1 Introduction

A projective simplicial toric variety \({\mathbb {P}}^d_{\Sigma }\) satisfies the Hodge Conjecture, i.e., every cohomology class in \(H^{p,p}({\mathbb {P}}^{d}_{\Sigma },{\mathbb {Q}})\) is a linear combination of algebraic cycles. On the one hand, by the Lefschetz hyperplane theorem, the Hodge conjecture holds true for every hypersurface and \(p<\frac{d-1}{2}\) and by the hard Lefschetz theorem also for \(p>\frac{d-1}{2}\). Moreover, by Theorem 1.1 in [3], when \(p=\frac{d-1}{2}\), \(d=2k+1\) and \({\mathbb {P}}_{\Sigma }^{2k+1}\) is an Oda variety with an ample class \(\beta \) such that \(k\beta -\beta _0\) is nef, where \(\beta _0\) is the anticanonical class, the Hodge conjecture with rational coefficients holds for a very general hypersurface in the linear system \(|\beta |\).

The notion of Oda varieties was introduced in [2]. Let us recall that the Cox ring of a toric variety \({\mathbb {P}}_\Sigma \) is graded over the class group \({\text {Cl}}({\mathbb {P}}_\Sigma )\), and that one has an injection \({\text {Pic}}({\mathbb {P}}_\Sigma ) \rightarrow {\text {Cl}}({\mathbb {P}}_\Sigma )\).

Definition 1.1

Let \({\mathbb {P}}_\Sigma \) be a toric variety with Cox ring S. \({\mathbb {P}}_\Sigma \) is said to be an Oda variety if the multiplication morphism \(S^{\alpha _1 }\otimes S^{\alpha _2} \rightarrow S^{\alpha _1+\alpha _2}\) is surjective whenever the classes \(\alpha _1\) and \(\alpha _2\) in \({\text {Pic}}({\mathbb {P}}_\Sigma )\) are ample and nef, respectively.

In [15] Mavlyutov proved a Lefschetz type theorem for quasi-smooth intersection subvarieties, and moreover using the “Cayley trick” he related the cohomology of a quasi-smooth subavariety \(X=X_{f_1}\cap \dots \cap X_{f_s}\subset {\mathbb {P}}^{d}_{\Sigma }\) to the cohomology of a quasi-smooth hypersurface \(Y\subset {\mathbb {P}}^{d+s-1}_{\Sigma }\). This allows us to prove a Noether–Lefschetz type theorem, namely:

Theorem 2.5. Let \({\mathbb {P}}^d_{\Sigma }\) be an Oda projective simplicial toric variety. For a very general quasi-smooth intersection subvariety X cut off by \(f_1,\dots f_s\) such that \(d+s=2(\ell +1)\) and

$$\begin{aligned} \sum _{i=1}^s \deg (f_i)- \beta _0 \end{aligned}$$

is nef, one has

$$\begin{aligned} H^{\ell +1-s,\ell +1-s}(X,{\mathbb {Q}})=i^*\left( H^{\ell +1-s,\ell +1-s}({\mathbb {P}}^d_{\Sigma },{\mathbb {Q}})\right) . \end{aligned}$$

From this one obtains the following result about the Hodge conjecture for quasi-smooth intersections.

Corollary 2.7. If \({\mathbb {P}}_{\Sigma }^d\) is an Oda projective simplicial toric variety, the Hodge Conjecture holds for a very general quasi-smooth intersection subvariety X cut off by \(f_1,\dots f_s\) such that \(d+s\) is even and \(\sum _{i=1}^s \deg (f_i)- \beta _0\) is nef.

Let T be the open subset of \(\vert \beta \vert \) corresponding to quasi-smooth hypersurfaces, and let \({\mathcal {H}}^{2k}=R^{2k}\pi _*{\mathbb {C}}\otimes _{\mathbb {C}}{\mathcal {O}}_T\) be the Hodge bundle on T; here \(\pi :{\mathcal {X}} \rightarrow T\) is the tautological family on T, and \(d=2k+1\). We restrict \({\mathcal {H}}^{2k}\) to a contractible open subset \(U\subset T\). The bundle \({\mathcal {H}}^{2k}\) has a Hodge decomposition

$$\begin{aligned} {\mathcal {H}}^{2k} = \bigoplus _{p+q=2k}{\mathcal {H}}^{p,q} \end{aligned}$$

but this is not holomorphic. On the other hand, the bundles that make up the Hodge filtration

