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

On the generic degree of two-parameter period mappings

Chongyao Chen and Haohua Deng Department of Mathematics, Duke University, Durham, North Carolina, 27708-0320 [email protected] [email protected]

Abstract. We present a method for computing the generic degree of a period map defined on a quasi-projective surface. As an application, we explicitly compute the generic degree of three period maps underlying families of Calabi-Yau 3-folds coming from toric hypersurfaces. As a consequence, we show that the generic Torelli theorem holds for these cases.

1. Introduction

The period map is an important tool for studying families of algebraic varieties. Many nice properties associated to smooth projective varieties varying in a family are recorded by the associated polarized variation of Hodge structures (PVHS).

Torelli-type problems characterize the extent to which one could distinguish non-isomorphic projective varieties via their Hodge structures. There are different types of Torelli theorems. The global Torelli theorem is the strongest, stating that any two non-isomorphic projective varieties in a family can be completely distinguished by their polarized Hodge structures. There are classical examples for the global Torelli theorem like the universal family of principle polarized abelian varieties and moduli space of marked polarized K3 surfaces. One of the most recent examples is the family of mirror quintic Calabi-Yau threefolds studied in [Fil23]. The infinitesimal Torelli theorem, on the other hand, states that the first-order deformation of a projective algebraic variety is mapped faithfully to its infinitesimal variation of Hodge structure. There are many more known examples for the infinitesimal Torelli theorem.

In this paper we consider the generic Torelli theorem, or more generally, the generic degree of a period map provided that the degree is well-defined. This is weaker than the global Torelli theorem in the sense that one has the global Torelli theorem only on a Zariski open subset, but it is still very useful.

We introduce the first main result of this paper. Let Φ:SΓ\D\Phi:S\rightarrow\Gamma\backslash D be a period map defined on a quasi-projective surface SS. We also choose a projective completion S¯\overline{S} of SS with S¯S\overline{S}-S being a simple normal crossing divisor. We show the following informally stated result:

Theorem 1.1.

The generic degree of Φ:SΓ\D\Phi:S\rightarrow\Gamma\backslash D can be computed from the limiting mixed Hodge structure (LMHS) types and monodromy matrices around boundary points sS¯Ss\in\overline{S}-S.

The main tool for our proof is Kato-Nakayama-Usui’s theory on the space of nilpotent orbits [KU08], [KNU13] which is based on the classical nilpotent and SL2\mathrm{SL}_{2}-orbit theorems [Sch73], [CKS86]. Since KNU’s construction always exists for one-parameter period maps, several authors have successfully computed the generic degree of some one-parameter period maps as applications of the theory, see for example [Usu08], [Shi09], [HK21]111Though the result in [Usu08] is correct, it has one flawed argument which may cause problems in higher-dimensional cases. The arguments in this paper successfully fix it..

Recently, C. Robles and the second author showed that KNU’s construction exists for any two-parameter period maps [DR23]. This allows us to prove Theorem 1.1 by only looking at some distinguished boundary points (and their image in the Kato–Nakayama–Usui space) like the one-parameter case, though the details are more complicated.

The second part of the paper is devoted to the study of specific examples. We study three different families of smooth Calabi-Yau threefolds with Hodge numbers (1,2,2,1)(1,2,2,1) that are hypersurfaces of complete toric varieties and compute the generic degree of each of these families using Theorem 1.1. These families are the primary examples of study on the subject of mirror symmetry [AGM94, Hos+95, CK99]. We denote them as the families V(2,29)V_{(2,29)}, V(2,38)V_{(2,38)} and V(2,86)V_{(2,86)} based on their Hodge diamonds. In particular, the family V(2,86)V_{(2,86)} is also known as the family of mirror octics.

For each of these Calabi-Yau hypersurfaces, we study the tautological family over the simplified moduli space. This family is a base change of the tautological family over complex moduli \mathcal{M} via a generically étale morphism ϕ:simp\phi:\mathcal{M}_{\mathrm{simp}}\rightarrow\mathcal{M}. One advantage of this family is that simp\mathcal{M}_{\mathrm{simp}} admits a natural projective completion given by a complete toric surface. Now, we study the VHS associated to these families and calculate boundary LMHS types and monodromy matrices required in Theorem 1.1.

First, the Picard-Fuchs system can be obtained from the GKZ system via Lemma 2.9, which justifies the factorization argument first made in [Hos+95] and rephrases it in the language of 𝒟\mathcal{D}-modules. Then one can calculate the discriminant locus from the Picard-Fuchs 𝒟\mathcal{D}-module. By blowing up the non-complete intersection points on the discriminant locus, we then obtain a smooth projective completion ¯simp\overline{\mathcal{M}}_{\mathrm{simp}} of simp\mathcal{M}_{\mathrm{simp}}, such that ¯simpsimp\overline{\mathcal{M}}_{\mathrm{simp}}-\mathcal{M}_{\mathrm{simp}} is a simple normal crossing divisor as desired. Next, one can obtain the Gauss-Manin connection represented in a chosen local basis from the Picard-Fuchs 𝒟\mathcal{D}-module. Then the log-monodromies represented in this basis are just the (Tate twisted) residues of the Gauss-Manin connection in suitable local coordinates. Meanwhile, the symplectic form can be obtained by calculating the intersection matrix of the local basis, which can be written in terms of the Yukawa couplings and nn-point functions. Finally, the types of the LMHS can be obtained via the Jordan normal form of the nilpotent operators and the polarized period relations [KPR19], as the first author did in [Che23].

As a consequence of computations and application of Theorem 1.1, we arrive at the second main theorem of this paper:

Theorem 1.2.

The period maps associated to the families V(2,29)V_{(2,29)}, V(2,38)V_{(2,38)}, and V(2,86)V_{(2,86)} (mirror octic) over simplified moduli have generic degree 22, 11, and 11 respectively.

Recall that a smooth variety is said to satisfy the generic Torelli theorem if the period map associated to the tautological family over complex moduli has generic degree 11. Theorem 1.2 and the computation of the generic degrees of ϕ:simp\phi:\mathcal{M}_{\mathrm{simp}}\rightarrow\mathcal{M} in Section 4 also imply:

Corollary 1.3.

The Calabi-Yau threefolds V(2,29)V_{(2,29)}, V(2,38)V_{(2,38)}, and V(2,86)V_{(2,86)} satisfy the generic Torelli theorem.

To the best of our knowledge, these are the first known examples of Calabi-Yau threefolds that satisfy the generic Torelli theorem with complex moduli larger than one dimension.

The article is organized as follows. In Section 2 we summarize all necessary materials in toric geometry and Hodge theory. In Section 3 we present and prove the main formula for computing generic degree. Section 4 is devoted to the study of the three families V(2,29),V(2,38)V_{(2,29)},V_{(2,38)} and V(2,86)V_{(2,86)}, including explicitly their boundary LMHS types and monodromy matrices. In Section 5 we will apply the main theorem proved in Section 3 and computational results from Section 4 to show Theorem 1.2.

Acknowledgement: The authors thank Colleen Robles for exchanges of many insightful ideas. The second author also thanks Matt Kerr for related discussions.

2. Background materials

2.1. Toric geometry

In this section, we fix the notations that will be used in the next section. These are standard and follow closely those in [CK99, CLS11].

2.1.1. Toric varieties

Denote the standard full-rank integral lattice group in n\mathbb{Q}^{n} as NN and its dual lattice as MM. We will use the k×kk\times k matrices An×kB(kn)×k\frac{A_{n\times k}}{B_{(k-n)\times k}} to record the toric data. The column vectors of AA form a finite set

{v1,v2,,vk}=:ΞN,\{v_{1},v_{2},\dots,v_{k}\}=:\Xi\subset N,

and the set of row vectors {w1,w2,,wr}\{w_{1},w_{2},\dots,w_{r}\} (r:=knr:=k-n) of BB generates the lattice of relations among Ξ\Xi

Λ:={l=(li)k:i=1klivi=0}\Lambda:=\left\{l=(l_{i})\in\mathbb{Z}^{k}:\sum_{i=1}^{k}l_{i}v_{i}=0\right\}

over \mathbb{Z}. In particular, we have ABt=0A\cdot B^{t}=0. These data can be viewed as recording either the information about a polyhedral fan Σ\Sigma or the information of an integral polytope Δ\Delta^{\circ}. Each of these points of view leads to one way to construct a toric variety, which we now briefly review.

For a strongly convex rational polyhedral fan ΣN:=N\Sigma\subset N_{\mathbb{R}}:=N\otimes_{\mathbb{Z}}\mathbb{R}, denote Σ(m)\Sigma(m) as the set of its mm-dimension cones. In this case, we use the columns of AA to record the primitive generators of elements in Σ(1)\Sigma(1). Denote σi1,,imΣ(m)\sigma_{i_{1},\dots,i_{m}}\in\Sigma(m) as the mm-dimensional cone generated by {vi1,vi2,,vim}\{v_{i_{1}},v_{i_{2}},\dots,v_{i_{m}}\}. The toric variety constructed from Σ\Sigma is denoted as XΣX_{\Sigma} and is called the GIT quotient construction. More precisely, let S=[x1,,xk]S=\mathbb{C}[x_{1},\dots,x_{k}].

Definition 2.1.

The Cox ideal IΣI_{\Sigma} is the ideal of SS generated by the monomials corresponding to each σΣ(n)\sigma\in\Sigma(n), defined by

xi1xi2xiα,α=kσ(1),+visσ(1).x_{i_{1}}x_{i_{2}}\dots x_{i_{\alpha}},\quad\alpha=k-\sigma(1),\,\mathbb{R}_{+}\cdot v_{i_{s}}\notin\sigma(1).

In particular, we have

XΣ\displaystyle X_{\Sigma} =ProjΣ(S)\displaystyle=\mathrm{Proj}_{\Sigma}(S)
=[Spec(S)𝕍(IΣ)]G,\displaystyle=[\mathrm{Spec(S)-\mathbb{V}(I_{\Sigma})}]\sslash G,

where the G:=()rG:=(\mathbb{C}^{*})^{r}-action on k\mathbb{C}^{k} is determined by the matrix BB as

G×k\displaystyle G\times\mathbb{C}^{k} k\displaystyle\rightarrow\mathbb{C}^{k}
(s1,,sr)×(x1,,xk)\displaystyle(s_{1},\dots,s_{r})\times(x_{1},\dots,x_{k}) (i=1rsiwi,1x1,,i=1rsiwi,kxk)\displaystyle\rightarrow\left(\prod_{i=1}^{r}s_{i}^{w_{i,1}}\cdot x_{1},\cdots,\prod_{i=1}^{r}s_{i}^{w_{i,k}}\cdot x_{k}\right)

This construction suggests that one could think about the toric variety as a generalization of weighted projected space. More precisely, the degree of a variable is replaced by a multi-degree, and the origin that got deleted before the quotient is replaced by 𝕍(IΣ)\mathbb{V}(I_{\Sigma}). For the variable xix_{i}, the multi-degree is deg(xi)=(w1,i,,wr,i)\mathrm{deg}(x_{i})=(w_{1,i},\dots,w_{r,i}). We note this multi-degree can be viewed as taking value in An1(XΣ)A_{n-1}(X_{\Sigma})\otimes\mathbb{Z}.

For an integral polytope Δ\Delta^{\circ} i.e., a polytope in k\mathbb{R}^{k} whose vertices lie in NN. The toric data records all the integral points inside the polytope Δ\Delta^{\circ}, i.e., Ξ=ΔN\Xi=\Delta^{\circ}\cap N. We will use Δi1,,im\Delta^{\circ}_{i_{1},\dots,i_{m}} to denote the m1m-1-dimensional face of Δ\Delta^{\circ} that has vi1,,vimv_{i_{1}},\dots,v_{i_{m}} as its vertices. Denote the toric variety constructed from it as Δ\mathbb{P}_{\Delta^{\circ}}, which is the Zariski closure of the open immersion

(2.1) ()nk(t1,,tn)[i=1ntir1,i:i=1ntir2,i::i=1ntirk,i].\begin{split}(\mathbb{C}^{*})^{n}&\hookrightarrow\mathbb{P}^{k}\\ (t_{1},\dots,t_{n})&\rightarrow\left[\prod_{i=1}^{n}t_{i}^{r_{1,i}}:\prod_{i=1}^{n}t_{i}^{r_{2,i}}:\cdots:\prod_{i=1}^{n}t_{i}^{r_{k,i}}\right].\end{split}

This also induces a ()n(\mathbb{C}^{*})^{n}-automorphism on Δ\mathbb{P}_{\Delta^{\circ}}. We will call this construction the projective embedding construction.

Each of the two constructions contains a top dimensional Zariski dense tori T:=()nT:=(\mathbb{C}^{*})^{n} inside the toric variety. For Δ\mathbb{P}_{\Delta^{\circ}} it is clear from the construction, whereas for XΣX_{\Sigma}, we have T=()k/GXΣT=(\mathbb{C}^{*})^{k}/G\hookrightarrow X_{\Sigma}.

2.1.2. Reflexive polytope and Baytrev mirror

The starting point of a Baytrev’s mirror construction is a reflexive polytope and its mirror, we first recall some definitions.

Definition 2.2.

For an arbitrary full-dimensional integral polytope ΔM\Delta\subset M_{\mathbb{R}}. For each codimension ss face FF of Δ\Delta, let σF\sigma_{F} be the ss-dimensional strongly convex rational polyhedral cone in NN_{\mathbb{R}} defined as

σF:={vN|m,vm,v,mF,mΔ}.\sigma_{F}:=\{v\in N_{\mathbb{R}}|\braket{m,v}\geq\braket{m^{\prime},v},\forall m\in F,m^{\prime}\in\Delta\}.

Then the normal fan ΣΔ\Sigma_{\Delta} of Δ\Delta is the union of all σF\sigma_{F}.

Definition 2.3.

Let Δ\Delta be the same in Definition 2.2, then ΔM\Delta\subset M_{\mathbb{R}} is called reflexive if for any facet (codimension-1 face) Γ\Gamma, there is a unique normal vector vΓNv_{\Gamma}\in N, such that

Δ={mM|m,vΓ1,ΓΔ}.\Delta=\{m\in M_{\mathbb{R}}|\braket{m,v_{\Gamma}}\geq-1,\forall\ \Gamma\leq\Delta\}.

and Int(Δ)N={0}\mathrm{Int}(\Delta)\cap N=\{0\}. The polar dual ΔN\Delta^{\circ}\subset N_{\mathbb{R}} of Δ\Delta is defined as

Δ:={nN|m,n1,mΔ}.\Delta^{\circ}:=\{n\in N_{\mathbb{R}}|\braket{m,n}\geq-1,\forall m\in\Delta\}.

If Δ\Delta is reflexive, Δ\Delta^{\circ} is also reflexive and we have (Δ)=Δ(\Delta^{\circ})^{\circ}=\Delta. Furthermore, Δ\Delta^{\circ} is the convex hull of vΓv_{\Gamma}, and {vΓ}\{v_{\Gamma}\} spans ΣΔ(1)\Sigma_{\Delta}(1). For each face FF of Δ\Delta, one can define its dual face to be

F:=FΓΓΔ,F^{\circ}:=\bigcap_{F\subset\Gamma}\Gamma\subset\Delta^{\circ},

where Γ\Gamma runs through all the facets of Δ\Delta that contains FF. One can easily show that (F)=F(F^{\circ})^{\circ}=F and dim(F)+dim(F)=dimΔ1\dim(F)+\dim(F^{\circ})=\dim\Delta-1.

These constructions also make the comparison of the two different constructions of the toric variety possible. In fact, we have ΔXΣΔ\mathbb{P}_{\Delta^{\circ}}\cong X_{\Sigma_{\Delta}}, and ΔXΣ\mathbb{P}_{\Delta}\cong X_{\Sigma^{\circ}}, where Σ:=ΣΔ\Sigma^{\circ}:=\Sigma_{\Delta^{\circ}}.

By Baytrev mirror construction, the mirror of XΣX_{\Sigma}, denote as XˇΣ\check{X}_{\Sigma} is XˇΣXΣΔ\check{X}_{\Sigma}\cong X_{\Sigma^{\circ}}\cong\mathbb{P}_{\Delta^{\circ}}. On the level of the Calabi-Yau hypersurface, we have Xˇ\check{X} as the anti-canonical hypersurface of XˇΣ\check{X}_{\Sigma} that is mirror to XX.

2.1.3. Automorphisms of XΣX_{\Sigma}

The automorphism group Aut(XΣ)\mathrm{Aut}(X_{\Sigma}) of XΣX_{\Sigma} is generated by three types of automorphisms that come from the TT-action induced from (2.1), the root symmetry, and the fan symmetry. The first two types of automorphism are continuous, together they generate the connected component Aut0(XΣ)\mathrm{Aut}_{0}(X_{\Sigma}) of Aut(XΣ)\mathrm{Aut}(X_{\Sigma}).

