations of Aomoto-Gel’fand hypergeometric functions and applications to Appell’s system F3F_{3} and Goursat’s system F23{}_{3}F_{2}. SIAM J. Math. Anal., 22(3):821–846, 1991.
  • [65] T. Sasaki and M. Yoshida. Linear differential equations modeled after hyperquadrics. Tohoku Mathematical Journal, 41(2), Jan. 1989.
  • [66] M. Schulze and U. Walther. Resonance equals reducibility for AA-hypergeometric systems. Algebra Number Theory, 6(3):527–537, 2012.
  • [67] I. Shimada. On elliptic K3K3 surfaces. Michigan Math. J., 47(3):423–446, 2000.
  • [68] J. Stienstra. Resonant hypergeometric systems and mirror symmetry. In Integrable systems and algebraic geometry (Kobe/Kyoto, 1997), pages 412–452. World Sci. Publ., River Edge, NJ, 1998.
  • [69] J. Stienstra. GKZ hypergeometric structures. In Arithmetic and geometry around hypergeometric functions, volume 260 of Progr. Math., pages 313–371. Birkhäuser, Basel, 2007.
  • [70] N. Suwa and N. Yui. Arithmetic of certain algebraic surfaces over finite fields. In Number theory (New York, 1985/1988), volume 1383 of Lecture Notes in Math., pages 186–256. Springer, Berlin, 1989.
  • [71] R. Vidūnas. Specialization of Appell’s functions to univariate hypergeometric functions. J. Math. Anal. Appl., 355(1):145–163, 2009.
  • [72] È. B. Vinberg. On automorphic forms on symmetric domains of type IV. Uspekhi Mat. Nauk, 65(3(393)):193–194, 2010.
  • [73] E. B. Vinberg. Some free algebras of automorphic forms on symmetric domains of type IV. Transform. Groups, 15(3):701–741, 2010.
  • [74] E. B. Vinberg. On the algebra of Siegel modular forms of genus 2. Trans. Moscow Math. Soc., pages 1–13, 2013.
  • [75] N. Yui. The arithmetic of certain Calabi-Yau varieties over number fields. In The arithmetic and geometry of algebraic cycles (Banff, AB, 1998), volume 548 of NATO Sci. Ser. C Math. Phys. Sci., pages 515–560. Kluwer Acad. Publ., Dordrecht, 2000.
  • [76] N. Yui. Arithmetic of certain Calabi-Yau varieties and mirror symmetry. In Arithmetic algebraic geometry (Park City, UT, 1999), volume 9 of IAS/Park City Math. Ser., pages 507–569. Amer. Math. Soc., Providence, RI, 2001.