$$\begin{aligned} F^p{\mathcal {H}}^{2k} = \bigoplus _{p=0}^{2k}{\mathcal {H}}^{2k-p,p} \end{aligned}$$

are holomorphic; to see this one can use the period map (which in particular we write for \(p=k\))

$$\begin{aligned} {\mathcal {P}}^{k,2k}:U \rightarrow {\text {Grass}}(b_k,H^{2k}(X_{u_0},{\mathbb {C}})) \end{aligned}$$

where \(b_k = \dim F^kH^{2k}(X_{u_0},{\mathbb {C}})\) for a fixed point \(u_0\in U\); this map sends \(f\in U\) to the subspace \(F^kH^{2k}(X_f,{\mathbb {C}}) \subset H^{2k}(X_f,{\mathbb {C}}) = H^{2k}(X_{u_0},{\mathbb {C}})\). This map is holomorphic (see [14] and [5, Prop. 3.4]). But, by the very definition of the period map (see also [17], Section 10.2.1 for the smooth case)

$$\begin{aligned} F^k{\mathcal {H}}^{2k} \simeq ({\mathcal {P}}^{k,2k})^*{\mathcal {U}}_k, \end{aligned}$$

where \({\mathcal {U}}_k\) is the tautological bundle on the Grassmannian \({\text {Grass}}(b_k,H^{2k}(X_{u_0},{\mathbb {C}}))\), so that the bundles \(F^k{\mathcal {H}}^{2k} \) are indeed holomorphic.

Pushing ahead the ideas developed in [5] and [4], let \(\lambda _f\) be a nonzero class in the primitive cohomology \(H^{k,k}(X_f,{\mathbb {Q}})/H^{k,k}({{\mathbb {P}}}_\Sigma ^{2k+1},{\mathbb {Q}})\), and let U be a contractible open subset of T around f, so that \({\mathcal {H}}^{2k}_{\vert U}\) is constant. Moreover, let \(\lambda \in {\mathcal {H}}^{2k}(U)\) be the section defined by \(\lambda _f\) and let \({\bar{\lambda }}\) be its image in \(({\mathcal {H}}^{2k}/F^k{\mathcal {H}}^{2k})(U)\). One has

Proposition 1.2

The local Noether–Lefschetz loci can be defined as

$$\begin{aligned} N^{k,\beta }_{\lambda ,U}:=\{G\in U \mid {\bar{\lambda }}_{G}=0\} \end{aligned}$$

where \(\beta =\deg (f)\).

The following result is Theorem 1.2 in [4].

Theorem. Let \({{\mathbb {P}}}_\Sigma ^{2k+1}\) be an Oda variety with an ample class \(\beta \) such that \(k\beta -\beta _0=n\eta \), where \(\beta _0\) is the anticanonical class, \(\eta \) is a primitive ample class, and \(n\in {\mathbb {N}}\). Let

$$\begin{aligned} m_\beta = \max \{i\in {\mathbb {N}}\,\,\vert \,i\eta \le \beta \}. \end{aligned}$$
(1)

For every positive \(\epsilon \) there is a positive \(\delta \) such that for every \(m\ge \max (\frac{1}{\delta },m_\beta )\) and \({\hat{d}}\in [1,m\delta ]\), and every nontrivial Hodge class \(\lambda \in F^k{\mathcal {H}}^{2k}(U)\) such that

$$\begin{aligned} {\text {codim}}N^{k,\beta }_{\lambda ,U} \le {\hat{d}}\frac{m_{\beta }^k}{k!}, \end{aligned}$$

for every \(f\in N_{\lambda ,U}^{k,\beta }\), there exists a k-dimensional variety \(V\subset X_f\) with \(\deg V\le (1+\epsilon ) {\hat{d}}\). Here \(\deg V\) is taken with respect to the ample divisor \(\eta \), i.e.,

$$\begin{aligned} \deg V = [V]\cdot \eta ^k. \end{aligned}$$

Based on this, in this paper we obtain the following result.

Theorem 4.2. Under the same hypotheses of the previous theorem, assume also that \({\text {Pic}}({{\mathbb {P}}}_\Sigma ^{2k+1})={\mathbb {Z}}\). Then, if \(V\subset X_f\) is a nonempty quasi-smooth intersection subvariety of \({{\mathbb {P}}}_\Sigma ^{2k+1}\) for some \(f\in N_{\lambda ,U}^{k,\beta }\), there exists \(c\in {\mathbb {Q}}^*\) such that \(\lambda _f=c\lambda _V\), where \(\lambda _V\) is the class of V in \(H_{\mathrm{prim}}^{k,k}(X_f,{\mathbb {Q}})\).