The root symmetry is defined as follows, for each of the variables xi,i=1,,kx_{i},i=1,\dots,k in the GIT quotient construction, if there exists a monomial xαx^{\alpha} in SS, such that xixαx_{i}\nmid x^{\alpha}, and deg(xi)=deg(xα)\mathrm{deg}(x_{i})=\mathrm{deg}(x^{\alpha}). Then the root pair (xi,xα)(x_{i},x^{\alpha}) defines an \mathbb{C}-automorphism on XΣX_{\Sigma}

×XΣXΣ(λ,[x1::xk])[x1::xi1:xi+λxα:xi+1::xk].\begin{split}\mathbb{C}\times X_{\Sigma}&\rightarrow X_{\Sigma}\\ (\lambda,[x_{1}:\dots:x_{k}])&\rightarrow[x_{1}:\dots:x_{i-1}:x_{i}+\lambda x^{\alpha}:x_{i+1}:\dots:x_{k}].\end{split}

Therefore, the group of root symmetry Autr(XΣ)\mathrm{Aut}_{r}(X_{\Sigma}) is the group generated by all the possible root pairs.

The last type of automorphism comes from the symmetry of the Σ\Sigma. More precisely, these automorphisms form a finite group Aut(Σ)\mathrm{Aut}(\Sigma), the subgroup of Aut(N)\mathrm{Aut}(N) that preserves Σ\Sigma.

2.1.4. Moduli spaces

There are three moduli spaces that are naturally associated to VV, i.e., the polynomial moduli, the simplified moduli, and the complex moduli. The polynomial moduli is

poly:=(L(ΔN))s/Aut(XΣ),\mathcal{M}_{\mathrm{poly}}:=\mathbb{P}(L(\Delta^{\circ}\cap N))^{s}/\mathrm{Aut}(X_{\Sigma^{\circ}}),

where L(ΔN)L(\Delta^{\circ}\cap N) is the vector space of Laurent polynomials associated to the finite set ΔN\Delta^{\circ}\cap N. Meanwhile, the superscript ss refers to the restriction to the smooth locus, which is quasi-projective. This space parameterized the complex structure of VV that can be realized as hypersurfaces of XΣX_{\Sigma}. The simplified is

simp:=(L((ΔN))0)s/T,\mathcal{M}_{\mathrm{simp}}:=\mathbb{P}(L((\Delta^{\circ}\cap N))_{0})^{s}/T,

where (ΔN))0(\Delta^{\circ}\cap N))_{0} is the subset of ΔN\Delta^{\circ}\cap N that excludes the points that lies in the interior of any facet of Δ\Delta^{\circ}.

The dominance theorem conjectured in [AGM93] and proved in [CK99] states that

Theorem 2.4.

The quotient map ϕ:simppoly\phi:\mathcal{M}_{\mathrm{simp}}\rightarrow\mathcal{M}_{\mathrm{poly}} is generically étale.

So if Autr(XΣ)=0\mathrm{Aut}_{r}(X_{\Sigma})=0, then the generic degree of ϕ\phi is the order of the subgroup of Aut(Σ)\mathrm{Aut}(\Sigma) that acts on simp\mathcal{M}_{\mathrm{simp}} non-trivially. Finally, the complex moduli \mathcal{M} is the moduli space of the complex structure. \mathcal{M} and poly\mathcal{M}_{\mathrm{poly}} are related by the following proposition

Proposition 2.5.

The following are equivalent:

  1. (1)

    =poly\mathcal{M}=\mathcal{M}_{poly}.

  2. (2)

    dim=dimpoly\dim\mathcal{M}=\dim\mathcal{M}_{poly}.

  3. (3)

    For any two-dimensional face of Δ\Delta, either it has no interior point, or its dual face in Δ\Delta^{\circ} has no interior point.

Proof.

See Proposition 6.1.3 of [CK99]. ∎

2.1.5. GKZ system

A GKZ AA-hypergeometric system τ(A,β)\tau(A,\beta) is a 𝒟n\mathcal{D}_{\mathbb{C}^{n}}-module that is determined by a tuple (A,β)(A,\beta), where AMm×n()A\in M_{m\times n}(\mathbb{Z}) and βm\beta\in\mathbb{Z}^{m}. For an introduction to the theory of 𝒟\mathcal{D}-module, we refer to [HT07].

For any Calabi-Yau hypersurface inside a toric variety whose toric data is recorded in (AB)(\frac{A}{B}), one can associate it with a GKZ system τ(A¯,β)\tau(\bar{A},\beta) that is constructed as follows. First we define the suspended fan Σ¯\bar{\Sigma} of Σ\Sigma, whose 1-dimensional cones are generated by Ξ¯:=(Ξ{0})×1\bar{\Xi}:=(\Xi\cup\{0\})\times 1. The toric data of the suspended fan will be recorded as (A¯B¯)(\frac{\bar{A}}{\bar{B}}). Then τ(A¯,β)\tau(\bar{A},\beta) is the cyclic 𝒟k+1\mathcal{D}_{\mathbb{C}^{k+1}}-module, determined by the left ideal IGKZI_{GKZ} in Weyl algebra, which is generated by

li>0λililj<0λjlj,lΛ,\prod_{l_{i}>0}\partial_{\lambda_{i}}^{l_{i}}-\prod_{l_{j}<0}\partial_{\lambda_{j}}^{l_{j}},\quad l\in\Lambda,

and

j=1k+1ri,jλjj+βi,i=1,,n+1,\sum_{j=1}^{k+1}r_{i,j}\lambda_{j}\partial_{j}+\beta_{i},\quad i=1,\dots,n+1,

where λi\lambda_{i} are the coordinates of k+1\mathbb{C}^{k+1}, and β=(0,,0,1)\beta=(0,\dots,0,-1). Now, recall the general result of the GKZ system

Theorem 2.6 (Hotta).

Let AMm×n()A\in M_{m\times n}(\mathbb{Z}), if (1,1,,1)(1,1,\dots,1) lies in the row span of AA, then τ(A,β)\tau(A,\beta) is regular holonomic for any βm\beta\in\mathbb{C}^{m}.

Proof.

See [Hot91]. ∎

Then the GKZ system we constructed above is always regular holonomic. Now consider, Let j:()k+1k+1j:(\mathbb{C}^{*})^{k+1}\hookrightarrow\mathbb{C}^{k+1} be the immersion, and π:()k+1()r\pi:(\mathbb{C}^{*})^{k+1}\rightarrow(\mathbb{C}^{*})^{r} be the projection under the action (2.1). Then p+jτ(A¯,β)p_{+}j^{*}\tau(\bar{A},\beta) is again regular holonomic by the general theory of 𝒟\mathcal{D}-modules, (see e.g. [HT07]).

Lemma 2.7.

p+jτ(A¯,β)p_{+}j^{*}\tau(\bar{A},\beta) is a regular holonomic and is determined by the left ideal sheaf GKZ𝒟()r\mathcal{I}_{GKZ}\subset\mathcal{D}_{(\mathbb{C}^{*})^{r}} that is locally generated by the following rr differential operators (i=1,,ri=1,\dots,r)

Pi:=j:wi,j>0k=1wi,j(l=1rwl,jδlkδj,k+1)zij:wi,j<0k=1wi,j(l=1rwl,jδlkδj,k+1),P_{i}:=\prod_{j:w_{i,j}>0}\prod_{k=1}^{w_{i,j}}(\sum_{l=1}^{r}w_{l,j}\delta_{l}-k-\delta_{j,k+1})-z_{i}\prod_{j:w_{i,j}<0}\prod_{k=1}^{w_{i,j}}(\sum_{l=1}^{r}w_{l,j}\delta_{l}-k-\delta_{j,k+1}),

where δl:=zlzl\delta_{l}:=z_{l}\partial_{z_{l}} with zlz_{l}, l=1,,rl=1,\dots,r be the coordinates on ()r(\mathbb{C}^{*})^{r}, and δi,j\delta_{i,j} is the Kronecker symbol.

Proof.

By [Gel+94], the singular locus (the principal A¯\bar{A}-deteminant) of jτ(A¯,β)j^{*}\tau(\bar{A},\beta) is away from the origin, so locally near origin, we have π\pi is smooth. Then by [HT07], the push forward along a smooth morphism is just a change of variables. For more detail see [Gel+94, CK99]. ∎

Remark 2.8.

The GKZ system can be viewed as a special case of the tautological system [LSY13]. In that regard, the suspension corresponds to including the Euler operator in the system and is the only part specialized to the anti-canonical section.

In what follows, we will refer to the 𝒟\mathcal{D}-module p+jτ(A¯,β)p_{+}j^{*}\tau(\bar{A},\beta) as the GKZ system, and denote it as τGKZ\tau_{GKZ}.

The singular locus of the GKZ system τGKZ\tau_{GKZ} is the so-called principal AA-determinant and it has a factorization as

EA=FΔD(ΔN)0Fm(F),E_{A}=\prod_{F\subset\Delta}D_{(\Delta^{\circ}\cap N)_{0}\cap F}^{m(F)},

where FF runs through all the faces of Δ\Delta, and D(ΔN)0FD_{(\Delta^{\circ}\cap N)_{0}\cap F} are the so-call AA-discriminants. We refer to [Gel+94, CK99] for the precise definition of the right-hand side. In practice, it is often easier to calculate the singular locus directly, which equals the projection of the characteristic variety minus the zero section.

2.1.6. Secondary fan

The space simp\mathcal{M}_{\mathrm{simp}} has a natural compactification given by the Chow quotient[KSZ91]

¯simp:=(L((ΔN))0)cT.\overline{\mathcal{M}}_{\mathrm{simp}}:=\mathbb{P}(L((\Delta^{\circ}\cap N))_{0})\sslash_{c}T.

This is an rr-dimensional toric variety whose toric data is denoted as (AsBs)(\frac{A^{s}}{B^{s}}), where the columns of AsA^{s} are the distinct primitive vectors of the columns of B¯\overline{B}. The fan Σs\Sigma^{s} of ¯simp\overline{\mathcal{M}}_{\mathrm{simp}} is called the secondary fan, and can be determined by calculating the GKZ decomposition, see for example [CK99]. In the case r2r\leq 2, the Σs\Sigma^{s} is completely determined by Σs(1)\Sigma^{s}(1).

The GKZ system τGKZ\tau_{GKZ} thus can be viewed as sitting on the dense tori ()r(\mathbb{C}^{*})^{r} of ¯simp\overline{\mathcal{M}}_{\mathrm{simp}}. Or one can consider the minimal extension L(τGKZ,()r)L(\tau_{GKZ},(\mathbb{C}^{*})^{r}) to ¯simp\overline{\mathcal{M}}_{\mathrm{simp}}. This is a holonomic 𝒟¯simp\mathcal{D}_{\overline{\mathcal{M}}_{\mathrm{simp}}}-module whose singular locus is contains inside

Sing(τGKZ)=EADv1sDvqs,\mathrm{Sing}(\tau_{GKZ})=E_{A}\cup D_{v_{1}^{s}}\cup\dots\cup D_{v_{q}^{s}},

where visv_{i}^{s}, i=1,,qi=1,\dots,q are the volumes of AsA^{s}, and DvisD_{v_{i}^{s}} is the toric divisor of ¯simp\overline{\mathcal{M}}_{\mathrm{simp}} associated to visv_{i}^{s}.

The geometry of ¯simpsimp\overline{\mathcal{M}}_{\mathrm{simp}}-\mathcal{M}_{\mathrm{simp}} is well-understood. In [Gel+94] the authors show EAE_{A} intersection of each DvisD_{v_{i}^{s}} at exactly one point with possible multiplicity. Meanwhile, the intersections between the toric divisors are recorded in the secondary fan. Therefore, by blowing up the tangencies at EADvisE_{A}\cap D_{v_{i}^{s}}, one arrives at a compactification of simp\mathcal{M}_{\mathrm{simp}} with normal crossing boundary divisors.

2.1.7. Picard-Fuchs system

Let Δ\mathbb{P}_{\Delta^{\circ}} be a toric variety with at most terminal singularity, and let VΔV\subset\mathbb{P}_{\Delta^{\circ}} be a smooth anti-canonical hypersurface. We are interested in the tautological family 𝒞\mathcal{C}\rightarrow\mathcal{M}. The Picard-Fuchs system τPF(ω)\tau_{PF}(\omega) for VV is a 𝒟\mathcal{D}_{\mathcal{M}}-module. At a generic non-singular point bb\in\mathcal{M}, the local holomorphic solution germ of the Picard-Fuchs system is a finite dimension \mathbb{C}-vector space that is spanned by the period functions of a fixed local section ω\omega of the relative canonical sheaf K𝒞/K_{\mathcal{C}/\mathcal{M}}.

If we further assume poly=\mathcal{M}_{\mathrm{poly}}=\mathcal{M}, then simp\mathcal{M}_{\mathrm{simp}} is a (possibly ramified) finite cover of \mathcal{M}. In this case, simp\mathcal{M}_{\mathrm{simp}} parameterized the isomorphism class of VV together with an isomorphism of Δ\mathbb{P}_{\Delta^{\circ}}. Thus simp\mathcal{M}_{\mathrm{simp}} still carries a tautological family 𝒞simp\mathcal{C}^{\prime}\rightarrow\mathcal{M}_{simp}, and one can also construct a Picard-Fuchs system by fixing a local section of K𝒞/simpK_{\mathcal{C}^{\prime}/\mathcal{M}_{\mathrm{simp}}}. In [BC94] a canonical invariant section of K𝒞/K_{\mathcal{C}^{\prime}/\mathcal{M}} is defined as

ω=Resf=0Ωf,\omega=\mathrm{Res}_{f=0}\frac{\Omega}{f},

where Ω\Omega is a TT-invariant meromorphic holomorphic mm-form on Δ\mathbb{P}_{\Delta^{\circ}} that does not depend on the base point.

The construction of the GKZ system implies [Hos+95, LSY13] that on simp\mathcal{M}_{\mathrm{simp}}, τPF\tau_{PF} is a quotient module of τGKZ\tau_{GKZ}. Since both of these simp\mathcal{M}_{\mathrm{simp}}-modules are cyclic with generator ω\omega, then they have a canonical filtration

FτPF,GKZ:=F𝒟simpω.F_{\bullet}\tau_{PF,GKZ}:=F_{\bullet}\mathcal{D}_{\mathcal{M}_{\mathrm{simp}}}\cdot\omega.

In particular, they both defines a \mathbb{C}-VHS on the non-singular locus, with (τPF,FτPF)(𝕍,Fm1𝕍)(\tau_{PF},F_{\bullet}\tau_{PF})\cong\mathcal{(}\mathbb{V},F^{m-1-\bullet}\mathbb{V}), where (𝕍,F𝕍)(\mathbb{V},F^{\bullet}\mathbb{V}) is the \mathbb{C}-VHS associated to the tautological family.

Now, for the case m=4m=4, i.e., VV is a Calabi-Yau 3-fold, we have

Lemma 2.9.

If τGKZ\tau_{GKZ} has holonomic rank less than 4r+44r+4, and it has a quotient module \mathcal{M} with holonomic tank 2r+22r+2, such that (,F)(\mathcal{M},F^{\bullet}\mathcal{M}) has a MUM point, where FF^{\bullet}\mathcal{M} is the induced filtration from the canonical filtration, then generically τPF\tau_{PF}\cong\mathcal{M}.

Proof.

For the definition of MUM point see Definition 2.14. As pointed out in [Che23], the existence of a MUM point implies the VHS is irreducible. Then since τPF\tau_{PF} is a quotient module of τGKZ\tau_{GKZ} with holonomic rank 2r+22r+2, thus generically it has to isomorphic to \mathcal{M}. ∎

We also note that in [HLZ16], the authors showed that for a smooth Calabi-Yau (m1)(m-1)-fold VΔV\subset\mathbb{P}_{\Delta}, the holonomic rank of τGKZ\tau_{GKZ} is equals to dim(Hm(ΔV,))\dim(H^{m}(\mathbb{P}_{\Delta^{\circ}}-V,\mathbb{C})), which is greater than dim(Hm1(V,))\dim(H^{m-1}(V,\mathbb{C})).

2.2. Hodge theory

Throughout this subsection we fix the tuple

(H,Q,{hp,q}p+q=l,D)(H_{\mathbb{Z}},Q,\{h^{p,q}\}_{p+q=l},D),

where (H,Q)(H_{\mathbb{Z}},Q) is an integral polarized lattice and {hp,q}p+q=l\{h^{p,q}\}_{p+q=l} is a set of Hodge numbers for some polarized Hodge structure of weight ll on (H,Q)(H_{\mathbb{Z}},Q) and DD is the corresponding period domain. We also denote D=G/KDˇ=G/PD=G_{\mathbb{R}}/K\subset\check{D}=G_{\mathbb{C}}/P where G:=Aut(H,Q)G:=\mathrm{Aut}(H,Q).

