Introduction
In [Reference YoshikawaYos04], a holomorphic torsion invariant of $K3$ surfaces with involution was introduced by the second-named author. The purpose of the present paper is to completely determine the structure of this invariant as a function on the moduli space of such $K3$ surfaces. We will express it using Borcherds products and Siegel modular forms. Let us explain our result in detail.
A pair consisting of a $K3$ surface and an anti-symplectic involution is called a $2$-elementary $K3$ surface. By Nikulin [Reference NikulinNik83], the deformation type of a $2$-elementary $K3$ surface is determined by the isometry class of the invariant lattice of the induced involution on the second integral cohomology. There exist $75$ deformation types, labeled by primitive $2$-elementary Lorentzian sublattices of the $K3$ lattice ${\Bbb L}_{K3}$. Let $M\subset {\Bbb L}_{K3}$ be one such sublattice of rank $r$. Its orthogonal complement $\Lambda = M^{\perp }\cap {\Bbb L}_{K3}$ is a $2$-elementary lattice of signature $(2,20-r)$. Let $\Omega _{\Lambda }^{+}$ be the Hermitian domain of type IV of dimension $20-r$ associated to $\Lambda$ and let $O^{+}(\Lambda )$ be the index $2$ subgroup of the orthogonal group of $\Lambda$ preserving $\Omega _{\Lambda }^{+}$. Via the period map, the moduli space of $2$-elementary $K3$ surfaces of type $M$ is isomorphic to the quotient
where ${\mathcal {D}}_{\Lambda }$ is the discriminant divisor. Hence ${\mathcal {M}}_{\Lambda }^{0}$ is a Zariski open subset of a modular variety of orthogonal type of dimension $20-r$.
In [Reference YoshikawaYos04], a holomorphic torsion invariant of $2$-elementary $K3$ surfaces was defined as follows. Let $(X,\iota )$ be a $2$-elementary $K3$ surface of type $M$. Write $X^{\iota }$ for the set of fixed points of $\iota$. Depending on $M$, $X^{\iota }$ is either empty or the disjoint union of smooth curves of total genus $g=g(M)$. (See § 3.2 for a formula for $g$.) When $X^{\iota }$ is empty, the corresponding type is unique and is called exceptional, in which case $(X,\iota )$ is the universal covering of an Enriques surface endowed with the non-trivial covering transformation. Take an $\iota$-invariant Ricci-flat Kähler form $\gamma$ on $X$ (cf. [Reference YauYau78]) and a holomorphic $2$-form $\eta \not =0$ on $X$. Let $\tau _{\textbf {Z}_{2}}(X,\gamma )(\iota )$ be the equivariant analytic torsion of $(X,\gamma )$ with respect to the $\iota$-action and let $\tau (X^{\iota },\gamma |_{X^{\iota }})$ be the analytic torsion of $(X^{\iota },\gamma |_{X^{\iota }})$ (cf. [Reference Ray and SingerRS73, Reference BismutBis95]). Then the real number
depends only on the isomorphism class of $(X,\iota )$, so that it gives rise to a function $\tau _{M}$ on ${\mathcal {M}}_{\Lambda }^{0}$. The goal of this paper is to give an explicit expression for $\tau _M$ in terms of modular forms. It turns out that $\tau _{M}$ is expressed by two types of modular forms: Borcherds products and Siegel modular forms. Let us explain these modular forms.
Let $\rho _{\Lambda }\colon {\rm Mp}_{2}(\textbf {Z})\to {\rm GL}(\textbf {C}[A_{\Lambda }])$ be the Weil representation attached to the discriminant group $A_{\Lambda }$ of $\Lambda$, where $\textbf {C}[A_{\Lambda }]$ is the group ring of $A_{\Lambda }$. By Borcherds [Reference BorcherdsBor98], given an $O^{+}(\Lambda )$-invariant elliptic modular form $f$ of type $\rho _{\Lambda }$ and of weight $1-(20-r)/2$ with integral Fourier expansion, we can take its Borcherds lift $\Psi _{\Lambda }(\cdot ,f)$. This is a (possibly meromorphic) automorphic form for $O^{+}(\Lambda )$ with infinite product expansion, and its Petersson norm $\|\Psi _{\Lambda }(\cdot ,f)\|$ descends to a function on ${\mathcal {M}}_{\Lambda }^{0}$. To express $\tau _{M}$, the Borcherds lift of the following elliptic modular form will be used. Let $\eta (\tau )$ be the Dedekind $\eta$-function and let $\theta _{{\Bbb A}_{1}^{+}}(\tau )$ be the theta series of the $A_{1}$-lattice. We put
This induces the following modular form of type $\rho _{\Lambda }$ (cf. [Reference BorcherdsBor00, Reference ScheithauerSch06, Reference YoshikawaYos13]):
where $|_{\gamma }$ is the Petersson slash operator and $\textbf {e}_{0}\in \textbf {C}[A_{\Lambda }]$ is the vector corresponding to $0\in A_{\Lambda }$. Except for two types, the Borcherds lift of $(2^{g-1}+\delta _{r,10})F_{\Lambda }$ will be used in the expression for $\tau _{M}$, where $\delta _{i,j}$ denotes the Kronecker delta.
On the other hand, Siegel modular forms also yield functions on ${\mathcal {M}}_{\Lambda }^{0}$. The period map for the fixed curves of $2$-elementary $K3$ surfaces induces a holomorphic map
where ${\mathcal {A}}_{g}$ is the Siegel modular variety of degree $g$. Then the pullback of the Petersson norm of a Siegel modular form by $J_{M}$ is a function on ${\mathcal {M}}_{\Lambda }^{0}$. The following Siegel modular forms of weights $2^{g+1}(2^{g}+1)$ and $2(2^{g}-1)(2^{g}+2)$ will be used to express $\tau _{M}$:
where $\theta _{a,b}$ is the Riemann theta constant with even characteristic $(a, b)$. Let $\|\chi _{g}\|^{2}$ and $\|\Upsilon _{g}\|^{2}$ be their Petersson norms. Hence $J_{M}^{*}\|\chi _{g}^{8}\|$ and $J_{M}^{*}\|\Upsilon _{g}\|$ are functions on ${\mathcal {M}}_{\Lambda }^{0}$. For convenience, if $M$ is exceptional, we set $g=1$ and $J_{M}^{*}\|\Upsilon _{g}\|=1$.
The main result of this paper is the following (Theorems 8.6, 9.2, 9.4, 9.5, 9.10).
Theorem 0.1 Let $M\subset {\Bbb L}_{K3}$ be a primitive $2$-elementary Lorentzian sublattice of rank $r$ with orthogonal complement $\Lambda$ and let $\delta \in \{0,1\}$ be the parity of its discriminant form. Then there exists a constant $C_{M}>0$ depending only on $M$ such that the following equality of functions on ${\mathcal {M}}_{\Lambda }^{0}$ holds:
(1) if $(r,\delta )\not =(2,0),(10,0)$, then
\[ \tau_{M}^{-2^{g}(2^{g}+1)} = C_{M} \| \Psi_{\Lambda}(\cdot,2^{g-1}F_{\Lambda}+f_{\Lambda}) \| \cdot J_{M}^{*} \| \chi_{g}^{8} \|; \](2) if $(r,\delta )=(2,0)$ or $(10,0)$, then
\[ \tau_{M}^{-(2^{g}-1)(2^{g}+2)} = C_{M} \| \Psi_{\Lambda}(\cdot,2^{g-1}F_{\Lambda}+f_{\Lambda}) \| \cdot J_{M}^{*} \| \Upsilon_{g} \|. \]
Here $f_{\Lambda }$ is the elliptic modular form of type $\rho _{\Lambda }$ given by $f_{\Lambda }=\delta _{r,10} F_{\Lambda }$ for $r\not =2$ and by (9.5), (9.19) below for $r=2$.
Case (1) covers the majority of types ($67$), while case (2) covers a further $8$ types. The formula for (1) does not hold in case (2) because $J_{M}^{*}\chi _{g}$ vanishes identically there. In [Reference YoshikawaYos04, Reference YoshikawaYos12], the automorphy of $\tau _{M}$ was proved (cf. (0.2) below), but the corresponding modular form was given explicitly only for the exceptional $M$. In [Reference YoshikawaYos13], the elliptic modular form $F_{\Lambda }$ was introduced and (1) was proved for $r\geq 10$. Theorem 0.1 completes this series of work. As a by-product of Theorem 0.1, we prove the quasi-affinity of ${\mathcal {M}}_{\Lambda }^{0}$ when $r>6$ (Theorem 9.16).
The invariant $\tau _{M}$ is closely related to the BCOV invariant $\tau _{\rm BCOV}$ of Calabi–Yau $3$-folds, which was introduced in [Reference Fang, Lu and YoshikawaFLY08] after the prediction of Bershadsky, Cecotti, Ooguri and Vafa [Reference Bershadsky, Cecotti, Ooguri and VafaBCOV94] on the mirror symmetry at genus $1$. On the moduli space of Calabi–Yau $3$-folds of Borcea and Voisin associated to $2$-elementary $K3$ surfaces of type $M$ and elliptic curves, one has the following equality of functions (cf. [Reference Yoshikawa, Bost, Hofer, Labourie, Le Jan, Ma and ZhangYos17]):
where ${\mathcal {C}}_{M}$ is a constant depending only on $M$. By Theorem 0.1 and (0.1), $\tau _{\rm BCOV}$ for Borcea–Voisin $3$-folds of type $M$ is given by the product of the Petersson norms of the modular forms $\Psi _{\Lambda }(\cdot ,F_{\Lambda })$, $\chi _{g}^{8}$ (or $\Upsilon _{g}$) and $\eta$. Since
is the genus $1$ string amplitude $F_{1}$ in the B-model (cf. [Reference Bershadsky, Cecotti, Ooguri and VafaBCOV94, Reference Fang, Lu and YoshikawaFLY08]), Theorem 0.1 gives an exact result of $F_{1}$ in the B-model for all Borcea–Voisin $3$-folds.
In Theorem 0.1 the automorphic form corresponding to $\tau _{M}$ splits into two factors. It is natural to ask if this factorization is realized at the level of holomorphic torsion invariants of $2$-elementary $K3$ surfaces. Thanks to the spin-$1/2$ bosonization formula [Reference Alvarez-Gaumé, Moore and VafaAMV86, Reference Bost and NelsonBN86, Reference FayFay92], we have an affirmative answer to this question. Let us introduce the following twisted version of $\tau _{M}$.
where $\Sigma$ runs over all ineffective theta characteristics on $X^{\iota }$ and $\tau (X^{\iota },\Sigma ;\gamma |_{X^{\iota }})$ is the analytic torsion of $\Sigma$ with respect to $\gamma |_{X^{\iota }}$. It turns out that $\tau _{M}^{{\rm spin}}(X,\iota )$ is independent of the choice of an $\iota$-invariant Ricci-flat Kähler form $\gamma$ and gives rise to a function $\tau _{M}^{{\rm spin}}$ on ${\mathcal {M}}_{\Lambda }^{0}$. Our second result is stated as follows (cf. Theorem 10.2):
Theorem 0.2 Let ${\mathcal {H}}_{\Lambda }\subset {\mathcal {M}}_{\Lambda }$ be the characteristic Heegner divisor (cf. § 2). Then there exists a constant $C'_{M}>0$ depending only on $M$ such that the following equality of functions on ${\mathcal {M}}_{\Lambda }^{0}\setminus {\mathcal {H}}_{\Lambda }$ holds:
We remark that if ${\mathcal {H}}_{\Lambda }\not =\emptyset$, then $\tau _{M}^{{\rm spin}}$ jumps along ${\mathcal {H}}_{\Lambda }$ and thus $\tau _{M}^{{\rm spin}}$ is a discontinuous function on ${\mathcal {M}}_{\Lambda }^{0}$. For an explicit relation between the constants in Theorems 0.1 and 0.2, see § 10.2 below. After Theorem 0.2, it is very likely that $C'_{M}$ can be determined up to an algebraic number by an Arakelov geometric study of $2$-elementary $K3$ surfaces with maximal Picard number. This subject will be discussed elsewhere. Theorem 0.2 can be interpreted as a formula for the equivariant determinant of Laplacian (with a certain correction term) on the space of invariant Ricci-flat metrics on a $K3$ surface with involution. In § 11 we use this interpretation to study an equivariant analogue of Borcherds’ conjecture [Reference BorcherdsBor98].
According to Theorem 0.2, the invariant $\tau _{M}^{{\rm spin}}$ is elliptic modular in the sense that it is the Borcherds lift of an elliptic modular form. When ${\mathcal {D}}_{\Lambda }\not =\emptyset$, the corresponding modular form is determined by $\tau _{M}^{{\rm spin}}$ in a canonical manner (cf. Theorem 10.5). In this way, there is a natural one-to-one correspondence between the holomorphic torsion invariants $\tau _{M}^{{\rm spin}}$ and the elliptic modular forms $2^{g-1}F_{\Lambda }+f_{\Lambda }$. A conceptual account of this unexpected elliptic modularity as well as the geometric meaning of the corresponding modular forms is strongly desired.
Let us explain the outline of the proof of Theorem 0.1. For the sake of simplicity, we explain only the major case $\delta =1$. Since $\tau _{M}$ was determined in [Reference YoshikawaYos13] when $g\leq 2$, we assume $3\leq g\leq 10$. The strategy is summarized as follows.
(a) Reduce Theorem 0.1 to determining the divisor of $J_{M}^{*}\chi _{g}^{8}$.
(b) Determine the support of the divisor of $J_{M}^{*}\chi _{g}^{8}$ for certain key lattices $M_{g,0}$.
(c) Determine inductively from (b) the support of the divisor of $J_{M}^{*}\chi _{g}^{8}$ for all $M$ and prove sharp estimates for its coefficients.
(d) Deduce Theorem 0.1 from the estimates in step (c).
Let us see each step in more detail.
(a) From the theory of Quillen metrics [Reference Bismut, Gillet and SouléBGS88, Reference BismutBis95, Reference MaMa00], the automorphy of $\tau _{M}$ follows [Reference YoshikawaYos04, Reference YoshikawaYos12]. There exist $\ell \in \textbf {Z}_{>0}$ and an automorphic form $\Phi _{M}$ on $\Omega _{\Lambda }^{+}$ with
(We will eventually see that $\ell$ can be taken to be $2^{g-1}(2^{g}+1)$.) By construction, the automorphic form $\Psi _{\Lambda }(\cdot ,2^{g-1} F_{\Lambda })^{\ell }\otimes J_{M}^{*}\chi _{g}^{8\ell }$ has the same weight as $\Phi _{M}^{2^{g-1}(2^{g}+1)}$. Hence by the Koecher principle, it is sufficient to show that
The divisor of $\Psi _{\Lambda }(\cdot ,2^{g-1} F_{\Lambda })$ is calculated by the theory of Borcherds products [Reference BorcherdsBor98], so the problem is reduced to calculating the divisor of $J_{M}^{*}\chi _{g}^{8}$.
Since $\delta =1$, the isometry class of $M$ is determined by $g$ and the number $k+1$ of components of the fixed curve. (See § 3.2 for the formula for the invariants $g=g(M)$ and $k=k(M)$.) Write $M_{g,k}$ for the lattice with these invariants and set $\Lambda _{g,k}=M_{g,k}^{\perp }$. Let ${\mathcal {D}}_{\Lambda _{g,k}}={\mathcal {D}}_{\Lambda _{g,k}}^{+}+{\mathcal {D}}_{\Lambda _{g,k}}^{-}$ be the decomposition according to the type of $(-2)$-vectors of $\Lambda _{g,k}$, and let ${\mathcal {H}}_{\Lambda _{g,k}}$ be the characteristic Heegner divisor (cf. § 2). We will proceed inductively on $k$.
(b) We first study the series $k=0$ by a geometric approach (§ 6). Curves with vanishing theta constants are characterized by the existence of certain half-canonical bundle. By analyzing bi-anticanonical sections of Del Pezzo surfaces with such property, we prove that the support of ${\rm div}(J_{M_{g,0}}^{*}\chi _{g}^{8})$ on ${\mathcal {M}}_{\Lambda _{g,0}}^{0}$ coincides with ${\mathcal {H}}_{\Lambda _{g,0}}$. Hence there exist integers $a_{g}, b_{g}, c_{g}\geq 0$ such that
(c) We have a natural inclusion of lattices $\Lambda _{g,k+1}\subset \Lambda _{g,k}$, which induces an inclusion of domains $i\colon \Omega _{\Lambda _{g,k+1}}\hookrightarrow \Omega _{\Lambda _{g,k}}$. Then we show that $J_{M_{g,k}}\circ i=J_{M_{g,k+1}}$ outside a locus of codimension $2$, that $i^{*}{\mathcal {H}}_{\Lambda _{g,k}}=2{\mathcal {H}}_{\Lambda _{g,k+1}}$ and that a similar relation holds between ${\mathcal {D}}_{\Lambda _{g,k}}^{\pm }$ and ${\mathcal {D}}_{\Lambda _{g,k+1}}^{\pm }$ (cf. §§ 2 and 3). This enables us to inductively extend (0.4) to the case $k\geq 1$ (§ 8):
where $a_{g,k}, b_{g,k}, c_{g,k}$ are integers satisfying
By using a formula in [Reference YoshikawaYos13] and the formula [Reference Teixidor i BigasTei88] for the theta-null divisor on the moduli space of curves, we also prove the estimates $a_{g}\geq 2^{2g-1}$ and $b_{g}\geq 2^{4}$.
(d) Substituting these relations and estimates in the left-hand side of (0.3), we obtain the desired inequality when $g<10$. In the case of $g=10$, an extra argument is required; see § 8. Note that ${\mathcal {H}}_{\Lambda }$ vanishes if $r\geq 10$, which explains why the proof of Theorem 0.1 is much simpler in the range $r\geq 10$, $(r,\delta )\not =(10,0)$ (cf. [Reference YoshikawaYos13]). As a by-product of the proof, we determine an explicit formula for the divisor of $J_{M}^{*}\chi _{g}^{8}$ as a Heegner divisor on $\Omega _{\Lambda }^{+}$ for all $M$, which is a result of independent interest. See Theorem 9.13 for the details.
This paper is organized as follows. Sections 1–5 are mainly preliminaries. In § 6 (respectively, § 7) we study the even theta characteristics of the fixed curve for $2$-elementary $K3$ surfaces with $\delta =1$ (respectively, $\delta =0$). In § 8 (respectively, § 9) we prove Theorem 0.1 when $\delta =1$ (respectively, $\delta =0$). In § 10 we introduce the twisted holomorphic torsion invariant $\tau _{M}^{{\rm spin}}$ and prove Theorem 0.2. In § 11 we study an equivariant analogue of Borcherds’ conjecture.
1. Lattices
A free $\textbf {Z}$-module of finite rank endowed with a non-degenerate, integral, symmetric bilinear form is called a lattice. The rank and signature of a lattice $L$ are denoted by $r(L)$ and ${\rm sign}(L)=(b^{+}(L),b^{-}(L))$, respectively. For a lattice $L=(\textbf {Z}^{r},\langle \cdot ,\cdot \rangle )$ and an integer $k\in \textbf {Z}\setminus \{0\}$, we define $L(k):=(\textbf {Z}^{r},k\langle \cdot ,\cdot \rangle )$. The group of isometries of $L$ is denoted by $O(L)$. The set of roots of $L$ is defined as $\Delta _{L}:=\{d\in L; \langle d,d\rangle =-2\}$. For the root systems $A_{k}$, $D_{k}$, $E_{k}$, the corresponding negative-definite root lattices are denoted by ${\Bbb A}_{k}$, ${\Bbb D}_{k}$, ${\Bbb E}_{k}$, respectively. We set ${\Bbb A}_{1}^{+}:={\Bbb A}_{1}(-1)$, etc. The hyperbolic plane ${\Bbb U}$ is the even unimodular lattice of signature $(1,1)$.
For an even lattice $L$, its dual lattice is denoted by $L^{\lor }$. The finite abelian group $A_{L}:=L^{\lor }/L$ is called the discriminant group of $L$, which is equipped with the $\textbf {Q}/2\textbf {Z}$-valued quadratic form $q_{L}$ called the discriminant form and the $\textbf {Q}/\textbf {Z}$-valued bilinear form $b_{L}$ called the discriminant bilinear form. The automorphism group of $(A_{L},q_{L})$ is denoted by $O(q_{L})$.
A lattice $L$ is $2$-elementary if there exists $l\in \textbf {Z}_{\geq 0}$ with $A_{L}\cong (\textbf {Z}/2\textbf {Z})^{\oplus l}$. For a $2$-elementary lattice $L$, we set $l(L):=\dim _{\mathbf{F}_{2}}A_{L}$. When $L$ is an even $2$-elementary lattice, the parity of $q_{L}$ is denoted by $\delta (L)\in \{0,1\}$. By Nikulin [Reference NikulinNik80, Theorem 3.6.2], the isometry class of an indefinite even $2$-elementary lattice $L$ is determined by the triplet $({\rm sign}(L),l(L),\delta (L))$. For an even $2$-elementary lattice $L$, there is a unique element $\textbf {1}_{L}\in A_{L}$, called the characteristic element, such that $b_{L}(\gamma ,\textbf {1}_{L})=q_{L}(\gamma )\mod \textbf {Z}$ for all $\gamma \in A_{L}$. Then $g(\textbf {1}_{L})=\textbf {1}_{L}$ for all $g\in O(q_{L})$. By definition, $\textbf {1}_{L}=0$ if and only if $\delta (L)=0$.
The $K3$ lattice is defined as the even unimodular lattice of signature $(3,19)$:
It is classical that ${\Bbb L}_{K3}$ is isometric to the second integral cohomology lattice of a $K3$ surface. For a sublattice $\Lambda \subset {\Bbb L}_{K3}$, we define $\Lambda ^{\perp }:=\{l\in {\Bbb L}_{K3};\,\langle l,\Lambda \rangle =0\}$.
A primitive $2$-elementary Lorentzian sublattice of ${\Bbb L}_{K3}$ isometric to ${\Bbb U}(2)\oplus {\Bbb E}_{8}(2)$ is said to be exceptional. (For the reason why this is exceptional, see § 3.2 below.) Its orthogonal complement in ${\Bbb L}_{K3}$, namely ${\Bbb U}\oplus {\Bbb U}(2)\oplus {\Bbb E}_{8}(2)$, is also said to be exceptional.
Proposition 1.1 The isometry classes of primitive $2$-elementary sublattices $\Lambda$ of ${\Bbb L}_{K3}$ with signature $(2,r(\Lambda )-2)$ consist of the $75$ classes in Table 1, where
Proof. See, for example, [Reference Finashin and KharlamovFK08, Tables 2 and 3, pp. 705–706]. Notice that the representative of each isometry class is not necessarily identical to the one in [Reference Finashin and KharlamovFK08, Tables 2 and 3, pp. 705–706].
For a primitive $2$-elementary Lorentzian sublattice $M\subset {\Bbb L}_{K3}$ and a root $d\in \Delta _{M^{\perp }}$, the smallest sublattice of ${\Bbb L}_{K3}$ containing $M$ and $d$ is given by
Then $[M\perp d]$ is again a primitive $2$-elementary Lorentzian sublattice of ${\Bbb L}_{K3}$ with $[M\perp d]^{\perp }=M^{\perp }\cap d^{\perp }$.
2. Domains of type IV and modular varieties of orthogonal type
In this section $\Lambda$ is assumed to be an even $2$-elementary lattice with ${\rm sign}(\Lambda )=(2,r(\Lambda )-2)$. We define the complex manifold $\Omega _{\Lambda }$ with projective $O(\Lambda )$-action by
Then $\Omega _{\Lambda }$ has two connected components $\Omega _{\Lambda }^{\pm }$, each of which is isomorphic to a bounded symmetric domain of type IV of dimension $r(\Lambda )-2$. The orthogonal modular variety ${\mathcal {M}}_{\Lambda }$ associated to $\Lambda$ is defined as the analytic space
where
We denote by ${\mathcal {M}}_{\Lambda }^{*}$ the Baily–Borel–Satake compactification of ${\mathcal {M}}_{\Lambda }$, which is an irreducible normal projective variety of dimension $r(\Lambda )-2$ with $\dim ({\mathcal {M}}_{\Lambda }^{*}\setminus {\mathcal {M}}_{\Lambda })\leq 1$.
2.1 Discriminant locus
For $\lambda \in \Lambda$ with $\langle \lambda ,\lambda \rangle <0$, we define
Then $H_{\lambda }$ is a non-zero divisor on $\Omega _{\Lambda }$. For any root $d\in \Delta _{\Lambda }$, we have the relation
The discriminant locus of $\Omega _{\Lambda }$ is the reduced divisor of $\Omega _{\Lambda }$ defined by
We define the $O(\Lambda )$-invariant Zariski open subset $\Omega _{\Lambda }^{0}$ of $\Omega _{\Lambda }$ by
We set
2.2 Some subloci of ${\mathcal {D}}_{\Lambda }$
We define the decomposition $\Delta _{\Lambda }=\Delta ^{+}_{\Lambda }\amalg \Delta ^{-}_{\Lambda }$ by
Then $\Delta ^{\pm }_{\Lambda }$ are $O(\Lambda )$-invariant. We define the $O(\Lambda )$-invariant reduced divisors ${\mathcal {D}}^{\pm }_{\Lambda }$ on $\Omega _{\Lambda }$ and the corresponding divisors $\bar {\mathcal {D}}^{\pm }_{\Lambda }$ on ${\mathcal {M}}_{\Lambda }$ by
Proposition 2.1 Let $d\in \Delta _{\Lambda }^{+}$. Let $i\colon \Omega _{\Lambda \cap d^{\perp }}=H_{d}\hookrightarrow \Omega _{\Lambda }$ be the inclusion. Then the following equalities of divisors on $\Omega _{\Lambda \cap d^{\perp }}$ hold:
Proof. Let $\delta \in \Delta _{\Lambda }\setminus \{\pm d\}$. Then $H_{d}\cap H_{\delta }\not =\emptyset$ if and only if $L:=\textbf {Z}d+\textbf {Z}\delta$ is negative-definite. Since the Gram matrix of $L$ with respect to the basis $\{d,\delta \}$ is given by $-2\binom {1\,a}{a\,1}$ where $a=-\langle \delta ,d/2\rangle \in \textbf {Z}$, we conclude that $H_{d}\cap H_{\delta }\not =\emptyset$ if and only if $a=0$, that is, $\delta \in \Delta _{\Lambda }\cap d^{\perp }=\Delta _{\Lambda \cap d^{\perp }}$. Since $\Delta _{\Lambda }^{\pm }\cap d^{\perp }=\Delta _{\Lambda \cap d^{\perp }}^{\pm }$, we get
This proves the result.
For $d\in \Delta _{\Lambda }$, we define a non-empty Zariski open subset $H^{0}_{d}\subset H_{d}$ by
We set
Then $\Omega _{\Lambda }^{0}\amalg {\mathcal {D}}_{\Lambda }^{0}$ is a Zariski open subset of $\Omega _{\Lambda }$, whose complement has codimension at least $2$ when $r(\Lambda )\geq 4$ and is empty when $r(\Lambda )\leq 3$.