In other words, \(\lambda _f\) is algebraic.

In his paper [11] A. Dan proves a form of our Theorem 4.2 for smooth hypersurfaces in odd-dimensional projective spaces \({\mathbb {P}}^{2k+1}\) which is not asymptotic. Although our result is more general in two ways, as we consider quasi-smooth intersections in toric varieties with \(h^{k,k}=1\) (for instance, weighted or fake projective spaces); however, our result is asymptotic.

In Sect. 3 we give an extension of the notion of Gorenstein ideal to Cox rings; this may have some interest on its own.

2 Very general quasi-smooth intersections

Let \(f_1,\dots ,f_s\) be homogeneous polynomials in the Cox ring \(S = {\mathbb {C}}[x_1,\dots ,x_n]\) of \({\mathbb {P}}_{\Sigma }^d\). Their zero locus \(V(f_1,\dots ,f_s)\) defines a closed subvariety \(X\subset {\mathbb {P}}^{d}_{\Sigma }\). Let \(U(\Sigma )= {\mathbb {A}}^n -Z(\Sigma )\), where \(Z(\Sigma )\) is the irrelevant locus, i.e., \(Z(\Sigma ) = {\text {Spec}}B\), where B is the irrelevant ideal.

Definition 2.1

[15] X is a codimension s quasi-smooth intersection if \(V(f_1,\dots ,f_s)\cap U(\Sigma )\) is either empty or a smooth intersection subvariety of codimension s in \(U(\Sigma )\).

This notion generalizes that of smooth complete intersection in a projective space. For \(s=1\) it reduces to the notion of quasi-smoooth hypersurface, see Def. 3.1 in [1]. If we regard \({\mathbb {P}}_{\Sigma }^d\) as an orbifold, then an intersection of hypersurfaces \(X_{f_1}\cap \dots \cap X_{f_s}\) is quasi-smooth when it is a sub-orbifold of \({\mathbb {P}}_{\Sigma }^d\) , see Prop 1.3 [15]; heuristically, “X has only singularities coming from the ambient variety.”

We also have a Lefschetz type theorem in this context.

Proposition 2.2

( [15] Proposition 1.4) Let \(X\subset {\mathbb {P}}_{\Sigma }^d\) be a closed subset, defined by homogeneous polynomials \(f_1,\dots f_s\in B\). Then the natural map \(i^{*}: H^i({\mathbb {P}}_{\Sigma }^d)\rightarrow H^i(X)\) is an isomorphism for \(i<d-s\) and an injection for \(i=d-s\). In particular, this is true if the hypersurfaces cut by the polynomials \(f_i\) are ample.

Hence if \(p\ne \frac{d-s}{2}\) every cohomology class in \(H^{p,p}(X)\) is a linear combination of algebraic cycles. So let us see what happens when \(p=\frac{d-s}{2}\). The idea is to relate the Hodge structure of a quasi-smooth intersection variety \(X=X_{f_1}\cap \dots \cap X_{f_s}\) in \({\mathbb {P}}^d_{\Sigma }\) with the Hodge structure of a quasi-smooth hypersurface Y in a toric variety \({\mathbb {P}}^{d+s-1}_{X,\Sigma }\) whose fan depends on X and \(\Sigma \).

Proposition 2.3

Let \(X=X_{f_1} \cap \dots \cap X_{f_s}\) be quasi-smooth intersection subvariety in \({\mathbb {P}}_{\Sigma }^d\) cut off by homogeneous polynomials \(f_1\dots f_s\). There exists a projective simplicial toric variety \({\mathbb {P}}^{d+s-1}_{X,\Sigma }\) and a quasi-smooth hypersurface \(Y\subset {\mathbb {P}}^{d+s-1}_{X,\Sigma }\) such that for \(p\ne \frac{d+s-1}{2}, \frac{d+s-3}{2} \)

$$\begin{aligned} H^{p-1,d+s-1-p}_{\mathrm{\small prim}}(Y)\simeq H^{p-s,d-p}_{\mathrm{\small prim}}(X). \end{aligned}$$