Let SS be a smooth quasi-projective variety, S¯\bar{S} be projective with S¯\S\bar{S}\backslash S a simple normal crossing divisor. We assume there is a \mathbb{Z}-local system 𝕍\mathbb{V} on SS which induces a polarized variation of Hodge structures (PVHS) of given type. This induces a period map:

(2.2) φ:SΓ\D\varphi:S\rightarrow\Gamma\backslash D

where DD is the classifying space of \mathbb{Z}-polarized Hodge structures of type (H,Q,{hp,q}p+q=l)(H_{\mathbb{Z}},Q,\{h^{p,q}\}_{p+q=l}), and ΓG=Aut(H,Q)\Gamma\leq G=\mathrm{Aut}(H_{\mathbb{Z}},Q) is the monodromy group.

These abstract settings have natural models in algebraic geometry: Suppose

(2.3) π:𝒳S\pi:\mathcal{X}\rightarrow S

is a smooth projective family where 𝒳\mathcal{X} is smooth and π\pi is a holomorphic proper submersion. Moreover suppose for any sSs\in S, the fiber Xs:=π1(s)X_{s}:=\pi^{-1}(s) is a smooth projective variety of fixed dimension dimXs=l\mathrm{dim}_{\mathbb{C}}X_{s}=l. Denote 𝒳\mathbb{Z}_{\mathcal{X}} as the \mathbb{Z}-constant sheaf on 𝒳\mathcal{X}, then

(2.4) 𝕍:=Rlπ(𝒳)\mathbb{V}:=R^{l}\pi_{*}(\mathbb{Z}_{\mathcal{X}})

is a \mathbb{Z}-local system over SS whose fiber at sSs\in S is H:=Hl(Xs,)/torsionH:=H^{l}(X_{s},\mathbb{Z})/\mathrm{torsion}. Take the Hodge decomposition

(2.5) H=H:=(Hl(Xs,)/torsion)=p+q=lHp,q(Xs)H_{\mathbb{C}}=H:=(H^{l}(X_{s},\mathbb{Z})/\mathrm{torsion})\otimes\mathbb{C}=\oplus_{p+q=l}H^{p,q}(X_{s})

on each fiber into consideration, this gives a variation of Hodge structure of weight ll on SS as well as a period map of the form (2.2).

For any sS¯s\in\overline{S}, a local neighborhood around sUS¯s\in U\subset\overline{S} satisfies US(Δ)k×ΔnkU\cap S\cong(\Delta^{*})^{k}\times\Delta^{n-k} on which we may consider the local monodromy operators {Ti}1ik\{T_{i}\}_{1\leq i\leq k} and their logarithms. One of the main ingredients is the following intepretation of Schmid’s nilpotent orbit theorem:

Theorem 2.10.

For every sS¯s\in\overline{S}, up to the action of Γ\Gamma there is a nilpotent orbit (σs,Fs)(\sigma_{s},F_{s}) canonically associated to ss. Here σs\sigma_{s} is the local monodromy nilpotent cone and FsDˇF_{s}\in\check{D}.

There are two ways to intepretate nilpotent orbit (σs,Fs)(\sigma_{s},F_{s}):

  1. (1)

    Suppose σs=N1,,Nr\sigma_{s}=\langle N_{1},...,N_{r}\rangle, for local coordinates {ti}\{t_{i}\} around ss, the map ψ(ti)=exp(log(ti)2πiNi)Fs\psi(t_{i})=\exp(\sum-\frac{\log(t_{i})}{2\pi i}N_{i})F_{s} approximates the local lift of φ\varphi at ss ([Sch73, Sec. 4]).

  2. (2)

    By [CK82], for every sS¯s\in\overline{S} there is a monodromy weight filtration Ws:=W(σs)W_{s}:=W(\sigma_{s}) associated to ss, such that (Ws,Fs,σs)(W_{s},F_{s},\sigma_{s}) is a limiting mixed Hodge structure (LMHS).

We give a better interpretation of (1) above. Suppose US¯U\subset\overline{S} is an open subset such that US(Δ)k×ΔlU\cap S\cong(\Delta^{*})^{k}\times\Delta^{l}. Let \mathfrak{H} be the upper-half plane {𝔪(z)>0}\{\mathfrak{Im}(z)>0\}. The local period map Φ\Phi on USU\cap S has the local lift:

(2.6) k×Δl{\mathfrak{H}^{k}\times\Delta^{l}}D{D}(Δ)k×Δl{(\Delta^{*})^{k}\times\Delta^{l}}Γ\D{\Gamma\backslash D}Φ~\scriptstyle{\tilde{\Phi}}Φ\scriptstyle{\Phi}

If we denote z=(zk),w=(wl)z=(z_{k}),w=(w_{l}) as coordinates on k\mathfrak{H}^{k}, Δl\Delta^{l},

(2.7) Φ~=exp(1jkzjNj)ψ(e2πiz,w),\tilde{\Phi}=\exp(\sum_{1\leq j\leq k}z_{j}N_{j})\psi(e^{2\pi iz},w),

where ψ(z,w)D\psi(z,w)\in D is holomorphic over Δk+l\Delta^{k+l}. Schmid’s nilpotent orbit theorem [Sch73, Thm. 4.12] also says:

Theorem 2.11.

For any wΔlw\in\Delta^{l}, (σ:=N1,,Nk,ψ(0,w))(\sigma:=\langle N_{1},...,N_{k}\rangle,\psi(0,w)) is a nilpotent orbit, and under the canonical metric dd on DD, as 𝔪(z)\mathfrak{Im}(z)\rightarrow\infty,

(2.8) d(exp(1jkzjNj)ψ(0,w),Φ~(e2πiz,w))eb𝔪(z)d(\exp(\sum_{1\leq j\leq k}z_{j}N_{j})\psi(0,w),\tilde{\Phi}(e^{2\pi iz},w))\sim e^{-b\mathfrak{Im}(z)}

for some b>0b>0.

For a general mixed Hodge structure (W,F)(W,F) on (H,Q)(H_{\mathbb{Z}},Q), there is a canonical splitting by Deligne:

(2.9) H=p,qHp,qH_{\mathbb{C}}=\oplus_{p,q}H^{p,q}

such that WlH=p+qlHp,qW_{l}H_{\mathbb{C}}=\oplus_{p+q\leq l}H^{p,q} and FlH=plHp,qF^{l}H_{\mathbb{C}}=\oplus_{p\geq l}H^{p,q}. We have the following definitions:

Definition 2.12.

The natural numbers {hp,q:=dimHp,q}\{h^{p,q}:=\mathrm{dim}H^{p,q}\} are called the Hodge numbers of the mixed Hodge structure. We say two mixed Hodge structures on (H,Q)(H_{\mathbb{Z}},Q) have the same type if they have the same Hodge numbers. In particular, we say a mixed Hodge structure (W,F)(W,F) has the type given by {hp,q}\{h^{p,q}\} if {hp,q}\{h^{p,q}\} are Hodge numbers of it.

Remark 2.13.

When σ=N1,N2\sigma=\langle N_{1},N_{2}\rangle and FDˇF\in\check{D} such that (σ,F)(\sigma,F) is a nilpotent orbit, we say (σ,F)(\sigma,F) is of type A|B|C\langle A|B|C\rangle if (N1,F),(σ,F),(N2,F)(N_{1},F),(\sigma,F),(N_{2},F) are of type A,B,CA,B,C correspondingly.

Therefore, the period map (2.2) associates each sS¯s\in\overline{S} a exp(σs,)Γ\exp(\sigma_{s,\mathbb{C}})\Gamma-class of LMHS (Ws,Fs,σs)(W_{s},F_{s},\sigma_{s}) which gives a type of LMHS. It is known that the type of LMHS at all sS¯s\in\overline{S} is constant along each stratum.

We refer to [KPR19] for classifying all possible Hodge diamonds of LMHS. For the sake of convenience, we list all possible LMHS types for the weight 33 period domain of Hodge type (1,r,r,1)(1,r,r,1), r1r\geq 1 in terms of Hodge-Deligne diagrams (See also [KPR19, Example 5.8]). In particular, the type IVr\mathrm{IV}_{r} LMHS is also called the Hodge-Tate type degeneration. The following definition from [CK99] will be useful in the rest of this paper.

Definition 2.14 (MUM point).

Let sS¯Ss\in\overline{S}-S. Then ss is a maximal unipotent monodromy (MUM) point if the following is true

  1. (1)

    sD1Drs\in D_{1}\cap\dots\cap D_{r}, where DiS¯SD_{i}\subset\overline{S}-S are distinct irreducible divisors of S¯\overline{S}.

  2. (2)

    Denote the monodromy operator associated to DiD_{i} as TiT_{i}. Then TiT_{i} are unipotent.

  3. (3)

    Denote Ni:=log(Ti)N_{i}:=\log(T_{i}), and the cone formed by them as σs\sigma_{s}. The LMHS (Ws,Fs,σs)(W_{s},F_{s},\sigma_{s}) associated to ss is Hodge-Tate.

  4. (4)

    Let e0e_{0} be the generator of (Ws)0(W_{s})_{0}, and e0,e1,,ere_{0},e_{1},\dots,e_{r} form a basis of (Ws)2(W_{s})_{2}. Denote (mij)(m_{ij}) as the r×rr\times r matrix defined via Niej=mije0N_{i}e_{j}=m_{ij}e_{0}. Then (mij)(m_{ij}) is invertible.

11rar-aaa11rar-aaa

          11aarar-a11rar-aaa

Figure 1. Type Ia\mathrm{I}_{a} and IVa\mathrm{IV}_{a} LMHS for 0ar0\leq a\leq r
1111aabb1111bbaa

          1111aabb1111bbaa

Figure 2. Type IIa\mathrm{II}_{a} and IIIa\mathrm{III}_{a} LMHS for a+b=h1a+b=h-1

3. Generic degree of two-parameter period mappings

In this section we assume SS is an algebraic surface admitting a projective compactification S^\hat{S} which is smooth and S^S\hat{S}-S is a normal crossing divisor. 𝕍S\mathbb{V}\rightarrow S is a PVHS with associated period map Φ:SΓ\D\Phi:S\rightarrow\Gamma\backslash D. We also assume the monodromy operator arouond each irreducible boundary divisor is unipotent which can always be done after a finite base change.

The main theorem of [DR23] provides a way to realize Φ\Phi as the restriction of a morphism between compact complex analytic spaces:

Theorem 3.1.

Suppose Γ\Gamma is neat, there exists a smooth compactification S¯S\overline{S}\supset S with simple normal crossing divisor S=S¯\S\partial S=\overline{S}\backslash S, and a logarithmic manifold Γ\DΣ\Gamma\backslash D_{\Sigma} parameterizing Γ\Gamma–conjugacy classes of nilpotent orbits on DD so that Γ\DΓ\DΣ\Gamma\backslash D\subset\Gamma\backslash D_{\Sigma} and the period map extends to a morphism ΦΣ:S¯Γ\DΣ\Phi_{\Sigma}:\overline{S}\to\Gamma\backslash D_{\Sigma} of logarithmic manifolds. The image ΦΣ(S¯)\Phi_{\Sigma}(\overline{S}) is a compact algebraic space.

Remark 3.2.

More precisely, S¯\overline{S} can be obtained by a (finite) sequence of blow-ups of S^\hat{S} along codimensional 22 boundary strata. In general for any given SS and S^\hat{S}, it is unrealistic to identify S¯\overline{S} exactly.

The image of ΦΣ(s)\Phi_{\Sigma}(s) for sS¯Ss\in\overline{S}-S can be described as follows. Let Σ\Sigma be the corresponding weak fan, (σs,Fs)(\sigma_{s},F_{s}) mod Γ\Gamma be the nilpotent orbit associated to ss by Φ\Phi via Schmid’s nilpotent orbit theorem. There is a unique minimal τsΣ\tau_{s}\in\Sigma such that σsτs\sigma_{s}\subset\tau_{s} and (τs,Fs)(\tau_{s},F_{s}) is a nilpotent orbit.

Let :=Img(Φ)Γ\D\wp:=\mathrm{Img}(\Phi)\subset\Gamma\backslash D, it is well-known that \wp is a complex analytic space (and quasi-projective by [BBT22]). Moreover, Theorem 3.1 implies there exists a compact complex analytic space ΣΓ\DΣ\wp_{\Sigma}\subset\Gamma\backslash D_{\Sigma} such that we have the following diagram in the category of complex analytic spaces:

(3.1) S{S}{\wp}S¯{\overline{S}}Σ{\wp_{\Sigma}}Φ\scriptstyle{\Phi}ΦΣ\scriptstyle{\Phi_{\Sigma}}

3.1. A degree-computing formula

Suppose \wp has dimension 22, then we have a well-defined degree for Φ\Phi and ΦΣ\Phi_{\Sigma} defined to be Card{Φ1(p)}\mathrm{Card}\{\Phi^{-1}(p)\} for a generic pp\in\wp, and deg(Φ)=deg(ΦΣ)\mathrm{deg}(\Phi)=\mathrm{deg}(\Phi_{\Sigma}). The advantage of computing deg(ΦΣ)\mathrm{deg}(\Phi_{\Sigma}) instead of deg(Φ)\mathrm{deg}(\Phi) is that we will be able to look at special boundary points.

Proposition 3.3.

Suppose sSs\in\partial S with σs\sigma_{s} has dimension 22, and moreover ss is not a branch point for ΦΣ\Phi_{\Sigma}, then deg(Φ)=deg(ΦΣ)\mathrm{deg}(\Phi)=\mathrm{deg}(\Phi_{\Sigma}) equals to Card{Φ1(Φ(s))}\mathrm{Card}\{\Phi^{-1}(\Phi(s))\}.

In general Γ\Gamma is not neat, and we will need to pass to a neat subgroup ΓΓ\Gamma^{{}^{\prime}}\leq\Gamma of finite index. There exists a finite analytic covering map φ:SS\varphi:S^{{}^{\prime}}\rightarrow S such that

(3.2) π1(S){\pi_{1}(S^{{}^{\prime}})}Γ{\Gamma^{{}^{\prime}}}π1(S){\pi_{1}(S)}Γ{\Gamma}ρ\scriptstyle{\rho}ρ\scriptstyle{\rho}

where ρ\rho is the monodromy representation. Lift the original PVHS to SS^{{}^{\prime}}:

(3.3) S{S^{{}^{\prime}}}Γ\D{\Gamma^{{}^{\prime}}\backslash D}S{S}Γ\D{\Gamma\backslash D}φ\scriptstyle{\varphi}Φ\scriptstyle{\Phi^{{}^{\prime}}}Φ\scriptstyle{\Phi}

Let :=Img(Φ)\wp^{{}^{\prime}}:=\mathrm{Img}(\Phi^{{}^{\prime}}). Since both S𝜑SS^{{}^{\prime}}\xrightarrow{\varphi}S and \wp^{{}^{\prime}}\rightarrow\wp have the same degree =[Γ:Γ]=[\Gamma:\Gamma^{{}^{\prime}}], deg(Φ)=deg(Φ)\mathrm{deg}(\Phi^{{}^{\prime}})=\mathrm{deg}(\Phi). In other words, passing Γ\Gamma to a finite-index neat subgroup does not change the generic degree.

Take any sS¯Ss\in\overline{S}-S with dim(σs)=2\mathrm{dim}(\sigma_{s})=2. Suppose for any tst\neq s, σsσt\sigma_{s}\neq\sigma_{t} mod Γ\Gamma. Take a small neighborhood sUS¯s\in U\subset\overline{S} with US(Δ)2U\cap S\simeq(\Delta^{*})^{2}. The restricted period map ΦUS\Phi_{U\cap S} can be decomposed as:

(3.4) USΦsΓs\DφsΓ\DU\cap S\xrightarrow{\Phi_{s}}\Gamma_{s}\backslash D\xrightarrow{\varphi_{s}}\Gamma\backslash D

where Γs\Gamma_{s} is generated by the local monodromy around ss and φs\varphi_{s} is the projection map.

Let s:=Φs(US)\wp_{s}:=\Phi_{s}(U\cap S). Since σs\sigma_{s} is non-degenerate, possibly after shrinking UU we may assume Φs\Phi_{s} and φs|s\varphi_{s}|_{\wp_{s}} are both proper morphisms between complex analytic spaces such that the (generic) degree of them are well-defined.

Proposition 3.4.

deg(Φ)=deg(Φs)deg(φs|s)\mathrm{deg}(\Phi)=\mathrm{deg}(\Phi_{s})\cdot\mathrm{deg}(\varphi_{s}|_{\wp_{s}}),

Proof.

This follows immediately from the sequence 3.4 and our assumption on σs\sigma_{s}. ∎

Note that by [KU08], Φs\Phi_{s} admits a Kato-Usui type completion

(3.5) Φs¯:UΓs\Ds.\overline{\Phi_{s}}:U\rightarrow\Gamma_{s}\backslash D_{s}.