2.3 Characteristic Heegner divisor
Set
We define the characteristic Heegner divisor of $\Omega _{\Lambda }$ as the reduced divisor
where $[\lambda ]:=\lambda +\Lambda \in A_{\Lambda }$. Since $\textbf {1}_{\Lambda }$ is $O(q_{\Lambda })$-invariant, ${\mathcal {H}}_{\Lambda }$ is $O(\Lambda )$-invariant. Since $\varepsilon _{\Lambda }\geq 0$ when $r(\Lambda )\leq 12$, we get ${\mathcal {H}}_{\Lambda }=0$ if $r(\Lambda )\leq 12$.
Proposition 2.2 Let $d\in \Delta _{\Lambda }^{+}$. Let $i\colon \Omega _{\Lambda \cap d^{\perp }}=H_{d}\hookrightarrow \Omega _{\Lambda }$ be the inclusion. Then the following equality of divisors on $\Omega _{\Lambda \cap d^{\perp }}$ holds:
Proof. Since $r(\Lambda )\leq 21$, we get $\varepsilon _{\Lambda }\geq -\frac {9}{2}$. Set $\Lambda ':=\Lambda \cap d^{\perp }$. Since $d\in \Delta _{\Lambda }^{+}$, we deduce from [Reference Finashin and KharlamovFK08, Proposition 3.1] that $\Lambda$ and $\Lambda '\oplus \textbf {Z}d$ have the same invariants $(r,l,\delta )$. Hence we get the orthogonal decomposition $\Lambda =\Lambda '\oplus \textbf {Z}d$. Let $\lambda \in \Lambda ^{\lor }$ be such that $\lambda ^{2}=\varepsilon _{\Lambda }$ and $[\lambda ]=\textbf {1}_{\Lambda }$. Then we can write $\lambda =\lambda '+a(d/2)$, where $\lambda '\in (\Lambda ')^{\lor }$ and $a=-\langle \lambda ,d\rangle \in \textbf {Z}$. Since $[\lambda ]=\textbf {1}_{\Lambda }$, we get $a\equiv 1\bmod 2$. Hence $a=2k+1$ for some $k\in \textbf {Z}$ and $\lambda '\in \textbf {1}_{\Lambda '}$. Since $H_{\lambda '}\not =\emptyset$ if and only if $(\lambda ')^{2}<0$, we get $i^{*}H_{\lambda }=H_{\lambda '}\not =\emptyset$ if and only if $0>(\lambda ')^{2}=\lambda ^{2}+a^{2}/2=(\varepsilon _{\Lambda }+\frac {1}{2})+2k(k+1)$. Since $\varepsilon _{\Lambda }\geq -\frac {9}{2}$ and hence $-2\leq ({-1-\sqrt {-2\varepsilon _{\Lambda }}})/{2}< ({-1+\sqrt {-2\varepsilon _{\Lambda }}})/{2}\leq 1$, we see that $(\varepsilon _{\Lambda }+\frac {1}{2})+2k(k+1)<0$ if and only if $k=0,-1$. This proves that
When $\lambda =\lambda '\pm (d/2)$, we get $(\lambda ')^{2}=\varepsilon _{\Lambda }+\frac {1}{2}=\varepsilon _{\Lambda '}$. This, together with (2.2), yields that
This proves the proposition.
3. $2$-elementary $K3$ surfaces and the Torelli map
3.1 $2$-elementary $K3$ surfaces
A $K3$ surface $X$ equipped with a holomorphic involution $\iota \colon X\to X$ is called a $2$-elementary $K3$ surface if $\iota$ is anti-symplectic:
The possible deformation types of $2$-elementary $K3$ surfaces were determined by Nikulin (see [Reference Alexeev and NikulinAN06, § 2.3] and the references therein). Let $\alpha \colon H^{2}(X,\textbf {Z})\cong {\Bbb L}_{K3}$ be an isometry of lattices. Set $H^{2}(X,\textbf {Z})_{\pm }:=\{l\in H^{2}(X,\textbf {Z});\,\iota ^{*}l=\pm l\}$ and
Then $M\subset {\Bbb L}_{K3}$ must be a primitive $2$-elementary Lorentzian sublattice. Conversely, for any primitive $2$-elementary Lorentzian sublattice $M\subset {\Bbb L}_{K3}$, there exists a $2$-elementary $K3$ surface with (3.2). For a $2$-elementary $K3$ surface $(X,\iota )$, the isometry class of $H^{2}(X,\textbf {Z})_{+}$ is called the type of $(X,\iota )$. By an abuse of notation, the sublattice itself $\alpha (H^{2}(X,\textbf {Z})_{+})\subset {\Bbb L}_{K3}$ is also called the type of $(X,\iota )$. Then there is a one-to-one correspondence between the deformation types of $2$-elementary $K3$ surfaces and the triplets $(r,l,\delta )$. Since the latter consists of $75$ points, there exist $75$ mutually distinct deformation types of $2$-elementary $K3$ surfaces. For a given primitive $2$-elementary Lorentzian sublattice $M\subset {\Bbb L}_{K3}$, the moduli space of $2$-elementary $K3$ surfaces of type $M$ is given as follows.
Let $(X,\iota )$ be a $2$-elementary $K3$ surface of type $M$ and let $\alpha \colon H^{2}(X,\textbf {Z})\cong {\Bbb L}_{K3}$ be an isometry satisfying (3.2). Since $H^{2,0}(X,\textbf {C})\subset H^{2}(X,\textbf {Z})_{-}\otimes \textbf {C}$ by (3.1), we get
Its $O(\Lambda )$-orbit is called the Griffiths period of $(X,\iota )$ and is denoted by
By [Reference YoshikawaYos04, Theorem 1.8] and [Reference YoshikawaYos13, Proposition 11.2], the coarse moduli space of $2$-elementary $K3$ surfaces of type $M$ is isomorphic to ${\mathcal {M}}_{\Lambda }^{0}$ via the period map $\bar {\pi }_{M}$. In the rest of this paper we identify the point $\bar {\pi }_{M}(X,\iota )\in {\mathcal {M}}_{\Lambda }^{0}$ with the isomorphism class of $(X,\iota )$.
3.2 The Torelli map for $2$-elementary $K3$ surfaces
3.2.1 The set of fixed points
For a $2$-elementary $K3$ surface $(X,\iota )$ of type $M$, set
Then $X^{\iota }=\emptyset$ if and only if $M$ is exceptional, that is, $M\cong {\Bbb U}(2)\oplus {\Bbb E}_{8}(2)$. In this case, the quotient $X/\iota$ is an Enriques surface and $(X,\iota )$ is the universal covering of an Enriques surface endowed with the non-trivial covering transformation. When $M$ is non-exceptional, by Nikulin [Reference NikulinNik83, Theorem 4.2.2], we have
for $M\not \cong {\Bbb U}(N)\oplus {\Bbb E}_{8}(2)$ $(N=1,2)$ and we have $X^{\iota }=C_{1}^{(1)}\amalg C_{2}^{(1)}$ for $M\cong {\Bbb U}\oplus {\Bbb E}_{8}(2)$, where $C^{(g)}$ is a projective curve of genus $g$ and $E_{i}\cong \mathbf{P}^{1}$ and
Since $r(\Lambda )=22-r(M)$ and $l(\Lambda )=l(M)$, we have the relations
As we defined $g(\Lambda )=(r(\Lambda )-l(\Lambda ))/2$ in Proposition 1.1, we have $g(M)=g(\Lambda )$. Notice that, when $M\cong {\Bbb U}(2)\oplus {\Bbb E}_{8}(2)$ and hence $X^{\iota }$ is empty, $g(M)=g(\Lambda )=1$ has no geometric meaning.
3.2.2 The Torelli map
For $g\geq 0$, let ${\frak S}_{g}$ be the Siegel upper half-space of degree $g$ and let ${\rm Sp}_{2g}(\textbf {Z})$ be the symplectic group of degree $2g$ over $\textbf {Z}$. We define
The Satake compactification of ${\mathcal {A}}_{g}$ is denoted by ${\mathcal {A}}_{g}^{*}$.
For a $2$-elementary $K3$ surface $(X,\iota )$ of type $M$, the period of $X^{\iota }$ is denoted by $\Omega (X^{\iota })\in {\mathcal {A}}_{g(M)}$. We define a map $\bar {J}_{M}^{0}\colon {\mathcal {M}}_{\Lambda }^{0}\to {\mathcal {A}}_{g(M)}$ by
Let $\Pi _{\Lambda }\colon \Omega _{\Lambda }\to {\mathcal {M}}_{\Lambda }$ be the projection. The Torelli map is the $O(\Lambda )$-equivariant holomorphic map $J_{M}^{0}\colon \Omega _{\Lambda }^{0}\to {\mathcal {A}}_{g(M)}$ defined by
In Theorem 3.3 below, we will extend $J_{M}^{0}$ to a certain Zariski open subset of $\Omega _{\Lambda }$ containing $\Omega _{\Lambda }^{0}\cup {\mathcal {D}}_{\Lambda }^{0}$ and prove its compatibility with respect to the inclusion $\Omega _{\Lambda \cap d^{\perp }}=H_{d}\hookrightarrow \Omega _{\Lambda }$ for $d\in \Delta _{\Lambda }$. For this, we introduce a stratification of $\Omega _{\Lambda }$.
3.2.3 A stratification of $\Omega _{\Lambda }$
For a primitive sublattice $L\subset \Lambda$ generated by $\Delta _{L}$, we define
Then $H_{L}\not =\emptyset$ if and only if $L$ is negative-definite, that is, $L$ is a root lattice. If $H_{L}\not =\emptyset$, then $H_{L}^{0}$ is a non-empty dense Zariski open subset of $H_{L}$. By definition, it is obvious that if $r(L)=r(L')$ and $L\not =L'$, then $H_{L}^{0}\cap H_{L'}^{0}=\emptyset$. Set
where $L$ runs over the set of all primitive root sublattices of $\Lambda$ of rank $k$. Then $\Omega _{\Lambda }^{\geq k}$ is a Zariski closed subset of $\Omega _{\Lambda }$ of pure codimension $k$, and $\Omega _{\Lambda }\setminus \Omega _{\Lambda }^{\geq k+1}$ is a dense Zariski open subset of $\Omega _{\Lambda }$, whose complement has codimension $k+1$. We have
For a root lattice ${\Bbb K}$, let ${\Bbb K}(\Lambda )$ be the set of primitive sublattices of $\Lambda$ isometric to ${\Bbb K}$. Since a root lattice of rank $2$ is either ${\Bbb A}_{1}^{\oplus 2}$ or ${\Bbb A}_{2}$, we have
We set
Then ${\mathcal {D}}_{\Lambda }^{1,+}\amalg {\mathcal {D}}_{\Lambda }^{1,-}=\Omega _{\Lambda }^{2}$ and we have the stratification
such that ${\mathcal {D}}_{\Lambda }\setminus \Omega _{\Lambda }^{\geq 3}={\mathcal {D}}_{\Lambda }^{0}\amalg {\mathcal {D}}_{\Lambda }^{1,+}\amalg {\mathcal {D}}_{\Lambda }^{1,-}$ is a divisor of $\Omega _{\Lambda }\setminus \Omega _{\Lambda }^{\geq 3}$.
Lemma 3.1 If $d\in \Delta _{\Lambda }^{+}$, then $H_{d}\setminus \Omega _{\Lambda }^{\geq 3}=\Omega _{\Lambda \cap d^{\perp }}^{0}\amalg {\mathcal {D}}_{\Lambda \cap d^{\perp }}^{0}$.
Proof. The proof consists of four steps.
Step 1. By the definition of ${\mathcal {D}}_{\Lambda }^{0}$, we have $H_{d}^{0}=H_{d}\cap {\mathcal {D}}_{\Lambda }^{0}$. Since $d\in \Delta _{\Lambda }^{+}$, we have $H_{d}\cap H_{\delta }\not =\emptyset$ for $\delta \in \Delta _{\Lambda }\setminus \{\pm d\}$ if and only if $\delta \in \Delta _{\Lambda \cap d^{\perp }}$. Hence $H_{d}\cap {\mathcal {D}}_{\Lambda }^{0}=H_{d}^{0}=H_{d}\setminus \bigcup _{\delta \in \Delta _{\Lambda }\setminus \{\pm d\}}H_{\delta } =H_{d}\setminus \bigcup _{\delta \in \Delta _{\Lambda \cap d^{\perp }}}H_{\delta }=\Omega _{\Lambda \cap d^{\perp }}^{0}$.
Step 2. Assume $H_{d}\cap {\mathcal {D}}_{\Lambda }^{1,-}\not =\emptyset$. Then there exist $L\in {\Bbb A}_{2}(\Lambda )$ and $[\eta ]\in H_{d}\cap H_{L}^{0}$. By the definition of $H_{L}^{0}$, we have $d\in \Delta _{L}$. Since $L\cong {\Bbb A}_{2}$, there exists $\delta \in \Delta _{L}$ with $\langle d,\delta \rangle =1$. Since $d\in \Delta _{\Lambda }^{+}$, this yields the contradiction $\langle d,\delta \rangle =2\langle (d/2),\delta \rangle \in 2\textbf {Z}$. This proves $H_{d}\cap {\mathcal {D}}_{\Lambda }^{1,-}=\emptyset$.
Step 3. Assume $H_{d}\cap {\mathcal {D}}_{\Lambda }^{1,+}\not =\emptyset$. Then there exists $L\in {\Bbb A}_{1}^{\oplus 2}(\Lambda )$ with $[\eta ]\in H_{d}\cap H_{L}^{0}$. By the definition of $H_{L}^{0}$, we get $d\in \Delta _{L}$. Hence $\Delta _{L}=\{\pm d,\pm \delta \}$ for some $\delta \in \Delta _{\Lambda \cap d^{\perp }}$. Thus
Since $d\in \Delta _{\Lambda }^{+}$, the third equality follows from the fact that $H_{d}\cap H_{\epsilon }\not =\emptyset$ for $\epsilon \in \Delta _{\Lambda }\setminus \{\pm d\}$ if and only if $\epsilon \in \Delta _{\Lambda \cap d^{\perp }}$. This proves $H_{d}\cap {\mathcal {D}}_{\Lambda }^{1,+}={\mathcal {D}}_{\Lambda \cap d^{\perp }}^{0}$.
Step 4. Since $H_{d}\cap \Omega _{\Lambda }^{0}=\emptyset$, the result follows from steps 1–3 and (3.4).
3.2.4 The local structure of $\Omega _{\Lambda }^{0}$ near ${\mathcal {D}}_{\Lambda }^{1}$
Let $\Delta \subset \textbf {C}$ be the unit disc and set $\Delta ^{*}:=\Delta \setminus \{0\}$. Let $L_{0}\subset \Delta ^{2}$ be the diagonal locus and set $L_{1}:=\{0\}\times \Delta$ and $L_{2}:=\Delta \times \{0\}$. Then $L_{0}$, $L_{1}$, $L_{2}$ are lines with $L_{i}\cap L_{j}=\{0\}$ for any $i\not =j$. We have $(\Delta ^{*})^{2}=\Delta ^{2}\setminus (L_{1}\cup L_{2})$. Set $L_{i}^{*}:=L_{i}\setminus \{(0,0)\}$.
Lemma 3.2 Let $[\eta ]\in {\mathcal {D}}_{\Lambda }^{1,+}\amalg {\mathcal {D}}_{\Lambda }^{1,-}$ and set $n:=\dim \Omega _{\Lambda }$. Then the following statements hold.
(1) If $[\eta ]\in {\mathcal {D}}_{\Lambda }^{1,+}$, then there is a neighborhood $U$ of $[\eta ]$ in $\Omega _{\Lambda }\setminus \Omega _{\Lambda }^{\geq 3}$ such that
\[ U\cap\Omega_{\Lambda}^{0}\cong(\Delta^{*})^{2}\times\Delta^{n-2}, \quad U\cap{\mathcal{D}}_{\Lambda}^{0}\cong(L_{1}^{*}\amalg L_{2}^{*})\times\Delta^{n-2}, \quad U\cap{\mathcal{D}}_{\Lambda}^{1,+}\cong\{(0,0)\}\times\Delta^{n-2}. \](2) If $[\eta ]\in {\mathcal {D}}_{\Lambda }^{1,-}$, then there is a neighborhood $U$ of $[\eta ]$ in $\Omega _{\Lambda }\setminus \Omega _{\Lambda }^{\geq 3}$ such that $U\setminus {\mathcal {D}}_{\Lambda }\cong (\Delta ^{2}\setminus L_{0}\cup L_{1}\cup L_{2})\times \Delta ^{n-2}$.
Proof. Since $[\eta ]\in {\mathcal {D}}_{\Lambda }^{1,+}\amalg {\mathcal {D}}_{\Lambda }^{1,-}$, there exist $L\in {\Bbb A}_{1}^{\oplus 2}(\Lambda )\amalg {\Bbb A}_{2}(\Lambda )$ and a neighborhood $U$ of $[\eta ]$ in $\Omega _{\Lambda }\setminus \Omega _{\Lambda }^{\geq 3}$ such that $U\cap {\mathcal {D}}_{\Lambda }=U\cap \bigcup _{d\in \Delta _{L}}H_{d}$.
(1) Assume $[\eta ]\in {\mathcal {D}}_{\Lambda }^{1,+}$. Then $L\in {\Bbb A}_{1}^{\oplus 2}(\Lambda )$. There exist $d_{1},d_{2}\in \Delta _{L}$ with $\langle d_{1},d_{2}\rangle =0$ such that $U\cap {\mathcal {D}}_{\Lambda }=U\cap (H_{d_{1}}\cup H_{d_{2}})$. Replacing $U$ by a smaller neighborhood if necessary, there exist a system of coordinates $(z_{1},z_{2},w)$, $w=(w_{1},\ldots ,w_{n-2})$, on $U$ such that $H_{d_{1}}={\rm div}(z_{1})$, $H_{d_{2}}={\rm div}(z_{2})$. The isomorphism $\psi \colon (U,[\eta ])\cong (\Delta ^{n},0)$ induced by $(z_{1},z_{2},w)$ has the desired property.
(2) Assume $[\eta ]\in {\mathcal {D}}_{\Lambda }^{1,-}$. Since $L\in {\Bbb A}_{2}(\Lambda )$, there exist $d_{0},d_{1},d_{2}\in \Delta _{L}$ with $d_{0}=d_{1}+d_{2}$ and $\langle d_{1},d_{2}\rangle =1$ such that $U\cap {\mathcal {D}}_{\Lambda }=U\cap (H_{d_{0}}\cup H_{d_{1}}\cup H_{d_{2}})$. Replacing $U$ by a smaller neighborhood if necessary, there exist a system of coordinates $(z_{1},z_{2},w)$ on $U$ such that $H_{d_{1}}={\rm div}(z_{1})$, $H_{d_{2}}={\rm div}(z_{2})$, $H_{d_{0}}={\rm div}(z_{1}+z_{2})$. The isomorphism $\psi \colon (U,[\eta ])\cong (\Delta ^{n},0)$ induced by $(z_{1},z_{2},w)$ has the desired property.
3.2.5 Inclusion of lattices and the Torelli map
Recall that for $d\in \Delta _{\Lambda }^{+}$, the equality of sets $H_{d}\setminus \Omega _{\Lambda }^{\geq 3}=\Omega _{\Lambda \cap d^{\perp }}^{0}\cup {\mathcal {D}}_{\Lambda \cap d^{\perp }}^{0}$ follows from Lemma 3.1. By Lemma 3.2 (1), $\Omega _{\Lambda }^{0}\cup {\mathcal {D}}_{\Lambda }^{0}\cup {\mathcal {D}}_{\Lambda }^{1,+}$ is a Zariski open subset of $\Omega _{\Lambda }\setminus \Omega _{\Lambda }^{\geq 3}$.
Theorem 3.3 $J_{M}^{0}$ extends to a holomorphic map from $\Omega _{\Lambda }^{0}\cup {\mathcal {D}}_{\Lambda }^{0}\cup {\mathcal {D}}_{\Lambda }^{1,+}$ to ${\mathcal {A}}_{g}^{*}$.
Proof. Set $n:=\dim \Omega _{\Lambda }$. Let $[\eta ]\in {\mathcal {D}}_{\Lambda }^{0}\cup {\mathcal {D}}_{\Lambda }^{1,+}$. By Lemma 3.2 (1), there is a neighborhood $U$ of $[\eta ]$ in $\Omega _{\Lambda }$ such that either $U\setminus ({\mathcal {D}}_{\Lambda }^{0}\cup {\mathcal {D}}_{\Lambda }^{1,+})\cong \Delta ^{*}\times \Delta ^{n-1}$ or $U\setminus ({\mathcal {D}}_{\Lambda }^{0}\cup {\mathcal {D}}_{\Lambda }^{1,+})\cong (\Delta ^{*})^{2}\times \Delta ^{n-2}$. By Borel [Reference BorelBor72], $J_{M}^{0}$ extends to a holomorphic map from $U$ to ${\mathcal {A}}_{g}^{*}$. Since $[\eta ]\in {\mathcal {D}}_{\Lambda }^{0}\cup {\mathcal {D}}_{\Lambda }^{1,+}$ is an arbitrary point, we get the result.
Remark 3.4 By Lemma 3.2 (2), Borel's extension theorem does not apply to $J_{M}^{0}$ near ${\mathcal {D}}_{\Lambda }^{1,-}$. This explains why $J_{M}$ does not extend to $\Omega _{\Lambda }\setminus \Omega _{\Lambda }^{\geq 3}$ in general.
Denote the extension of $J_{M}^{0}$ by
and call it again the Torelli map. By [Reference YoshikawaYos13, Theorem 2.5], the equality
holds for all $d\in \Delta _{\Lambda }$. The following refinement is crucial for the proof Theorem 0.1.
Theorem 3.5 If $d\in \Delta _{\Lambda }^{+}$, then
Proof. Since both $J_{M}|_{H_{d}\setminus \Omega _{\Lambda }^{\geq 3}}$ and $J_{[M\perp d]}|_{\Omega _{\Lambda \cap d^{\perp }}^{0}\cup {\mathcal {D}}_{\Lambda \cap d^{\perp }}^{0}}$ are holomorphic maps from $\Omega _{\Lambda \cap d^{\perp }}^{0}\cup {\mathcal {D}}_{\Lambda \cap d^{\perp }}^{0}$ to ${\mathcal {A}}_{g}^{*}$ by Lemma 3.1 and Theorem 3.3, it suffices to prove the equality on $\Omega _{\Lambda \cap d^{\perp }}^{0}$. Since this was proved in [Reference YoshikawaYos13, Theorem 2.5], we get the result.
3.3 Hyperelliptic linear system
This subsection is the technical basis for §§ 6 and 7. We advise the reader to skip it for the moment and return when necessary. We prepare some tools to realize a given 2-elementary $K3$ surface as a double cover of $\mathbf{P}^2$ or a Hirzebruch surface.
We will use the following notation. For $n\geq 0$ let
be the $n$th Hirzebruch surface, equipped with the natural projection $\pi :\mathbf{F}_{n}\to \mathbf{P}^1$. When $n>0$, denote by $\Sigma \subset \mathbf{F}_{n}$ its unique $(-n)$-section. We write $L_{a,b}$ for the line bundle on $\mathbf{F}_{n}$ of $\pi$-degree $a$ with $(L_{a,b}, \Sigma )=b$. In particular, we have $\pi ^{\ast }\mathcal {O}_{\mathbf{P}^1}(1)\simeq L_{0,1}$, $\mathcal {O}_{\mathbf{F}_{n}}(\Sigma )\simeq L_{1,-n}$ and $K_{\mathbf{F}_{n}}\simeq L_{-2,-2+n}$.
Let $(X, \iota )$ be a 2-elementary $K3$ surface. A line bundle $L$ on $X$ with $(L, L)=2d>0$ is called hyperelliptic [Reference Saint-DonatSai74] if the linear system $|L|$ contains a smooth hyperelliptic member. In that case, $L$ is base point free and every smooth member of $|L|$ is hyperelliptic of genus $d+1$. The associated morphism
is generically two-to-one onto its image, mapping a smooth member of $|L|$ to a rational normal curve in a hyperplane of $|L|^{\vee }$. According to Saint-Donat [Reference Saint-DonatSai74, § 5], we have the following possibilities for the image surface $\phi _L(X)$.
(I) $\phi _L(X)$ coincides with $|L|^{\vee }\simeq \mathbf{P}^2$: this is the case $d=1$.
(II) $\phi _L(X)$ is a Veronese surface in $|L|^{\vee }\simeq \mathbf{P}^5$: this happens when $L=2L'$ for $L'\in {\rm Pic}(X)$ with $(L', L')=2$.
(III) $\phi _L(X)$ is a rational normal scroll, that is, the embedding image of $\mathbf{F}_n$ by a line bundle $L_{1,m}$ with $m>0$ and $n+2m=d$.
(IV) $\phi _L(X)$ is a cone over a rational normal curve, that is, the image of $\mathbf{F}_d$ by the bundle $L_{1,0}$. In this case $\phi _L$ lifts to a morphism $X\to \mathbf{F}_d$, and we must have $2\leq d\leq 4$.
We now assume that the hyperelliptic bundle $L$ is $\iota$-invariant, in the sense that there exists an isomorphism
Although $\iota$ may not necessarily act on $L$ equivariantly, it does so on the morphism (3.6). We then obtain an $\iota$-equivariant morphism
with $Y=\mathbf{P}^2$ in cases (I) and (II), and $Y=\mathbf{F}_n$, $\mathbf{F}_d$ in cases (III) and (IV).
It will be useful to have a purely lattice-theoretic method for finding an $\iota$-invariant hyperelliptic bundle. This is based on the following lemmas. Denote by $H_+=H^2(X, \textbf {Z})_+$ the invariant lattice of $(X, \iota )$.
Lemma 3.6 Let $L\in H_+$ be nef with $(L, L)=2d>0$. Assume that:
(a) there exists $E\in {\rm Pic}(X)$ with $(E, E)\geq 0$ and $(E, L)=2$; and
(b) there is no $F\in H_+$ with $(F, F)=0$ and $(F, L)=1$.
Then $L$ is hyperelliptic.
Proof. We first show that $L$ is base point free. Otherwise, by [Reference MayerMay72, Proposition 5] the linear system $|L|$ would be of the form $|(d+1)F|+\Gamma$ where $F$ is a smooth elliptic curve and $\Gamma$ is a $(-2)$-curve with $(F, \Gamma )=1$. Since $\iota$ acts on $|L|$, the class of $F$ is $\iota$-invariant and then would violate assumption $({\rm b})$. Hence $L$ is free, and a general member $C\in |L|$ is smooth and irreducible of genus $d+1$. We show that $C$ is hyperelliptic. Since $g(C)=2$ when $d=1$, we may assume $d>1$. Let $E$ be a divisor as in assumption $({\rm a})$. Since $h^0(-E)=0$, we have $h^0(E)\geq 2$ by the Riemann–Roch inequality. Consider the exact sequence
Since $(E-L, L)<0$, we have $h^0(E-L)=0$ by the nefness of $L$. Hence $h^0(E|_C)\geq h^{0}(E)\geq 2$ and $E|_C$ gives a $g^1_2$ on $C$.