Proof

One constructs \({\mathbb {P}}^{d+s-1}_{X,\Sigma }\) via the so-called “Cayley trick”. Let \(L_1,\dots , L_s\) be the line bundles associated to the quasi-smooth hypersurfaces \(X_1,\dots X_s\), and so let \({\mathbb {P}}(E)\) be the projective bundle of \(E=L_1\oplus \dots \oplus L_s\). It turns out that \({\mathbb {P}}(E)\) is a \(d+s-1\)- dimensional projective simplicial toric variety whose Cox ring is

$$\begin{aligned} {\mathbb {C}}[x_1,\dots ,x_n,y_1,\dots y_s] \end{aligned}$$

where \(S={\mathbb {C}}[x_1,\dots , x_n]\) is the Cox ring of \({\mathbb {P}}^{d}_{\Sigma }\). The hypersurface Y is cut off by the polynomial \(F=y_1f_1+\dots + y_sf_s\) and is quasi-smooth by Lemma 2.2 in [15]. Moreover, combining Theorem 10.13 in [1] and Theorem 3.6 in [15], we have that

$$\begin{aligned} {H^{p-1,d+s-1-p}_{\mathrm{\small prim}}(Y)}\simeq { R(F)_{(d+s-p)\beta -\beta _0}} \simeq H^{p-s,d-p}_{\mathrm{\small prim}}(X) \end{aligned}$$

for \(p\ne \frac{d+s-1}{2}, \frac{d+s-3}{2} \) as desired. \(\square \)

Here R(F) is the Jacobian ring of Y, i.e., the quotient of the Cox ring

$$\begin{aligned} R(F) = {\mathbb {C}}[x_1,\dots ,x_n,y_1,\dots y_s]/J(F), \end{aligned}$$

where J(F) is the ideal generated by the derivatives of F, see [1].

Remark 2.4

With the same notation of Proposition 2.3, note that we have a well defined map

$$\begin{aligned} \begin{array}{rcl} \phi : |\beta _1|\times \dots \times |\beta _s| &{}\rightarrow &{}|\beta |\\ (f_1,\dots ,f_s)&{}\mapsto &{} f_1y_1+\dots +f_sy_s. \end{array} \end{aligned}$$

Moreover, by the Noether-Lefschetz theorem \( NL_{\beta }\) is a countable union of closed sets \(\bigcup _i {C_i}\) and hence \(\bigcup \phi ^{-1}(C_i)\) is too.

We have a Noether-Lefschetz type theorem, namely,

Theorem 2.5

Let \({\mathbb {P}}^d_{\Sigma }\) be an Oda projective simplicial toric variety. Then for a very general quasi-smooth intersection subvariety X cut off by \(f_1,\dots f_s\) such that \(d+s=2(l+1)\) and \(\sum _{i=1}^s \deg (f_i)-\beta _0\) is nef, one has that

$$\begin{aligned} H^{l+1-s,l+1-s}(X,{\mathbb {Q}})=i^*\left( H^{l+1-s,l+1-s}({\mathbb {P}}^d_{\Sigma },{\mathbb {Q}})\right) \end{aligned}$$

So we get a natural generalization of the Noether-Lefschetz loci.

Definition 2.6

The Noether-Lefschetz locus \(NL_{\beta _1,\dots ,\beta _s}\) of quasi-smooth intersection varieties is the locus of \(s-\)tuples \((f_1,\dots ,f_s)\) such that \(X=X_{f_1}\cap \dots X_{f_s}\) is quasi-smooth intersection with \(f_i\in |\beta _i|\) and \(H^{l+1-s,l+1-s}(X,{\mathbb {Q}})\ne i^*\left( H^{l+1-s,l+1-s}({\mathbb {P}}^d_{\Sigma },{\mathbb {Q}})\right) \).

Now we consider the Hodge conjecture for very general quasi-smooth intersection subvarieties in \({\mathbb {P}}^d_{\Sigma }\).

Corollary 2.7

If \({\mathbb {P}}_{\Sigma }^d\) is a Oda projective simplicial toric variety, the Hodge Conjecture holds for a very general quasi-smooth intersection subvariety X cut off by \(f_1,\dots f_s\) such that \(d+s=2(l+1)\) and \(\sum _{i=1}^s \deg (f_i)- \beta _0\) is nef.