The assumption dim(σs)=2\mathrm{dim}(\sigma_{s})=2 implies Φs\Phi_{s} is proper up to a possible shrinking of UU, therefore by [Usu06], s¯:=Φs¯(U)\overline{\wp_{s}}:=\overline{\Phi_{s}}(U) admits a structure of complex analytic space. Therefore we also have:

Proposition 3.5.

deg(Φ)=deg(Φs¯)deg(φs¯|s¯)\mathrm{deg}(\Phi)=\mathrm{deg}(\overline{\Phi_{s}})\cdot\mathrm{deg}(\overline{\varphi_{s}}|_{\overline{\wp_{s}}}),

Remark 3.6.

In 3.5, ss is chosen in the unknown space S¯\overline{S}. However, since S¯\overline{S} and S^\hat{S} are birationally equivalent, we may apply Proposition 3.5 on the original space S^\hat{S}.

To use the formula, we need to compute the two degrees on the right hand side separately. deg(Φs¯)\mathrm{deg}(\overline{\Phi_{s}}) will be computed via coordinate interpretation and we will do it for specific examples in Section 4-5. In the next subsection we show how to find deg(φs¯|s¯)\mathrm{deg}(\overline{\varphi_{s}}|_{\overline{\wp_{s}}}).

3.2. The local-to-global map for Kato-Nakayama-Usui spaces

The following two lemmas are critical to our argument:

Lemma 3.7 ([KU08], Prop. 7.4.3).

Suppose ΓG\Gamma\leq G_{\mathbb{Z}} is neat and (σ,F)(\sigma,F) is a nilpotent orbit. If γΓ\gamma\in\Gamma satisfies γ(σ,F)=(σ,F)\gamma\cdot(\sigma,F)=(\sigma,F), then γΓσ:=exp(σ)Γ\gamma\in\Gamma_{\sigma}:=\exp(\sigma_{\mathbb{C}})\cap\Gamma.

Lemma 3.8 ([KU08], Thm. A(iv)).

Assume Γ\Gamma is neat, then Γσ\DσΓ\DΣ\Gamma_{\sigma}\backslash D_{\sigma}\hookrightarrow\Gamma\backslash D_{\Sigma} is a local homeomorphism.

These two lemmas and the diagram (3.3) imply that to calculate deg(φs¯|s¯)\mathrm{deg}(\overline{\varphi_{s}}|_{\overline{\wp_{s}}}) for non-neat Γ\Gamma (with a finite-index neat subgroup Γ\Gamma^{{}^{\prime}} chosen), it is enough to consider the set

Γσ,tor:={γΓΓ,γ(σs,Fs)=(σs,Fs),γnΓσfor somen}\Gamma_{\sigma,\mathrm{tor}}:=\{\gamma\in\Gamma-\Gamma^{{}^{\prime}},\ \gamma(\sigma_{s},F_{s})=(\sigma_{s},F_{s}),\ \gamma^{n}\in\Gamma_{\sigma}\ \text{for some}\ n\in\mathbb{Z}\}

and its induced automorphism group on the nilpotent orbit (σs,Fs)(\sigma_{s},F_{s}).

Suppose σs=N1,N2\sigma_{s}=\langle N_{1},N_{2}\rangle. Any finite-order automorphism η\eta of the nilpotent orbit (σs,Fs)(\sigma_{s},F_{s}) is a combination of the following actions:

  1. (1)

    Rescaling NiN_{i} by a root of unity, and fixes some chosen base point FsF_{s};

  2. (2)

    Permutes N1N_{1} and N2N_{2}, and fixes some chosen base point FsF_{s}.

In the case η=Adμ\eta=\mathrm{Ad}_{\mu} for some μΓσ,tor\mu\in\Gamma_{\sigma,\mathrm{tor}}, it must not rescale NiN_{i} by a root of unity other than 11 because of rationality and positivity. Therefore, the only case η\eta could be non-trivial is η\eta flips the cone. Combining with the fact that Adμ\mathrm{Ad}_{\mu} does not change the LMHS type, we have:

Proposition 3.9.

deg(φs¯|s¯)=1\mathrm{deg}(\overline{\varphi_{s}}|_{\overline{\wp_{s}}})=1 or 22. It is 22 if and only if there exists μΓ\mu\in\Gamma whose adjoint action on (σs,Fs)(\sigma_{s},F_{s}) preserves the nilpotent orbit but flips the boundary of the cone. In particular, if (σs,Fs)(\sigma_{s},F_{s}) has LMHS type A|B|C\langle A|B|C\rangle with ACA\neq C, we must have deg(φs¯|s¯)=1\mathrm{deg}(\overline{\varphi_{s}}|_{\overline{\wp_{s}}})=1.

Remark 3.10.

The same arguments show for a nilpotent orbit (σ,F)(\sigma,F) with dim(σ)=n\mathrm{dim}(\sigma)=n, its automorphism group induced from any arithmetic group ΓG\Gamma\in G_{\mathbb{Q}} is isomorphic to a subgroup of SnS_{n}.

4. Examples: Calabi-Yau 3-folds in toric variety

In this section, we will study the geometric VHS coming from the tautological family of Calabi-Yau hypersurface inside the toric variety over the simplified moduli simp\mathcal{M}_{\mathrm{simp}}. The results in this section will be used to compute the generic degree of the period maps

Φsimp:simpΓ\D.\Phi_{\mathrm{simp}}:\mathcal{M}_{\mathrm{simp}}\rightarrow\mathrm{\Gamma}\backslash D.

Moreover, we will also compute the generic degree of the map

ϕ:simp.\phi:\mathcal{M}_{\mathrm{simp}}\rightarrow\mathcal{M}.

Then the generic Torelli theorem for the Calabi-Yau hypersurface in question will be established if

deg(Φsimp)=deg(ϕ).\deg(\Phi_{\mathrm{simp}})=\deg(\phi).

The subsections will be named after the Hodge type of the Calabi-Yau 3-folds, namely in subsection V(a,b)V_{(a,b)}, we will study a Calabi-Yau 3-fold with Hodge numbers (h2,1,h1,1)=(a,b)(h^{2,1},h^{1,1})=(a,b). These Hodge numbers can be easily computed from the toric data by [Bat94].

4.1. V(2,29)V_{(2,29)}

We study the smooth Calabi-Yau 3-folds VV with Hodge numbers (h2,1,h1,1)=(2,29)(h^{2,1},h^{1,1})=(2,29), which is the anti-canonical hypersurface of the toric variety XΣΔX_{\Sigma^{\circ}}\cong\mathbb{P}_{\Delta^{\circ}}. The toric data for Δ\Delta^{\circ} and Σ\Sigma is

(4.1) Ξv1v2v3v4v5v6r1100012r2010111r3001122r4000033w1100011w2011100.\begin{array}[]{c|cccccc}\Xi&v_{1}&v_{2}&v_{3}&v_{4}&v_{5}&v_{6}\\ \hline\cr r_{1}&1&0&0&0&1&-2\\ r_{2}&0&1&0&-1&1&-1\\ r_{3}&0&0&1&-1&2&-2\\ r_{4}&0&0&0&0&3&-3\\ \hline\cr w_{1}&1&0&0&0&1&1\\ w_{2}&0&1&1&1&0&0\end{array}.

The polytope Δ\Delta^{\circ} can be visualized as follows: Ξ\Xi is naturally divided into two sets {v1,v5,v6}\{v_{1},v_{5},v_{6}\} and {v2,v3,v4}\{v_{2},v_{3},v_{4}\} where the points in both sets lie in the same plane, the 9 facets of Δ\Delta then come from choosing two points from each set. After suspension, the toric data becomes

(4.2) Ξ¯v¯1v¯2v¯3v¯4v¯5v¯6v¯7r¯11000120r¯20101110r¯30011220r¯40000330r¯51111111w¯11000113w¯20111003.\begin{array}[]{c|ccccccc}\bar{\Xi}&\bar{v}_{1}&\bar{v}_{2}&\bar{v}_{3}&\bar{v}_{4}&\bar{v}_{5}&\bar{v}_{6}&\bar{v}_{7}\\ \hline\cr\bar{r}_{1}&1&0&0&0&1&-2&0\\ \bar{r}_{2}&0&1&0&-1&1&-1&0\\ \bar{r}_{3}&0&0&1&-1&2&-2&0\\ \bar{r}_{4}&0&0&0&0&3&-3&0\\ \bar{r}_{5}&1&1&1&1&1&1&1\\ \hline\cr\bar{w}_{1}&1&0&0&0&1&1&-3\\ \bar{w}_{2}&0&1&1&1&0&0&-3\end{array}.

4.1.1. Moduli spaces

The toric data shows that (ΔN)0=ΔN(\Delta^{\circ}\cap N)_{0}=\Delta^{\circ}\cap N, i.e., there is no integral point lies in the interior of codimension 2 face. Therefore, we have =poly\mathcal{M}=\mathcal{M}_{\mathrm{poly}}.

From the multi-degree, it is clear that Δ\mathbb{P}_{\Delta^{\circ}} has no root symmetry. Then the generic degree of the map ϕ:polysimp\phi:\mathcal{M}_{\mathrm{poly}}\rightarrow\mathcal{M}_{\mathrm{simp}} is |Aut(Σ)/Autt(Σ)||\mathrm{Aut}(\Sigma)/\mathrm{Aut}^{t}(\Sigma)|, where Autt(Σ)\mathrm{Aut}^{t}(\Sigma) is the normal subgroup of Aut(Σ)\mathrm{Aut}(\Sigma) that acts trivially on poly\mathcal{M}_{\mathrm{poly}}.

Denote Aut(Ξ)\mathrm{Aut}(\Xi) as the subgroup of Aut(N)\mathrm{Aut}(N) that permutes Ξ\Xi, then a direct calculation shows Aut(Ξ)=C2×C3S3\mathrm{Aut}(\Xi)=C_{2}\times C_{3}\rtimes S_{3}, which acts on Ξ\Xi as follows: the C2C_{2} factor determines whether the permutation is even or odd, the set {v2,v3,v4}\{v_{2},v_{3},v_{4}\} is sent to {v2,v3,v4}\{v_{2},v_{3},v_{4}\} or {v1,v5,v6}\{v_{1},v_{5},v_{6}\} respectively for the two cases. Then the S3S_{3} factor determines the image of v2,v3,v4v_{2},v_{3},v_{4} under the automorphism, and since the parity of the permutation is already determined, there are only C3C_{3} choices for the image of v1,v5,v6v_{1},v_{5},v_{6}. From this description, it is also clear that every element in Aut(Ξ)\mathrm{Aut}(\Xi) permits the codimension-1 cones of Σ\Sigma, thus we have Aut(Σ)=Aut(Ξ)\mathrm{Aut}(\Sigma)=\mathrm{Aut}(\Xi). As a subgroup of S6S_{6} which acts on Ξ\Xi by permutation, we have Aut(Σ)\mathrm{Aut}(\Sigma) is generated by the permutations a:=(234),b:=(34)(56),c:=(12)(35)(46)a:=(234),b:=(34)(56),c:=(12)(35)(46).

A generic point in (L(ΔN))\mathbb{P}(L(\Delta^{\circ}\cap N)) is given by

f=λ1t1+λ2t2+λ3t3+λ4t21t31+λ5t1t2t32t43+λ6t12t21t32t43+λ7,f=\lambda_{1}t_{1}+\lambda_{2}t_{2}+\lambda_{3}t_{3}+\lambda_{4}t^{-1}_{2}t_{3}^{-1}+\lambda_{5}t_{1}t_{2}t_{3}^{2}t_{4}^{3}+\lambda_{6}t_{1}^{-2}t_{2}^{-1}t_{3}^{-2}t_{4}^{-3}+\lambda_{7},

Then we have

af=λ1t1+λ3t2+λ4t3+λ2t21t31+λ5t1t2t32t43+λ6t12t21t32t43+λ7,a\cdot f=\lambda_{1}t_{1}+\lambda_{3}t_{2}+\lambda_{4}t_{3}+\lambda_{2}t^{-1}_{2}t_{3}^{-1}+\lambda_{5}t_{1}t_{2}t_{3}^{2}t_{4}^{3}+\lambda_{6}t_{1}^{-2}t_{2}^{-1}t_{3}^{-2}t_{4}^{-3}+\lambda_{7},

which can be realized as a TT-action

(t1,t2,t3,t4)(t1,λ21λ3t2,λ31λ4t3,λ21/3λ31/3λ42/3t4),(t_{1},t_{2},t_{3},t_{4})\rightarrow(t_{1},\lambda_{2}^{-1}\lambda_{3}t_{2},\lambda_{3}^{-1}\lambda_{4}t_{3},\lambda_{2}^{1/3}\lambda_{3}^{1/3}\lambda^{-2/3}_{4}t_{4}),

similarly, bb can be realized as

(t1,t2,t3,t4)(t1,t2,λ31λ4t3,λ32/3λ42/3λ51/3λ61/3t4).(t_{1},t_{2},t_{3},t_{4})\rightarrow(t_{1},t_{2},\lambda_{3}^{-1}\lambda_{4}t_{3},\lambda_{3}^{2/3}\lambda^{-2/3}_{4}\lambda_{5}^{-1/3}\lambda_{6}^{1/3}t_{4}).

However, cc can not be realized as a TT-action, therefore, we have Autt(Σ)=C3S3\mathrm{Aut}^{t}(\Sigma)=C_{3}\rtimes S_{3}, and the generic degree of ϕ\phi is

degg(ϕ)=|Aut(Σ)/Autt(Σ)|=2.\deg_{g}(\phi)=|\mathrm{Aut}(\Sigma)/\mathrm{Aut}^{t}(\Sigma)|=2.

The toric data for the secondary fan Σs\Sigma^{s} is

(4.3) Ξsv1sv2sv3sr1s101r2s011w1s111.\begin{array}[]{c|ccc}\Xi^{s}&v_{1}^{s}&v_{2}^{s}&v_{3}^{s}\\ \hline\cr r_{1}^{s}&1&0&-1\\ r_{2}^{s}&0&1&-1\\ \hline\cr w_{1}^{s}&1&1&1\\ \end{array}.

Therefore, ¯simp=XΣs2=Proj[Z1,Z2,Z3]\overline{\mathcal{M}}_{\mathrm{simp}}=X_{\Sigma^{s}}\cong\mathbb{P}^{2}=\mathrm{Proj}\mathbb{C}[Z_{1},Z_{2},Z_{3}], and the /2\mathbb{Z}/2\mathbb{Z}-action extend to ¯simp\overline{\mathcal{M}}_{\mathrm{simp}} as Z1Z2Z_{1}\leftrightarrow Z_{2} that ramifies at Z1=Z2Z_{1}=Z_{2}.

4.1.2. Picard-Fuchs system

By Lemma 2.7, the GKZ system is generated by

P1:=δ13+z1(3δ1+3δ2+1)(3δ1+3δ2+2)(3δ1+3δ2+3),\displaystyle P_{1}:=\delta_{1}^{3}+z_{1}(3\delta_{1}+3\delta_{2}+1)(3\delta_{1}+3\delta_{2}+2)(3\delta_{1}+3\delta_{2}+3),
P2:=δ23+z2(3δ1+3δ2+1)(3δ1+3δ2+2)(3δ1+3δ2+3).\displaystyle P_{2}:=\delta_{2}^{3}+z_{2}(3\delta_{1}+3\delta_{2}+1)(3\delta_{1}+3\delta_{2}+2)(3\delta_{1}+3\delta_{2}+3).

The canonical affine chart is Spec[z1,z2]\mathrm{Spec}[z_{1},z_{2}], with

z1=λ1λ5λ6λ72,z2=λ2λ3λ4λ72,z_{1}=\frac{\lambda_{1}\lambda_{5}\lambda_{6}}{\lambda_{7}^{2}},\quad z_{2}=\frac{\lambda_{2}\lambda_{3}\lambda_{4}}{\lambda_{7}^{2}},

then Dv1s=𝕍(z1)D_{v_{1}^{s}}=\mathbb{V}(z_{1}), Dv1s=𝕍(z2)D_{v_{1^{s}}}=\mathbb{V}(z_{2}). In the projective coordinates, this is the chart with Z30Z_{3}\neq 0, and zi=XiX3,i=1,2z_{i}=\frac{X_{i}}{X_{3}},i=1,2.

By Lemma, 2.9 and the comment after (4.6), the Picard-Fuchs system on simp\mathcal{M}_{\mathrm{simp}} is generated by the two differential operators

P1:=δ13+z1(3δ1+3δ2+1)(3δ1+3δ2+2)(3δ1+3δ2+3),\displaystyle P_{1}:=\delta_{1}^{3}+z_{1}(3\delta_{1}+3\delta_{2}+1)(3\delta_{1}+3\delta_{2}+2)(3\delta_{1}+3\delta_{2}+3),
P2:=(δ12δ1δ2+δ22)+3(z1+z2)(3δ1+3δ2+1)(3δ1+3δ2+2).\displaystyle P_{2}:=(\delta_{1}^{2}-\delta_{1}\delta_{2}+\delta_{2}^{2})+3(z_{1}+z_{2})(3\delta_{1}+3\delta_{2}+1)(3\delta_{1}+3\delta_{2}+2).