Conditions (a) and (b) are purely arithmetic. It is always possible to meet the nefness condition by the following lemma.
Lemma 3.7 Let $W(X)$ be the Weyl group of ${\rm Pic}(X)$ generated by the reflections with respect to $(-2)$-vectors in ${\rm Pic}(X)$. Let $L\in H_+$ be a line bundle with $(L, L)\geq 0$ and $(L, L_0)>0$ for some ample class $L_0$. Then there exists $w\in W(X)$ such that $w\circ \iota =\iota \circ w$ and $w(L)$ is nef.
Proof. The same argument as in [Reference Barth, Hulek, Peters and Van de VenBHPV04, Proposition 21.1] applies with a few minor modifications. We leave this to the reader.
Thus we can obtain an $\iota$-equivariant morphism (3.7) by just finding a vector in $H_+$ with the arithmetic conditions (a) and (b). The $\iota$-equivariant Weyl group action as in Lemma 3.7 would then carry this vector to a class of hyperelliptic bundles.
We are interested in when $\iota$ acts by the covering transformation of $\phi :X\to Y$, in which case $(X, \iota )$ may be recovered from $Y$ and the branch curve of $\phi$. Let $g$ denote the genus of the main component of the fixed curve $X^{\iota }$.
Lemma 3.8 The involution $\iota$ acts trivially on $Y$ when:
(1) $g\geq 3$ in the case of $Y=\mathbf{P}^2$;
(2) $g\geq 4$ in the case of $Y=\mathbf{P}^1\times \mathbf{P}^1$; and
(3) $g\geq n+2$ in the case of $Y=\mathbf{F}_n$ with $n>0$.
Proof. We may assume that $g>0$. Let $B\subset Y$ be the branch curve of $\phi$, which belongs to $|{-}2K_Y|$. Suppose that $\iota$ acts non-trivially on $Y$; then the genus $g$ component of $X^{\iota }$ is the normalization of the double cover of a curve component $D$ of $Y^{\iota }$ branched over $B|_D$. When $Y=\mathbf{P}^2$, $D$ must be a line and so intersects with $B$ at six points. When $Y=\mathbf{P}^1\times \mathbf{P}^1$, $D$ is either a ruling fiber or a smooth bidegree $(1, 1)$ curve, which satisfies $(D, B)\leq 8$. Finally, let $Y=\mathbf{F}_n$ with $n>0$. If $\iota$ acts non-trivially on the $(-n)$-section $\Sigma$, then $D$ is a ruling fiber so that $(D, -2K_Y)=4$. If $\iota$ acts trivially on $\Sigma$, we have $Y^{\iota }=H+\Sigma$ for a smooth $H\in |L_{1,0}|$ because $\iota$ must preserve every fiber of $\mathbf{F}_n$ and hence induces a non-trivial involution on every fiber. Then $(\Sigma ,B)=(\Sigma , -2K_Y)\leq 2$ and $(H,B)=(H, -2K_Y)=2n+4$. This gives us the estimate $2g+2\leq (D, B)\leq 2n+4$.
This criterion is coarse, but will suffice for our purpose.
4. Automorphic forms on the period domain
4.1 Siegel modular forms
Recall that the line bundle on ${\mathcal {A}}_{g}$ associated with the automorphic factor ${\rm Sp}_{2g}(\textbf {Z})\ni \binom {A\,B}{C\,D}\mapsto \det (C\Omega +D)\in {\mathcal {O}}({\frak S}_{g})$, $\Omega \in {\frak S}_{g}$, is called the Hodge line bundle on ${\mathcal {A}}_{g}$ and is denoted by ${\mathcal {F}}_{g}$ in this paper. A holomorphic section of ${\mathcal {F}}_{g}^{\otimes q}$ is identified with a Siegel modular form of weight $q$, and ${\mathcal {F}}_{g}^{\otimes q}$ is equipped with the Hermitian metric $\|\cdot \|_{{\mathcal {F}}_{g}^{\otimes q}}$ called the Petersson norm: for any Siegel modular form $S$ of weight $q$, we define $\|S(\Omega )\|_{{\mathcal {F}}_{g}^{\otimes q}}^{2}:=(\det \Im \,\Omega )^{q}|S(\Omega )|^{2}$.
In this paper, the following Siegel modular forms on ${\frak S}_{g}$ play crucial roles:
Here
is the theta constant with even characteristic $(a,b)$, where $a,b\in \{0,1/2\}^{g}$ and $4{}^{t}ab\equiv 0\bmod 2$. For $g=0$, we set $\chi _{0}=\Upsilon _{0}=1$. Note that $\Upsilon _{g}(\Omega )$ is the elementary symmetric polynomial of degree $2^{g-1}(2^{g}+1)-1=(2^{g-1}+1)(2^{g}-1)$ in the even theta constants $\theta _{a,b}(\Omega )^{8}$. By [Reference IgusaIgu72, Corollary, p. 176, and Theorem 3, p. 182], $\chi _{g}^{8}$ (respectively, $\Upsilon _{g}$) is a Siegel modular form of weight $2^{g+1}(2^{g}+1)$ (respectively, $2(2^{g}-1)(2^{g}+2)$). The locus of vanishing theta-null $\theta _{{\rm null},g}$ is the reduced divisor on ${\mathcal {A}}_{g}$ defined by $\chi _{g}$
Lemma 4.1 There exist at least two distinct vanishing even theta constants at $\Omega \in {\frak S}_{g}$ if and only if $\chi _{g}(\Omega )=\Upsilon _{g}(\Omega )=0$. In particular, a smooth projective curve $C$ of genus $g$ has at least two effective even theta characteristics if and only if $\chi _{g}(\Omega (C))=\Upsilon _{g}(\Omega (C))=0$.
Proof. Assume $\chi _{g}(\Omega )=\Upsilon _{g}(\Omega )=0$. Since $\chi _{g}(\Omega )=0$, there is an even pair $(a,b)$ with $\theta _{a,b}(\Omega )=0$. Then $\Upsilon _{g}(\Omega )=\prod _{(c,d)\not =(a,b)}\theta _{c,d}(\Omega )^{8}$, where $(c,d)\in \{0,1/2\}^{2g}$ runs over all even pairs distinct from $(a,b)$. Since $\Upsilon _{g}(\Omega )=0$, we get $\theta _{c,d}(\Omega )=0$ for some even pair $(c,d)\not =(a,b)$. Thus $\theta _{a,b}(\Omega )=\theta _{c,d}(\Omega )=0$. The converse is trivial. The second assertion follows from the Riemann singularity theorem [Reference Arbarello, Cornalba, Griffiths and HarrisACGH85, p. 226].
For the proof of Theorem 0.1(1) and (2), we need an estimate for the vanishing order of $\Upsilon _{g}$ for certain ordinary singular families of curves.
Lemma 4.2 Let $p\colon {\mathcal {C}}\to \Delta$ be an ordinary singular family of curves of genus $g>0$ with irreducible $C_{0}:=p^{-1}(0)$. Namely, $p\colon {\mathcal {C}}\to \Delta$ is a proper surjective holomorphic function from a complex surface ${\mathcal {C}}$ to the unit disc $\Delta$ without critical points on ${\mathcal {C}}\setminus C_{0}$ and with a unique, non-degenerate critical point on $C_{0}$. Assume that $\chi _{g}(\Omega (C_{t}))=0$ and $\Upsilon _{g}(\Omega (C_{t}))\not =0$ for all $t\in \Delta ^{*}$ and that $\chi _{g-1}(\Omega (\hat {C}_{0}))\not =0$, where $\hat {C}_{0}$ is the normalization of $C_{0}$. Then there exists $h(t)\in {\mathcal {O}}(\Delta )$ such that
Proof. We follow [Reference YoshikawaYos13, Proof of Lemma 4.1]. For $\Omega \in {\frak S}_{g}$, write $\Omega =\binom {z\,{}^{t}\omega }{\omega \,Z}$, where $z\in {\frak H}$, $\omega \in \textbf {C}^{g-1}$, $Z\in {\frak S}_{g-1}$. For $t\in \Delta ^{*}$, we can express
where $\psi (t)$ is a holomorphic function on $\Delta$ with values in complex symmetric $g\times g$ matrices such that $\psi (0)=\binom {\psi _{0}\,{}^{t}\omega _{0}}{\omega _{0}\, Z_{0}}$, $Z_{0}\in {\frak S}_{g-1}$, $\Omega (\hat {C}_{0})=[Z_{0}]\in {\mathcal {A}}_{g-1}$.
By the assumption $\chi _{g}(\Omega (C_{t}))=0$, $\Upsilon _{g}(\Omega (C_{t}))\not =0$ for all $t\in \Delta ^{*}$ and Lemma 4.1, $C_{t}$ has a unique effective even theta characteristic for $t\in \Delta ^{*}$. By fixing a marking of a reference curve, there is a unique even pair $(a,b)$, $a,b\in \{0,1/2\}^{g}$, such that $\theta _{a,b}(\Omega (C_{t}))=0$ on $\Delta$ and $\theta _{c,d}(\Omega (C_{t}))\not =0$ on $\Delta ^{*}$ for all even $(c,d)\not =(a,b)$. Write $a=(a_{1},a')$ and $b=(b_{1},b')$. If $a_{1}=0$, then we get $\theta _{a',b'}(Z_{0})=0$ for the even pair $(a',b')$, $a',b'\in \{0,1/2\}^{g-1}$, by Yoshikawa [Reference YoshikawaYos13, (4.4)], which contradicts the assumption $\chi _{g-1}(Z_{0})=\chi _{g-1}(\Omega (\hat {C}_{0}))\not =0$. Thus $a_{1}=1/2$.
By [Reference YoshikawaYos13, (4.3), (4.4)], there is a holomorphic function $F_{a,b}(\zeta ,\omega ,Z)$ such that
Hence there is a holomorphic function $\phi (\zeta ,\omega ,Z)$ such that
Since $\Upsilon _{g}(\Omega )$ is a Siegel modular form and hence $\Upsilon _{g}(\Omega +A)=\Upsilon _{g}(\Omega )$, $\phi (\zeta ,\omega ,Z)$ is an even function in $\zeta$. There exists a holomorphic function $h(t)\in {\mathcal {O}}(\Delta )$ such that
This, together with [Reference YoshikawaYos13, (4.7)], implies the result.
As a consequence of Lemma 4.2, we get the following result.
Lemma 4.3 Let $M\subset {\Bbb L}_{K3}$ be a primitive $2$-elementary Lorentzian sublattice and set $\Lambda =M^{\perp }$. Let $\gamma \colon \Delta \to {\mathcal {M}}_{\Lambda }$ be a holomorphic curve with $\gamma (\Delta ^{*})\subset {\mathcal {M}}_{\Lambda }^{0}$ intersecting $\bar {\mathcal {D}}_{\Lambda }^{0}$ transversally at $\gamma (0)\in \bar {\mathcal {D}}_{\Lambda }^{0}$. Let $\tilde {\gamma }\colon \Delta \to \Omega _{\Lambda }$ be its lift with $\gamma (t^{2})=\Pi _{\Lambda }\circ \tilde {\gamma }(t)$ and let $d\in \Delta _{\Lambda }^{-}$ be such that $\tilde {\gamma }(0)\in H_{d}$. If $\chi _{g}({J}_{M}(\tilde {\gamma }(t)))=0$ and $\Upsilon _{g}({J}_{M}(\tilde {\gamma }(t)))\not =0$ for all $t\in \Delta ^{*}$ and if $\chi _{g-1}(J_{[M\perp d]}(\tilde {\gamma }(0)))\not =0$, then
Proof. Let $\bar {J}_{M}\colon {\mathcal {M}}_{\Lambda }^{0}\cup \bar {\mathcal {D}}_{\Lambda }^{0}\to {\mathcal {A}}_{g}^{*}$ be the extension of $\bar {J}_{M}^{0}\colon {\mathcal {M}}_{\Lambda }^{0}\to {\mathcal {A}}_{g}$. By [Reference YoshikawaYos13, Theorem 2.3(1) and (2)], there is an ordinary singular family of curves $p\colon {\mathcal {C}}\to \Delta$ of genus $g$ with irreducible $C_{0}$ and with period map $\bar {J}_{M}\circ \gamma$. By Lemma 4.2, we get
which, together with $J_{M}(\gamma (t^{2}))=\bar {J}_{M}(\tilde {\gamma }(t))$, yields the result.
4.2 Automorphic forms on $\Omega _{\Lambda }$
Let $M\subset {\Bbb L}_{K3}$ be a primitive $2$-elementary Lorentzian sublattice and set $\Lambda =M^{\perp }$ as before. Let $q\in \textbf {Z}_{>0}$ be such that ${\mathcal {F}}_{g}^{\otimes q}$ extends to a very ample line bundle on ${\mathcal {A}}_{g}^{*}$. Let $i\colon \Omega _{\Lambda }^{0}\cup {\mathcal {D}}_{\Lambda }^{0}\hookrightarrow \Omega _{\Lambda }$ be the inclusion and define $\lambda _{M}^{q}$ as the trivial extension of $J_{M}^{*}{\mathcal {F}}_{g(M)}^{\otimes q}$ from $\Omega _{\Lambda }^{0}\cup {\mathcal {D}}_{\Lambda }^{0}$ to $\Omega _{\Lambda }$, that is,
Since $\Omega _{\Lambda }\setminus (\Omega _{\Lambda }^{0}\cup {\mathcal {D}}_{\Lambda }^{0})$ has codimension $2$ in $\Omega _{\Lambda }$, $\lambda _{M}^{q}$ is an $O(\Lambda )$-equivariant invertible sheaf on $\Omega _{\Lambda }$. On $\Omega _{\Lambda }^{0}$, $\lambda _{M}^{q}$ is equipped with the Hermitian metric
Fix $l_{\Lambda }\in \Lambda \otimes \textbf {R}$ with $\langle l_{\Lambda },l_{\Lambda }\rangle \geq 0$. Define $j_{\Lambda }(\gamma ,\cdot )\in {\mathcal {O}}^{*}_{\Omega _{\Lambda }}$, $\gamma \in O(\Lambda )$ and $K_{\Lambda }(\cdot )\in C^{\infty }(\Omega _{\Lambda })$ by
Let $p,q\in \textbf {Z}$. Then $F\in H^{0}(\Omega _{\Lambda },\lambda _{M}^{q})$ is called an automorphic form on $\Omega _{\Lambda }$ for $O(\Lambda )$ of weight $(p,q)$ if it satisfies the following functional equation on $\Omega _{\Lambda }$:
The notion of automorphic forms on $\Omega _{\Lambda }^{+}$ for $O^{+}(\Lambda )$ of weight $(p,q)$ is defined in the same way. In the rest of this paper the vector space of automorphic forms on $\Omega _{\Lambda }$ for $O(\Lambda )$ of weight $(p,q)$ is identified with the vector space of automorphic forms on $\Omega _{\Lambda }^{+}$ for $O^{+}(\Lambda )$ of weight $(p,q)$ via the restriction map
We define the Petersson norm of an automorphic form $F$ on $\Omega _{\Lambda }$ for $O(\Lambda )$ of weight $(p,q)$ as the $O(\Lambda )$-invariant $C^{\infty }$ function on $\Omega _{\Lambda }^{0}$ defined as
5. The invariant $\tau _{M}$ and its automorphic property
Let $(X,\iota )$ be a $2$-elementary $K3$ surface of type $M$. Let $\gamma$ be an $\iota$-invariant Kähler form on $X$. The Laplacian acting on $(0,q)$-forms on $X$ is denoted by $\square _{0.q}$. Write $\sigma (\square _{0,q})$ for the spectrum of $\square _{0,q}$ and $E_{0,q}(\lambda )$ for eigenspace of $\square _{0,q}$ corresponding to $\lambda \in \sigma (\square _{0,q})$. Since $\iota$ preserves $\gamma$, $\iota$ acts on $E_{0,q}(\lambda )$. The equivariant zeta function of $\square _{0,q}$ is defined as the following convergent series for $s\in \textbf {C}$ with $\Re s\gg 0$:
Then $\zeta _{0,q}(\iota )(s)$ extends meromorphically to $\textbf {C}$ and is holomorphic at $s=0$. We define the equivariant analytic torsion [Reference BismutBis95] of $(X,\gamma )$ as
Let $\eta \in H^{0}(X,K_{X})\setminus \{0\}$ and set $\|\eta \|_{L^{2}}^{2}:=(2\pi )^{-2}\int _{X}\eta \wedge \bar {\eta }$. For a compact Kähler manifold $(V,\omega )$, define ${\rm Vol}(V,\omega ):=(2\pi )^{-\dim V}\int _{V}\omega ^{\dim V}/(\dim V)!$.
When $X^{\iota }\not =\emptyset$, write $X^{\iota }=\amalg _{i}C_{i}$ for the decomposition into the connected components. Let $\tau (C_{i},\gamma |_{C_{i}})$ be the analytic torsion [Reference Ray and SingerRS73] of $(C_{i},\gamma |_{C_{i}})$. We define
When $X^{\iota }=\emptyset$, we set $\tau (X^{\iota },\gamma |_{X^{\iota }})={\rm Vol}(X^{\iota },\gamma |_{X^{\iota }})=1$. Let $c_{1}(X^{\iota },\gamma |_{X^{\iota }})$ be the first Chern form of $(X^{\iota },\gamma |_{X^{\iota }})$. By [Reference YoshikawaYos04, Theorem 5.7], the real number
is determined by the isomorphism class of $(X,\iota )$ and hence the period $\bar {\pi }_{M}(X,\iota )$. For the arithmetic counterpart of the invariant $\tau _{M}(X,\iota )$, we refer to [Reference Maillot and RösslerMR09].
We set $\Lambda =M^{\perp }$ and we regard $\tau _{M}$ as the $O(\Lambda )$-invariant function on $\Omega _{\Lambda }^{0}$ or equivalently the function on ${\mathcal {M}}_{\Lambda }^{0}$ defined by
As an application of the theory of (equivariant) Quillen metrics [Reference Bismut, Gillet and SouléBGS88, Reference BismutBis95, Reference MaMa00], the automorphy of $\tau _{M}$ has been established.
Theorem 5.1 [Yos04, Yos13, Yos12] There exist $\ell \in \textbf {Z}_{>0}$ and an automorphic form $\Phi _{M}$ on $\Omega _{\Lambda }$ for $O(\Lambda )$ of weight $(\ell (r(M)-6),4\ell )$ with
In the rest of this paper we determine $\Phi _{M}$ for all $M$. Since this was done for exceptional $M$ in [Reference YoshikawaYos04], $M$ is assumed to be non-exceptional in what follows.
6. The locus of vanishing theta-null: the case $\delta =1$
In the rest of this paper, for a primitive $2$-elementary Lorentzian sublattice $M\subset {\Bbb L}_{K3}$, we write $r$, $l$, $\delta$, $g$, $k$ for $r(M)$, $l(M)$, $\delta (M)$, $g(M)$, $k(M)$, respectively, when there is no possibility of confusion. Recall that these invariants are introduced in §§ 1 and 3.2.1. We write
Let ${\frak M}_{g}$ be the moduli space of smooth curves of genus $g\geq 1$. In this section and the next we study the Torelli map $J_M: \Omega _{\Lambda }^0\to \mathcal {A}_g$ from the geometric point of view. We here view the Torelli map rather as a morphism
from the moduli space of 2-elementary $K3$ surfaces to that of curves, which associates to $(X, \iota )$ the genus $g$ component of $X^{\iota }$. Our main purpose here is to describe the inverse image of a certain geometric locus in $\frak {M}_g$ as a Heegner divisor of $\mathcal {M}_{\Lambda }^{0}$ in a few cases. This will be the first and necessary step toward a more complete description, Theorem 9.13, which will be obtained in the final part of the paper.
Recall that the characteristic Heegner divisor $\mathcal {H}_{\Lambda }$ of $\Omega _{\Lambda }$ was defined in § 2.3, and that the theta-null divisor $\theta _{{\rm null},g}$ of ${\mathcal {A}}_{g}$ was defined in § 4.1. We denote by
the reduced algebraic divisor corresponding to $\mathcal {H}_{\Lambda }$, and by
the reduced theta-null divisor in ${\frak M}_{g}$. It is well known that ${\frak M}_{g}'\subset {\frak M}_{g}$ is the locus of curves $C$ having an effective even theta characteristic, namely an effective line bundle $L$ with $L^{\otimes 2}\simeq K_C$ and $h^0(L)$ even.
In the present section we treat the moduli spaces in the following two series:
• $k=0$, $\delta =1$, $3\leq g\leq 10$;
• $k=1$, $6\leq g\leq 9$.
Notice that in the second series we have $\delta =0$ only when $g=6$. We will prove the following theorem.
Theorem 6.1 If $3\leq g\leq 10$ and $(k,\delta )=(0,1)$, the Heegner divisor $\bar {\mathcal {H}}_{\Lambda }$ is irreducible and the following set-theoretic equality of (reduced) divisors of $\mathcal {M}_{\Lambda }^0$ holds:
The same assertion also holds for the second series when $\delta =1$, but the following weaker version will suffice for the rest of the paper.
Proposition 6.2 If $6\leq g\leq 9$ and $k=1$, $\bar {\mu }_{\Lambda }(\mathcal {M}_{\Lambda }^0)$ is not contained in ${\frak M}_{g}'$.
These results will be used in § 8. Theorem 6.1 will be proved in §§ 6.1–6.3, and Proposition 6.2 in § 6.5.
6.1 Proof of Theorem 6.1: the strategy
Let us first explain the outline of the proof of Theorem 6.1, reducing it to the construction of certain elliptic curves.
As the first step we see the irreducibility of $\bar {\mathcal {H}}_{\Lambda }$, which holds in a wider range.
Lemma 6.3 When $\delta =1$ and $r\leq 9$, $\bar {\mathcal {H}}_{\Lambda }$ is an irreducible divisor of $\mathcal {M}_{\Lambda }$.
Proof. This is restated as the property that vectors $l\in \Lambda ^{\vee }$ with $(l, l)=\varepsilon _{\Lambda }$ and $[l]= {\boldsymbol 1}_{\Lambda }$ are all equivalent under $O^+(\Lambda )$. Consider the vector $l'=2l$ in $\Lambda$. Since $ {\boldsymbol 1}_{\Lambda }$ is of order $2$, $l'$ is primitive in $\Lambda$ and satisfies
When $(r, l)\ne (9, 9)$, the lattice $\Lambda$ contains $\mathbb {U}\oplus \mathbb {U}$, and then we can resort to the Eichler criterion (cf. [Reference ScattoneSca87]) which says that the $O^+_0(\Lambda )$-equivalence class of a primitive vector $l'\in \Lambda$ depends only on the norm $(l', l')$ and the element $[l'/{\rm div}(l')]\in A_{\Lambda }$. Hence the above vectors $l'=2l$ are all $O^+_0(\Lambda )$-equivalent. When $(r, l)=(9, 9)$, $\bar {\mathcal {H}}_{\Lambda }$ is defined by $(-2)$-vectors $l'\in \Lambda$ with $\Lambda =\textbf {Z}l'\oplus (l')^{\perp }$. The isometry class of $(l')^{\perp }$ is uniquely determined by [Reference NikulinNik80], namely $(l')^{\perp }\simeq {\Bbb U}^{\oplus 2}\oplus {\Bbb E}_{8}(2)$, so that these $(-2)$-vectors $l'$ are all ${\rm O}^{+}(\Lambda )$-equivalent.
With the irreducibility of $\bar {\mathcal {H}}_{\Lambda }$ verified, the proof of Theorem 6.1 is reduced to showing the non-emptiness of $\bar {\mu }_{\Lambda }^{-1}({\frak M}_{g}')$ and the inclusion
We only need to verify (6.1) outside a locus of codimension at least $2$ of $\mathcal {M}_{\Lambda }^{0}$. Our approach will be based on the following geometric observation.
Proposition 6.4 Let $(X,\iota )$ be a $2$-elementary $K3$ surface. If $X$ has a smooth elliptic curve $E$ with $E+\iota (E) \sim X^{\iota }$, then the period of $(X, \iota )$ is contained in the Heegner divisor $\bar {\mathcal {H}}_{\Lambda }$.
Proof. Let $H_{\pm }$ denote the $\iota$-(anti-)invariant lattices of $(X, \iota )$. The cycle $D_{\pm }:=E\pm \iota (E)$ is contained in $H_{\pm }$, respectively. We will show that $D_{-}$ satisfies
The presence of such an anti-invariant cycle in the Picard lattice means that the period of $(X, \iota )$ lies in $\bar {\mathcal {H}}_{\Lambda }$.
Since $E$ is an elliptic curve and hence its class in $H^{2}(X,\textbf {Z})$ is isotropic, the first equality of (6.2) follows from
The second property in (6.2) holds because
To see the last equality of (6.2), we note that the anti-isometry $\lambda :A_{H_+}\to A_{H_-}$ induced from the relation $H_+=(H_-)^{\perp }\cap H^2(X,\textbf {Z})$ maps $[D_+/2]$ to $[D_-/2]$ because $D_+/2+D_-/2=E$ is contained in $H^2(X,\textbf {Z})$. Since $[D_+/2]$ is the characteristic element of $A_{H_+}$ by Lemma 6.5 below, so is $[D_-/2]$ in $A_{H_-}$.
Lemma 6.5 For any 2-elementary $K3$ surface $(X, \iota )$ the element $[X^{\iota }/2]\in A_{H_+}$ is the characteristic element.
Proof. Let $f\colon X\to Y$ be the quotient map by $\iota$, and let $B\subset Y$ be the branch curve. Every element of the dual lattice $H_+^{\vee }$ can be written as $f^{\ast }L/2$ for some $L\in {\rm Pic}(Y)$. Then
We can see that $(L+K_Y, L)\in 2\textbf {Z}$ from the Riemann–Roch formula. Therefore
which means that $X^{\iota }/2 \in H_+^{\vee }$ and that $[X^{\iota }/2]$ is characteristic.
In §§ 6.2 and 6.3, we will construct an elliptic curve $E$ as in Proposition 6.4. The non-emptiness of $\bar {\mu }_{\Lambda }^{-1}({\frak M}_{g}')$ will be seen in the course of proof.
6.2 Proof of Theorem 6.1: the case $g=3$
We begin with the case $g=3$. Recall that a smooth curve of genus $3$ has an effective even theta characteristic precisely when it is hyperelliptic. So let $(X, \iota )$ be a 2-elementary $K3$ surface with $(g, k)=(3, 0)$ such that $X^{\iota }$ is hyperelliptic. Consider the degree $4$ line bundle $L=\mathcal {O}_{X}(X^{\iota })$, which is hyperelliptic in the sense of § 3.3. By Saint-Donat's classification, $L$ defines a degree $2$ morphism
onto a quadric $Q\subset \mathbf{P}^3$. We may assume that $L$ is ample, because the locus where $X^{\iota }$ is hyperelliptic and $\mathcal {O}_{X}(X^{\iota })$ is non-ample has codimension at least $2$ in $\mathcal {M}_{\Lambda }^0$. In that case $Q$ is smooth.