For simplicity, we rescale the local coordinates as z1,2:=33z1,2z_{1,2}^{\prime}:=3^{3}\cdot z_{1,2}, and rename z1,2z_{1,2}^{\prime} as the new z1,2z_{1,2}. The Picard-Fuchs operators then become

(4.4) P1:=δ13+z1(δ1+δ2+13)(δ1+δ2+23)(δ1+δ2+1),\displaystyle P_{1}:=\delta_{1}^{3}+z_{1}(\delta_{1}+\delta_{2}+\frac{1}{3})(\delta_{1}+\delta_{2}+\frac{2}{3})(\delta_{1}+\delta_{2}+1),
P2:=(δ12δ1δ2+δ22)+(z1+z2)(δ1+δ2+13)(δ1+δ2+23).\displaystyle P_{2}:=(\delta_{1}^{2}-\delta_{1}\delta_{2}+\delta_{2}^{2})+(z_{1}+z_{2})(\delta_{1}+\delta_{2}+\frac{1}{3})(\delta_{1}+\delta_{2}+\frac{2}{3}).

The Picard-Fuchs ideal sheaf is denoted as II, which in the canonical affine affine chart is generated by P1P_{1} and P2P_{2}. The associated 𝒟\mathcal{D}-module is /2\mathbb{Z}/2\mathbb{Z}-equivariant under z1z2z_{1}\leftrightarrow z_{2}, 12\partial_{1}\leftrightarrow\partial_{2}. More precisely, P2P_{2} is invariant under the /2\mathbb{Z}/2\mathbb{Z}-action, and P1P_{1} changes to (δ1+δ2)P2P1(\delta_{1}+\delta_{2})P_{2}-P_{1}, so II is /2\mathbb{Z}/2\mathbb{Z}-invariant.

4.1.3. Discriminant locus

The principal AA-determinant EAE_{A} is irreducible, and is thus equals to the AA-discriminant DAD_{A}

DA=\displaystyle D_{A}= z13+3z12z2+3z12+3z1z2221z1z2+3z1+z23+3z22+3z2+1\displaystyle z_{1}^{3}+3z_{1}^{2}z_{2}+3z_{1}^{2}+3z_{1}z_{2}^{2}-21z_{1}z_{2}+3z_{1}+z_{2}^{3}+3z_{2}^{2}+3z_{2}+1
=\displaystyle= (z1+z2+1)327z1z2\displaystyle(z_{1}+z_{2}+1)^{3}-27z_{1}z_{2}

so the discriminant locus is contained inside

Disc=DA(B¯B)=DADv1sDv2sDv3s.Disc=D_{A}\cup(\bar{B}-B)=D_{A}\cup D_{v_{1}^{s}}\cup D_{v_{2}^{s}}\cup D_{v_{3}^{s}}.

In the canonical affine chart DAD_{A} intersect Dv1sD_{v_{1}^{s}} at (1,0)(-1,0) with 3-tangency, and by the /2\mathbb{Z}/2\mathbb{Z} symmetry, DAD_{A} intersect Dv2sD_{v_{2}^{s}} at (0,1)(0,-1) with 3-tangency.

In order to make DiscDisc only have normal crossing singularities, we need to blow-up these tangencies 3 times each. Since we will only be interested in the local monodromy, we only need to study the geometry at z1=0,z2=1z_{1}=0,z_{2}=-1. Denote the exceptional divisors of the three blow-ups as E1,E2,E3E_{1},E_{2},E_{3}, then we want to find 4 Picard-Fuchs ideals corresponding to the normal crossings

Dv1E3,E1E2,E2E3DAE3,D_{v_{1}}\cap E_{3},\quad E_{1}\cap E_{2},\quad E_{2}\cap E_{3}\quad D_{A}\cap E_{3},

First, we make a change of coordinate of (4.4) by z1z1z_{1}\rightarrow z_{1}, z2z21z_{2}\rightarrow z_{2}-1. Then the local affine coordinates systems for each of the normal crossing is given by

(s,t)=(z1z23,z2),\displaystyle(s,t)=\left(\frac{z_{1}}{z_{2}^{3}},z_{2}\right),\quad (s,t)=(z22z1,z1z2),\displaystyle(s,t)=\left(\frac{z_{2}^{2}}{z_{1}},\frac{z_{1}}{z_{2}}\right),
(s,t)=(z23z1,z1z22),\displaystyle(s,t)=\left(\frac{z_{2}^{3}}{z_{1}},\frac{z_{1}}{z_{2}^{2}}\right),\quad (s,t)=(z1z23127,z2)\displaystyle(s,t)=\left(\frac{z_{1}}{z_{2}^{3}}-\frac{1}{27},z_{2}\right)

With s=0s=0 gives the left divisor and t=0t=0 gives the right divisor.

4.1.4. Gauss-Manin connection and nilpotent cones

By the local Torrelli theorem of Calabi-Yau manifolds, we can take a global multi-valued frame

(4.5) w=(Ω,δ1Ω,δ2Ω,δ1δ2Ω,δ12Ω,δ12δ2Ω).w=(\Omega,\delta_{1}\Omega,\delta_{2}\Omega,\delta_{1}\delta_{2}\Omega,\delta_{1}^{2}\Omega,\delta_{1}^{2}\delta_{2}\Omega).

The Gauss-Manin connection can be easily calculated from the Picard-Fuchs ideal by finding the

δ1w=w[0000z1h1,51z1h1,611000z1h2,51z1h2,610000z1h3,51z1h3,610010z1h4,51z1h4,610100z1h5,51z1h5,610001z1h6,51z1h6,61]=:wR1,δ2w=w[00z2h1,32z2h1,420z2h1,6200z2h2,32z2h2,420z2h2,6210z2h3,32z2h3,420z2h3,62011+z2h4,32z2h4,420z2h4,62001z2h4,420z2h5,620001+z2h5,421z2h6,62]=:wR2,\begin{split}&\nabla_{\delta_{1}}w=w\left[\begin{array}[]{rrrrrr}0&0&0&0&z_{1}h^{1}_{1,5}&z_{1}h^{1}_{1,6}\\ 1&0&0&0&z_{1}h^{1}_{2,5}&z_{1}h^{1}_{2,6}\\ 0&0&0&0&z_{1}h^{1}_{3,5}&z_{1}h^{1}_{3,6}\\ 0&0&1&0&z_{1}h^{1}_{4,5}&z_{1}h^{1}_{4,6}\\ 0&1&0&0&z_{1}h^{1}_{5,5}&z_{1}h^{1}_{5,6}\\ 0&0&0&1&z_{1}h^{1}_{6,5}&z_{1}h^{1}_{6,6}\end{array}\right]=:w\cdot R_{1},\\ &\nabla_{\delta_{2}}w=w\left[\begin{array}[]{rrrrrr}0&0&z_{2}h^{2}_{1,3}&z_{2}h^{2}_{1,4}&0&z_{2}h^{2}_{1,6}\\ 0&0&z_{2}h^{2}_{2,3}&z_{2}h^{2}_{2,4}&0&z_{2}h^{2}_{2,6}\\ 1&0&z_{2}h^{2}_{3,3}&z_{2}h^{2}_{3,4}&0&z_{2}h^{2}_{3,6}\\ 0&1&1+z_{2}h^{2}_{4,3}&z_{2}h^{2}_{4,4}&0&z_{2}h^{2}_{4,6}\\ 0&0&-1&z_{2}h^{2}_{4,4}&0&z_{2}h^{2}_{5,6}\\ 0&0&0&1+z_{2}h^{2}_{5,4}&1&z_{2}h^{2}_{6,6}\end{array}\right]=:w\cdot R_{2},\end{split}

where hl,mi(z1,z2)h^{i}_{l,m}(z_{1},z_{2}) are rational functions that is holomorphic at z1=z2=0z_{1}=z_{2}=0.

Denote T1,T2T_{1},T_{2} as the log-monodromy matrix around a small loop of the divisors z1=0z_{1}=0 and z2=0z_{2}=0, represented on Flim0F^{0}_{\mathrm{lim}}. Denote the (Tate twisted) log-monodormy matrix as N¯i=12πilog(Ti)\overline{N}_{i}=\frac{1}{2\pi i}\log(T_{i}), then we have N¯i=(Reszi=0Ri)|z=0\overline{N}_{i}=(\mathrm{Res}_{z_{i}=0}R_{i})|_{z=0}, and more precisely

(4.6) N¯1=[000000100000000000001000010000000100],N¯2=[000000000000100000011000001000000110].\overline{N}_{1}=\left[\begin{array}[]{cccccc}0&0&0&0&0&0\\ 1&0&0&0&0&0\\ 0&0&0&0&0&0\\ 0&0&1&0&0&0\\ 0&1&0&0&0&0\\ 0&0&0&1&0&0\\ \end{array}\right],\quad\overline{N}_{2}=\left[\begin{array}[]{cccccc}0&0&0&0&0&0\\ 0&0&0&0&0&0\\ 1&0&0&0&0&0\\ 0&1&1&0&0&0\\ 0&0&-1&0&0&0\\ 0&0&0&1&1&0\\ \end{array}\right].

Then the nilpotent cone associated to z1=z2=0z_{1}=z_{2}=0 is generated by N¯1,N¯2\overline{N}_{1},\overline{N}_{2}, and a simple calculation shows that this is a MUM point. Therefore, the Picard-Fuchs system (4.4) is indeed the one we want by Lemma 2.9.

4.1.5. Local monodromy and LMHS type

By the same method as in [Che23], the types of LMHS of the rest possible two-dimensional nilpotent cones can be determined up to some ambiguity by calculating the Jordan normal form of the nilpotent operators of the boundary and using the polarization relations. The result is

(4.7) Cone(Dv1s,Dv2s)\displaystyle\mathrm{Cone}(D_{v_{1}^{s}},D_{v_{2}^{s}}) =III0|IV2|III0\displaystyle=\braket{III_{0}}{IV_{2}}{III_{0}}
(4.8) Cone(Dv1s,E3)\displaystyle\mathrm{Cone}(D_{v_{1}^{s}},E_{3}) =III0|IV2|IV1\displaystyle=\braket{III_{0}}{IV_{2}}{IV_{1}}
(4.9) Cone(E1,E2)\displaystyle\mathrm{Cone}(E_{1},E_{2}) =IV1|IV2(IV1)|IV1\displaystyle=\braket{IV_{1}}{IV_{2}(IV_{1})}{IV_{1}}
(4.10) Cone(E2,E3)\displaystyle\mathrm{Cone}(E_{2},E_{3}) =IV1|IV2(IV1)|IV1\displaystyle=\braket{IV_{1}}{IV_{2}(IV_{1})}{IV_{1}}
(4.11) Cone(DA,E3)\displaystyle\mathrm{Cone}(D_{A},E_{3}) =I1|IV2|IV1\displaystyle=\braket{I_{1}}{IV_{2}}{IV_{1}}

where there is another copy of cone (4.8)-(4.11) by the 2\mathbb{Z}_{2}-symmetry of the 𝒟\mathcal{D}-module. The possible ambiguity has been indicated inside the parathesis. Since the monodromy around Dv3sD_{v_{3}^{s}} is finite, we have excluded the cones associated with it.

4.1.6. Yukawa coupling and symplectic form

In this subsection, we will identify ω\omega as 1. The symplectic form QQ is determined by the intersection matrix Qij=VwiwjQ_{ij}=\int_{V}w_{i}\cup w_{j}, which in turn is determined by the nn-point functions

Kij:=VΩδ1iδ2jΩ,i+j=n.K^{ij}:=\int_{V}\Omega\wedge\delta_{1}^{i}\delta_{2}^{j}\Omega,\quad i+j=n.

When n=3n=3, these are the so-called Yukawa couplings. Rewrite the Picard-Fuchs operators

P1:=\displaystyle P_{1}:= (1+z1)δ13+3z1δ12δ2+3z1δ1δ22+z1δ23+2z1δ12+4z1δ1δ2+2z1δ22\displaystyle(1+z_{1})\delta_{1}^{3}+3z_{1}\delta_{1}^{2}\delta_{2}+3z_{1}\delta_{1}\delta_{2}^{2}+z_{1}\delta_{2}^{3}+2z_{1}\delta_{1}^{2}+4z_{1}\delta_{1}\delta_{2}+2z_{1}\delta_{2}^{2}
+119z1δ1+119z1δ2+29z1\displaystyle+\frac{11}{9}z_{1}\delta_{1}+\frac{11}{9}z_{1}\delta_{2}+\frac{2}{9}z_{1}
P2:=\displaystyle P_{2}:= (z1+z2+1)δ12+(2z1+2z21)δ1δ2+(z1+z2+1)δ22\displaystyle(z_{1}+z_{2}+1)\delta_{1}^{2}+(2z_{1}+2z_{2}-1)\delta_{1}\delta_{2}+(z_{1}+z_{2}+1)\delta_{2}^{2}
+(z1+z2)δ1+(z1+z2)δ2+29(z1+z2).\displaystyle+(z_{1}+z_{2})\delta_{1}+(z_{1}+z_{2})\delta_{2}+\frac{2}{9}(z_{1}+z_{2}).

Together with the Griffith transversality, from VΩP1Ω=0\int_{V}\Omega\wedge P_{1}\Omega=0 we get

(1+z1)K3,0+3z1K2,1+3z1K1,2+z1K0,3=0.(1+z_{1})K^{3,0}+3z_{1}K^{2,1}+3z_{1}K^{1,2}+z_{1}K^{0,3}=0.

Similarly, from VΩδiP2Ω=0\int_{V}\Omega\wedge\delta_{i}P_{2}\Omega=0, i=1,2i=1,2, we get

(z1+z2+1)K3,0+(2z1+2z21)K2,1+(z1+z2+1)K1,2=0,\displaystyle(z_{1}+z_{2}+1)K^{3,0}+(2z_{1}+2z_{2}-1)K^{2,1}+(z_{1}+z_{2}+1)K^{1,2}=0,
(z1+z2+1)K2,1+(2z1+2z21)K1,2+(z1+z2+1)K0,3=0.\displaystyle(z_{1}+z_{2}+1)K^{2,1}+(2z_{1}+2z_{2}-1)K^{1,2}+(z_{1}+z_{2}+1)K^{0,3}=0.

Therefore, K3,0K^{3,0} determines all the other Yukawa couplings. More precisely, we have

K21\displaystyle K^{21} =2z12z1z2z1+z22+2z2+13z1(z1+z22)K30,\displaystyle=\frac{-2z_{1}^{2}-z_{1}z_{2}-z_{1}+z_{2}^{2}+2z_{2}+1}{3z_{1}(z_{1}+z_{2}-2)}K^{30},
K12\displaystyle K^{12} =2z22z1z2z2+z12+2z1+13z1(z1+z22)K30,\displaystyle=\frac{-2z_{2}^{2}-z_{1}z_{2}-z_{2}+z_{1}^{2}+2z_{1}+1}{3z_{1}(z_{1}+z_{2}-2)}K^{30},
K03\displaystyle K^{03} =z2z1K30.\displaystyle=\frac{z_{2}}{z_{1}}K^{30}.

We note these results are consistent with the /2\mathbb{Z}/2\mathbb{Z} symmetry. Next, from VΩδ1P1Ω=0\int_{V}\Omega\wedge\delta_{1}P_{1}\Omega=0, we have

(4.12) 0=(1+z1)K4,0+3z1K3,1+3z1K2,2+z1K1,3+3z1K3,0+7z1K2,1+5z1K1,2+z1K0,3,\begin{split}0=&(1+z_{1})K^{4,0}+3z_{1}K^{3,1}+3z_{1}K^{2,2}+z_{1}K^{1,3}+3z_{1}K^{3,0}\\ &+7z_{1}K^{2,1}+5z_{1}K^{1,2}+z_{1}K^{0,3},\end{split}

similarly, from VΩδ12P2Ω\int_{V}\Omega\wedge\delta_{1}^{2}P_{2}\Omega, we have

(4.13) 0=(z1+z2+1)K4,0+(2z1+2z21)K31+(z1+z2+1)K2,2+(3z1+z2)K3,0+(5z1+z2)K2,1+2z1K1,2.\begin{split}0=&(z_{1}+z_{2}+1)K^{4,0}+(2z_{1}+2z_{2}-1)K^{31}+(z_{1}+z_{2}+1)K^{2,2}\\ &+(3z_{1}+z_{2})K^{3,0}+(5z_{1}+z_{2})K^{2,1}+2z_{1}K^{1,2}.\end{split}

To simplify the notation, we denote i,j|l,m=Vδ1iδ2jΩδ1lδ2mΩ\braket{i,j}{l,m}=\int_{V}\delta_{1}^{i}\delta_{2}^{j}\Omega\wedge\delta_{1}^{l}\delta_{2}^{m}\Omega. Then Ki,j=0,0|i,jK^{i,j}=\braket{0,0}{i,j}, and we have i,j|l,m=l,m|i,j\braket{i,j}{l,m}=-\braket{l,m}{i,j}. From δ12K2,0=0\delta_{1}^{2}K^{2,0}=0, we have K4,0=21,0|3,0K^{4,0}=-2\braket{1,0}{3,0}, and since δ1K3,0=1,0|3,0+K4,0\delta_{1}K^{3,0}=\braket{1,0}{3,0}+K^{4,0}, we have K4,0=2δ1K3,0K^{4,0}=2\delta_{1}K^{3,0}. Similarly, from δ1δ2K2,0=δ12K1,1=0\delta_{1}\delta_{2}K^{2,0}=\delta_{1}^{2}K^{1,1}=0 we get

1,0|2,1+0,1|3,0+1,1|2,0+K3,1=0,\displaystyle\braket{1,0}{2,1}+\braket{0,1}{3,0}+\braket{1,1}{2,0}+K^{3,1}=0,
21,0|2,11,1|2,0+K3,1=0.\displaystyle 2\braket{1,0}{2,1}-\braket{1,1}{2,0}+K^{3,1}=0.

combining with δ1K2,1=1,0|2,1+K3,1\delta_{1}K^{2,1}=\braket{1,0}{2,1}+K^{3,1}, and δ2K3,0=0,1|3,0+K3,1\delta_{2}K^{3,0}=\braket{0,1}{3,0}+K^{3,1}, we have

K3,1=32δ1K2,1+12δ2K3,0.K^{3,1}=\frac{3}{2}\delta_{1}K^{2,1}+\frac{1}{2}\delta_{2}K^{3,0}.

By symmetry, we have

K1,3=32δ2K1,2+12δ1K0,3,K0,4=2δ2K0,3.K^{1,3}=\frac{3}{2}\delta_{2}K^{1,2}+\frac{1}{2}\delta_{1}K^{0,3},\quad K^{0,4}=2\delta_{2}K^{0,3}.

A similar calculation shows

K2,2=δ1K1,2+δ2K2,1,K^{2,2}=\delta_{1}K^{1,2}+\delta_{2}K^{2,1},

Then substitute everything back into (4.12) and (4.13), we arrive at the system of first order PDEs

(4.14) δ1K3,0=α1(z1+z22)DAK3,0δ2K3,0=α2(z1+z22)DAK3,0\begin{split}&\delta_{1}K^{3,0}=\frac{\alpha_{1}}{(z_{1}+z_{2}-2)D_{A}}K^{3,0}\\ &\delta_{2}K^{3,0}=\frac{\alpha_{2}}{(z_{1}+z_{2}-2)D_{A}}K^{3,0}\end{split}

where

α1=\displaystyle\alpha_{1}= z142z13z2+4z1318z12z2+9z12+2z1z23+6z1z22+6z1z2+2z1\displaystyle-z_{1}^{4}-2z_{1}^{3}z_{2}+4z_{1}^{3}-18z_{1}^{2}z_{2}+9z_{1}^{2}+2z_{1}z_{2}^{3}+6z_{1}z_{2}^{2}+6z_{1}z_{2}+2z_{1}
+z24+z233z225z22,\displaystyle+z_{2}^{4}+z_{2}^{3}-3z_{2}^{2}-5z_{2}-2,
α2=\displaystyle\alpha_{2}= z2(2z13+6z12z230z12+6z1z226z1z2+42z1+2z23\displaystyle-z_{2}(2z_{1}^{3}+6z_{1}^{2}z_{2}-30z_{1}^{2}+6z_{1}z_{2}^{2}-6z_{1}z_{2}+42z_{1}+2z_{2}^{3}
3z2212z27),\displaystyle-3z_{2}^{2}-12z_{2}-7),

With the help of Wolfram Mathematica®{}^{{\text{\tiny\textregistered}}} we get

K3,0=cz1(z1+z22)DA,c0,K^{3,0}=\frac{cz_{1}}{(z_{1}+z_{2}-2)D_{A}},\quad c\neq 0,

Next, the intersection matrix is given by

(4.15) Q=[00000K2,10001,0|1,11,0|2,01,0|2,10000,1|1,10,1|2,00,1|2,101,1|2,01,1|2,102,0|2,10]Q=\left[\begin{array}[]{cccccc}0&0&0&0&0&K^{2,1}\\ 0&0&0&\braket{1,0}{1,1}&\braket{1,0}{2,0}&\braket{1,0}{2,1}\\ 0&0&0&\braket{0,1}{1,1}&\braket{0,1}{2,0}&\braket{0,1}{2,1}\\ &*&*&0&\braket{1,1}{2,0}&\braket{1,1}{2,1}\\ &*&*&*&0&\braket{2,0}{2,1}\\ &*&*&*&*&0\\ \end{array}\right]

where the * entries are determined by the rest by the antisymmetricity of QQ. By the previous calculation, we have

(4.16) Q=[00000K2,1000K2,1K3,0δ1K2,1K3,1000K1,2K2,1δ2K2,1K2,202δ1K2,1K3,11,1|2,102,0|2,10],Q=\left[\begin{array}[]{cccccc}0&0&0&0&0&K^{2,1}\\ 0&0&0&-K^{2,1}&-K^{3,0}&\delta_{1}K^{2,1}-K^{3,1}\\ 0&0&0&-K^{1,2}&-K^{2,1}&\delta_{2}K^{2,1}-K^{2,2}\\ &*&*&0&2\delta_{1}K^{2,1}-K^{3,1}&\braket{1,1}{2,1}\\ &*&*&*&0&\braket{2,0}{2,1}\\ &*&*&*&*&0\\ \end{array}\right],

Now, the entry 1,1|2,1\braket{1,1}{2,1} can be determined by the following. First, from

(4.17) 0=\displaystyle 0= δ13K3,0=3,0|0,2+32,0|1,2+31,0|2,2+K3,2,\displaystyle\delta_{1}^{3}K^{3,0}=\braket{3,0}{0,2}+3\braket{2,0}{1,2}+3\braket{1,0}{2,2}+K^{3,2},
0=\displaystyle 0= δ12δ2K1,1=2,1|1,1+2,0|1,2+21,1|2,1+21,0|2,2\displaystyle\delta_{1}^{2}\delta_{2}K^{1,1}=\braket{2,1}{1,1}+\braket{2,0}{1,2}+2\braket{1,1}{2,1}+2\braket{1,0}{2,2}
+0,1|3,1+K3,2,\displaystyle+\braket{0,1}{3,1}+K^{3,2},
δ12K1,2=\displaystyle\delta_{1}^{2}K^{1,2}= 2,0|1,2+21,0|2,2+K3,2,\displaystyle\braket{2,0}{1,2}+2\braket{1,0}{2,2}+K^{3,2},
δ1δ2K2,1=\displaystyle\delta_{1}\delta_{2}K^{2,1}= 1,1|2,1+1,0|2,2+0,1|3,1+K3,2,\displaystyle\braket{1,1}{2,1}+\braket{1,0}{2,2}+\braket{0,1}{3,1}+K^{3,2},
δ22K3,0=\displaystyle\delta_{2}^{2}K^{3,0}= 0,2|3,0+20,1|3,1+K3,2,\displaystyle\braket{0,2}{3,0}+2\braket{0,1}{3,1}+K^{3,2},

we get

1,1|2,1=K3,2δ12K1,232δ1δ2K2,112δ22K3,0,\braket{1,1}{2,1}=K^{3,2}-\delta_{1}^{2}K^{1,2}-\frac{3}{2}\delta_{1}\delta_{2}K^{2,1}-\frac{1}{2}\delta_{2}^{2}K^{3,0},

and similarly one has

2,0|2,1=\displaystyle\braket{2,0}{2,1}= K4,143δ1K3,116δ2K4,0\displaystyle K^{4,1}-\frac{4}{3}\delta_{1}K^{3,1}-\frac{1}{6}\delta_{2}K^{4,0}
=\displaystyle= K4,12δ12K2,1δ1δ2K3,0.\displaystyle K^{4,1}-2\delta_{1}^{2}K^{2,1}-\delta_{1}\delta_{2}K^{3,0}.

Therefore, these two entries are determined by the 5-point functions K3,2,K4,1K^{3,2},K^{4,1}. To calculate these 5-point functions, we consider 0=V1δ1iδ2jP10=\int_{V}1\wedge\delta_{1}^{i}\delta_{2}^{j}P_{1}, i+j=2i+j=2, and 0=V1δ1iδ2jP20=\int_{V}1\wedge\delta_{1}^{i}\delta_{2}^{j}P_{2}, i+j=3i+j=3. These 7 equations can be used to rewrite the 6 different 5-point functions into 4-point and 3-point functions. More precisely, the first 6 equations read

0=\displaystyle 0= (z1+1)K5,0+3z1K4,1+4z1K4,0+3z1K3,2+10z1K3,1+56z1K3,09\displaystyle(z_{1}+1)K^{5,0}+3z_{1}K^{4,1}+4z_{1}K^{4,0}+3z_{1}K^{3,2}+10z_{1}K^{3,1}+\frac{56z_{1}K^{3,0}}{9}
+z1K2,3+8z1K2,2+110z1K2,19+2z1K1,3+7z1K1,2+z1K0,3\displaystyle+z_{1}K^{2,3}+8z_{1}K^{2,2}+\frac{110z_{1}K^{2,1}}{9}+2z_{1}K^{1,3}+7z_{1}K^{1,2}+z_{1}K^{0,3}
0=\displaystyle 0= (z1+1)K4,1+3z1K3,2+3z1K3,1+3z1K2,3+7z1K2,2\displaystyle(z_{1}+1)K^{4,1}+3z_{1}K^{3,2}+3z_{1}K^{3,1}+3z_{1}K^{2,3}+7z_{1}K^{2,2}
+29z1K2,19+z1K1,4+5z1K1,3+47z1K1,29+z1K0,4+2z1K0,3\displaystyle+\frac{29z_{1}K^{2,1}}{9}+z_{1}K^{1,4}+5z_{1}K^{1,3}+\frac{47z_{1}K^{1,2}}{9}+z_{1}K^{0,4}+2z_{1}K^{0,3}
0=\displaystyle 0= (z1+1)K3,2+3z1K2,3+2z1K2,2+3z1K1,4+4z1K1,3\displaystyle(z_{1}+1)K^{3,2}+3z_{1}K^{2,3}+2z_{1}K^{2,2}+3z_{1}K^{1,4}+4z_{1}K^{1,3}
+11z1K1,29+z1K0,5+2z1K0,4+11z1K0,39\displaystyle+\frac{11z_{1}K^{1,2}}{9}+z_{1}K^{0,5}+2z_{1}K^{0,4}+\frac{11z_{1}K^{0,3}}{9}
0=\displaystyle 0= (z1+z2+1)K5,0+(2z1+2z21)K4,1+(4z1+z2)K4,0\displaystyle(z_{1}+z_{2}+1)K^{5,0}+(2z_{1}+2z_{2}-1)K^{4,1}+(4z_{1}+z_{2})K^{4,0}
+(z1+z2+1)K3,2+(7z1+z2)K3,1+56z1+2z29K3,0\displaystyle+(z_{1}+z_{2}+1)K^{3,2}+(7z_{1}+z_{2})K^{3,1}+\frac{56z_{1}+2z_{2}}{9}K^{3,0}
+3z1K2,2+9z1K2,1+3z1K1,2\displaystyle+3z_{1}K^{2,2}+9z_{1}K^{2,1}+3z_{1}K^{1,2}
0=\displaystyle 0= (z1+z2+1)K4,1+z2K4,0+(2z1+2z21)K3,2+(3z1+3z2)K3,1\displaystyle(z_{1}+z_{2}+1)K^{4,1}+z_{2}K^{4,0}+(2z_{1}+2z_{2}-1)K^{3,2}+(3z_{1}+3z_{2})K^{3,1}
+z2K3,0+(z1+z2+1)K2,3+(5z1+2z2)K2,2+29z1+11z29K2,1\displaystyle+z_{2}K^{3,0}+(z_{1}+z_{2}+1)K^{2,3}+(5z_{1}+2z_{2})K^{2,2}+\frac{29z_{1}+11z_{2}}{9}K^{2,1}
+2z1K1,3+4z1K1,2+z1K0,3\displaystyle+2z_{1}K^{1,3}+4z_{1}K^{1,2}+z_{1}K^{0,3}
0=\displaystyle 0= (z1+z2+1)K3,2+2z2K3,1+z2K3,0+(2z1+2z21)K2,3\displaystyle(z_{1}+z_{2}+1)K^{3,2}+2z_{2}K^{3,1}+z_{2}K^{3,0}+(2z_{1}+2z_{2}-1)K^{2,3}
+(2z1+5z2)K2,2+4z2K2,1+(z1+z2+1)K1,4+(3z1+3z2)K1,3\displaystyle+(2z_{1}+5z_{2})K^{2,2}+4z_{2}K^{2,1}+(z_{1}+z_{2}+1)K^{1,4}+(3z_{1}+3z_{2})K^{1,3}
+11z1+29z29K1,2+z1K0,4+z1K0,3\displaystyle+\frac{11z_{1}+29z_{2}}{9}K^{1,2}+z_{1}K^{0,4}+z_{1}K^{0,3}

and from these equations we can solve K4,1K^{4,1} and K3,2K^{3,2}, which in turn gives 1,1|2,1\braket{1,1}{2,1} and 2,0|2,1\braket{2,0}{2,1}. The result is

1,1|2,1\displaystyle\braket{1,1}{2,1} =cz1α127(z1+z22)4DA2,\displaystyle=-\frac{cz_{1}\alpha_{1}}{27(z_{1}+z_{2}-2)^{4}D_{A}^{2}},
2,0|2,1\displaystyle\braket{2,0}{2,1} =cz1α227(z1+z22)4DA2,\displaystyle=-\frac{cz_{1}\alpha_{2}}{27(z_{1}+z_{2}-2)^{4}D_{A}^{2}},

where α1,2[z1,z2]\alpha_{1,2}\in\mathbb{Z}[z_{1},z_{2}] are two complicated degree 6 polynomials. Therefore, at z1=z2=0z_{1}=z_{2}=0, the symplectic form is given by

Q=c12[000001000100000110011000001000100000],Q=\frac{c}{12}\cdot\left[\begin{array}[]{cccccc}0&0&0&0&0&1\\ 0&0&0&-1&0&0\\ 0&0&0&-1&-1&0\\ 0&1&1&0&0&0\\ 0&0&1&0&0&0\\ -1&0&0&0&0&0\\ \end{array}\right],

By rescale the local basis ww by (c12)12\left(\frac{c}{12}\right)^{\frac{1}{2}}, we can get rid of the coefficient c12\frac{c}{12}. These calculations are done with the help of Matlab®{}^{{\text{\tiny\textregistered}}}.

4.2. V(2,38)V_{(2,38)}

The toric data for this case is given by

(4.18) Ξv1v2v3v4v5r110014r201025r300112r400033w111000w200111.\begin{array}[]{c|ccccccc}\Xi&v_{1}&v_{2}&v_{3}&v_{4}&v_{5}\\ \hline\cr r_{1}&1&0&0&1&-4\\ r_{2}&0&1&0&2&-5\\ r_{3}&0&0&1&1&-2\\ r_{4}&0&0&0&3&-3\\ \hline\cr w_{1}&1&1&0&0&0\\ w_{2}&0&0&1&1&1\end{array}.

Then, we have XΣX_{\Sigma} is a quotient of [3,3,1,1,1]\mathbb{P}[3,3,1,1,1] by /3\mathbb{Z}/3\mathbb{Z}, and the Baytrev mirror of XΣX_{\Sigma} is Δ\mathbb{P}_{\Delta^{\circ}}. However, since XΣX_{\Sigma} has more than terminal singularity, we need to conduct a toric resolution by adding an additional column (1,1,0,0)t(-1,-1,0,0)^{t}. We will denote the new fan still as Σ\Sigma. Then the suspended toric data is

(4.19) Ξ¯v¯1v¯2v¯3v¯4v¯5v¯6v¯7r¯11001410r¯20102510r¯30011200r¯40003300r¯51111111w¯11100013w¯20011130.\begin{array}[]{c|ccccccc}\bar{\Xi}&\bar{v}_{1}&\bar{v}_{2}&\bar{v}_{3}&\bar{v}_{4}&\bar{v}_{5}&\bar{v}_{6}&\bar{v}_{7}\\ \hline\cr\bar{r}_{1}&1&0&0&1&-4&-1&0\\ \bar{r}_{2}&0&1&0&2&-5&-1&0\\ \bar{r}_{3}&0&0&1&1&-2&0&0\\ \bar{r}_{4}&0&0&0&3&-3&0&0\\ \bar{r}_{5}&1&1&1&1&1&1&1\\ \hline\cr\bar{w}_{1}&1&1&0&0&0&1&-3\\ \bar{w}_{2}&0&0&1&1&1&-3&0\end{array}.