Claim 6.6 $\iota$ acts on $Q$ by switching the two rulings on it.
Proof. First note that $\iota$ acts on $Q$ non-trivially, for the branch curve of $\phi$ is a member of $|{-}2K_{Q}|$ and hence has genus $9$. Then $\iota$ fixes the curve $\phi (X^{\iota })$, which by the definition of $\phi$ has bidegree $(1, 1)$. It is easily verified that any non-trivial involution of $Q$ fixing a smooth bidegree $(1, 1)$ curve must switch the two rulings.
We choose a ruling line $l$ on $Q$ and put $E = \phi ^{\ast }l$. Then $E$ is a smooth elliptic curve on $X$ for a general choice of $l$ and satisfies the linear equivalence
By Proposition 6.4, we get the inclusion (6.1).
The non-emptiness of $\bar {\mu }_{\Lambda }^{-1}({\frak M}_{3}')$ can be seen by reversing this construction: choose a bidegree $(4, 4)$ curve $B\subset \mathbf{P}^1\times \mathbf{P}^1$ preserved by the switch involution of $\mathbf{P}^1\times \mathbf{P}^1$ and take the double cover $X\to \mathbf{P}^1\times \mathbf{P}^1$ branched over $B$. The switch involution can be lifted to $X$ so that its fixed curve is the preimage of the diagonal of $\mathbf{P}^1\times \mathbf{P}^1$, which is hyperelliptic of genus $3$. Thus Theorem 6.1 is proved in the case $g=3$.
6.3 Proof of Theorem 6.1: the case $4\leq g\leq 10$
We next treat the case $4\leq g\leq 10$ in Theorem 6.1. Let $(X, \iota )$ be a 2-elementary $K3$ surface with $4\leq g\leq 10$, $k=0$ and $\delta =1$. Let $Y=X/\iota$ be the quotient surface and $C\subset Y$ be the branch ($-2K_Y$)-curve. The quotient map $X\to Y$ gives the canonical identification
The anti-canonical model $\bar {Y}\subset \mathbf{P}^{g-1}$ of $Y$ is a Gorenstein del Pezzo surface of degree $g-1$, and $Y$ is its minimal resolution. Note that if $X\to \bar {X}$ is the contraction of $(-2)$-curves disjoint from $X^{\iota }$, we naturally have $\bar {Y}\simeq \bar {X}/\iota$. We may view $C$ also as lying on the smooth locus of $\bar {Y}$. By the adjunction formula we have $-K_Y|_C\simeq K_C$, and the restriction map $|{-}K_{Y}|\to |K_{C}|$ is isomorphic because $h^{0}(K_{Y})=h^{1}(K_{Y})=0$. Therefore the composition of inclusions
gives the canonical embedding of $C$, and we can identify $\mathbf{P}^{g-1}$ with $|K_C|^{\vee }$. This also shows that $C$ is non-hyperelliptic.
Lemma 6.7 If $C$ has an effective even theta characteristic $L$, then $h^0(L)=2$.
Proof. If $h^0(L)\ne 2$, then $h^0(L)\geq 4$. By Clifford's theorem we have
This holds only when $g\geq 8$ and $h^0(L)=4$. Actually the case $g=8$ can be excluded because a blow-down of $Y$ presents $C$ as a plane sextic with two double points and hence $C$ has Clifford index $2$. The case $g=10$, where $C$ is a smooth plane sextic, is treated in [Reference Arbarello, Cornalba, Griffiths and HarrisACGH85, Exercise VI. B-3]. In the case $g=9$, presenting $C$ as plane sextic with a node or cusp, we can argue similarly.
Proposition 6.8 There exists a locus ${\mathcal {Z}}\subset \mathcal {M}_{\Lambda }^{0}$ of codimension at least $2$ with the following property. When $(X, \iota )$ lies outside ${\mathcal {Z}}$, the curve $C$ has a theta characteristic $L$ with $h^0(L)=2$ if and only if $C$ can be cut out from $\bar {Y}$ by a quadric $Q\subset \mathbf{P}^{g-1}$ of rank $3$. In this case, $L$ is base point free.
Proof. The proof consists of five steps.
Step 1. Assume that $C\subset \bar {Y}$ is cut out by a quadric $Q$ of rank $3$. The vertex of $Q$, a $(g-4)$-plane, is disjoint from $\bar {Y}$; otherwise $C$ would be singular. The pencil of $(g-3)$-planes in $Q$ gives $C$ a theta characteristic $L$ with $h^0(L)\geq 2$, which is free because $C$ is disjoint from the vertex. We have $h^0(L)=2$, for $|L|=|K_C-L|$ is identified with the linear system of hyperplanes of $\mathbf{P}^{g-1}$ containing a $(g-3)$-plane, which is a pencil.
Step 2. Conversely, suppose that $C$ has a complete half-canonical pencil $|L|$. Choose a basis $\alpha , \beta$ of $H^{0}(L)$ and write $|L|=|L_0|+D_0$ with $|L_0|$ the free part and $D_0$ the fixed part. The divisor $D_0$ is defined by $\alpha =\beta =0$. If we set
then $C$ is contained in the rank $3$ quadric $Q\subset \mathbf{P}^{g-1}$ defined by $u_1u_2=u_3^2$. The vertex of $Q$ is the $(g-4)$-plane $V$ defined (set-theoretically) by $(u_1=u_2=0)\cap Q$. The free part $|L_0|$ is given by the pencil of $(g-3)$-planes in $Q$ through $V$, and the fixed part $D_0$ is defined by
In order to show that this quadric $Q$ cuts out $C$ from $\bar {Y}$, it suffices to prove $\bar {Y}\not \subset Q$, for then $C$ and $Q|_{\bar {Y}}$ are both ($-2K_{\bar {Y}}$)-curves on $\bar {Y}$.
Step 3. Suppose the contrary: $\bar {Y}\subset Q$. We then have three linearly independent sections $\tilde {u}_1, \tilde {u}_2, \tilde {u}_3\in H^0(-K_Y)$ on $Y$ with $\tilde {u}_1\tilde {u}_2=\tilde {u}_3^2$ and $u_i=\tilde {u}_i|_{C}$. Choose a blow-down $\pi :Y\to \mathbf{P}^2$ and let $Z\subset \mathbf{P}^{2}$ be the blown-up points (which possibly contain infinitely near ones). We have $\# Z = 10-g$. The image $\pi (C)$ of $C$ is a plane sextic having double points at $Z$ and no other singularities. Via the mapping
we can identify $|{-}K_Y|$ with the linear system of plane cubics through $Z$, and $D$ is recovered from $\Gamma =\pi (D)$ by $D=\pi ^{\ast }\Gamma -\pi ^{-1}(Z)$. Now by our assumption $\tilde {u}_1\tilde {u}_2=\tilde {u}_3^{2}$, the divisors of $\tilde {u}_1$, $\tilde {u}_2$, $\tilde {u}_3$ correspond to three linearly independent cubics $\Gamma _1$, $\Gamma _2$, $\Gamma _3$ with $\Gamma _1+\Gamma _2=2\Gamma _3$. This equality can hold only when
for some distinct lines $l_1, l_2, l$. Let $p=l_1\cap l_2$. The net spanned by $\Gamma _1, \Gamma _2, \Gamma _3$ consists of splitting cubics $l+l'+l''$ where $l', l''$ are lines through $p$. In particular, its base locus is $p\cup l$. Recall that the restriction of this net to $C$, after the transformation (6.5), is equal to $\mathbf{P}\langle u_1, u_2, u_3\rangle = \mathbf{P}{\rm Sym}^2H^0(L)$.
Step 4. For a plane curve $\Gamma$ we write $\hat {\Gamma }$ for its strict transform in $Y$. We also denote $E=\pi ^{-1}(p)$. We observe the following:
(1) $Z$ lies on $p\cup l$;
(2) $p\in Z$ and $p\notin l$;
(3) the free part $|L_0|$ is given by the projection from $p$;
(4) the fixed part $D_0$ is given by
(6.6)\begin{equation} 2D_0 = (\hat{l}+E)|_{C}. \end{equation}
Statement (1) is obvious. We can see (3) by noticing that $\phi _{L_0}:C\to \mathbf{P}^1$ coincides with the projection from $p$ after composing them with the conic embedding $\mathbf{P}^1\hookrightarrow \mathbf{P}^2$, the composition being the resolution of the rational map defined by ${\rm Sym}^2H^0(L)$. Under (2), we have ${\rm div}(\tilde {u}_i)=2\hat {l}_i+E+\hat {l}$ for $i=1, 2$, so that (4) follows from (6.4). It remains to see (2). Firstly, if $p\notin \pi (C)$, then $L_0\sim \pi ^{*}\mathcal {O}_{\mathbf{P}^2}(1)|_{C}$ which is absurd. If $p\in \pi (C)$ but $p\notin Z={\rm Sing}(\pi (C))$, then $L_0\sim \pi ^{*}\mathcal {O}_{\mathbf{P}^2}(1)|_{C} - \pi ^{-1}(p)$ by (3) and $D_0\ni \pi ^{-1}(p)$ by (6.4). It follows that $L\geq \pi ^{*}{\mathcal {O}}_{\mathbf{P}^{2}}(1)|_{C}$, and hence ${\dim } |L| \geq 2$, a contradiction. Therefore $p\in Z$. Next assume $p\in l\cap Z$. Then both $\Gamma _1$ and $\Gamma _2$ have multiplicity $3$ at $p$, so that ${\rm div}(\tilde {u}_i)=2\hat {l}_i+2E+\hat {l}$ for $i=1, 2$. Then $2D_{0}=(2E+\hat {l})|_{C}$ by (6.4). Since $L_{0}\sim \pi ^{*}\mathcal {O}_{\mathbf{P}^2}(1)|_{C}-E|_{C}$ by (3), we have $L\geq \pi ^{*}{\mathcal {O}}_{\mathbf{P}^{2}}(1)|_{C}$, the same contradiction as before. This verifies (2).
Step 5. Now since the $9-g$ points $Z\backslash p$ lie on $l$ and since no four points of $Z$ can be collinear by the nefness of $-K_{Y}$, we must have $|Z\setminus p|\leq 3$ and thus $6 \leq g\leq 9$. The right-hand side of (6.6) is divisible by $2$ only if $\pi (C)$ is totally tangent to $l$ outside $Z={\rm Sing}(\pi (C))$ and $\pi (C)$ has a cusp at $p$. The cusp condition defines a divisor in the moduli space; when $7\leq g \leq 9$ (respectively, $g=6$), the tangency condition (respectively, the collinear condition on the three points $Z\setminus p$) defines another divisor. These two divisors, both Heegner and irreducible, define a codimension $2$ locus ${\mathcal {Z}}$ in the moduli space. If $(X,\iota )\not \in {\mathcal {Z}}$, then (6.6) cannot hold and thus $\bar {Y}\not \subset Q$. By step 2, this implies that $C$ is cut out from $\bar {Y}$ by $Q$ when $(X,\iota )\not \in {\mathcal {Z}}$.
Since the intersection of a general Del Pezzo surface $Y\subset \mathbf{P}^{g-1}$ and a quadric $Q\subset \mathbf{P}^{g-1}$ of rank $3$ is a smooth curve, we see the non-emptiness of $\bar {\mu }_{\Lambda }^{-1}({\frak M}_{g}')$.
Now let $C\subset \bar {Y}$ be cut out by a quadric $Q$ of rank $3$. We take the double cover
branched over $Q$. The (contracted) quotient map $\bar {X}\to \bar {Y}$ by $\iota$ can be identified with the restriction $\pi ^{-1}(\bar {Y})\to \bar {Y}$ of $\pi$. We again denote by $\iota$ the covering transformation of $\pi : \tilde {Q} \to \mathbf{P}^{g-1}$. We also view $\bar {X}\subset \mathbf{P}^{g}$ naturally. Note that $\tilde {Q}$ is a quadric of rank $4$ in $\mathbf{P}^g$ and hence is the cone over a smooth quadric surface $Q_0\simeq \mathbf{P}^1\times \mathbf{P}^1$ with vertex $\mathbf{P}^{g-4}$. Let $f\colon \tilde {Q}\dashrightarrow Q_0$ be the projection from the vertex. Then the pencils $f^{\ast }|\mathcal {O}_{Q_0}(1, 0)|$ and $f^{\ast }|\mathcal {O}_{Q_0}(0, 1)|$ are $\mathbf{P}^1$-families of $(g-2)$-planes that sweep out $\tilde {Q}$. As is easily verified, $\iota$ switches these two rulings. Restricted to $\bar {X}=\pi ^{-1}(\bar {Y})$, the pencil $f^{\ast }|\mathcal {O}_{Q_0}(1, 0)|$ induces an elliptic fibration on $X$, say $|E|$. The cycle $E+\iota (E)$ is a hyperplane section of $\bar {X}\subset \mathbf{P}^{g}$. On the other hand, $X^{\iota }\subset \bar {X}$ is the ramification divisor of $\bar {X}\to \bar {Y}$ and hence also cut out by a hyperplane of $\mathbf{P}^g$. Therefore $E+\iota (E)$ is linearly equivalent to $X^{\iota }$ on $X$. By Proposition 6.4, Theorem 6.1 is proved in the case $4\leq g\leq 10$.
Remark 6.9 In the above construction, the given half-canonical pencil on $C\simeq X^{\iota }$ coincides with the restriction of the elliptic fibration $|E|$: this follows because $|E|$ is the restriction of a $\mathbf{P}^{g-2}$-ruling on $\tilde {Q}$, which in turn is a component in the pullback of the family of hyperplanes in $\mathbf{P}^{g-1}$ that cut out doubly from $Q$ the $\mathbf{P}^{g-3}$-ruling.
Remark 6.10 A general member of $\bar {\mu }_{\Lambda }^{-1}({\frak M}_{g}')$ has a unique effective even theta characteristic, because the period of those $(X, \iota )$ having several elliptic curve classes $[E]$ with $E+\iota (E)\sim X^{\iota }$ lies in (the image of) the intersection of at least two components of $\mathcal {H}_{\Lambda }$.
6.4 Trigonal curves
This subsection is a prerequisite for the subsequent §§ 6.5 and 7. A smooth projective curve $C$ of genus $g\geq 5$ is said to be trigonal if it has a degree $3$ morphism to $\mathbf{P}^1$. Such a morphism, if it exists, is unique up to ${\rm Aut}(\mathbf{P}^1)$. Let us summarize some properties of trigonal curves (cf. [Reference Martens and SchreyerMS86]).
It is classically known that a trigonal curve $C$ can be canonically embedded in a Hirzebruch surface. This is due to the fact that the canonical model of $C$ is contained in a unique rational normal scroll, namely the image of a Hirzebruch surface $\mathbf{F}_n$ by a bundle $L_{1,m}$ with $m>0$ (see § 3.3 for the notation). The integer $n$ is called the scroll invariant, and $m$ the Maroni invariant of $C$. As a curve on $\mathbf{F}_n$, $C$ belongs to the linear system $|L_{3,b}|$ with $b=m-n+2$ by the adjunction formula. The trigonal map of $C$ is given by the restriction of the projection $\mathbf{F}_n\to \mathbf{P}^1$. We have the genus formula $g=3n+2b-2$, which gives the relation of $n$ and $m$. These (equivalent) invariants give a stratification of the moduli space of trigonal curves. By the canonicity of the embedding $C\subset \mathbf{F}_n$, the isomorphism classes of trigonal curves of genus $g$ and Maroni invariant $m$ correspond to the ${\rm Aut}(\mathbf{F}_n)$-orbits in the locus of smooth curves in $|L_{3,b}|$.
Maroni described the variety $W_d^r(C)\subset {\rm Pic}^d(C)$ of line bundles $L$ of degree $d$ and $h^0(L)\geq r+1$ (see [Reference Martens and SchreyerMS86, Proposition 1]). We need his description in the case $d=g-1$.
Proposition 6.11 (Maroni) Let $T=L_{0,1}|_C$ be the trigonal bundle and write $W_+=rT+W_{g-1-3r}(C)$, where $W_{g-1-3r}(C):=W_{g-1-3r}^{0}(C)$. Then
Using this description, we can give a geometric characterization of trigonal curves having effective theta characteristics.
Lemma 6.12 A trigonal curve $C$ of scroll invariant $n$ and Maroni invariant $m$ has a theta characteristic $L$ with $h^0(L)\geq r+1$ if and only if there exists a curve $H\in |L_{1,m-2r}|$ on $\mathbf {F}_n$ such that $H|_C=2D$ for a divisor $D$ of degree $g-1-3r$ on $C$. In that case, the divisor $rT+D$ gives such a theta characteristic.
Proof. A line bundle $L$ is a theta characteristic with $h^0(L)\geq r+1$ if and only if it is a fixed point of the residual involution on $W_{g-1}^r(C)$. Such $L$ should be contained in $W_+ \cap (K_C-W_+)$ by Proposition 6.11, so we can write $L=rT+D$ for some $D\in W_{g-1-3r}(C)$. Since $L\sim K_{C}-L$, we have $2D\sim K_C-2rT$. The restriction map $|L_{1,m-2r}|\to |K_C-2rT|$ is isomorphic because $h^i(K_{\mathbf{F}_n}-L_{0,2r})=0$ for $i=0, 1$. Thus there exists $H\in |L_{1,m-2r}|$ with $H|_C=2D$. Conversely, if we have such $H$ and $D$, then $L=rT+D$ is a theta characteristic with $h^{0}(L)\geq r+1$ by Proposition 6.11.
6.5 Proof of Proposition 6.2
We will show that a general member $C$ of $\bar {\mu }_{\Lambda }(\mathcal {M}_{\Lambda }^{0})$ has no effective even theta characteristic when $k=1$ and $6\leq g\leq 9$, using the description of $C$ given in [Reference MaMa15].
We first consider the case $\delta =1$, where $6\leq g \leq 9$. In this case $C$ is a general trigonal curve of genus $g$ with Maroni invariant $2$, which can be realized as a general member of the linear system $|L_{3,10-g}|$ on $\mathbf{F}_{g-6}$ [Reference MaMa15, Corollaries 6.2 and 7.6]. By Lemma 6.12, $C$ has a theta characteristic $L$ with $h^0(L)\geq 2$ if and only if there exists a curve $H\in |L_{1,0}|$ totally bitangent to $C$. However, it is readily seen that a general member of $|L_{3,10-g}|$ has no such tangent curve.
When $\delta =0$, we have $g=6$. In this case $C$ is a plane quintic [Reference MaMa15, Corollary 7.3]. Then our assertion follows from the classical fact that any smooth plane quintic has no effective even theta characteristic: indeed, according to [Reference Arbarello, Cornalba, Griffiths and HarrisACGH85, p. 211], we have
The residual involution acts on this $W_5^1(C)$ by switching $p$ and $q$. It has no fixed point other than $\mathcal {O}_{\mathbf{P}^2}(1)|_C$, which is an odd theta characteristic. Thus Proposition 6.2 is proved.
7. The locus of vanishing theta-null: the case $\delta =0$
We continue the geometric study of the Torelli map, still viewed as a morphism $\bar {\mu }_{\Lambda }\colon \mathcal {M}_{\Lambda }^{0}\to \frak {M}_{g}$ between the moduli spaces. In this section we treat the following two series:
• $r=2, \ \delta =0 \ (g=9, 10)$;
• $r=10, \ \delta =0, \ 4\leq g \leq 6$.
In these cases, $\bar {\mu }_{\Lambda }(\mathcal {M}_{\Lambda }^{0})$ is contained in $\frak {M}_{g}'$ because a general member of $\bar {\mu }_{\Lambda }(\mathcal {M}_{\Lambda }^{0})$ possesses a rather apparent effective even theta characteristic.
For the first series, we will show that an analogue of Theorem 6.1 holds by replacing $\frak {M}_{g}'$ with the (reduced) divisor
of $\frak {M}_{g}'$, where $\Upsilon _{g}$ is the Siegel modular form introduced in § 4.1. Geometrically this locus parametrizes curves having at least two effective even theta characteristics (see Lemma 4.1).
Proposition 7.1 When $(r, \delta )=(2, 0)$ and $g=10$, the Heegner divisor $\bar {\mathcal {H}}_{\Lambda }$ is irreducible and equal to $\bar {\mu }_{\Lambda }^{-1}(\frak {M}_{10}'')$. In particular, the genus $10$ component of $X^{\iota }$ has a unique effective even theta characteristic if the period of $(X,\iota )$ lies outside $\bar {\mathcal {H}}_{\Lambda }$.
In the case $(r, l, \delta )=(2, 2, 0)$, the Heegner divisor $\bar {\mathcal {H}}_{\Lambda }$ is reducible, reflecting the fact that norm $-4$ vectors $l\in \Lambda$ are divided into two classes according to whether ${\rm div}(l)=1$ or $2$. We accordingly write $\bar {\mathcal {H}}_{\Lambda } = \bar {\mathcal {H}}_1 + \bar {\mathcal {H}}_2$.
Proposition 7.2 When $(r, \delta )=(2, 0)$ and $g=9$, the component $\bar {\mathcal {H}}_1$ is irreducible and equal to $\bar {\mu }_{\Lambda }^{-1}(\frak {M}_{9}'')$. In particular, $X^{\iota }$ has a unique effective even theta characteristic if the period of $(X,\iota )$ lies outside $\bar {\mathcal {H}}_{\Lambda }$.
For the second series, we will prove the following lemmas.
Lemma 7.3 If $(r, \delta )=(10, 0)$ and $g=4, 5$, then $\bar {\mu }_{\Lambda }(\mathcal {M}_{\Lambda }^0)$ is not contained in the hyperelliptic locus.
Lemma 7.4 If $(r, \delta )=(10, 0)$ and $g=6$, then a general member of $\bar {\mu }_{\Lambda }(\mathcal {M}_{\Lambda }^0)$ has exactly one effective even theta characteristic.
In § 9 these results will be used to prove Theorem 0.1 (2), (3). Lemmas 7.3 and 7.4 will be strengthened in Corollary 9.14 by an argument of modular form (a geometric proof is also possible).
7.1 Proof of Proposition 7.1
Proposition 7.1 is proved along similar lines to Theorem 6.1 (as is Proposition 7.2). We begin by checking the irreducibility of the Heegner divisor $\bar {\mathcal {H}}_{\Lambda }$, which is defined by norm $-4$ vectors in $\Lambda$.
Lemma 7.5 When $(r, l, \delta )=(2, 0, 0)$, $\bar {\mathcal {H}}_{\Lambda }$ is irreducible.
Proof. Since $\Lambda \simeq \mathbb {U}^{\oplus 2}\oplus \mathbb {E}_8^{\oplus 2}$ is unimodular, by the Eichler criterion the $O^{+}(\Lambda )$-equivalence class of a primitive vector of $\Lambda$ is determined by its norm.
We describe the members of $\mathcal {M}_{\Lambda }^{0}$ following the construction in [Reference MaMa15, § 6.1]. We consider curves on the Hirzebruch surface $\mathbf{F}_4$. Let $U\subset |L_{3,0}|$ be the open locus of smooth curves. We have a morphism $p\colon U \to \mathcal {M}_{\Lambda }^{0}$ by associating to $C\in U$ the double cover of $\mathbf{F}_4$ branched over $C+\Sigma$.
Lemma 7.6 There exists a geometric quotient $U/{\rm Aut}(\mathbf{F}_4)$ and the period map $p$ descends to a biregular isomorphism $\mathcal {P}\colon U/{\rm Aut}(\mathbf{F}_4) \to \mathcal {M}_{\Lambda }^{0}$. In particular, the Torelli map $\mathcal {M}_{\Lambda }^{0}\to \frak {M}_{g}$ is given by $\mathcal {P}^{-1}$ and is injective, with image the trigonal locus of Maroni invariant $2$.
Proof. The $p$-fibers are the ${\rm Aut}(\mathbf{F}_4)$-orbits because the ${\rm Aut}(\mathbf{F}_4)$-orbits correspond to the isomorphism classes of trigonal curves of Maroni invariant $2$, and the Torelli map recovers these isomorphism classes. Since $\mathcal {M}_{\Lambda }^{0}$ is normal, [Reference Mumford, Fogarty and KirwanMFK94, Proposition 0.2] tells us that the image of $p$ is identified with the geometric quotient of $U$ by ${\rm Aut}(\mathbf{F}_4)$. It remains to show the surjectivity of $p$.
Let $(X, \iota )$ be an arbitrary member of $\mathcal {M}_{\Lambda }^{0}$, and let $\{{\frak e},{\frak f}\}$ be the natural hyperbolic basis of its invariant lattice $H_+\simeq \mathbb {U}$. The vector $2({\frak e}+{\frak f})$ satisfies the arithmetic conditions in Lemma 3.6 and hence gives an $\iota$-invariant hyperelliptic bundle of degree $8$. This defines a generically two-to-one morphism $X\to Y$ where $Y=\mathbf{P}^{2}$ or $\mathbf{F}_{2n}$ with $n\leq 2$, on which $\iota$ acts by the covering transformation by Lemma 3.8. Among these possibilities of $Y$, the branch ($-2K_{Y}$)-curve can contain a component of genus $10$ only when $Y=\mathbf{P}^{2}$ or $\mathbf{F}_{4}$. The case $Y=\mathbf{P}^{2}$ cannot happen because it would be $(r, l, \delta )=(1, 1, 1)$ in that case. Hence $Y=\mathbf{F}_{4}$. Since $|{-}2K_{\mathbf{F}_{4}}|=\Sigma +|L_{3,0}|$ and since $L_{3,0}$ has arithmetic genus $10$, the branch curve should be of the form $\Sigma +C$ with smooth $C\in |L_{3,0}|$. Thus $(X, \iota )=p(C)$.
Let $C$ be a member of $U$. Since $C$ is disjoint from $\Sigma$, the bundle $L_{1,-4}|_{C}$ is trivial so that $K_C\simeq L_{1,2}|_C\simeq L_{0,6}|_C$. Hence $L_{0,3}|_C$ is a theta characteristic with $h^{0}(L_{0,3}|_C)=4$. This shows that $\bar {\mu }_{\Lambda }(\mathcal {M}_{\Lambda }^{0})\subset \frak {M}_{10}'$. We are interested in the locus $\bar {\mu }_{\Lambda }(\mathcal {M}_{\Lambda }^{0})\cap \frak {M}_{10}''$ where $C$ has another effective even theta characteristic.
Lemma 7.7 The curve $C$ has an effective even theta characteristic different from $L_{0,3}|_C$ if and only if there exists a smooth member $H$ of $|L_{1,0}|$ such that $H|_C=2D$ for some divisor $D$ of degree $6$ on $C$.
Proof. By Lemma 6.12, $C$ has a theta characteristic $L$ with $h^0(L)\geq 2$ if and only if there exists $H\in |L_{1,0}|$ with $H|_C$ divisible by $2$, in which case $L$ is given by $(H|_C)/2+L_{0,1}|_C$. If $H$ is singular, it contains $\Sigma$ as a component and hence can be written as $H=\Sigma +\sum _{i=1}^{4}F_i$ for some $L_{0,1}$-fibers $F_1, \ldots , F_4$. Since $H|_C=\sum _{i=1}^{4}F_i|_C$, after renumbering we must have $F_1=F_2$ and $F_3=F_4$. Thus $L=L_{0,3}|_C$ in this case.