4.2.1. Moduli spaces

Since Δ345\Delta^{\circ}_{345} has an integral interior point v7v_{7}, in order to show poly=\mathcal{M}_{\mathrm{poly}}=\mathcal{M}, we need to show the dual face of Δ345\Delta^{\circ}_{345} has no integral interior point. A direct calculation gives that the dual face is the interval between v1,v2Mv^{\circ}_{1},v^{\circ}_{2}\in M, with v1=(1,2,1,1)v^{\circ}_{1}=(-1,2,-1,-1), and v2=(2,1,1,0)v^{\circ}_{2}=(2,-1,-1,0). Thus, we have poly=\mathcal{M}_{\mathrm{poly}}=\mathcal{M}. It is also clear from the muti-degrees that Autr(Δ)\mathrm{Aut}_{r}(\mathbb{P}_{\Delta^{\circ}}) is trivial.

The fan symmetry of Aut(Σ)\mathrm{Aut}(\Sigma) the group S3S_{3} which is generated by (12)(34)(12)(34) and (345)(345). A generic point in (L(ΔN))\mathbb{P}(L(\Delta^{\circ}\cap N)) is given by

f=λ1t1+λ2t2+λ3t3+λ4t1t22t3t43+λ5t14t25t32t43+λ6t11t21+λ7,f=\lambda_{1}t_{1}+\lambda_{2}t_{2}+\lambda_{3}t_{3}+\lambda_{4}t_{1}t_{2}^{2}t_{3}t_{4}^{3}+\lambda_{5}t_{1}^{-4}t_{2}^{-5}t_{3}^{-2}t_{4}^{-3}+\lambda_{6}t_{1}^{-1}t_{2}^{-1}+\lambda_{7},

One can easily check both (12)(34)(12)(34) and (345)(345) can be realized as a TT action, thus Aut(Σ)\mathrm{Aut}(\Sigma) acts on simp\mathcal{M}_{\mathrm{simp}} trivially. Therefore, we have simp=\mathcal{M}_{\mathrm{simp}}=\mathcal{M}.

The secondary fan Σs\Sigma^{s} is

(4.20) AsBs=Ξsv1sv2sv3sv4sr1s1011r2s0130w1s1001w2s0311.\frac{A^{s}}{B^{s}}=\begin{array}[]{c|cccc}\Xi^{s}&v_{1}^{s}&v_{2}^{s}&v_{3}^{s}&v_{4}^{s}\\ \hline\cr r_{1}^{s}&1&0&1&-1\\ r_{2}^{s}&0&1&-3&0\\ \hline\cr w_{1}^{s}&1&0&0&1\\ w_{2}^{s}&0&3&1&1\\ \end{array}.

Then ¯simpΔs\overline{\mathcal{M}}_{\mathrm{simp}}\cong\mathbb{P}_{\Delta^{s}} is a smooth toric varies. More concretely, we have ΔsProjBs[Z1,Z2,Z3,Z4]\mathbb{P}_{\Delta^{s}}\cong\mathrm{Proj}_{B^{s}}\mathbb{C}[Z_{1},Z_{2},Z_{3},Z_{4}], with ZiZ_{i} has multi-degree (w1,is,w2,is)(w_{1,i}^{s},w_{2,i}^{s}). Then the canonical affine coordinates are

z1=λ1λ2λ6λ73=Z1Z3Z4,z2=λ3λ4λ5λ63=Z2Z33.z_{1}=\frac{\lambda_{1}\lambda_{2}\lambda_{6}}{\lambda_{7}^{3}}=\frac{Z_{1}Z_{3}}{Z_{4}},\quad z_{2}=\frac{\lambda_{3}\lambda_{4}\lambda_{5}}{\lambda_{6}^{3}}=\frac{Z_{2}}{Z_{3}^{3}}.

4.2.2. Picard-Fuchs equations and discriminant locus

P1=δ1(δ13δ2)+3z1(3δ1+1)(3δ1+2)\displaystyle P_{1}=\delta_{1}(\delta_{1}-3\delta_{2})+3z_{1}(3\delta_{1}+1)(3\delta_{1}+2)
P2=δ23z2(δ13δ2)(δ13δ21)(δ13δ22)\displaystyle P_{2}=\delta_{2}^{3}-z_{2}(\delta_{1}-3\delta_{2})(\delta_{1}-3\delta_{2}-1)(\delta_{1}-3\delta_{2}-2)

For simplicity, we rescale the local coordinates as z1,2:=33z1,2z_{1,2}^{\prime}:=3^{3}\cdot z_{1,2}, and rename z1,2z_{1,2}^{\prime} as the new z1,2z_{1,2}. The Picard-Fuchs operators become

P1=δ1(δ13δ2)+z1(δ1+13)(δ1+23)\displaystyle P_{1}=\delta_{1}(\delta_{1}-3\delta_{2})+z_{1}(\delta_{1}+\frac{1}{3})(\delta_{1}+\frac{2}{3})
P2=δ23z2(13δ1δ2)(13δ1δ213)(13δ1δ223)\displaystyle P_{2}=\delta_{2}^{3}-z_{2}(\frac{1}{3}\delta_{1}-\delta_{2})(\frac{1}{3}\delta_{1}-\delta_{2}-\frac{1}{3})(\frac{1}{3}\delta_{1}-\delta_{2}-\frac{2}{3})

The singular locus contains inside

DAD1Dv1sDv2sDv3sDv4sD_{A}\cup D_{1}\cup D_{v_{1}^{s}}\cup D_{v_{2}^{s}}\cup D_{v_{3}^{s}}\cup D_{v_{4}^{s}}

where DA=(1+z1)3+z13z2D_{A}=(1+z_{1})^{3}+z_{1}^{3}z_{2}, D1=1+z2D_{1}=1+z_{2}. In this affine chart, we have Dv1sD_{v_{1}^{s}} intersects Dv2sD_{v_{2}^{s}} transversely at origin; The DAD_{A} intersects Dv2s=𝕍(z2)D_{v_{2}^{s}}=\mathbb{V}(z_{2}) at (z1,z2)=(1,0)(z_{1},z_{2})=(-1,0) with 3-tangency and thus requires to be blown up three times, with the exceptional divisors denoted as E1,E2,E3E_{1},E_{2},E_{3}; The D1D_{1} intersects Dv1sD_{v_{1}^{s}} transversally at one point (z1,z2)=(0,1)(z_{1},z_{2})=(0,-1). Now, there are intersections outside this affine chart. Now, switch back to the original z1,z2z_{1},z_{2} coordinates before the rescaling and let

z1=Z1Z213Z4,z2=Z3Z213,z_{1}^{\prime}=\frac{Z_{1}Z_{2}^{\frac{1}{3}}}{Z_{4}},\quad z_{2}^{\prime}=\frac{Z_{3}}{Z_{2}^{\frac{1}{3}}},

we have

z1=z1z2,z2=z23,z_{1}=z_{1}^{\prime}z_{2}^{\prime},\quad z_{2}=z_{2}^{\prime-3},

Since the coordinates change is étale away from the boundaries, the Picard-Fuchs ideal becomes

P1=δ1δ2+3z1z2(3δ1+1)(3δ1+2)P2=z23(δ1δ2)3δ2(δ21)(δ22)\begin{split}&P_{1}=\delta_{1}^{\prime}\delta_{2}^{\prime}+3z_{1}^{\prime}z_{2}^{\prime}(3\delta_{1}^{\prime}+1)(3\delta_{1}^{\prime}+2)\\ &P_{2}=z_{2}^{\prime 3}(\delta_{1}^{\prime}-\delta_{2}^{\prime})^{3}-\delta_{2}^{\prime}(\delta_{2}^{\prime}-1)(\delta_{2}^{\prime}-2)\end{split}

Similar, by a rescale z1′′:=34z1z_{1}^{\prime\prime}:=3^{4}\cdot z_{1}^{\prime}, z2′′:=31z2z_{2}^{\prime\prime}:=3^{-1}\cdot z_{2}^{\prime}, and rename z1,2′′z_{1,2}^{\prime\prime} as the new z1,2z_{1,2}^{\prime}, then DA=(1+z1z2)3+z13D_{A}=(1+z_{1}^{\prime}z_{2}^{\prime})^{3}+z_{1}^{\prime 3}, D1=1+z23D_{1}=1+z_{2}^{-3}, and

P1=δ1δ2+z1z2(δ1+13)(δ1+23)P2=27z23(δ1δ2)3δ2(δ21)(δ22),\begin{split}&P_{1}=\delta_{1}^{\prime}\delta_{2}^{\prime}+z_{1}^{\prime}z_{2}^{\prime}(\delta_{1}^{\prime}+\frac{1}{3})(\delta_{1}^{\prime}+\frac{2}{3})\\ &P_{2}=27z_{2}^{\prime 3}(\delta_{1}^{\prime}-\delta_{2}^{\prime})^{3}-\delta_{2}^{\prime}(\delta_{2}^{\prime}-1)(\delta_{2}^{\prime}-2),\end{split}

In this affine chart we have additionally two cones that correspond to Dv2sD_{v_{2}^{s}} intersects Dv3sD_{v_{3}^{s}} transversely at origin, and DAD_{A} intersects Dv3sD_{v_{3}^{s}} transversely at (1,0)(-1,0). The last affine chart that we would want to consider is

z1′′=z11,z2′′=z2,z_{1}^{\prime\prime}=z_{1}^{-1},\quad z_{2}^{\prime\prime}=z_{2},

where ziz_{i} are the original ziz_{i} before rescale. Then the Picard-Fuchs ideal is

P1=z1δ1(δ1+3δ2)+3(3δ11)(3δ12)\displaystyle P_{1}=z_{1}\delta_{1}(\delta_{1}+3\delta_{2})+3(3\delta_{1}-1)(3\delta_{1}-2)
P2=δ23+z2(δ1+3δ2)(δ1+3δ2+1)(δ1+3δ2+2),\displaystyle P_{2}=\delta_{2}^{3}+z_{2}(\delta_{1}+3\delta_{2})(\delta_{1}+3\delta_{2}+1)(\delta_{1}+3\delta_{2}+2),

Similar, by a rescale z1′′′:=33z1z_{1}^{\prime\prime\prime}:=3^{-3}\cdot z_{1}^{\prime}, z2′′:=33z2z_{2}^{\prime\prime}:=3^{3}\cdot z_{2}^{\prime}, and rename z1,2′′z_{1,2}^{\prime\prime} as the new z1,2z_{1,2}^{\prime}, then DA=(1+z1)3+z2D_{A}=(1+z_{1})^{3}+z_{2}, D1=1+z2D_{1}=1+z_{2}, and

P1=z1δ1(δ1+3δ2)+(δ113)(δ123)\displaystyle P_{1}=z_{1}\delta_{1}(\delta_{1}+3\delta_{2})+(\delta_{1}-\frac{1}{3})(\delta_{1}-\frac{2}{3})
P2=δ23+z2(13δ1+δ2)(13δ1+δ2+13)(13δ1+δ2+23),\displaystyle P_{2}=\delta_{2}^{3}+z_{2}(\frac{1}{3}\delta_{1}+\delta_{2})(\frac{1}{3}\delta_{1}+\delta_{2}+\frac{1}{3})(\frac{1}{3}\delta_{1}+\delta_{2}+\frac{2}{3}),

This time, DAD_{A} intersects D1D_{1} and Dv4sD_{v_{4}^{s}} at a triple intersection point (0,1)(0,-1). Thus we need to blow up this point once, and the exceptional divisor is denoted as E0E_{0}. Since Dv4sD_{v_{4}^{s}} has finite monodromy, the only new possible two-dimension cones in this chart are those associated to DAE0D_{A}\cap E_{0} and D1E0D_{1}\cap E_{0}.

Remark 4.1.

All the rescales in this section can be done simultaneously in the homogeneous coordinates ZiZ_{i} by rescaling Z1,Z2Z_{1},Z_{2} by a 333^{3} factor.

4.2.3. Nilpotent cone and types of LMHS

Similarly, we can take the global multi-valued frame

w=(Ω,δ1Ω,δ2Ω,δ1δ2Ω,δ22Ω,δ22δ1Ω),w=(\Omega,\delta_{1}\Omega,\delta_{2}\Omega,\delta_{1}\delta_{2}\Omega,\delta_{2}^{2}\Omega,\delta_{2}^{2}\delta_{1}\Omega),

and under this basis, the nilpotent cone at z1=z2=0z_{1}=z_{2}=0 are represented as

N¯1=[000000100000000000031000000000000310],N¯2=[000000000000100000010000001000000100].\overline{N}_{1}=\left[\begin{array}[]{cccccc}0&0&0&0&0&0\\ 1&0&0&0&0&0\\ 0&0&0&0&0&0\\ 0&3&1&0&0&0\\ 0&0&0&0&0&0\\ 0&0&0&3&1&0\\ \end{array}\right],\quad\overline{N}_{2}=\left[\begin{array}[]{cccccc}0&0&0&0&0&0\\ 0&0&0&0&0&0\\ 1&0&0&0&0&0\\ 0&1&0&0&0&0\\ 0&0&1&0&0&0\\ 0&0&0&1&0&0\\ \end{array}\right].

The possible 2-dimensional cones with the types of LMHS are given as follows

(4.21) Cone(Dv1s,Dv2s)\displaystyle\mathrm{Cone}(D_{v_{1}^{s}},D_{v_{2}^{s}}) =IV1|IV2|III0\displaystyle=\braket{IV_{1}}{IV_{2}}{III_{0}}
(4.22) Cone(Dv2s,E3)\displaystyle\mathrm{Cone}(D_{v_{2}^{s}},E_{3}) =III0|III0(IV2)|III0\displaystyle=\braket{III_{0}}{III_{0}(IV_{2})}{III_{0}}
(4.23) Cone(E1,E2)\displaystyle\mathrm{Cone}(E_{1},E_{2}) =III0|III0(IV2)|III0\displaystyle=\braket{III_{0}}{III_{0}(IV_{2})}{III_{0}}
(4.24) Cone(E2,E3)\displaystyle\mathrm{Cone}(E_{2},E_{3}) =III0|III0(IV2)|III0\displaystyle=\braket{III_{0}}{III_{0}(IV_{2})}{III_{0}}
(4.25) Cone(DA,E3)\displaystyle\mathrm{Cone}(D_{A},E_{3}) =I1|III0(IV2)|III0\displaystyle=\braket{I_{1}}{III_{0}(IV_{2})}{III_{0}}
(4.26) Cone(Dv1s,D1)\displaystyle\mathrm{Cone}(D_{v_{1}^{s}},D_{1}) =IV1|IV2|I1\displaystyle=\braket{IV_{1}}{IV_{2}}{I_{1}}
(4.27) Cone(Dv2s,Dv3s)\displaystyle\mathrm{Cone}(D_{v_{2}^{s}},D_{v_{3}^{s}}) =III0|III0(IV2)|I1\displaystyle=\braket{III_{0}}{III_{0}(IV_{2})}{I_{1}}
(4.28) Cone(Dv3s,DA)\displaystyle\mathrm{Cone}(D_{v_{3}^{s}},D_{A}) =I1|I2(I1)|I1\displaystyle=\braket{I_{1}}{I_{2}(I_{1})}{I_{1}}
(4.29) Cone(Dv1s,Dv3s)\displaystyle\mathrm{Cone}(D_{v_{1}^{s}},D_{v_{3}^{s}}) =IV1|IV2|I1\displaystyle=\braket{IV_{1}}{IV_{2}}{I_{1}}
(4.30) Cone(DA,E0)\displaystyle\mathrm{Cone}(D_{A},E_{0}) =I1|III0(IV2)|III0\displaystyle=\braket{I_{1}}{III_{0}(IV_{2})}{III_{0}}
(4.31) Cone(D1,E0)\displaystyle\mathrm{Cone}(D_{1},E_{0}) =I1|III0(IV2)|III0\displaystyle=\braket{I_{1}}{III_{0}(IV_{2})}{III_{0}}

4.3. Symplectic form

Similar to the calculation in Section 4.1.6, one can calculate the symplectic form by calculating the Yukawa couplings. In particular, form VΩδiP1Ω\int_{V}\Omega\wedge\delta_{i}P_{1}\Omega and VΩP2Ω\int_{V}\Omega\wedge P_{2}\Omega, we have

(1+z1)K3,03K2,1=0,\displaystyle(1+z_{1})K^{3,0}-3K^{2,1}=0,
(1+z1)K2,13K1,2=0,\displaystyle(1+z_{1})K^{2,1}-3K^{1,2}=0,
z2K3,09z2K2,1+27z2K1,227(1+z2)K0,3=0,\displaystyle z_{2}K^{3,0}-9z_{2}K^{2,1}+27z_{2}K^{1,2}-27(1+z_{2})K^{0,3}=0,

where the last equation gives

K0,3=z2(1+3z1+3z12)27(1+z2)K3,0.K^{0,3}=\frac{z_{2}(1+3z_{1}+3z_{1}^{2})}{27(1+z_{2})}K^{3,0}.

Substitute these into VΩδ12P1Ω=0\int_{V}\Omega\wedge\delta_{1}^{2}P_{1}\Omega=0, VΩδ2P2Ω=0\int_{V}\Omega\wedge\delta_{2}P_{2}\Omega=0, we have

δz1K3,0=z1(6z1+3z12z2+3z12+3)DAK3,0,\displaystyle\delta_{z_{1}}K^{3,0}=\frac{z_{1}(6z_{1}+3z_{1}^{2}z_{2}+3z_{1}^{2}+3)}{D_{A}}K^{3,0},
δz2K3,0=z13z2DAK3,0.\displaystyle\delta_{z_{2}}K^{3,0}=\frac{z_{1}^{3}z_{2}}{D_{A}}K^{3,0}.

Therefore, we have K3,0=cDAK^{3,0}=c\cdot D_{A}, where cc is some constant. Similar to the calculation in (4.16), the symplectic form is given as

Q=c9[000001000100000310013000001000100000].Q=\frac{c}{9}\cdot\left[\begin{array}[]{cccccc}0&0&0&0&0&1\\ 0&0&0&-1&0&0\\ 0&0&0&-3&-1&0\\ 0&1&3&0&0&0\\ 0&0&1&0&0&0\\ -1&0&0&0&0&0\\ \end{array}\right].

4.4. V(2,86)V_{(2,86)}, the mirror octic

This is the smooth Calabi-Yau 3-fold mirror to the the smooth anti-canonical hypersurface in [1,1,2,2,2]\mathbb{P}[1,1,2,2,2]. This family is very similar to the case of V(2,38)V_{(2,38)}. In particular, the simplified moduli space is the same as the complex moduli space. The mirror symmetry of this family is studied in detail in [Can+94, CK99].

The first author calculated the nilpotent fan for the family in [Che23]. The two-dimensional nilpotent cones are

σ12=IV2|IV2|II1,\displaystyle\sigma_{12}=\braket{IV_{2}}{IV_{2}}{II_{1}}, σ13=IV2|IV2|I1,\displaystyle\sigma_{13}=\braket{IV_{2}}{IV_{2}}{I_{1}},
σ34=I1|I2|I1,\displaystyle\sigma_{34}=\braket{I_{1}}{I_{2}}{I_{1}}, σ45=I1|II1|II0,\displaystyle\sigma_{45}=\braket{I_{1}}{II_{1}}{II_{0}},
σ52=II0|II1|II1,\displaystyle\sigma_{52}=\braket{II_{0}}{II_{1}}{II_{1}}, σ63=I1|I2|I1,\displaystyle\sigma_{63}=\braket{I_{1}}{I_{2}}{I_{1}},
σ67=I1|I2|I1.\displaystyle\sigma_{67}=\braket{I_{1}}{I_{2}}{I_{1}}.

Under the following basis of F0limF^{0}_{\mathrm{lim}}

w=(Ω,δ1Ω,δ2Ω,δ1δ2Ω,δ12Ω,δ12δ2Ω),w=(\Omega,\delta_{1}\Omega,\delta_{2}\Omega,\delta_{1}\delta_{2}\Omega,\delta_{1}^{2}\Omega,\delta_{1}^{2}\delta_{2}\Omega),

the nilpotent operators N¯1,N¯2\overline{N}_{1},\overline{N}_{2} that generates the cone σ12\sigma_{12} are represented in this basis as

(4.32) N¯1=[000000100000000000001000010000000120],N¯2=[000000000000100000010000000000000010],\overline{N}_{1}=\left[\begin{array}[]{cccccc}0&0&0&0&0&0\\ 1&0&0&0&0&0\\ 0&0&0&0&0&0\\ 0&0&1&0&0&0\\ 0&1&0&0&0&0\\ 0&0&0&1&2&0\\ \end{array}\right],\quad\overline{N}_{2}=\left[\begin{array}[]{cccccc}0&0&0&0&0&0\\ 0&0&0&0&0&0\\ 1&0&0&0&0&0\\ 0&1&0&0&0&0\\ 0&0&0&0&0&0\\ 0&0&0&0&1&0\\ \end{array}\right],

and the symplectic form is

(4.33) Q=c[000001000120000010010000021000100000],Q=c\cdot\left[\begin{array}[]{cccccc}0&0&0&0&0&1\\ 0&0&0&-1&-2&0\\ 0&0&0&0&-1&0\\ 0&1&0&0&0&0\\ 0&2&1&0&0&0\\ -1&0&0&0&0&0\end{array}\right],

where cc is some constant.

5. Computing the generic degree

In this section we show the main Theorem 1.2 by explicitly computing deg(Φs¯)\mathrm{deg}(\overline{\Phi_{s}}) (see Proposition 3.5) around a specific boundary point sS¯Ss\in\overline{S}-S. This will fullfill the proof of Theorem 1.2.

5.1. V(2,29)V_{(2,29)}

The period map is Φ:simp=:SΓ\D\Phi:\mathcal{M}_{\mathrm{simp}}=:S\rightarrow\Gamma\backslash D. We choose a minimal resolution simp¯\overline{\mathcal{M}_{\mathrm{simp}}} of 2\mathbb{P}^{2} such that simp¯simp\overline{\mathcal{M}_{\mathrm{simp}}}-\mathcal{M}_{\mathrm{simp}} is a normal crossing divisor.

Take s=[1:0:0]2s=[1:0:0]\in\mathbb{P}^{2}. Since locally around 2simp\mathbb{P}^{2}-\mathcal{M}_{\mathrm{simp}} is normal-crossing, we may think ss as a codimension 22 boundary point in simp¯simp\overline{\mathcal{M}_{\mathrm{simp}}}-\mathcal{M}_{\mathrm{simp}}.

Types listed in (4.7) - (4.11) imply (σs,Fs)(\sigma_{s},F_{s}) is the only local nilpotent orbit with type III0|IV2|III0\langle\mathrm{III}_{0}|\mathrm{IV}_{2}|\mathrm{III}_{0}\rangle, and deg(φs¯|s¯)=2\mathrm{deg}(\overline{\varphi_{s}}|_{\overline{\wp_{s}}})=2 as a consequence of Sections 4.1.1 and 4.1.2. Therefore to show the Theorem 1.2 for V(2,29)V_{(2,29)}, we only have to show deg(Φs¯)=1\mathrm{deg}(\overline{\Phi_{s}})=1.

The matrices of the monodromy cone σs=N1,N2\sigma_{s}=\langle N_{1},N_{2}\rangle are computed in 4.6. In a local lifting Φ~U\tilde{\Phi}_{U} around sUΔ2s\in U\simeq\Delta^{2}, denote l(zi):=log(zi)2πil(z_{i}):=\frac{\log(z_{i})}{2\pi i}, the local lifted period map (2.7) takes the form:

(5.1) Φ~U(z1,z2)=exp(l(z1)N1+l(z2)N2)ψU(z1,z2),(z1,z2)Δ2.\tilde{\Phi}_{U}(z_{1},z_{2})=\exp(l(z_{1})N_{1}+l(z_{2})N_{2})\psi_{U}(z_{1},z_{2}),\ (z_{1},z_{2})\in\Delta^{2}.

Let Fs:=ψU(0,0)DˇF_{s}:=\psi_{U}(0,0)\in\check{D}, under the basis (4.5) which we denote as (e1,,e6)(e_{1},...,e_{6}), we make suppose

(5.2) F3ψU(z1,z2)=(1,ζ1(z1,z2),,ζ5(z1,z2)).F^{3}\psi_{U}(z_{1},z_{2})=\langle(1,\zeta_{1}(z_{1},z_{2}),...,\zeta_{5}(z_{1},z_{2}))\rangle.

By computation we have

(5.3) F3Φ~U(z1,z2)=(1,ζ1(z1,z2)+l(z1),ζ2(z1,z2)+l(z2),,,)F^{3}\tilde{\Phi}_{U}(z_{1},z_{2})=\langle(1,\zeta_{1}(z_{1},z_{2})+l(z_{1}),\zeta_{2}(z_{1},z_{2})+l(z_{2}),*,*,*)\rangle

where *’s are irrelevant terms. This implies we may construct the map UΔ2U\rightarrow\Delta^{2} by

(5.4) (z1,z2)(exp(2πiF3Φ~U(z1,z2),e5),exp(2πiF3Φ~U(z1,z2),e4))\displaystyle(z_{1},z_{2})\rightarrow(\exp(2\pi i\langle F^{3}\tilde{\Phi}_{U}(z_{1},z_{2}),e_{5}\rangle),\exp(2\pi i\langle F^{3}\tilde{\Phi}_{U}(z_{1},z_{2}),e_{4}\rangle))
(5.5) =(z1exp(2πiζ1(z1,z2)),z2exp(2πiζ2(z1,z2))).\displaystyle=(z_{1}\exp(2\pi i\zeta_{1}(z_{1},z_{2})),z_{2}\exp(2\pi i\zeta_{2}(z_{1},z_{2}))).

Since the Jacobian matrix of this map at (0,0)(0,0) is invertible, (0,0)(0,0) is not a branched point and (5.4) is a local isomorphism. This implies the desired deg(Φs¯)=1\mathrm{deg}(\overline{\Phi_{s}})=1.

5.2. V(2,38)V_{(2,38)}

We use the boundary point (4.21) and its corresponding monodromy matrices (4.2.3). Using the same notations as the last subsection, the map (5.3) now read as:

(5.6) F3Φ~U(z1,z2)=(1,ζ1(z1,z2)+l(z2),ζ2(z1,z2)+l(z1),,,),F^{3}\tilde{\Phi}_{U}(z_{1},z_{2})=\langle(1,\zeta_{1}(z_{1},z_{2})+l(z_{2}),\zeta_{2}(z_{1},z_{2})+l(z_{1}),*,*,*)\rangle,

and the analog of map (5.4) is read as:

(5.7) (z1,z2)(z2exp(2πiζ1(z1,z2)),z1exp(2πiζ2(z1,z2))).(z_{1},z_{2})\rightarrow(z_{2}\exp(2\pi i\zeta_{1}(z_{1},z_{2})),z_{1}\exp(2\pi i\zeta_{2}(z_{1},z_{2}))).

The same argument shows deg(Φs¯)=1\mathrm{deg}(\overline{\Phi_{s}})=1. Note that among monodromy cones (4.21) - (4.31), the LMHS type of (4.21) is unique, and (4.21) does not have any non-trivial torsion automorphisms other than rescaling because of Proposition 3.9, thus deg(φs¯|s¯)=1\mathrm{deg}(\overline{\varphi_{s}}|_{\overline{\wp_{s}}})=1.

5.3. V(2,86)V_{(2,86)}

We use the boundary point corresponds to the cone σ12\sigma_{12} whose monodromy matrices are (4.32). Again using the same notations as the last subsection, the map (5.3) now read as:

(5.8) F3Φ~U(z1,z2)=(1,ζ1(z1,z2)+l(z1),ζ2(z1,z2)+l(z2),,,),F^{3}\tilde{\Phi}_{U}(z_{1},z_{2})=\langle(1,\zeta_{1}(z_{1},z_{2})+l(z_{1}),\zeta_{2}(z_{1},z_{2})+l(z_{2}),*,*,*)\rangle,

and the analog of map (5.4) is read as:

(5.9) (z1,z2)(z1exp(2πiζ1(z1,z2)),z2exp(2πiζ2(z1,z2))).(z_{1},z_{2})\rightarrow(z_{1}\exp(2\pi i\zeta_{1}(z_{1},z_{2})),z_{2}\exp(2\pi i\zeta_{2}(z_{1},z_{2}))).

The same argument shows deg(Φs¯)=1\mathrm{deg}(\overline{\Phi_{s}})=1. The same reasons as those for V(2,38)V_{(2,38)} show that deg(φs¯|s¯)=1\mathrm{deg}(\overline{\varphi_{s}}|_{\overline{\wp_{s}}})=1. The proof of Theorem 1.2 is now completed.

References

  • [AGM93] Paul S Aspinwall, Brian R Greene and David R Morrison “The monomial-divisor mirror map” In arXiv preprint alg-geom/9309007, 1993
  • [AGM94] Paul S Aspinwall, Brian R Greene and David R Morrison “Calabi-Yau moduli space, mirror manifolds and spacetime topology change in string theory” In Nuclear Physics B 416.2 Elsevier, 1994, pp. 414–480
  • [Bat94] V Batyrev “Dual polyhedra and mirror symmetry for Calabi–Yau hypersurfaces in toric varieties” In J. Algebraic Geom. 3, 1994, pp. 493
  • [BBT22] Benjamin Bakker, Yohan Brunebarbe and Jacob Tsimerman “o-minimal GAGA and a conjecture of Griffiths” In Inventiones mathematicae 232, 2022, pp. 163–228
  • [BC94] Victor V Batyrev and David A Cox “On the Hodge structure of projective hypersurfaces in toric varieties” In Duke Math. J. 76.1, 1994, pp. 293–338
  • [Can+94] Philip Candelas et al. “Mirror symmetry for two-parameter models (I)” In Nuclear Physics B 416.2 Elsevier, 1994, pp. 481–538
  • [Che23] Chongyao Chen “Completion of a period map of hodge type (1, 2, 2, 1)” In arXiv preprint arXiv:2311.10212, 2023
  • [CK82] Eduardo Cattani and Aroldo Kaplan “Polarized mixed Hodge structures and the local monodromy of a variation of Hodge structure” In Inventiones mathematicae 67, 1982, pp. 101–115
  • [CK99] David A Cox and Sheldon Katz “Mirror symmetry and algebraic geometry” American Mathematical Society Providence, RI, 1999
  • [CKS86] Eduardo Cattani, Aroldo Kaplan and Wilfried Schmid “Degeneration of Hodge Structures” In Annals of Mathematics 123.03, 1986, pp. 457–535
  • [CLS11] David A. Cox, John B. Little and Henry K. Schenck “Toric Varieties” 124, Graduate Studies in Mathematics American Mathematical Society, 2011
  • [DR23] Haohua Deng and Colleen Robles “Completion of two-parameter period maps by nilpotent orbits”, 2023 arXiv:2312.00542 [math.AG]
  • [Fil23] Simion Filip “Global properties of some weight 3 variations of Hodge structure” In Proceedings of the 8th European Congress of Mathematics, 2023
  • [Gel+94] Israel M Gelfand et al. “A-discriminants” Springer, 1994
  • [HK21] Tatsuki Hayama and Atsushi Kanazawa “Degenerating Hodge structure of one–parameter family of Calabi–Yau threefolds” In Asian Journal of Mathematics 25.1, 2021, pp. 31–42
  • [HLZ16] An Huang, Bong H Lian and Xinwen Zhu “Period integrals and the Riemann–Hilbert correspondence” In Journal of differential geometry 104.2 Lehigh University, 2016, pp. 325–369
  • [Hos+95] Shinobu Hosono, Albrecht Klemm, S Thiesen and Shing-Tung Yau “Mirror symmetry, mirror map and applications to Calabi-Yau hypersurfaces” In Communications in Mathematical Physics 167 Springer, 1995, pp. 301–350
  • [Hot91] Ryoshi Hotta “Equivariant D-modules” In arXiv preprint math/9805021 Citeseer, 1991
  • [HT07] Ryoshi Hotta and Toshiyuki Tanisaki “D-modules, perverse sheaves, and representation theory” Springer Science & Business Media, 2007
  • [KNU13] Kazuya Kato, Chikara Nakayama and Sampei Usui “Classifying spaces of degenerating mixed Hodge structures, III: Spaces of nilpotent orbits” In J. Algebraic Geom. 22, 2013, pp. 671–772
  • [KPR19] M Kerr, G Pearlstein and C Robles “Polarized relations on horizontal SL(2)’s” In Documenta Mathematica 24, 2019, pp. 1295–1360
  • [KSZ91] Mikhail M Kapranov, Bernd Sturmfels and Andrei V Zelevinsky “Quotients of toric varieties” In Mathematische Annalen 290 Springer, 1991, pp. 643–655
  • [KU08] Kazuya Kato and Sampei Usui “Classifying Spaces of Degenerating Polarized Hodge Structures” 169, Annals of Mathematics Studies Princeton University Press, 2008
  • [LSY13] Bong H Lian, Ruifang Song and Shing-Tung Yau “Periodic integrals and tautological systems” In Journal of the European Mathematical Society 15.4, 2013, pp. 1457–1483
  • [Sch73] Wilfried Schmid “Variation of Hodge Structure: The Singularities of the Period Mapping” In Inventiones mathematicae 22, 1973, pp. 211–320
  • [Shi09] Kennichiro Shirakawa “Generic Torelli theorem for one-parameter mirror families to weighted hypersurfaces” In Proc. Japan Acad. Ser. A Math. Sci. 85.10, 2009, pp. 167–170
  • [Usu06] Sampei Usui “Images of extended period maps” In J. Algebraic Geom. 15, 2006, pp. 603–621
  • [Usu08] Sampei Usui “Generic Torelli theorem for quintic-mirror family” In Proc. Japan Acad. Ser. A Math. Sci. 84.8, 2008, pp. 143–146