We can now complete the proof of Proposition 7.1. Since $\bar {\mu }_{\Lambda }({\mathcal {M}}_{\Lambda }^{0})\subset {\frak M}'_{10}$, the inverse image $\bar {\mu }_{\Lambda }^{-1}({\frak M}''_{10})$ is a divisor in ${\mathcal {M}}_{\Lambda }^{0}$. It can be easily checked with Lemma 7.7 that $\bar {\mu }_{\Lambda }({\mathcal {M}}_{\Lambda }^{0})\cap {\frak M}''_{10}$ is non-empty. Hence it remains to prove the inclusion $\bar {\mu }_{\Lambda }^{-1}({\frak M}''_{10})\subset \bar {\mathcal {H}}_{\Lambda }$. Let $(X,\iota )$ be a $2$-elementary $K3$ surface with $(g, k)=(10, 1)$ such that $\bar {\mu }_{\Lambda }(X,\iota )\in {\frak M}''_{10}$. By Lemmas 7.6 and 7.7, we have $(X, \iota )=p(C)$ for a curve $C$ as in Lemma 7.7. Let $H$ be a smooth $L_{1,0}$-curve with $H|_C$ divisible by $2$. The pullback of $H$ by the covering map $\pi : X\to \mathbf{F}_4$ splits into two $(-2)$-curves: $\pi ^{\ast }H=E+\iota (E)$. Therefore $X$ has the $\iota$-anti-invariant cycle
of norm $-16$. Let $H_-\subset H^{2}(X,\textbf {Z})$ be the anti-invariant lattice of $(X, \iota )$. Since
and since $H_{-}$ is unimodular, $D_{-}$ is divisible by $2$ in $H_-$. Therefore ${\rm Pic}(X)$ contains the anti-invariant cycle $D_{-}/2$ of norm $-4$, which implies that the period of $(X, \iota )$ lies in $\bar {{\mathcal {H}}}_{\Lambda }$. This proves Proposition 7.1.
7.2 Proof of Proposition 7.2
In this subsection we prove Proposition 7.2. We first explain the decomposition of the Heegner divisor $\bar {\mathcal {H}}_{\Lambda }$. Recall that $\bar {\mathcal {H}}_{\Lambda }$ is defined by norm $-4$ vectors $l$ in $\Lambda \simeq \mathbb {U}\oplus \mathbb {U}(2)\oplus \mathbb {E}_8^{\oplus 2}$. Since $\Lambda$ cannot contain $\langle -4 \rangle$ as an orthogonal direct summand, we have either ${\rm div}(l)=1$ or $2$. By the Eichler criterion, each type of norm $-4$ vectors consists of a single $O^{+}_{0}(\Lambda )$-orbit (in the case ${\rm div}(l)=2$, $[l/2]\in A_{\Lambda }$ is the unique element of norm $\equiv 1$ mod $2\textbf {Z}$). We accordingly obtain the decomposition
where $\bar {\mathcal {H}}_{i}$ is defined by those $l$ with ${\rm div}(l)=i$, and each $\bar {\mathcal {H}}_i$ is irreducible.
We next recall a (well-known) construction of members of $\mathcal {M}_{\Lambda }^0$. Let $\tilde {U}$ be the parameter space of smooth $(2, 4)$ complete intersections in $\mathbf{P}^3$. For each $C\in \tilde {U}$ the quadric containing it is unique; $\tilde {U}$ is thus stratified according to whether the quadric is smooth or a quadratic cone. In the first case $C$ is a smooth bidegree $(4, 4)$ curve on $Y=\mathbf{P}^1 \times \mathbf{P}^1$, and in the second case $C$ is a smooth $L_{4,0}$-curve on $Y=\mathbf{F}_2$. We have a period map $p:\tilde {U}\to \mathcal {M}_{\Lambda }^0$ by associating to $C$ the double cover of $Y$ branched over $C$.
Lemma 7.8 The period map $p$ descends to a biregular isomorphism $\tilde {U}/{\rm PGL}_4 \to \mathcal {M}_{\Lambda }^0$ from the geometric quotient $\tilde {U}/{\rm PGL}_4$.
Proof. This is similar to Lemma 7.6, so we only indicate minimal ingredients of the argument: (1) using the natural norm $4$ vector in the invariant lattice $\mathbb {U}(2)$, we can realize a given $(X, \iota )$ as a double cover of a quadric so that $p$ is surjective; (2) the $p$-fibers are ${\rm PGL}_4$-orbits either by an argument modeled in § 4.3 of [Reference MaMa15] or by observing that the ${\rm PGL}_4$-orbits correspond to the isomorphism classes of curves (cf. [Reference Arbarello, Cornalba, Griffiths and HarrisACGH85, Exercise IV. F-2]).
In the proof of Proposition 7.2 we restrict ourselves to the generic case, namely the smooth quadric case. The quadratic cone case can be dealt with similarly. So let $Y=\mathbf{P}^1 \times \mathbf{P}^1$ and consider the open locus $U\subset |\mathcal {O}_{Y}(4, 4)|$ of smooth bidegree $(4, 4)$ curves. If $C\in U$, then $K_C\simeq \mathcal {O}_{Y}(2, 2)|_C$ and the restriction map $|\mathcal {O}_{Y}(2, 2)| \to |K_C|$ is isomorphic. In particular, an effective divisor $D$ of degree $8$ on $C$ satisfies $2D \sim K_C$ if and only if there exists a bidegree $(2, 2)$ curve $H$ on $Y$ with $H|_C=2D$. We have one apparent theta characteristic, $\mathcal {O}_{Y}(1, 1)|_C$, which has $h^{0}=4$. This is the case where $H$ is a double bidegree $(1, 1)$ curve. By [Reference Arbarello, Cornalba, Griffiths and HarrisACGH85, Exercise IV. F-2], any other effective even theta characteristic of $C$, if it exists, must satisfy $h^{0}=2$.
Now let $V\subset U$ be the locus where $C$ has a theta characteristic with $h^0=2$. By the same argument as in § 7.1, the proof of Proposition 7.2 is reduced to showing the inclusion $p(V)\subset \bar {\mathcal {H}}_1$. So let $C\in V$ and $(X, \iota )=p(C)$. We write $\pi \colon X\to Y$ for the covering map. The curve $C$ admits a one-dimensional family $\{ H_{t} \}_{t\in \mathbf{P}^1}$ of ‘totally tangent’ curves of bidegree $(2, 2)$, that is, $H_t|_C=2D_t$ for some divisors $D_t$ of degree $8$ on $C$, which are not double bidegree $(1, 1)$ curves. Then $\{ D_{t} \}_{t\in \mathbf{P}^1}$ is a (complete) half-canonical pencil of $C$. Note that this pencil can also be obtained by picking up $t=0$ and considering the linear system of bidegree $(2, 2)$ curves passing through $D_0$, which intersect $C$ at $D_0+D_t$.
Claim 7.9 A general member of $\{ H_{t} \}_{t\in \mathbf{P}^1}$ is smooth.
Proof. If $H_t$ is reducible, its irreducible components are smooth rational curves intersecting $C$ transversely at at most two points and tangent to $C$ elsewhere. Their pullback to $X$ splits into two $(-2)$-curves. So if a general member is reducible, then the $K3$ surface $X$ would be covered by rational curves, which is absurd. By the same reason, a general (irreducible) member cannot be singular.
Let $H$ be a general member of $\{ H_{t} \}_{t\in \mathbf{P}^1}$. Since $H$ is totally tangent to $C$ at eight points, its pullback to $X$ splits into two smooth elliptic curves: $\pi ^{\ast } H = E + \iota (E)$. Hence $(X, \iota )$ possesses the $\iota$-anti-invariant cycle of norm $-16$:
Claim 7.10 $D_{-}$ is divisible by $2$ in the anti-invariant lattice $H_-$ and satisfies $(D_{-}/2, H_-)=\textbf {Z}$.
Proof. Since $D_{-}/2 = E - \pi ^{\ast }\mathcal {O}_Y(1, 1)$ is contained in $H^{2}(X, \textbf {Z})$, we have $D_{-}/2\in H_{-}$. If $(D_{-}/2, H_-)\ne \textbf {Z}$, then $(D_{-}/2, H_-)\subset 2\textbf {Z}$ so that $D_{-}/4$ would be contained in the dual lattice $H_{-}^{\vee }$. Recall that the discriminant forms $A_{H_{+}}$, $A_{H_{-}}$ are isometric to $A_{\mathbb {U}(2)}$. As elements of $A_{H_-}$ and $A_{H_+}$, $[D_{-}/4]$ and $[\pi ^{\ast }\mathcal {O}_Y(1, 1)/2]$ are respectively the unique elements of norm $\equiv 1$ mod $2\textbf {Z}$. By Nikulin [Reference NikulinNik80], then $E/2 = D_{-}/4 + \pi ^{\ast }\mathcal {O}_Y(1, 1)/2$ would be contained in $H^{2}(X, \textbf {Z})$. This contradicts the well-known fact that the class of a smooth elliptic curve is primitive in ${\rm Pic}(X)$.
To sum up, if a smooth bidegree $(4, 4)$ curve $C\subset Y$ has a theta characteristic with $h^{0}=2$, then the associated 2-elementary $K3$ surface $(X, \iota )$ has an anti-invariant cycle $D_{-}/2$ of norm $-4$ and with $(D_{-}/2, H_{-})=\textbf {Z}$ in its Picard lattice. Hence the period of $(X, \iota )$ lies in the component $\bar {\mathcal {H}}_1$ of $\bar {\mathcal {H}}_{\Lambda }$. This finishes the proof of Proposition 7.2.
7.3 Proof of Lemmas 7.3 and 7.4
Lemma 7.3 is an immediate consequence of the following known description of general members of $\bar {\mu }_{\Lambda }(\mathcal {M}_{\Lambda }^{0})$. When $g=4$, they are general curves in ${\frak M}'_{4}$ by Ma [Reference MaMa15, Corollary 9.10]; when $g=5$, they are general trigonal curves with vanishing theta-null by Kondō [Reference KondōKon94]. Since these curves are not hyperelliptic, Lemma 7.3 follows.
For the proof of Lemma 7.4 we use the generic description given in [Reference MaMa15, Corollary 7.11]. Let $C$ be a smooth curve on $\mathbf{F}_{2}$ belonging to the linear system $|L_{3,1}|$ such that the $L_{0,1}$-fiber $F$ through the point $C\cap \Sigma$ intersects $C$ with multiplicity $3$ there. By taking the resolution of the double cover of $\mathbf{F}_2$ branched over $C+F+\Sigma$, we obtain a 2-elementary $K3$ surface with $(r, l, \delta )=(10, 0, 0)$. This construction covers general members of $\mathcal {M}_{\Lambda }^{0}$, so a general member of $\bar {\mu }_{\Lambda }(\mathcal {M}_{\Lambda }^{0})$ is a curve $C$ as above.
Denote $p:=C\cap \Sigma$ and $T:=L_{0,1}|_C$. Since
the divisor $T+2p$ gives a theta characteristic of $C$ with $h^0(T+2p)\geq 2$. By Lemma 6.12, we have $h^0(T+2p)=2$. Conversely, suppose we have a theta characteristic $L$ on $C$ with $h^0(L)\geq 2$. By Lemma 6.12 we can find a curve $H\in |L_{1,-1}|$ with $H|_C=2D$ for some divisor $D$ satisfying $L\sim T+D$. Since $|L_{1,-1}|=\Sigma +|L_{0,1}|$, $H$ is of the form $H=\Sigma +F'$ for some $F'\in |L_{0,1}|$. The condition $(\Sigma +F')|_C=2D$ forces $F'$ to pass through $p=\Sigma \cap C$. Therefore $F=F'$ and $2D=4p$. Thus $L$ is uniquely determined as $L\sim T+2p$. This proves Lemma 7.4.
8. The structure of $\tau _{M}$: the case $\delta =1$
In this section we determine the structure of $\Phi _{M}$ when $\delta =1$.
8.1 Borcherds products for $2$-elementary lattices
Recall that the Dedekind $\eta$-function and the Jacobi theta series $\theta _{{\Bbb A}_{1}^{+}+\epsilon /2}(\tau )$, $(\epsilon =0,1)$ are the holomorphic functions on the complex upper half-plane ${\frak H}$:
Let $\Lambda \subset {\Bbb L}_{K3}$ be a primitive $2$-elementary sublattice of signature $(2,r(\Lambda )-2)$. We set
Let $\{\textbf {e}_{\gamma }\}_{\gamma \in A_{\Lambda }}$ be the standard basis of the group ring $\textbf {C}[A_{\Lambda }]$. For $0\leq j\leq 3$, set $\textbf {v}_{j}:=\sum _{\gamma \in A_{\Lambda }, q_{\Lambda }(\gamma )\equiv j/2}\textbf {e}_{\gamma }$. By [Reference YoshikawaYos13, Definition 7.6, Theorem 7.7], the $\textbf {C}[A_{\Lambda }]$-valued function
is a modular form for ${\rm Mp}_{2}(\textbf {Z})$ of weight $1-b^{-}(\Lambda )/2$ with respect to the Weil representation $\rho _{\Lambda }\colon {\rm Mp}_{2}(\textbf {Z})\to {\rm GL}(\textbf {C}[A_{\Lambda }])$ attached to $\Lambda$ [Reference BorcherdsBor98], where ${\rm Mp}_{2}(\textbf {Z})$ is the metaplectic cover of ${\rm SL}_{2}(\textbf {Z})$. By [Reference YoshikawaYos13, (7.9)], the principal part of $F_{\Lambda }$ is given by
By (8.1), we easily see that ${\mathcal {P}}_{\leq 0}[F_{\Lambda }]=0$ if and only if $(r(\Lambda ),\delta (\Lambda ))=(16,0)$.
Let $\ell \in \textbf {Z}_{>0}$ be such that $2^{r(\Lambda )-16}|\ell$ for all $\Lambda$. Then $\ell F_{\Lambda }(\tau )$ has integral Fourier expansion at $+i\infty$. Define $\Psi _{\Lambda }^{\ell }$ as the Borcherds lift of $\ell F_{\Lambda }(\tau )$ (cf. [Reference BorcherdsBor98, Theorem 13.3]):
If $r(\Lambda )\leq 16$, then $\Psi _{\Lambda }(\cdot ,F_{\Lambda })$ is well defined and $\Psi _{\Lambda }^{\ell }=\Psi _{\Lambda }(\cdot ,\ell F_{\Lambda })=\Psi _{\Lambda }(\cdot ,F_{\Lambda })^{\ell }$ in the ordinary sense. Since $O(\Lambda )$ (equivalently, $O(q_{\Lambda })$) acts trivially on $\ell F_{\Lambda }$ by [Reference YoshikawaYos13, Theorem 7.7, (2)], $\Psi _{\Lambda }^{\ell }$ is an automorphic form on $\Omega _{\Lambda }$ for $O^{+}(\Lambda )$ by [Reference BorcherdsBor98, Theorem 13.3]. Recall that the divisors ${\mathcal {D}}_{\Lambda }^{+}$, ${\mathcal {D}}_{\Lambda }^{-}$ and ${\mathcal {H}}_{\Lambda }$ were introduced in §§ 2.2 and 2.3.
Theorem 8.1 The weight and the divisor of $\Psi _{\Lambda }^{\ell }$ are given as follows.
(1) If $r(\Lambda )\leq 20$, then
\begin{gather*} {\rm wt}(\Psi_{\Lambda}^{\ell}) = \begin{cases} \begin{array}{@{}ll} (16-r(\Lambda))(2^{g(\Lambda)}+1)\ell & (r(\Lambda)\not=12,20), \\ (16-r(\Lambda))(2^{g(\Lambda)}+1)\ell-8(1-\delta(\Lambda))\ell & (r(\Lambda)=12), \\ (16-r(\Lambda))(2^{g(\Lambda)}+1)\ell-(28-r(\Lambda))2^{15-r(\Lambda)}(1-\delta(\Lambda))\ell & (r(\Lambda)=20), \end{array} \end{cases}\\ {\rm div}(\Psi_{\Lambda}^{\ell}) = \ell\, \{ {\mathcal{D}}_{\Lambda}^{-} + (2^{g(\Lambda)}+1) {\mathcal{D}}_{\Lambda}^{+} - 2^{16-r(\Lambda)} {\mathcal{H}}_{\Lambda}\}. \end{gather*}(2) If $r(\Lambda )=21$, then
\[ \begin{array}{c} {\rm wt}(\Psi_{\Lambda}^{\ell}) = (16-r(\Lambda))(2^{g(\Lambda)}+1)\ell = -5^{3}\cdot 41\ell,\\ \begin{aligned} {\rm div}(\Psi_{\Lambda}^{\ell}) & = \ell [ {\mathcal{D}}_{\Lambda}^{-} + (2^{g(\Lambda)}+1) {\mathcal{D}}_{\Lambda}^{+} - 2^{16-r(\Lambda)}\{ {\mathcal{H}}_{\Lambda} + (28-r(\Lambda)) {\mathcal{D}}_{\Lambda}^{+}\} ]\\ & = 2^{-5}\ell\cdot \{ 32{\mathcal{D}}_{\Lambda}^{-} + 3\cdot17\cdot643\cdot{\mathcal{D}}_{\Lambda}^{+} - {\mathcal{H}}_{\Lambda} \}. \end{aligned} \end{array} \]
Proof. By using (8.1), the result follows from [Reference BorcherdsBor98, Theorem 13.3].
The Petersson norm $\|\Psi _{\Lambda }^{\ell }\|=\|\Psi _{\Lambda }(\cdot ,\ell F_{\Lambda })\|$ is an $O(\Lambda )$-invariant function on $\Omega _{\Lambda }$. We identify $\|\Psi _{\Lambda }^{\ell }\|$ with the corresponding function on ${\mathcal {M}}_{\Lambda }$ and set
Then $\|\Psi _{\Lambda }(\cdot ,F_{\Lambda })\|$ is independent of the choice of $\ell \in \textbf {Z}_{>0}$ with $2^{r(\Lambda )-16}|\ell$. If $r(\Lambda )\leq 16$, then $\|\Psi _{\Lambda }(\cdot ,F_{\Lambda })\|$ is the ordinary Petersson norm of $\Psi _{\Lambda }(\cdot ,F_{\Lambda })$.
8.2 The structure of $\Phi _{M}$: the case $\delta =1$
Write $M_{g,k}$ for a primitive $2$-elementary Lorentzian sublattice of ${\Bbb L}_{K3}$ such that
Then $M_{g,0}\cong {\Bbb A}_{1}^{+}\oplus {\Bbb A}_{1}^{\oplus 10-g}$ and $(r,l,\delta )=(11-g,11-g,1)$ for $M_{g,0}$. Set
Lemma 8.2 There exist mutually perpendicular roots $d_{1},\ldots ,d_{k}\in \Delta _{\Lambda _{g,0}}^{+}$ with
In particular, if $M_{g,k}$ exists, one has $\Omega _{\Lambda _{g,k}}=\Omega _{\Lambda _{g,k-1}}\cap H_{d_{k}}$.
By Theorem 6.1, there exist integers $a_{g},b_{g},c_{g}\in \textbf {Z}_{\geq 0}$ for $3\leq g\leq 10$ with
Lemma 8.3 The following inequalities and equality hold:
Proof. We get $a_{g}>0$ by [Reference YoshikawaYos13, Proposition 4.2 (2)] and $b_{g}>0$ by Theorem 6.1. Let $6\leq g\leq 9$. Recall that the Zariski open subset ${\mathcal {D}}_{\Lambda }^{0,+}$ (respectively, ${\mathcal {D}}_{\Lambda }^{0,-}$) of ${\mathcal {D}}_{\Lambda }^{+}$ (respectively, ${\mathcal {D}}_{\Lambda }^{-}$) was defined in § 2.2. By (3.5) and Proposition 6.2, we get $J_{M_{g,0}}({\mathcal {D}}_{\Lambda _{g,0}}^{0,+})=J_{M_{g,1}}(\Omega _{\Lambda _{g,1}}^{0})\not \subset \theta _{{\rm null},g}$, which implies $c_{g}=0$ for $6\leq g\leq 9$. Let $1\leq g\leq 5$. By [Reference YoshikawaYos13, Proposition 4.2 (1)], $J_{M_{g,g}}(\Omega _{\Lambda _{g,g}}^{0})\not \subset \theta _{{\rm null},g}$. Since $J_{M_{g,g-1}}({\mathcal {D}}_{\Lambda _{g,g-1}}^{0,+})=J_{M_{g,g}}(\Omega _{\Lambda _{g,g}}^{0})$ by (3.5) and Lemma 8.2, we get $J_{M_{g,g-1}}(\Omega _{\Lambda _{g,g-1}}^{0})\not \subset \theta _{{\rm null},g}$ because ${\mathcal {D}}_{\Lambda _{g,g-1}}^{0,+}\subset \Omega _{\Lambda _{g,g-1}}$. By (3.5) and Lemma 8.2 again, we get $J_{M_{g,g-2}}({\mathcal {D}}_{\Lambda _{g,g-2}}^{0,+})=J_{M_{g,g-1}}(\Omega _{\Lambda _{g,g-1}}^{0})\not \subset \theta _{{\rm null},g}$. In the same way, we get inductively $J_{M_{g,k}}({\mathcal {D}}_{\Lambda _{g,k}}^{0,+})\not \subset \theta _{{\rm null},g}$ for $k\leq g-1$. This proves $c_{g}=0$ for $3\leq g\leq 5$.
Proposition 8.4 If $3\leq g\leq 10$, the following equality of divisors on $\Omega _{\Lambda _{g,k}}$ holds:
Proof. When $g=10$, $M_{g,k}$ makes sense only for $k=0$ by Proposition 1.1 . Hence the assertion is obvious by (8.2) when $g=10$. We must prove (8.3) when $g\leq 9$. We will prove by induction the existence of integers $a_{g,k},b_{g,k}\in \textbf {Z}$ with
When $k=0$, the assertion follows from (8.2) and Lemma 8.3. We assume (8.4) for $M_{g,k}$. By Lemma 8.2, there exists $d\in \Delta ^{+}_{\Lambda _{g,k}}$ such that $\Lambda _{g,k}=\Lambda _{g,k+1}\oplus \textbf {Z}d$. Then $\Omega _{\Lambda _{g,k+1}}=H_{d}\cap \Omega _{\Lambda _{g,k}}$. Let $i\colon \Omega _{\Lambda _{g,k+1}}=H_{d}\cap \Omega _{\Lambda _{g,k}}\hookrightarrow \Omega _{\Lambda _{g,k}}$ be the inclusion. By the compatibility of Torelli maps (3.5), we get the equality of holomorphic maps
Hence we get the equality of holomorphic sections on $\Omega _{\Lambda _{g,k+1}}^{0}\cup {\mathcal {D}}_{\Lambda _{g,k+1}}^{0}$:
Since $\Omega _{\Lambda _{g,k+1}}\setminus (\Omega _{\Lambda _{g,k+1}}^{0}\cup {\mathcal {D}}_{\Lambda _{g,k+1}}^{0})$ has codimension at least $2$ in $\Omega _{\Lambda _{g,k+1}}$, we deduce from (8.4) and (8.6) the equality of divisors on $\Omega _{\Lambda _{g,k+1}}$:
Here the last equality follows from Propositions 2.1 and 2.2. This proves (8.4) for $M_{g,k+1}$. By induction, (8.4) holds for all $M_{g,k}$.
Since $a_{g,k+1}=a_{g,k}$ and $b_{g,k+1}=2b_{g,k}$ for $k\geq 0$ by (8.4), (8.7), we get
for $g\leq 9$ and $k\geq 0$. This proves the result.
Proposition 8.5 If $3\leq g\leq 10$, then the following inequalities hold:
Proof. The inequality $a_{g}\geq 2^{2g-1}$ follows from [Reference YoshikawaYos13, Proposition 4.2 (2)]. (To confirm $J_{M}(\Omega _{\Lambda }^{0})\not \subset \theta _{{\rm null},g}$, it is assumed there that either $r>10$ or $(r,\delta )=(10,1)$. Since the same proof of [Reference YoshikawaYos13, Proposition 4.2 (2)] works under the same assumption $J_{M}(\Omega _{\Lambda }^{0})\not \subset \theta _{{\rm null},g}$ and since $J_{M_{g,0}}(\Omega _{\Lambda _{g,0}}^{0})\not \subset \theta _{{\rm null},g}$ for $3\leq g\leq 10$ by Theorem 6.1, the same conclusion as in [Reference YoshikawaYos13, Proposition 4.2 (2)] still holds for $M_{g,0}$.) Let us prove the second inequality.
Let $j\colon {\frak M}_{g}\ni C\to {\rm Jac}(C)\in {\mathcal {A}}_{g}$ be the Torelli map. By [Reference Teixidor i BigasTei88, Proposition 3.1] or [Reference TsuyumineTsu91, Proof of Theorem 1, p. 542], we get the equality of divisors on ${\frak M}_{g}$:
Set $\Lambda :=\Lambda _{g,0}$ and ${\mu }_{\Lambda }^{0}:=\bar {\mu }_{\Lambda }\circ \Pi _{\Lambda }|_{\Omega _{\Lambda }^{0}}$, where $\Pi _{\Lambda }\colon \Omega _{\Lambda }\to {\mathcal {M}}_{\Lambda }$ is the projection. Since $J_{M_{g,0}}|_{\Omega _{\Lambda }^{0}}=j\circ {\mu }_{\Lambda }^{0}$, we get by (8.9) the following equality of divisors on $\Omega _{\Lambda }^{0}$:
Since ${\rm Supp}(({\mu }_{\Lambda }^{0})^{*}{\frak M}'_{g})\subset {\mathcal {H}}_{\Lambda }\cap \Omega _{\Lambda }^{0}$ and since $\bar {\mathcal {H}}_{\Lambda }$ is irreducible by Theorem 6.1, there exists $\beta _{g}\in \textbf {Z}_{>0}$ such that the following equality of divisors on $\Omega _{\Lambda }^{0}$ holds:
By (8.10), (8.11), we get the equality of divisors on $\Omega _{\Lambda }^{0}$:
Comparing (8.2) with (8.12), we get $b_{g}=2^{4}\beta _{g}\geq 2^{4}$. This completes the proof.
Theorem 8.6 There exists a constant $C_{g,k,\ell }$ depending only on $g$, $k$, $\ell$ such that the following equality of automorphic forms on $\Omega _{\Lambda _{g,k}}$ holds:
In particular, there exists a constant $C_{g,k}$ depending only on $g$, $k$ such that
Proof. Since the result was proved in [Reference YoshikawaYos13, Theorem 9.1] in the case $g\leq 2$, we assume $g\geq 3$ in what follows. Set $M:=M_{g,k}$ and $\Lambda :=\Lambda _{g,k}$. Then $r(\Lambda )\geq 5$ and hence ${\mathcal {M}}_{\Lambda }^{*}\setminus {\mathcal {M}}_{\Lambda }$ has codimension at least $2$ in ${\mathcal {M}}^{*}$. By definition, $M=M_{g,k}$ has invariants $r(M)=11+k-g$ and $l(M)=11-k-g$, so that $r(\Lambda )=11-k+g$ and $l(\Lambda )=11-k-g$.
Since ${\rm wt}(\chi _{g})=2^{g-2}(2^{g}+1)$ and ${\rm wt}(\Psi _{\Lambda }^{\ell })=\{16-r(\Lambda )\}(2^{g}+1)\ell$ by Theorem 8.1, we get
Since ${\rm wt}(\Phi _{\Lambda })=\ell (16-r(\Lambda ),4)$ by Theorem 5.1, we get by (8.13) that
By (8.14),
descends to a meromorphic function on ${\mathcal {M}}_{\Lambda }$. Since ${\mathcal {M}}_{\Lambda }^{*}$ is normal and since $\dim {\mathcal {M}}_{\Lambda }^{*}\setminus {\mathcal {M}}_{\Lambda }\leq \dim {\mathcal {M}}_{\Lambda }^{*}-2$, $\varphi _{\Lambda }$ extends to a meromorphic function on ${\mathcal {M}}_{\Lambda }^{*}$.
(1) Let $r(M)\geq 2$. Hence $r(\Lambda )\leq 20$. By Theorem 8.1 and Proposition 8.4,
Since ${\rm div}(\Phi _{\Lambda })=\ell {\mathcal {D}}_{\Lambda }$, we deduce from (8.15) that
Since $a_{g}\geq 2^{2g-1}$ and $b_{g}\geq 2^{4}$ by Proposition 8.5, the divisor of $\varphi _{\Lambda }$ is effective by (8.16). Since $\varphi _{\Lambda }$ is a holomorphic function on ${\mathcal {M}}_{\Lambda }^{*}$, $\varphi _{\Lambda }$ must be a constant function on ${\mathcal {M}}_{\Lambda }^{*}$, which implies that
This completes the proof when $r(M)\geq 2$.
(2) Assume $r(M)=1$. Then $M={\Bbb A}_{1}^{+}$ and $r(\Lambda )=21$, $g=10$, $k=0$, $\delta =1$. By Theorem 8.1 (2), Proposition 8.4 and the equality $(g-1)+16-r(\Lambda )=4$, we get
Since ${\rm div}(\Phi _{\Lambda })=\ell {\mathcal {D}}_{\Lambda }$, we get by (8.18) that
Since $a_{g}\geq 2^{2g-1}$ and $b_{g}\geq 2^{4}$ by Proposition 8.5, it follows from (8.19) that $\varphi _{\Lambda }$ is a non-zero holomorphic function on ${\mathcal {M}}_{\Lambda }\setminus \bar {\mathcal {D}}_{\Lambda }^{+}$. By Lemma 8.7 below, $\varphi _{\Lambda }$ is a non-zero constant. This completes the proof of Theorem 8.6.
Lemma 8.7 When $M\simeq {\Bbb A}_{1}^{+}$, any holomorphic function on ${\mathcal {M}}_{\Lambda }\setminus \bar {\mathcal {D}}_{\Lambda }^{+}$ is a constant.
Proof. Let $U\subset |\mathcal {O}_{\mathbf{P}^2}(6)|$ be the space of smooth plane sextics, and let $V\subset |\mathcal {O}_{\mathbf{P}^2}(6)|$ be that of sextics with at most one node. By the stability criterion for plane sextics [Reference ShahSha80], we have a geometric quotient $V/{\rm PGL}_3$ of $V$ by ${\rm PGL}_3$, which contains $U/{\rm PGL}_3$ as an open set. It is well known that ${\mathcal {M}}_{\Lambda }^{0}$ is isomorphic to $U/{\rm PGL}_3$ by associating to a smooth plane sextic the double covers of $\mathbf{P}^2$ branched over it. Shah [Reference ShahSha80] has shown that this isomorphism extends to an open embedding $V/{\rm PGL}_3 \hookrightarrow \mathcal {M}_{\Lambda }$ and that its image is contained in ${\mathcal {M}}_{\Lambda }\setminus \bar {\mathcal {D}}_{\Lambda }^{+}$. Hence a holomorphic function on ${\mathcal {M}}_{\Lambda }\setminus \bar {\mathcal {D}}_{\Lambda }^{+}$ gives one on $V/{\rm PGL}_3$, which in turn is pulled back to $V$. Since the complement of $V$ in $|\mathcal {O}_{\mathbf{P}^2}(6)|$ is of codimension $2$ in $|\mathcal {O}_{\mathbf{P}^2}(6)|$, a holomorphic function on $V$ extends to $|\mathcal {O}_{\mathbf{P}^2}(6)|$ holomorphically and so is a constant.
Remark 8.8 Let $(g,k)=(10,0)$. Since $\varphi _{\Lambda }$ is constant when $\Lambda =({\Bbb A}_{1}^{+})^{\perp }$, we get
by (8.19). In particular, $J_{{\Bbb A}_{1}^{+}}^{*}\chi _{10}$ vanishes on ${\mathcal {D}}_{\Lambda }^{+}$. Since ${\Bbb U}=[{\Bbb A}_{1}^{+}\oplus \textbf {Z}d]$ for any $d\in \Delta _{\Lambda }^{+}$, this, together with (3.5), implies that $J_{\Bbb U}^{*}\chi _{10}$ vanishes identically on $\Omega _{{\Bbb U}^{\perp }}$.
9. The structure of $\tau _{M}$: the case $\delta =0$
In this section $M\subset {\Bbb L}_{K3}$ is assumed to be a primitive $2$-elementary Lorentzian sublattice with $\delta =0$. As before, we set $\Lambda =M^{\perp }$.
9.1 The structure of $\Phi _{M}$: the case $r\not =2,10$ and $\delta =0$
Lemma 9.1 Let $\Lambda \cong {\Bbb U}\oplus {\Bbb U}(k)\oplus {\Bbb D}_{4}\oplus {\Bbb E}_{8}$ with $k=1,2$. Then the following equality of divisors on $\Omega _{\Lambda }$ holds:
Proof. Recall that the lattice $\Lambda _{g,k}$ was defined in § 8. Since $\Lambda _{g,k-1}\cong \Lambda \oplus {\Bbb A}_{1}$ by comparing the invariants $(r,l,\delta )$, there is a root $d\in \Delta _{\Lambda _{g,k-1}}^{+}$ with $\Lambda _{g,k-1}\cap d^{\perp }\cong \Lambda$. Since $d\in \Delta _{\Lambda _{g,k-1}}^{+}$ we get, by (8.3), (8.8), (8.17) and Propositions 2.1 and 2.2,
To get the last equality, we used $g-1=k+4$, ${\mathcal {D}}_{\Lambda }^{+}=0$ and ${\mathcal {H}}_{\Lambda }={\mathcal {D}}_{\Lambda }$ for $\Lambda$, where the last two equalities follow from $\varepsilon _{\Lambda }=-2$, $\delta (\Lambda )=0$ and $\textbf {1}_{\Lambda }=0$.
Theorem 9.2 If $r\not =2,10$ and $\delta =0$, there is a constant $C_{M,\ell }>0$ depending only on $M$ and $\ell$ such that the following equality of automorphic forms on $\Omega _{\Lambda }$ holds:
In particular, there is a constant $C_{M}>0$ depending only on $M$ such that the following equality of automorphic forms on $\Omega _{\Lambda }$ holds:
Proof. For $r>10$, the result was proved in [Reference YoshikawaYos13, Theorem 9.1]. We may assume $2 < r < 10$ and $\delta =0$. By Proposition 1.1, we get $r=6$ and $\Lambda \cong {\Bbb U}\oplus {\Bbb U}(k)\oplus {\Bbb D}_{4}\oplus {\Bbb E}_{8}$, $k=1,2$. Since $\delta (\Lambda )=0$, we have ${\mathcal {D}}_{\Lambda }^{+}=0$ and $\textbf {1}_{\Lambda }=0$. Since $r(\Lambda )=16$, we have $\varepsilon _{\Lambda }=-2$, which, together with $\textbf {1}_{\Lambda }=0$, yields that ${\mathcal {H}}_{\Lambda }={\mathcal {D}}_{\Lambda }$. By Theorem 8.1, we get ${\rm wt}(\Psi _{\Lambda }^{\ell })=0$ and
which implies that $\Psi _{\Lambda }$ is a non-zero constant function on $\Omega _{\Lambda }$.
Set $\varphi _{\Lambda }:=\Psi _{\Lambda }^{2^{g-1}\ell }\otimes J_{M}^{*}\chi _{g}^{8\ell }/\Phi _{M}^{2^{g-1}(2^{g}+1)}$. Since $\Psi _{\Lambda }$ is a non-zero constant, we deduce from Lemma 9.1 and ${\rm wt}(\Phi _{M})=(0,4\ell )$, ${\rm div}(\Phi _{M})=\ell {\mathcal {D}}_{\Lambda }$ that
Hence $\varphi _{\Lambda }$ is a holomorphic function on ${\mathcal {M}}_{\Lambda }$ by (9.1) and extends holomorphically to ${\mathcal {M}}_{\Lambda }^{*}$. Thus $\varphi _{\Lambda }$ is a non-zero constant.
9.2 The structure of $\Phi _{M}$: the case $(r,\delta )=(10,0)$
In this subsection we assume that $M$ is non-exceptional and
Then $0\leq l\leq 8$ and $2\leq g\leq 6$. Since $J_{M}^{*}\chi _{g}$ vanishes identically on $\Omega _{\Lambda }$ (e.g. [Reference YoshikawaYos13, Proposition 9.3]), Theorem 8.6 does not hold in this case. Identify ${\frak M}_{g}$ with its image by the Torelli map $j\colon {\frak M}_{g}\hookrightarrow {\mathcal {A}}_{g}$. Then $J_{M}\colon \Omega _{\Lambda }^{0}\to {\mathcal {A}}_{g}$ is identified with the map $\mu _{\Lambda }=j^{-1}\circ J_{M}\colon \Omega _{\Lambda }^{0}\to {\frak M}_{g}$. Write ${\frak H}_{{\rm hyp},g}\subset {\frak M}_{g}$ for the hyperelliptic locus.
Proposition 9.3 Let $M$ be non-exceptional with $(r,\delta )=(10,0)$. Then $J_{M}^{*}\Upsilon _{g}$ does not vanish identically on $\Omega _{\Lambda }^{0}$. Moreover, for any $d\in \Delta _{\Lambda }$, $J_{[M\perp d]}^{*}\chi _{g-1}$ is nowhere vanishing on $\Omega _{\Lambda \cap d^{\perp }}^{0}$.
Proof. For the first assertion, it suffices to prove $\mu _{\Lambda }(\Omega _{\Lambda }^{0})\not \subset {\rm div}(\Upsilon _{g})\cap {\frak M}_{g}$. Since $\Upsilon _{2}$ is nowhere vanishing on the diagonal locus of ${\frak S}_{2}$ and since $\mu _{\Lambda }(\Omega _{\Lambda }^{0})$ is the image of the diagonal locus by the projection ${\frak S}_{2}\to {\mathcal {A}}_{2}$, we get ${\rm div}(\Upsilon _{2})\cap \mu _{\Lambda }(\Omega _{\Lambda }^{0})=\emptyset$. Similarly, we have ${\rm div}(\Upsilon _{3})\cap \mu _{\Lambda }(\Omega _{\Lambda }^{0})=\emptyset$ by [Reference IgusaIgu67, Lemma 11]. Let $g=4$. Since $\mu _{\Lambda }(\Omega _{\Lambda }^{0})\subset {\rm div}(\chi _{4})$, the inclusion $\mu _{\Lambda }(\Omega _{\Lambda }^{0})\subset {\rm div}(\Upsilon _{4})\cap {\frak M}_{4}$ would imply $\mu _{\Lambda }(\Omega _{\Lambda }^{0})\subset {\rm div}(\chi _{4})\cap {\rm div}(\Upsilon _{4})\cap {\frak M}_{4}$. Since the right-hand side coincides with ${\frak H}_{\rm hyp,4}$ by [Reference IgusaIgu82, Corollary, p. 544], this last inclusion contradicts Lemma 7.3. Thus $\mu _{\Lambda }(\Omega _{\Lambda }^{0})\not \subset {\rm div}(\Upsilon _{4})\cap {\frak M}_{4}$. Let $g=5$. Let $F_{5}$ be the Schottky form in genus $5$ (cf. [Reference Grushevsky and Salvati-ManniGS11, p. 1018]), whose zero divisor characterizes the (closure of) trigonal locus of ${\frak M}_{5}$ (cf. [Reference Grushevsky and Salvati-ManniGS11, Corollary 18]). By [Reference KondōKon94], a general point of $\mu _{\Lambda }(\Omega _{\Lambda }^{0})$ is contained in the intersection of the theta-null divisor and the trigonal locus. Then the inclusion $\mu _{\Lambda }(\Omega _{\Lambda }^{0})\subset {\rm div}(\Upsilon _{5})\cap {\frak M}_{5}$ would imply
Since the right-hand side coincides with ${\frak H}_{\rm hyp,5}$ by [Reference Fontanari and PascoluttiFP12, p. 67], this last inclusion contradicts Lemma 7.3. Thus $\mu _{\Lambda }(\Omega _{\Lambda }^{0})\not \subset {\rm div}(\Upsilon _{5})\cap {\frak M}_{5}$. When $g=6$, we get $\mu _{\Lambda }(\Omega _{\Lambda }^{0})\not \subset {\rm div}(\Upsilon _{6})\cap {\frak M}_{6}$ by Lemmas 4.1 and 7.4. This proves the first assertion. Since $r([M\perp d])>10$, the second assertion follows from [Reference YoshikawaYos13, Proposition 4.2 (1)]. This completes the proof.
Theorem 9.4 Let $M$ be non-exceptional with $(r,\delta )=(10,0)$. Then there is a constant $C_{M,\ell }>0$ depending only on $M$ and $\ell$ such that the following equality of automorphic forms on $\Omega _{\Lambda }$ holds:
In particular, there is a constant $C_{M}>0$ depending only on $M$ such that
Proof. Since $(r,\delta )=(10,0)$, we get $(r(\Lambda ),\delta (\Lambda ))=(12,0)$. By [Reference YoshikawaYos13, Theorem 8.1],
Since $J_{M}^{*}\Upsilon _{g}$ does not vanish identically on $\Omega _{\Lambda }^{0}$ by Proposition 9.3 and since $\bar {\mathcal {D}}_{\Lambda }$ is irreducible, there exist $a\in \textbf {Z}_{\geq 0}$ and an effective divisor ${\mathcal {E}}_{\Lambda }$ on $\Omega _{\Lambda }$ such that
Set $\varphi _{\Lambda }:=\Psi _{\Lambda }^{(2^{g-1}+1)\ell }\otimes J_{M}^{*}\Upsilon _{g}^{\ell }/\Phi _{M}^{(2^{g-1}+1)(2^{g}-1)}$. Comparing (9.2), (9.3) and ${\rm wt}(\Phi _{M})=(4\ell ,4\ell )$, ${\rm div}(\Phi _{M})=\ell {\mathcal {D}}_{\Lambda }$, we get
By Proposition 9.3, we can apply Lemma 4.3 to a general curve $\gamma \colon \Delta \to {\mathcal {M}}_{\Lambda }$ intersecting $\bar {\mathcal {D}}_{\Lambda }^{0}$ transversally. Since $a\geq 2(2^{2(g-1)}-1)$ by Lemma 4.3, we get ${\rm div}(\varphi _{\Lambda })\geq 0$. By the Koecher principle, $\varphi _{\Lambda }$ is a non-zero constant.
In the rest of this section we determine $\Phi _{M}$ for the remaining $M$, that is, those $M$ with $(r,\delta )=(2,0)$. Then, either $M\cong {\Bbb U}$ or ${\Bbb U}(2)$.
9.3 The structure of $\Phi _{\Bbb U}$
In this subsection we set
Then $g=10$ and $J_{M}^{*}\chi _{10}$ vanishes identically on $\Omega _{\Lambda }^{0}$.
Let $E_{4}(\tau )=\theta _{{\Bbb E}_{8}^{+}}(\tau )=1+240q+\cdots$ be the Eisenstein series of weight $4$ (or equivalently, the theta series of ${\Bbb E}_{8}^{+}$) and set
Then $f_{\Lambda }(\tau )$ is a modular form of weight $-8$. We prove the following theorem.
Theorem 9.5 There exists a constant $C_{M,\ell }>0$ such that
In particular, there is a constant $C_{M}>0$ depending only on $M=\Bbb U$ such that
For the proof of Theorem 9.5, we first prove the following lemma.
Lemma 9.6 $J_{M}^{*}\Upsilon _{10}$ is nowhere vanishing on $\Omega _{\Lambda }^{0}\setminus {\mathcal {H}}_{\Lambda }$.
Proof. Let $(X,\iota ,\alpha )$ be an arbitrary marked $2$-elementary $K3$ surface of type $M={\Bbb U}$ with period in $\Omega _{\Lambda }\setminus {\mathcal {H}}_{\Lambda }$. Let $C$ be the component of genus $10$ of $X^{\iota }$. Fix a symplectic basis of $H_{1}(C,\textbf {Z})$, so that $\Omega (C)\in {\frak S}_{10}$, where $\Omega (C)$ is the period of $C$ with respect to the symplectic basis. By Proposition 7.1, there is a unique even pair $(a_{0},b_{0})$, $a_{0},b_{0}\in \{0,\frac {1}{2}\}^{10}$ such that $\theta _{a_{0},b_{0}}(\Omega (C))=0$ and $\theta _{a,b}(\Omega (C))\not =0$ for all even pairs $(a,b)$ with $(a,b)\not =(a_{0},b_{0})$. Hence we get
This proves the lemma.
By Lemma 9.6, there exist $\alpha ,\beta \in \textbf {Z}_{>0}$ such that
To prove Theorem 9.5, we must determine $\beta$.
We use the following notation. Set
Let $d\in \Delta ^{+}_{L}$ be a generator of ${\Bbb A}_{1}$. Then $\Lambda =L\cap d^{\perp }$, $L^{\perp }={\Bbb A}_{1}^{+}$ and $A_{L}=\{0,\textbf {1}_{L}\}$, where $\textbf {1}_{L}=[d/2]$. As before, we make the identification
Let $[\eta ]\in H_{d}^{0}$ be an arbitrary point. Let $U\cong \Delta ^{19}$ be a small neighborhood of $[\eta ]$ in $\Omega _{L}$ equipped with a system of coordinates $(t,s_{1},\ldots ,s_{18})$ such that $U\cap {\mathcal {D}}_{L}=U\cap H_{d}=U\cap \Omega _{\Lambda }=\{(t,s)\in U;\,t=0\}\cong \Delta ^{18}$. Since $d\in \Delta ^{+}_{L}$ and hence $g(L)=g(\Lambda )=10$ by [Reference YoshikawaYos13, Lemma 11.5], the Torelli map $J_{{\Bbb A}_{1}^{+}}$ is a holomorphic map from $U$ to ${\mathcal {A}}_{10}$ by Theorem 3.5. Let $\Pi \colon {\frak S}_{10}\to {\mathcal {A}}_{10}$ be the projection. Since $U$ is contractible, $J_{{\Bbb A}_{1}^{+}}\colon U\to {\mathcal {A}}_{10}$ is liftable. Namely, there exists a holomorphic map $\tilde {J}_{{\Bbb A}_{1}^{+}}\colon U\to {\frak S}_{10}$ such that
Since $\tilde {J}_{{\Bbb A}_{1}^{+}}$ takes its values in ${\frak S}_{10}$, the value $\theta _{a,b}(\tilde {J}_{{\Bbb A}_{1}^{+}}(t,s))$ makes sense for all $(s,t)\in U$ and even $(a,b)$, $a,b\in \{0,1/2\}^{10}$. Since $\tilde {J}_{{\Bbb A}_{1}^{+}}^{*}\chi _{10}=\prod _{(a,b)\,{\rm even}}\tilde {J}_{{\Bbb A}_{1}^{+}}^{*}\theta _{a,b}$ is nowhere vanishing on $U\setminus ({\mathcal {D}}_{L}\cup {\mathcal {H}}_{L})$ and since there exists by Proposition 7.1 a unique even theta constant $\theta _{a_{0},b_{0}}$ vanishing identically on $H_{d}$, we get the following:
• $U\cap H_{d}$ is a component of $U\cap {\rm div}(\tilde {J}_{{\Bbb A}_{1}^{+}}^{*}\theta _{a_{0},b_{0}})$;
• $U\cap H_{d}$ is not a component of $U\cap {\rm div}(\tilde {J}_{{\Bbb A}_{1}^{+}}^{*}\theta _{a,b})$ for any even $(a,b)\not =(a_{0},b_{0})$.
Thus there exist $c\in \textbf {Z}_{>0}$, $c_{a,b}(\lambda )\in \textbf {Z}_{\geq 0}$ such that, for $(a,b)=(a_{0},b_{0})$,
and such that, for all $(a,b)\not =(a_{0},b_{0})$,
Let $s_{d}\in O(L)$ be the reflection with respect to $d$. Since $J_{{\Bbb A}_{1}^{+}}\circ s_{d}=J_{{\Bbb A}_{1}^{+}}$, we have $s_{d}^{*}\circ (\tilde {J}_{{\Bbb A}_{1}^{+}})^{*}=(\tilde {J}_{{\Bbb A}_{1}^{+}})^{*}$, which implies the following equality for every even pair $(a,b)$:
Lemma 9.7 Let $\lambda \in L^{\lor }$ be such that $\lambda ^{2}=-9/2$, $|\langle \lambda ,d\rangle |=1$, $[\lambda ]=\textbf {1}_{L}$. If $c_{a_{0},b_{0}}(\lambda )>0$, then $c_{a,b}(\lambda )=0$ for all $(a,b)\not =(a_{0},b_{0})$.
Proof. Assume $c_{a_{0},b_{0}}(\lambda )>0$ and $c_{a',b'}(\lambda )>0$ for some even $(a',b')\not =(a_{0},b_{0})$. Then, for every $2$-elementary $K3$ surface $(X,\iota )$ of type ${\Bbb A}_{1}^{+}$ whose period lies in $U\cap H_{\lambda }\setminus H_{d}$, $X^{\iota }$ has two distinct effective even half canonical bundles corresponding to $(a_{0},b_{0})$ and $(a',b')$. This contradicts Remark 6.10.
Recall that $i\colon \Omega _{\Lambda }\hookrightarrow \Omega _{L}$ is the inclusion induced by that of lattices $\Lambda =L\cap d^{\perp }\subset L$. On $H_{d}^{0}\cap U$, set
By (3.5), we have $J_{M}=\Pi \circ \tilde {J}_{M}$.
Since $U\cap H_{d}\subset H_{d}^{0}=\Omega _{\Lambda }^{0}$, we have the following equality of divisors on $U\cap \Omega _{\Lambda }^{0}$ by (9.6):
Lemma 9.8 For every $\lambda \in L^{\lor }$ with $\lambda ^{2}=-9/2$, $\langle \lambda ,d\rangle =\pm 1$, $\lambda \equiv \textbf {1}_{L}\bmod L$, the following equalities hold:
Proof. Since $\tilde {J}_{{\Bbb A}_{1}^{+}}^{*}(\chi _{10}/\theta _{a_{0},b_{0}})^{8}|_{U\cap H_{d}} = \tilde {J}_{{\Bbb A}_{1}^{+}}^{*}\Upsilon _{10}|_{U\cap H_{d}}$ by the definitions of $\Upsilon _{g}$ and $\theta _{a_{0},b_{0}}$, we get the following equality of divisors on $U\cap \Omega _{\Lambda }^{0}$:
where the second equality follows from (9.10). On the other hand we get, by (9.8),
which, together with (2.2), yields the following equality of divisors on $U$:
Substituting (9.11), (9.13) into (9.12) and comparing the coefficients of $H_{\mu }$, we get
Since $c_{a,b}(\mu +{d}/{2})=c_{a,b}(\mu -{d}/{2})$ by (9.9), we get $\sum _{(a,b)\not =(a_{0},b_{0})}c_{a,b}(\lambda )=\beta /2$ by (9.14). If $c_{a_{0},b_{0}}(\lambda )>0$, then $\sum _{(a,b)\not =(a_{0},b_{0})}c_{a,b}(\lambda )=0$ by Lemma 9.7. Since $\beta >0$, this contradicts the equality $\sum _{(a,b)\not =(a_{0},b_{0})}c_{a,b}(\lambda )=\beta /2$. Thus $c_{a_{0},b_{0}}(\lambda )=0$.
Lemma 9.9 One has the equality $\beta =2^{5}$.
Proof. By (9.7), (9.8) and Lemma 9.8, we get on $U$,
which yields the equality of divisors ${\rm div}(J_{{\Bbb A}_{1}^{+}}^{*}\chi _{10}^{8})=c {\mathcal {D}}_{L}^{+}+(\beta /2)\cdot {\mathcal {H}}_{L}$ on $U$. By Remark 8.8, we get $c=c_{10}=2^{4}\cdot 7$ and $\beta /2=b_{10}=2^{4}$. Thus $\beta =2^{5}$.
Proof Proof of Theorem 9.5 By Theorem 8.1 (1), we have
By [Reference BorcherdsBor98, Theorem 13.3], we get
Set $\varphi _{\Lambda }:=\Psi _{\Lambda }(\cdot ,2^{g-1}F_{\Lambda } +f_{\Lambda })^{\ell }\otimes \Upsilon _{g}^{\ell }/\Phi _{M}^{(2^{g-1}+1)(2^{g}-1)}$. Since ${\rm wt}(\Phi _{M})=(-4\ell ,4\ell )$ and ${\rm div}(\Phi _{M})=\ell {\mathcal {D}}_{\Lambda }$, we deduce from (9.17), (9.18) that ${\rm wt}(\varphi _{\Lambda })=(0,0)$ and
where we used Lemma 9.9 to get the second equality. By Lemma 9.6 and Theorem 8.6 for $[M\perp d]$, $d\in \Delta _{\Lambda }$, Lemma 4.3 applies to a general curve $\gamma \colon \Delta \to {\mathcal {M}}_{\Lambda }$ intersecting $\bar {\mathcal {D}}_{\Lambda }^{0}$ transversally. Since $\alpha \geq 2(2^{18}-1)$ by Lemma 4.3, we get ${\rm div}(\varphi _{\Lambda })\geq 0$. As before, this implies that $\varphi _{\Lambda }$ is a constant.
9.4 The structure of $\Phi _{{\Bbb U}(2)}$
In this subsection we set
Then $g=9$ and $J_{M}^{*}\chi _{9}$ vanishes identically. Let $\{{\frak e},{\frak f}\}$ be a basis of $M={\Bbb U}(2)$ with ${\frak e}^{2}={\frak f}^{2}=0$ and $\langle {\frak e},{\frak f}\rangle =2$. Hence $({\frak e}+{\frak f})/2\in A_{\Lambda }$ is the unique element with non-zero norm. Let $\textbf {e}_{00},\textbf {e}_{01},\textbf {e}_{10},\textbf {e}_{11}$ be the standard basis of $\textbf {C}[A_{\Lambda }]=\textbf {C}[A_{{\Bbb U}(2)}]$, where $\textbf {e}_{\alpha \beta }$ corresponds to $(\alpha \,{\frak e}+\beta \,{\frak f})/2\in A_{\Lambda }$. Applying the construction of [Reference BorcherdsBor00, Proof of Lemma 11.1] and [Reference ScheithauerSch06, Theorem 6.2] to the modular form $\eta (\tau )^{-8}\eta (2\tau )^{-8}$, we define
Then $f_{\Lambda }(\tau )$ is an $O(A_{\Lambda })$-invariant modular form of weight $-8$ and of type $\rho _{\Lambda }$ with principal part
Since $O(q_{\Lambda })$ preserves $({\frak e}+{\frak f})/2$, the Heegner divisor of $\Omega _{\Lambda }$,
is $O^{+}(\Lambda )$-invariant. By [Reference BorcherdsBor98, Theorem 13.3], the Borcherds lift $\Psi _{\Lambda }(\cdot ,f_{\Lambda })$ is an automorphic form on $\Omega _{\Lambda }$ for $O^{+}(\Lambda )$ such that
In this subsection we prove the following theorem.
Theorem 9.10 There exists a constant $C_{M,\ell }>0$ such that
In particular, there is a constant $C_{M}>0$ depending only on $M={\Bbb U}(2)$ such that
Define the reduced divisor ${\mathcal {H}}_{1}$ on $\Omega _{\Lambda }$ as
The divisor $\bar {\mathcal {H}}_{1}\subset {\mathcal {M}}_{\Lambda }$ in § 7.2 is obtained as the quotient $\bar {\mathcal {H}}_{1}={\mathcal {H}}_{1}/O(\Lambda )$.
Lemma 9.11 $J_{M}^{*}\Upsilon _{9}$ is nowhere vanishing on $\Omega _{\Lambda }^{0}\setminus {\mathcal {H}}_{1}$.
Proof. By using Proposition 7.2 instead of Proposition 7.1, the proof is parallel to that of Lemma 9.6.
By Lemma 9.11, there exist $\alpha ,\beta \in \textbf {Z}_{>0}$ such that
Lemma 9.12 One has the inequality $\beta \geq 2^{4}$.
Proof. Let $\lambda \in \Lambda$ be an arbitrary vector such that $\lambda ^{2}=-4$ and ${\rm div}(\lambda )=1$. We set
Then $H_{\lambda }^{0}$ is a non-empty Zariski open subset of $H_{\lambda }$. Let $[\eta ]\in H_{\lambda }^{0}$ be an arbitrary point. Let $U\cong \Delta ^{18}$ be a small neighborhood of $[\eta ]$ in $\Omega _{\Lambda }$ such that $U\cap ({\mathcal {H}}_{1}\cup {\mathcal {D}}_{\Lambda })=U\cap H_{\lambda }^{0}\cong \Delta ^{17}$. Since $U$ is small enough, there is a marked family of $2$-elementary $K3$ surfaces $(p\colon ({\mathcal {X}},\iota )\to U,\alpha )$ of type $M={\Bbb U}(2)$, whose period map is the inclusion $U\hookrightarrow \Omega _{\Lambda }$. Set ${\mathcal {C}}:={\mathcal {X}}^{\iota }$. Then $p\colon {\mathcal {C}}\to U$ is a family of smooth curves of genus $9$. Set $C_{t}:=p^{-1}(t)\cap {\mathcal {C}}$ for $t\in U$. The period map $J_{M}|_{U}$ is a holomorphic map from $U$ to ${\mathcal {A}}_{9}$ such that $J_{M}(t)=\Omega (C_{t})$. Since $U$ is contractible, the local system $R^{1}(p|_{\mathcal {C}})_{*}\textbf {Z}$ is trivial and admits a symplectic basis. Hence $J_{M}\colon U\to {\mathcal {A}}_{9}$ lifts to a holomorphic map $\tilde {J}_{M}\colon U\to {\frak S}_{9}$ such that $J_{M}=\Pi \circ \tilde {J}_{M}$, where $\Pi \colon {\frak S}_{9}\to {\mathcal {A}}_{9}$ is the projection.
Since $\tilde {J}_{M}$ takes its values in ${\frak S}_{9}$, the value of the theta constant $\theta _{a,b}(\tilde {J}_{M}(t))$ makes sense for all $t\in U$ and for every even pair $(a,b)$, $a,b\in \{0,1/2\}^{9}$. Moreover, since the family $p\colon {\mathcal {C}}\to U$ admits a level $4l$ structure for any $l\in \textbf {Z}_{>0}$, the square root $\tilde {J}_{M}^{*}\sqrt {\theta _{a,b}}$ is a well-defined holomorphic section of a holomorphic line bundle on $U$ for every even pair $(a,b)$ by [Reference TsuyumineTsu91, Theorem 1]. Since any holomorphic line bundle on $U$ is trivial, we may regard $\tilde {J}_{M}^{*}\sqrt {\theta _{a,b}}\in {\mathcal {O}}(U)$.
We deduce from Proposition 7.2 the existence of a unique even pair $(a_{0},b_{0})$ and at least one even pair $(a_{1},b_{1})$ with the following properties:
(1) $\tilde {J}_{M}^{*}\theta _{a_{0},b_{0}}$ vanishes identically on $U$;
(2) set-theoretically, $U\cap {\rm div}(\tilde {J}_{M}^{*}\theta _{a_{1},b_{1}})=U\cap H_{\lambda }$.
By (2) and the fact that $\tilde {J}_{M}^{*}\sqrt {\theta _{a_{1},b_{1}}}\in {\mathcal {O}}(U)$, there exists $c\in \textbf {Z}_{>0}$ such that on $U$,
By (1) and the definition of $\Upsilon _{9}$, we have
Setting $E:={\rm div}(\prod _{(a,b)\not =(a_{0},b_{0}),(a_{1},b_{1})}\tilde {J}_{M}^{*}\theta _{a,b}^{8})$, we get the following equality of divisors on $U$:
Thus we get the desired inequality $\beta \geq 16c\geq 2^{4}$.
Proof Proof of Theorem 9.10 By Theorem 8.1 (2), we get
Since ${\rm wt}(J_{M}^{*}\Upsilon _{9})=(0,2^{2}(2^{8}+1)(2^{9}-1))$ and since ${\rm div}(J_{M}^{*}\Upsilon _{9})=\alpha {\mathcal {D}}_{\Lambda }+\beta {\mathcal {H}}_{1}$ by (9.22), we get
Since ${\rm wt}(\Phi _{M})=(-4\ell ,4\ell )$ and ${\rm div}(\Phi _{M})=\ell {\mathcal {D}}_{\Lambda }$,
is a meromorphic function on ${\mathcal {M}}_{\Lambda }$ with divisor
By Lemma 9.11 and Theorem 8.6 for $[M\perp d]$, $d\in \Delta _{\Lambda }$, Lemma 4.3 applies to a general curve $\gamma \colon \Delta \to {\mathcal {M}}_{\Lambda }$ intersecting $\bar {\mathcal {D}}_{\Lambda }^{0}$ transversally. Since $\alpha \geq 2(2^{16}-1)$ and $\beta \geq 2^{4}$ by Lemmas 4.3 and 9.12, we get ${\rm div}(\varphi _{\Lambda })\geq 0$. Thus $\varphi _{\Lambda }$ is a constant.
9.5 The divisors of $J_{M}^{*}\chi _{g}^{8}$ and $J_{M}^{*}\Upsilon _{g}$
We now summarize the formulas for the divisors of $J_{M}^{*}\chi _{g}^{8}$ and $J_{M}^{*}\Upsilon _{g}$ obtained so far. We also give a geometric interpretation of this result in terms of log Del Pezzo surfaces.
Theorem 9.13 Let $M$ be a non-exceptional primitive $2$-elementary Lorentzian sublattice of ${\Bbb L}_{K3}$ with $\Lambda =M^{\perp }$ and invariants $(r,l,\delta )$. Then the following statements hold.
(1) If $g=0$ (i.e. $r+l=22$) then ${\rm div}(J_{M}^{*}\chi _{g}^{8})=0$.
(2) If $r\geq 2$, $\delta =1$ and $1\leq g\leq 9$, then ${\rm div}(J_{M}^{*}\chi _{g}^{8})=2^{2g-1}{\mathcal {D}}_{\Lambda }^{-}+2^{k+4}{\mathcal {H}}_{\Lambda }$.
(3) If $(r,\delta )=(1,1)$, then ${\rm div}(J_{M}^{*}\chi _{g}^{8})=2^{19}{\mathcal {D}}_{\Lambda }^{-}+2^{4}\cdot 7 {\mathcal {D}}_{\Lambda }^{+}+2^{4}{\mathcal {H}}_{\Lambda }$.
(4) If $\delta =0$ and $r\not =2,10$, then ${\rm div}(J_{M}^{*}\chi _{g}^{8})=2^{g-1}(2^{g}+1) {\mathcal {D}}_{\Lambda }$.
(5) If $(r,\delta )=(2,0)$ or $(10,0)$, then $J_{M}^{*}\chi _{g}^{8}$ vanishes identically on $\Omega _{\Lambda }$.
(6) If $(r,\delta )=(10,0)$, then ${\rm div}(J_{M}^{*}\Upsilon _{g})=2(2^{2(g-1)}-1) {\mathcal {D}}_{\Lambda }$.
(7) If $(r,l,\delta )=(2,0,0)$, then ${\rm div}(J_{M}^{*}\Upsilon _{10})=2(2^{18}-1) {\mathcal {D}}_{\Lambda }+2^{5} {\mathcal {H}}_{\Lambda }$.
(8) If $(r,l,\delta )=(2,2,0)$, then ${\rm div}(J_{M}^{*}\Upsilon _{9})=2(2^{16}-1) {\mathcal {D}}_{\Lambda }+2^{4}\,\{{\mathcal {H}}_{\Lambda }-{\mathcal {H}}_{\Lambda }(-1,\textbf {e}_{11})\}$.
Proof. Assertion (1) is obvious since $\chi _{g}=1$ for $g=0$. For $g=1,2$ (respectively, $3\leq g\leq 9$), we get (2) by [Reference YoshikawaYos13, Proposition 4.2 (2), (3)] (respectively, Lemma 8.3, Proposition 8.4, (8.17)). We get (3) by Proposition 8.4, (8.17) and (8.20). When $r > 10$ (respectively, $2 < r < 10$), we get (4) by [Reference YoshikawaYos13, (9.3)] and the equality $a=E=0$ in [Reference YoshikawaYos13, Proof of Theorem 9.1] (respectively, Lemma 9.1). We get (5) by [Reference YoshikawaYos13, Proposition 9.3] when $r=10$. When $r=2$, $\delta =0$, there are two possible cases: $g=10$ and $g=9$. For $g=10$, (5) was proved in Remark 8.8. For $g=9$, (5) was proved in § 7.2. This proves (5). We get (6) by (9.3) and the equality $a=2(2^{2(g-1)}-1)$ in the proof of Theorem 9.4. We get (7) by (9.6), since $\beta =2^{5}$ and $\alpha =2(2^{18}-1)$ in the proof of Theorem 9.5. We get (8) by (9.22), since the equalities $\alpha =2(2^{16}-1)$ and $\beta =2^{4}$ follow from the proof of Theorem 9.10. This completes the proof.
Corollary 9.14 For $(r,l,\delta )=(10,2,0),(10,4,0)$, $J_{M}(\Omega _{\Lambda }^{0})$ is disjoint from the hyperelliptic locus. For $(r,l,\delta )=(10,0,0)$, every member of $J_{M}(\Omega _{\Lambda }^{0})$ has exactly one effective even theta characteristic.
Proof. Set-theoretically, ${\frak H}_{{\rm hyp},4}$ (respectively, ${\frak H}_{{\rm hyp},5}$) is given by ${\rm div}(\chi _{4})\cap {\rm div}(\Upsilon _{4})$ (respectively, ${\rm div}(\chi _{5})\cap {\rm div}(\Upsilon _{5})\cap {\rm div}(F_{5})$) on ${\frak M}_{4}$ (respectively, ${\frak M}_{5}$). Since $J_{M}^{*}\Upsilon _{4}$ (respectively, $J_{M}^{*}\Upsilon _{5}$) is nowhere vanishing on $\Omega _{\Lambda }^{0}$ by Theorem 9.13 (6) for $(r,l,\delta )=(10,4,0)$ (respectively, $(10,2,0)$), we get the first assertion. When $(r,l,\delta )=(10,0,0)$, $J_{M}^{*}\chi _{6}$ vanishes identically on $\Omega _{\Lambda }^{0}$ and $J_{M}^{*}\Upsilon _{6}$ is nowhere vanishing on $\Omega _{\Lambda }^{0}$ by Theorem 9.13 (5), (6). This, together with Lemma 4.1, implies the second assertion.
Let us give a geometric interpretation of Theorem 9.13, where we use the notion of log Del Pezzo surfaces of index at most $2$. We refer to [Reference Alexeev and NikulinAN06] for this notion. Let $S$ be a log Del Pezzo surface of index at most $2$. By [Reference Alexeev and NikulinAN06, Theorem 1.5], the bi-anticanonical system of $S$ contains a smooth member. For any smooth member $C\in |-2K_{S}|$, one can canonically associate a $2$-elementary $K3$ surface $(X_{(S,C)},\iota _{(S,C)})$ whose quotient $X_{(S,C)}/\iota _{(S,C)}$ is the right resolution of $S$ (see [Reference Alexeev and NikulinAN06, § 2.1]). We define the invariant $\delta (S)\in \{0,1\}$ as that of the $2$-elementary lattice $H^{2}(X_{(S,C)},\textbf {Z})_{+}$. Then $\delta (S)$ is independent of the choice of a smooth member $C\in |-2K_{S}|$.
Corollary 9.15 Let $S$ be a log Del Pezzo surface of index at most $2$ and let $C\in |-2K_{S}|$ be a smooth member. If $S\not \cong \mathbf{F}_{0},\mathbf{P}(1,1,2)$, then the following statements hold.
(1) When $(\rho (S),\delta (S))\not =(2,0),(10,0)$, $C$ has an effective even theta characteristic if and only if the period of $(X_{(S,C)},\iota _{(S,C)})$ lies in the characteristic Heegner divisor.
(2) When $(\rho (S),\delta (S))=(2,0)$ or $(10,0)$, $C$ always has an effective even theta characteristic. Moreover, $C$ has at least two effective even theta characteristics if and only if the period of $(X_{(S,C)},\iota _{(S,C)})$ lies in the characteristic Heegner divisor.
Proof. Since the period of $C$ is exactly the image of $(X_{(S,C)},\iota _{(S,C)})$ by the Torelli map, the result follows from Theorem 9.13.
9.6 The quasi-affinity of ${\mathcal {M}}_{\Lambda }^{0}$
As an application of the results in this and the previous section, we obtain the quasi-affinity of ${\mathcal {M}}_{\Lambda }^{0}$ for a wide range of $\Lambda$ as follows.
Theorem 9.16 If $\Lambda$ is a primitive $2$-elementary sublattice of ${\Bbb L}_{K3}$ with $r(\Lambda )<16$ and signature $(2,r(\Lambda )-2)$, then ${\mathcal {M}}_{\Lambda }^{0}$ is quasi-affine.
Proof. By [Reference YoshikawaYos13, Proposition 2.2], $\bar {J}_{M}^{0}$ extends to a meromorphic map from ${\mathcal {M}}_{\Lambda }^{*}$ to ${\mathcal {A}}_{g}^{*}$. We regard ${\mathcal {M}}_{\Lambda }^{0}$ as a Zariski open subset of a subvariety of ${\mathcal {M}}_{\Lambda }^{*}\times {\mathcal {A}}_{g}^{*}$ via the embedding ${\rm id}_{{\mathcal {M}}_{\Lambda }^{0}}\times \bar {J}_{M}^{0}$. It suffices to prove the existence of a meromorphic section of an ample line bundle on ${\mathcal {M}}_{\Lambda }^{*}\times {\mathcal {A}}_{g}^{*}$, which is nowhere vanishing on ${\mathcal {M}}_{\Lambda }^{0}$ [Reference GrothenidieckGro61, Proposition 5.1.2]. Let $\lambda _{\Lambda }$ be the Hodge bundle on ${\mathcal {M}}_{\Lambda }^{*}$. By Baily and Borel, the line bundle $\lambda _{\Lambda }^{\otimes a}\boxtimes {\mathcal {F}}_{g}^{\otimes b}$ on ${\mathcal {M}}_{\Lambda }^{*}\times {\mathcal {A}}_{g}^{*}$ is ample if $a>0$ and $b>0$. Under the assumption $r>6$, it follows from Theorems 8.6, 9.2 and 9.4 that $\Phi _{M}^{\nu }$ is a meromorphic section of $\lambda _{\Lambda }^{\otimes a}\boxtimes {\mathcal {F}}_{g}^{\otimes b}$ for some $a,b,\nu \in {\mathbf {Z}}_{>0}$. Since $\Phi _{M}^{\nu }$ is nowhere vanishing on ${\mathcal {M}}_{\Lambda }^{0}$ by Theorem 5.1, $\Phi _{M}^{\nu }$ is a desired section.
By [Reference Borcherds, Katzarkov, Pantev and Shepherd-BarronBKPS98], it is known that ${\mathcal {M}}_{\Lambda }^{0}$ is quasi-affine when $\Lambda ={\Bbb U}^{\oplus 2}\oplus {\Bbb E}_{8}^{\oplus 2}\oplus {\Bbb A}_{1}$, ${\Bbb U}^{\oplus 2}\oplus {\Bbb E}_{8}^{\oplus 2}$. This, together with Theorem 9.16, implies that ${\mathcal {M}}_{\Lambda }^{0}$ is quasi-affine, possibly except for $12$ isometry classes of primitive $2$-elementary sublattices of ${\Bbb L}_{K3}$.
10. Spin-$1/2$ bosonization formula and a factorization of $\tau _{M}$
In this section we introduce a twisted version $\tau _{M}^{{\rm spin}}$ of $\tau _{M}$ and give its explicit formula purely in terms of Borcherds products. The relations (10.5) and (10.6) below provide a factorization of $\tau _{M}$ at the level of holomorphic torsion invariants.
10.1 Spin-$1/2$ bosonization formula
Let $C$ be a smooth projective curve of genus $g$ and let $\Sigma$ be a theta characteristic. The pair $(C,\Sigma )$ is called a spin curve. A theta characteristic $\Sigma$ is ineffective if $h^{0}(\Sigma )=0$. Let $\omega$ be a Kähler form on $C$. Then $\Sigma$ is equipped with the Hermitian metric induced by $\omega$. Let $\tau (C,\Sigma ;\omega )$ be the analytic torsion of $\Sigma$ with respect to $\omega$. Recall that ${\rm vol}(C,\omega )=\int _{C}\omega /2\pi$. We set
By the anomaly formula for Quillen metrics [Reference Bismut, Gillet and SouléBGS88], if $\Sigma$ is ineffective, $\tilde {\tau }_{g}(C,\Sigma )$ is independent of the choice of a Kähler form $\omega$ on $C$. Thus we get an invariant $\tilde {\tau }_{g}$ of ineffective spin curves of genus $g$. In this subsection we recall the spin-$1/2$ bosonization formula [Reference Alvarez-Gaumé, Moore and VafaAMV86, Reference Bost and NelsonBN86, Reference FayFay92], which gives an explicit formula for $\tilde {\tau }_{g}$ viewed as a function on the moduli space of ineffective spin curves of genus $g$ with level $2$ structure.
Let $V$ be a fixed symplectic vector space of rank $2g$ over $\mathbf{F}_{2}$ equipped with a fixed symplectic basis $\{{\frak e}_{1},\ldots ,{\frak e}_{g},{\frak f}_{1},\ldots ,{\frak f}_{g}\}$. A level $2$ structure on $C\in {\frak M}_{g}$ is defined as an isomorphism of symplectic vector spaces $\alpha \colon V\cong H_{1}(C,\mathbf{F}_{2})$, where $H_{1}(C,\mathbf{F}_{2})$ is equipped with the intersection pairing. Let ${\rm Alb}(C)[2]$ (respectively, ${\rm Pic}^{0}(C)[2]$) be the $2$-division points of the Albanese variety ${\rm Alb}(C)$ (respectively, Picard variety ${\rm Pic}^{0}(C)$). By the canonical isomorphism $H_{1}(C,\mathbf{F}_{2})\cong \frac {1}{2}H_{1}(C,\textbf {Z})/H_{1}(C,\textbf {Z})\cong {\rm Alb}(C)[2]$ and the Abel–Jacobi isomorphism ${\rm Pic}^{0}(C)\cong {\rm Alb}(C)$, a level $2$ structure on $C$ is identified with a symplectic basis of ${\rm Pic}^{0}(C)[2]$ with respect to the Weil pairing.
Let ${\frak M}_{g}(2)$ be the moduli space of projective curves of genus $g$ with level $2$ structure and let $p\colon {\frak M}_{g}(2)\to {\frak M}_{g}$ be the natural projection. Let ${\mathcal {S}}_{g}^{+}$ be the moduli space of even spin curves of genus $g$ and let $\pi \colon {\mathcal {S}}_{g}^{+}\to {\frak M}_{g}$ be the natural projection. We define ${\mathcal {S}}_{g}^{+}(2)$ as the fiber product ${\mathcal {S}}_{g}^{+}\times _{{\frak M}_{g}}{\frak M}_{g}(2)$. The projection from ${\mathcal {S}}_{g}^{+}(2)$ to ${\frak M}_{g}(2)$ (respectively, ${\mathcal {S}}_{g}^{+}$) is denoted again by $p$ (respectively, $\pi$). The covering $p\colon {\mathcal {S}}_{g}^{+}(2)\to {\frak M}_{g}(2)$ of degree $2^{g-1}(2^{g}+1)$ is trivial as follows. On $(C,\alpha )\in {\frak M}_{g}(2)$, there is a distinguished even theta characteristic $\kappa \in {\rm Pic}^{g-1}(C)$ called Riemann's constant (e.g. [Reference FayFay92, p. 6 and Lemma 1.5]). For every even pair $(a,b)$, $a,b\in \{0,1/2\}^{g}$, we define a section $\sigma _{a,b}\colon {\frak M}_{g}(2)\to {\mathcal {S}}_{g}^{+}(2)$ by $\sigma _{a,b}(C,\alpha ):=(C,\kappa \otimes \chi _{a,b},\alpha )$, where $\chi _{a,b}\in {\rm Pic}^{0}(C)[2]$ is the point corresponding to $\sum _{i}2a_{i}{\frak e}_{i}+\sum _{j}2b_{j}{\frak f}_{j}\in \mathbf{F}_{2}^{2g}$ via the isomorphism $\mathbf{F}_{2}^{2g}\cong {\rm Pic}^{0}(C)[2]$ induced by $\alpha$. In this way, we get a decomposition ${\mathcal {S}}_{g}^{+}(2)=\amalg _{(a,b)\,{\rm even}}\sigma _{a,b}({\frak M}_{g}(2))$. We set ${\mathcal {S}}_{g}^{+,0}(2):=\amalg _{(a,b)\,{\rm even}}\sigma _{a,b}({\frak M}_{g}(2)\setminus {\rm div}(\theta _{a,b}))$ and ${\mathcal {S}}_{g}^{+,0}:=\pi ({\mathcal {S}}_{g}^{+,0}(2))=\bigcup _{(a,b)\,{\rm even}} \pi (\sigma _{a,b}({\frak M}_{g}(2)\setminus {\rm div}(\theta _{a,b})))$. Since $h^{0}(\kappa \otimes \chi _{a,b})=0$ if and only if $\theta _{a,b}(\Omega (C))\not =0$ for $(C,\alpha )\in {\frak M}_{g}(2)$, $\tilde {\tau }_{g}$ is a function on ${\mathcal {S}}_{g}^{+,0}$.
On the other hand, for every even pair $(a,b)$, the theta constant $\theta _{a,b}$ is a section of a certain line bundle on ${\frak M}_{g}(2)$ and its Petersson norm $\|\theta _{a,b}\|$ is a function on ${\frak M}_{g}(2)$. For $g=0$, we define $\|\theta _{a,b}\|:=1$. Let $\zeta _\textbf {Q}(s)$ be the Riemann zeta function. By the spin-$1/2$ bosonization formula ([Reference Alvarez-Gaumé, Moore and VafaAMV86, Reference Bost and NelsonBN86], [Reference FayFay92, Theorem 4.9 (i), p. 94 (4.58), p. 97 (5.7)]), the following equality of functions on ${\frak M}_{g}(2)\setminus {\rm div}(\theta _{a,b})$ holds:
where ${\frak c}_{g}$ is evaluated by the arithmetic Riemann–Roch theorem [Reference Gillet and SouléGS92] for $g=0$, by Kronecker's limit formula and Ray and Singer's formula [Reference Ray and SingerRS73, Theorem 4.1] for $g=1$, and by Wentworth's formula [Reference WentworthWen12, (1.1) and Corollary 1.1] for $g\geq 2$. In other words,
for all $(C,\alpha )\in {\frak M}_{g}(2)\setminus {\rm div}(\theta _{a,b})$ and even pairs $(a,b)$. Notice that the Laplacians (respectively, volume) in [Reference WentworthWen12] differ (respectively, differs) from ours by the scaling factor $2$ (respectively, $2\pi$). Hence Wentworth's formula [Reference WentworthWen12, (1.1) and Corollary 1.1] reads
where $\zeta _{{\mathcal {O}}_{C}}(s)$ (respectively, $\zeta _{\Sigma }(s)$) is the spectral zeta function of the Laplacian $(\bar {\partial }+\bar {\partial }^{*})^{2}$ acting on the smooth sections of ${\mathcal {O}}_{C}$ (respectively, $\Sigma$) and ${\rm Area}(C,\omega ):=\int _{C}\omega$. Since
by [Reference FayFay92, p. 37, line 16] and since ${\rm Vol}(C,\omega )={\rm Area}(C,\omega )/2\pi$, we get the value ${\frak c}_{g}$ in (10.1).
Let us extend the definition of $\tilde {\tau }_{g}$ to disconnected curves as follows. A line bundle on a disconnected curve is a theta characteristic if it is a componentwise theta characteristic. Similarly, a theta characteristic on a disconnected curve is ineffective if it is componentwise ineffective. In what follows, for a disjoint union of smooth projective curves $C=\amalg _{i\in I}C_{i}$ with $g(C_{i}):=g_{i}$ and an ineffective theta characteristic $\Sigma =\{\Sigma _{i}\}_{i\in I}$ on $C$, we define
where $g:=\sum _{i\in I}g_{i}$ is the total genus of $C$.
10.2 A factorization of $\tau _{M}$
We introduce the following twisted version of $\tau _{M}$.
Definition 10.1 Let $(X,\iota )$ be a $2$-elementary $K3$ surface of type $M$ and let $\gamma$ be an $\iota$-invariant Kähler form on $X$. If $M\not \cong {\Bbb U}(2)\oplus {\Bbb E}_{8}(2)$, define
where $\Sigma$ runs over all ineffective theta characteristics on $X^{\iota }$. If $M\cong {\Bbb U}(2)\oplus {\Bbb E}_{8}(2)$ and hence $X^{\iota }=\emptyset$, define
Recall that the vector-valued modular form $f_{\Lambda }$ of type $\rho _{\Lambda }$ was defined by (9.5) and (9.19) when $r=22-r(\Lambda )=2$. We extend its definition to the case $r\not =2$ by setting
Then Theorem 0.1 is interpreted as the modularity of $\tau _{M}^{{\rm spin}}$ as follows.
Theorem 10.2 There exists a constant $C'_{M}>0$ depending only on $M$ such that the following equality of functions on ${\mathcal {M}}_{\Lambda }^{0}\setminus {\mathcal {H}}_{\Lambda }$ holds:
Proof. For a $2$-elementary $K3$ surface $(X,\iota )$ of type $M$ with $(r,l,\delta )\not =(10,10,0),(10,8,0)$, recall that $X^{\iota }$ consists of a curve of genus $g=g(M)$ and $k=k(M)$ smooth rational curves (cf. § 3.2).
Case 1. If $(r,\delta )\not =(2,0),(10,0)$ and $X^{\iota }$ has no effective even theta characteristics, we get
with
Here the first equality of (10.5) follows from Definition 10.1 and the second follows from (10.2). Comparing Theorem 0.1 and (10.5), we get (10.4) with $(C'_{M})^{2}:=c_{M}^{2}/C_{M}$ in this case.
Case 2. If $(r,\delta )=(2,0),(10,0)$ with $(r,l,\delta )\not =(10,10,0),(10,8,0)$ and if $X^{\iota }$ has a unique effective even theta characteristic corresponding to a theta constant $\theta _{a_{0},b_{0}}(\Omega (X^{\iota }))$ with respect to a suitable level $2$ structure, we get in the same way
with
By Theorem 0.1 and (10.6), we get (10.4) with $(C'_{M})^{2}:=c_{M}^{2}/C_{M}$ in this case.
Case 3. If $(r,l,\delta )=(10,10,0)$, then $X^{\iota }=\emptyset$. Since we defined $g=1$ in this case, we get (10.4) with $(C'_{M})^{2}:=C_{M}^{-1}$ by Definition 10.1 and Theorem 0.1. If $(r,l,\delta )=(10,8,0)$, then $X^{\iota }$ consists of two disjoint elliptic curves. In the same way as in (10.6), we get (10.4) with $(C'_{M})^{2}:=c_{M}^{2}/C_{M}$, $c_{M}=({\frak c}_{1}^{-2})^{9}=(4\pi )^{-18}$ in this case. This completes the proof.
Remark 10.3 Assume ${\mathcal {H}}_{\Lambda }\not =\emptyset$. As the period of a $2$-elementary $K3$ surface of type $M$ approaches a point of ${\mathcal {H}}_{\Lambda }$, one of the ineffective even theta characteristics on its fixed curve becomes effective in the limit and the value $\tau _{M}^{{\rm spin}}$ jumps there. Because of this jumping, $\tau _{M}^{{\rm spin}}$ is a discontinuous function on ${\mathcal {M}}_{\Lambda }^{0}$. Since $\|\Psi _{\Lambda }(\cdot ,2^{g-1}F_{\Lambda }+f_{\Lambda })\|$ is also discontinuous along ${\mathcal {H}}_{\Lambda }$ by [Reference SchoferSch09, Theorem 1.1 (i)], it is an interesting problem to compare $\tau _{M}^{{\rm spin}}$ and $\|\Psi _{\Lambda }(\cdot ,2^{g-1}F_{\Lambda }+f_{\Lambda })\|$ on the locus ${\mathcal {H}}_{\Lambda }$.
Remark 10.4 As suggested by a referee, it is possible to express the invariants $\tau _{M}(X,\iota )$ and $\tau _{M}^{{\rm spin}}(X,\iota )$ in terms of X. Ma's orbifold analytic torsions [Reference MaMa05] for $X/\iota$ and the twisted sector $\Sigma (X/\iota )=X^{\iota }$. Here $X/\iota$ is a smooth surface with non-trivial orbifold structure. In this direction, it is possible to construct a holomorphic torsion invariant for log-Enriques surfaces by making use of orbifold analytic torsion and give its explicit formula as an automorphic function on the moduli space (in a forthcoming paper by Dai and Yoshikawa).
10.3 A uniqueness of elliptic modular form corresponding to $\tau _{M}^{{\rm spin}}$
Set
Since $\tau _{M}^{{\rm spin}}=C'_{M}\|\Phi _{M}^{{\rm spin}}\|^{-1/2}$, $\Phi _{M}^{{\rm spin}}$ can be identified with $\tau _{M}^{{\rm spin}}$. In this subsection we study the uniqueness of elliptic modular form whose Borcherds lift is $\Phi _{M}^{{\rm spin}}$.
For a modular form $\varphi (\tau )$ of type $\rho _{\Lambda }$ with weight $1-b^{-}(\Lambda )/2$, we write
for its Fourier expansion. The principal part of $\varphi$ is the Laurent polynomial defined as
Notice that we used the notation ${\mathcal {P}}_{\leq 0}[\varphi ]=\sum _{\gamma \in A_{\Lambda }}\textbf {e}_{\gamma }\sum _{n\in \gamma ^{2}/2+\textbf {Z}, n\leq 0}c_{\gamma }(n;\varphi )q^{n}$ in the previous sections. Obviously, ${\mathcal {P}}_{\leq 0}[\varphi ]-{\mathcal {P}}[\varphi ]\in \textbf {C}[A_{\Lambda }]$ is the constant term of $\varphi$.
Similarly, for a Heegner divisor $H=\sum _{\gamma \in A_{\Lambda }}\sum _{n\in \gamma ^{2}/2+\textbf {Z}, n<0}a_{\gamma }(n)\,H(n,\gamma )$ on $\Omega _{\Lambda }$, where $H(n,\gamma ):=\sum _{\lambda \in (\gamma +\Lambda )/\pm 1, \lambda ^{2}=2n}H_{\lambda }$, $n\in \textbf {Q}$, $\gamma \in A_{\Lambda }$, we define
Comparing (8.1), (9.5), (9.20) with Theorem 8.1, (9.18), (9.24), we have the equality
if $\Lambda \not \cong {\Bbb U}(2)^{\oplus 2},({\Bbb A}_{1}^{+})^{\oplus 2}$ or equivalently ${\mathcal {D}}_{\Lambda }\not =0$. When $\Lambda \cong {\Bbb U}(2)^{\oplus 2}$ or $({\Bbb A}_{1}^{+})^{\oplus 2}$, we have ${\mathcal {P}}[2^{g-1}F_{\Lambda }+f_{\Lambda }]\not =0$ but ${\frak P}[{\rm div}(\Phi _{M}^{{\rm spin}})]=0$. Thus (10.8) does not hold in these two cases. Except for them, the elliptic modular form $2^{g-1}F_{\Lambda }+f_{\Lambda }$ is characterized uniquely by the holomorphic torsion invariant $\tau _{M}^{{\rm spin}}$ as follows.
Theorem 10.5 If $\Lambda \not \cong {\Bbb U}(2)^{\oplus 2},({\Bbb A}_{1}^{+})^{\oplus 2}$, then there exists a unique $O(q_{\Lambda })$-invariant elliptic modular form $\varphi _{\Lambda }$ of type $\rho _{\Lambda }$ with weight $1-b^{-}(\Lambda )/2$ such that
In particular, the $O(q_{\Lambda })$-invariance and (10.9) characterize $2^{g-1}F_{\Lambda }+f_{\Lambda }$ uniquely.
Proof. Let $\varphi _{\Lambda }$ be an $O(q_{\Lambda })$-invariant modular form satisfying (10.9). Set $\psi :=\varphi _{\Lambda }-(2^{g-1}F_{\Lambda }+f_{\Lambda })$. This is a modular form of type $\rho _{\Lambda }$ and weight $1-b^{-}(\Lambda )/2$ which is holomorphic at the cusp, is $O(q_{\Lambda })$-invariant, and satisfies $c_{0}(0; \psi )=0$. We must prove $\psi =0$. When $b^{-}(\Lambda )>2$, $\psi$ has negative weight and hence $\psi =0$.
Let $b^{-}(\Lambda )=2$. Then $\psi$ has weight $0$, so it must be a constant vector of $\textbf {C}[A_{\Lambda }]$. From [Reference NikulinNik80, Lemma 3.9.1] and [Reference SkoruppaSko08, Theorem 1], we deduce that ${\rm Mp}_{2}(\textbf {Z})\times O(q_{\Lambda })$-invariant vectors in ${\mathbf {C}}[A_{\Lambda }]$ are scalar multiples of $\nu _{\Lambda }\textbf {e}_{0}+\textbf {v}_{\Lambda }+\nu _{\Lambda }\delta (\Lambda )\textbf {e}_{\textbf {1}_{\Lambda }}$ for some $\nu _{\Lambda }\in \textbf {Z}_{>0}$, where ${\mathbf {v}}_{\Lambda } =\sum _{\gamma \ne 0,\textbf {1}_{\Lambda },\gamma ^{2}\equiv 0}\textbf {e}_{\gamma }$. Since $c_{0}(0; \psi )=0$, we have $\psi =0$.
Let $b^{-}(\Lambda )=1$. By [Reference SkoruppaSko08, Theorems 5 and 9] there is a canonical isomorphism between the space of modular forms of type $\rho _{\Lambda }$ and weight $1/2$ with the space of ${\rm Mp}_{2}({\mathbf {Z}})$-invariant vectors in ${\mathbf {C}}[A_{{\Bbb A}_{1}}]\otimes {\mathbf {C}}[A_{\Lambda }] \simeq {\mathbf {C}}[A_{{\Bbb A}_{1}\oplus \Lambda }]$. By [Reference SkoruppaSko08, Theorem 1], the latter is generated by the vectors $I_{U}=\sum _{\gamma \in U}\textbf {e}_{\gamma }$ where $U$ runs over self-dual isotropic subgroups of $A_{{\Bbb A}_{1}\oplus \Lambda }$. In this isomorphism the modular form corresponding to $I_{U}$ is given by $\sum _{\gamma \in U}\theta _{{\Bbb A}_{1}^{+}+\gamma _{1}}\textbf {e}_{\gamma _{2}}$ where $\gamma =(\gamma _{1}, \gamma _{2})\in A_{{\Bbb A}_{1}\oplus \Lambda }$ (see [Reference SkoruppaSko08, Theorem 8]). Now we have $\Lambda \cong {\Bbb A}_{1}^{+}\oplus {\Bbb U}$ or $({\Bbb A}_{1}^{+})^{\oplus 2}\oplus {\Bbb A}_{1}$. In the first case, $A_{{\Bbb A}_{1}\oplus \Lambda }$ contains a unique non-zero isotropic element, and the corresponding modular form $\varphi$ has $c_{0}(0; \varphi )\ne 0$. In the second case, $A_{{\Bbb A}_{1}\oplus \Lambda }$ contains exactly two isotropic subgroups of rank $2$, which can be switched by an element of $O(q_{\Lambda })$. If $\varphi _{1}$ and $\varphi _{2}$ are the corresponding modular forms, then $\psi$ must be a scalar multiple of $\varphi _{1}+\varphi _{2}$. Again we have $c_{0}(0; \varphi _{1}+\varphi _{2})\ne 0$, so $\psi =0$.
11. An equivariant analogue of Borcherds’ conjecture
In this section we study an equivariant analogue of Borcherds’ conjecture [Reference BorcherdsBor98]. Let us briefly explain this conjecture. Let $X_{K3}$ be the oriented $4$-manifold underlying a $K3$ surface. Let ${\mathcal {E}}$ be the set of Ricci-flat Riemannian metrics on $X_{K3}$ with normalized volume $1$ (cf. [Reference YauYau78]). For $\gamma \in {\mathcal {E}}$, let $\Delta _{\gamma }$ be the Laplacian of $(X_{K3},\gamma )$ acting on $C^{\infty }(X_{K3})$. Let $\det \Delta _{\gamma }$ be the regularized determinant of $\Delta _{\gamma }$. Then the assignment $\det \Delta \colon {\mathcal {E}}\ni \gamma \to \det \Delta _{\gamma }\in {\mathbf {R}}$ is a function on ${\mathcal {E}}$. In [Reference BorcherdsBor98, Example 15.2], Borcherds conjectured that $\det \Delta$ is given by the automorphic form $\Phi _{{\Bbb L}_{K3}}(\cdot ,1,E_{4}/\eta ^{24})$ on $G({\Bbb L}_{K3})$, the period space of ${\mathcal {E}}$, where $E_{4}(\tau )$ is the Eisenstein series of weight $4$. To our knowledge, this conjecture is still open. Here, instead of the original Borcherds conjecture, we study its equivariant analogue.
Let $\iota \colon X_{K3}\to X_{K3}$ be a $C^{\infty }$ involution. We define the lattices $H^{2}(X_{K3},{\mathbf {Z}})_{\pm }$ as in the preceding sections. Then $\iota$ is called hyperbolic if $H^{2}(X_{K3},{\mathbf {Z}})_{+}$ is Lorentzian. Let ${\mathcal {E}}^{\iota }$ be the set of $\iota$-invariant Ricci-flat Riemannian metrics on $X_{K3}$ with volume $1$. Since we are interested in an equivariant analogue of Borcherds’ conjecture, throughout this section, we restrict our consideration to those involutions $\iota$ satisfying
By [Reference YoshikawaYos08, Propisitions 3.4, 3.6], if $\iota$ is hyperbolic, then (11.1) is equivalent to the existence of a complex structure $I$ on $X_{K3}$ such that $\iota$ is an anti-symplectic holomorphic involution on $(X_{K3},I)$. In particular, if $\iota$ is hyperbolic with (11.1), then $X_{K3}^{\iota }$ is a disjoint union of (possibly empty) smooth compact real surfaces.
Let $\iota$ be a hyperbolic involution on $X_{K3}$ with (11.1). Its type is defined as the isometry class of $H^{2}(X_{K3},{\mathbf {Z}})_{+}$. Let $M$ be the type of $\iota$ and set $\Lambda :=M^{\perp _{{\Bbb L}_{K3}}}$. Then $M$ and $\Lambda$ are primitive $2$-elementary sublattices of ${\Bbb L}_{K3}$. To formulate an equivariant analogue of Borcherds’ conjecture, we construct two functions on ${\mathcal {E}}^{\iota }$.
Let $\gamma \in {\mathcal {E}}^{\iota }$. Let $C^{\infty }(X_{K3})_{\pm }$ be the $\pm 1$-eigenspace of the $\iota$-action on $C^{\infty }(X_{K3})$. Since $\Delta _{\gamma }$ preserves $C^{\infty }(X_{K3})_{\pm }$, we can define $\Delta _{\gamma ,\pm }:=\Delta _{\gamma }|_{C^{\infty }(X_{K3})_{\pm }}$. Let $\zeta _{\pm }(s)$ be the spectral zeta function of $\Delta _{\gamma ,\pm }$. The equivariant determinant of $\Delta _{\gamma }$ is defined as (cf. [Reference BismutBis95])
Assume $X_{K3}^{\iota }\not =\emptyset$. Let $S_{\gamma }$ be a spinor bundle on the fixed point set $(X_{K3}^{\iota }, \gamma |_{X^{\iota }_{K3}})$. Let $D_{S_{\gamma }}$ be the Dirac operator acting on $C^{\infty }(S_{\gamma })$. Let $\det D_{S_{\gamma }}^{2}$ be the regularized determinant of $D_{S_{\gamma }}^{2}$. If $\zeta _{D_{S_{\gamma }}^{2}}(s)$ denotes the spectral zeta function of $D_{S_{\gamma }}^{2}$, then
When $X_{K3}^{\iota }=\emptyset$, we define $\det D_{S_{\gamma }}^{2}:=1$. As an equivariant analogue of the function $\det \Delta$ on ${\mathcal {E}}$, we consider the following function on ${\mathcal {E}}^{\iota }$.
Definition 11.1 For $\gamma \in {\mathcal {E}}^{\iota }$, define
where $S_{\gamma }$ runs over the spinor bundles on $(X_{K3}^{\iota }, \gamma |_{X^{\iota }_{K3}})$ with $\ker D_{S_{\gamma }}=0$.
Let us construct an automorphic function on the period space of ${\mathcal {E}}^{\iota }$. By [Reference YoshikawaYos08, Proposition 3.6], for example, there exists a hyperkähler structure $(I,J,K)$ on $(X_{K3},\gamma )$ such that
By [Reference YoshikawaYos08, Lemma 3.17], the pair of conjugate points of ${\mathcal {M}}_{\Lambda }$ defined as
is independent of the choice of a triplet $(I,J,K)$ satisfying (11.2) and a marking, that is, an isometry $\alpha \colon H^{2}(X_{K3},{\mathbf {Z}})\to {\Bbb L}_{K3}$ satisfying $\alpha (H^{2}(X,{\mathbf {Z}})_{+})=M$. The pair of conjugate points $\pi _{\iota }(\gamma )\in {\mathcal {M}}_{\Lambda }$ is called the period of $\gamma \in {\mathcal {E}}^{\iota }$.
Lemma 11.2 $\pi _{\iota }^{*}\|\Psi _{\Lambda }(\cdot ,2^{g-1}F_{\Lambda }+f_{\Lambda })\|$ is a well-defined function on ${\mathcal {E}}^{\iota }$.
Proof. Write $\Lambda ={\Bbb U}(-N)\oplus L$, $N\in \{1,2\}$, where $L$ is a Lorentzian lattice. Set $C_{L}:=\{x\in L\otimes {\mathbf {R}};\,\langle x,x\rangle >0\}$. Since $L$ is Lorentzian, $C_{L}$ consists of two components $C_{L}^{\pm }$ with $C_{L}^{-}=-C_{L}^{+}$. Then $L\otimes {\mathbf {R}}+\sqrt {-1}C_{L}\subset L\otimes {\mathbf {C}}$ is isomorphic to $\Omega _{\Lambda }$ via the map
Since $\exp _{N}\circ (-1_{L})=(1_{{\Bbb U}(-N)}\oplus -1_{L})\circ \exp _{N}$ and since $-1_{L}$ exchanges the components of $C_{L}$, $1_{{\Bbb U}(-N)}\oplus -1_{L}\in O(\Lambda )$ exchanges the components of $\Omega _{\Lambda }$.
Set $\eta :=\alpha (\omega _{J}+\sqrt {-1}\omega _{K})$. Let $z\in L\otimes {\mathbf {R}}+\sqrt {-1}C_{L}$ be such that $[\eta ]=\exp _{N}(z)$. Then $\bar {\eta }=\alpha (\omega _{J}-\sqrt {-1}\omega _{K})$ and $[\bar {\eta }]=\exp _{N}(\bar {z})$. Let $\Omega _{\Lambda }^{+}$ be the component of $\Omega _{\Lambda }$ such that $[\eta ]\in \Omega _{\Lambda }^{+}$. Let $\Omega _{\Lambda }^{-}$ be the remaining component. Then $\Omega _{\Lambda }^{-}=\overline {\Omega _{\Lambda }^{+}}$ and $[\bar {\eta }]\in \Omega _{\Lambda }^{-}$. Since $(1_{{\Bbb U}(-N)}\oplus -1_{L})[\bar {\eta }] = [(1_{{\Bbb U}(-N)}\oplus -1_{L})(\exp _{N}(\bar {z}))] = [\exp _{N}(-\bar {z})] \in \Omega _{\Lambda }^{+}$, the point of $\Omega _{\Lambda }^{+}/O^{+}(\Lambda )$ corresponding to $[\bar {\eta }]$ is represented by $[\exp _{N}(-\bar {z})]$.
For simplicity, write $\Psi (\cdot )$ for $\Psi _{\Lambda }(\cdot ,2^{g-1}F_{\Lambda }+f_{\Lambda })$. Let $w$ be its weight. Then $\Psi \in {\mathcal {O}}(L\otimes {\mathbf {R}}+\sqrt {-1}C_{L}^{+})$. By the definition of Petersson norm (cf. § 4.2),
Since $\Psi$ is a Borcherds product, it is expressed as a Fourier series
with $a(\lambda )\in {\mathbf {Z}}$. Since $\overline {e^{2\pi i\langle l,z\rangle }}=e^{2\pi i\langle l,-\bar {z}\rangle }$ and $a(l)\in {\mathbf {Z}}$ for all $l\in L^{\lor }$ and hence $\overline {\Psi (z)}=\Psi (-\bar {z})$, we deduce from (11.4) and (11.5) that $\|\Psi ([\eta ])\|^{2}=\|\Psi ([\bar {\eta }])\|^{2}$.
Now we can formulate an equivariant analogue of Borcherds’ conjecture as the coincidence of the two functions $\tau _{\iota }^{{\rm spin}}$ and $\pi _{\iota }^{*}\|\Psi _{\Lambda }(\cdot ,2^{g-1}F_{\Lambda }+f_{\Lambda })\|$ on ${\mathcal {E}}^{\iota }$. By Theorem 10.2, we have an affirmative answer to this problem.
Theorem 11.3 Let $\iota$ be a hyperbolic involution on $X_{K3}$ with ${\mathcal {E}}^{\iota }\not =\emptyset$. Then the following equality of functions on $\pi _{\iota }^{-1}({\mathcal {M}}_{\Lambda }^{0}\setminus {\mathcal {H}}_{\Lambda })\subset {\mathcal {E}}^{\iota }$ holds:
where $C''_{M}$ is a constant depending only on $M$.
Proof. Let $\gamma \in {\mathcal {E}}^{\iota }$. Let $(I,J,K)$ be a hyperkähler structure on $(X_{K3},\gamma )$ with (11.2) and set $X_{I}:=(X_{K3},I)$. Then $(X_{I},\iota )$ is a $2$-elementary $K3$ surface of type $M$. By [Reference YoshikawaYos08, Lemma 4.3], we have
On the other hand, it is classical that $D_{S_{\gamma }}$ can be identified with the Dolbeault operator $\sqrt {2}(\bar {\partial }+\bar {\partial }^{*})$ acting on $A^{0,*}(X_{K3}^{\iota },\Sigma _{S_{\gamma }})$, where $\Sigma _{S_{\gamma }}$ is the theta characteristic on $X_{K3}^{\iota }$ corresponding to $S_{\gamma }$. Here $X_{K3}^{\iota }$ is endowed with the complex structure induced by $\gamma$. Since $D_{S_{\gamma }}^{2}=2(\bar {\partial }+\bar {\partial }^{*})^{2}$ by this identification, we have $\zeta _{D_{S_{\gamma }}^{2}}(s)=2^{-s+1}\zeta _{\Sigma _{S_{\gamma }}}(s)$, where $\zeta _{\Sigma _{S_{\gamma }}}(s)$ is the spectral zeta function of $(\bar {\partial }+\bar {\partial }^{*})^{2}|_{A^{0,0}(\Sigma _{S_{\gamma }})}$ as in § 10.1. Hence
By (11.6), (11.7) and the definitions of $\tau ^{{\rm spin}}_{M}$ and $\tau ^{{\rm spin}}_{\iota }$, we get
where $N(\gamma ,\iota )$ is the number of ineffective spinor bundles on $(X_{K3}^{\iota },\gamma )$.
Let $\gamma \in \pi _{\iota }^{-1}({\mathcal {M}}_{\Lambda }^{0}\setminus {\mathcal {H}}_{\Lambda })$. Since $N(\gamma ,\iota )$ is a constant function on $\pi _{\iota }^{-1}({\mathcal {M}}_{\Lambda }^{0}\setminus {\mathcal {H}}_{\Lambda })$ by Theorem 9.13 and (11.3) and since $\pi _{\iota }(\gamma )$ is given by the pair of $\bar {\pi }_{M}(X_{I},\iota )$ and its conjugate point by (11.3), the result follows from Theorem 10.2 and (11.8).
As in [Reference YoshikawaYos08] and [Reference YoshikawaYos13, § 10], we obtain, as a corollary of Theorem 11.3, an interpretation of Theorem 10.2 on the mirror side, that is, in terms of real $K3$ surfaces. Recall that a pair consisting of a $K3$ surface and an anti-holomorphic involution is called a real $K3$ surface. The set of fixed points of the involution on a real $K3$ surface is the set of real points. A holomorphic $2$-form on a real $K3$ surface is said to be defined over ${\mathbf {R}}$ if it is mapped to its complex conjugation by the involution. In view of mirror symmetry for $K3$ surfaces with involution [Reference Gross and WilsonGW97, § 2], the following corollary is a counterpart of Theorem 10.2 in mirror symmetry.
Corollary 11.4 Let $(Y,\sigma )$ be a real $K3$ surface. Let $M$ be the type of $\sigma$ and let $\alpha$ be a marking with $\alpha (H^{2}(X,{\mathbf {Z}})_{+})=M$. Let $\gamma$ be a $\sigma$-invariant Ricci-flat Kähler metric on $Y$ with volume $1$. Let $\omega _{\gamma }$ be the Kähler form of $\gamma$ and let $\eta _{\gamma }$ be a holomorphic $2$-form on $Y$ defined over ${\mathbf {R}}$ such that $\eta _{\gamma }\wedge \bar {\eta }_{\gamma }=2\omega _{\gamma }^{2}$. Then
where $\Im (\eta _{\gamma })+\sqrt {-1}\omega _{\gamma }\in H^{2}(Y,{\mathbf {R}})+\sqrt {-1}{\mathcal {K}}_{Y}$ is a point of the complexified Kähler cone of the Kähler surface $(Y,\omega _{\gamma })$ with $B$-field $\Im (\eta _{\gamma })$.
Proof. Set $I'=K$, $J'=-J$, $K'=I$, where $(I,J,K)$ is a hyperkähler structure on $(X_{K3},\gamma )$ satisfying (11.2) for $\sigma$. Then $(I',J',K')$ is a hyperkähler structure on $(X_{K3},\gamma )$ such that $\sigma _{*}I'=-I'\sigma _{*}$, $\sigma _{*}J'=-J'\sigma _{*}$, $\sigma _{*}K'=K'\sigma _{*}$. Set $Y=(X_{K3},I')$. Then $\sigma$ is an anti-holomorphic involution on $Y$. On the other hand, $\sigma$ is an anti-symplectic holomorphic involution on $(X_{K3},K')$. By [Reference Gross and WilsonGW97, p. 514], we see that $\Re (\eta _{\gamma })$ and $\Im (\eta _{\gamma })+\sqrt {-1}\omega _{\gamma }$ are a Kähler form and a holomorphic $2$-form on $(X_{K3},K')$, respectively. From this interpretation and Theorem 11.3, the result follows.
Is it possible to prove Corollary 11.4 without passing through algebraic geometry? Such a proof will provide a new understanding of Theorem 10.2.
Acknowledgements
The first-named author is partially supported by JSPS KAKENHI grants 17K14158 and 20H00112. The second-named author is partially supported by JSPS KAKENHI grants 16H03935 and 16H06335. He is grateful to Professor Jean-Michel Bismut for helpful discussions and to Professor Riccardo Salvati Manni for answering his questions. The authors are grateful to the referee for a careful reading of the paper and helpful comments.