Proof

First note that by Thereom 4.1 in [12] the projective simplicial toric variety \({\mathbb {P}}_{X,\Sigma }^{2l+1}\) is Oda and since X is very general the quasi-smooth hypersurface Y is very general as well. So applying the Noether-Lefschetz theorem one has that \(h^{l,l}_{\mathrm{prim}}(Y)=0= h^{l+1-s,l+1-s}_{\mathrm{prim}}(X)\) or equivalently every \((l+1-s,+1-s)\) cohomology class is a linear combination of algebraic cycles. \(\square \)

3 Cox-Gorenstein ideals

We shall need a partial generalization of Macaulay’s theorem (see e.g. Thm. 6.19 in [18] for the classical theorem). This generalization is basically contained in the work of Cox and Cattani-Cox-Dickenstein [7, 9].

Let S be the Cox ring of a complete simplicial toric variety \({\mathbb {P}}_\Sigma \). This is graded over the effective classes in the class group \({\text {Cl}}({\mathbb {P}}_\Sigma )\) and [8]

$$\begin{aligned} S^\alpha \simeq H^0({\mathbb {P}}_\Sigma ,{\mathcal {O}}_{{\mathbb {P}}_\Sigma }(\alpha )). \end{aligned}$$

As \({\mathcal {O}}_{{\mathbb {P}}_\Sigma }(\alpha )\) is coherent and \({\mathbb {P}}_\Sigma \) is complete, each \(S^\alpha \) is finite-dimensional over \({\mathbb {C}}\); in particular, \(S^0\simeq {\mathbb {C}}\).

Lemma 3.1

For every effective \(N\in {\text {Cl}}({\mathbb {P}}_\Sigma )\), the set of classes \(\alpha \in {\text {Cl}}({\mathbb {P}}_\Sigma )\) such that \(N-\alpha \) is effective is finite.

Proof

Since the torsion submodule of \({\text {Cl}}({\mathbb {P}}_\Sigma )\) is finite, we may assume that \({\text {Cl}}({\mathbb {P}}_\Sigma )\) is free. Then the exact sequence

$$\begin{aligned} 0 \rightarrow M \rightarrow {\text {Div}}_{{\mathbb {T}}}({\mathbb {P}}_\Sigma ) \rightarrow {\text {Cl}}({\mathbb {P}}_\Sigma ) \rightarrow 0 \end{aligned}$$

splits, and we may identify \({\text {Cl}}({\mathbb {P}}_\Sigma )\) with a free subgroup of \( {\text {Div}}_{{\mathbb {T}}}({\mathbb {P}}_\Sigma )\), generated by a subset \(\{D_1,\dots ,D_r\}\) of \({\mathbb {T}}\)-invariant divisors. A class in \({\text {Cl}}({\mathbb {P}}_\Sigma )\) is effective if and only its coefficients on this basis are nonnegative, whence the claim follows. \(\square \)

We shall give a definition of Cox-Gorenstein ideal of the Cox rings which generalizes to toric varieties the definition given by Otwinowska in [16] for projective spaces. Let \(B\subset S\) be the irrelevant ideal, and for a graded ideal \(I\subset B\), denote by \(V_{{\mathbb {T}}}(I)\) the corresponding closed subscheme of \({\mathbb {P}}_\Sigma \).

Definition 3.2

A graded ideal I of S contained in B is said to be a Cox-Gorentstein ideal of socle degree \(N\in {\text {Cl}}({\mathbb {P}}_\Sigma )\) if

  1. 1.

    there exists a \({\mathbb {C}}\)-linear form \(\Lambda \in (S^N)^\vee \) such that for all \(\alpha \in {\text {Cl}}({\mathbb {P}}_\Sigma )\)

    $$\begin{aligned} I^\alpha =\{f\in S^\alpha \,\vert \, \Lambda (fg) = 0 \ \hbox {for all} \ g\in S^{N-\alpha }\}; \end{aligned}$$
    (2)
  2. 2.

    \(V_{{\mathbb {T}}}(I)=\emptyset \).

Remark 3.3

Cox-Gorenstein ideals need not be Artinian. Property 2 in this definition replaces that condition.

Proposition 3.4

Let \(R=S/I\). If I is Cox-Gorenstein then

  1. 1.

    \(\dim _{\mathbb {C}}R^N = 1\);

  2. 2.

    the natural bilinear morphism

    $$\begin{aligned} R^\alpha \times R^{N-\alpha } \rightarrow R^N\simeq {\mathbb {C}}\end{aligned}$$
    (3)

    is nondegenerate whenever \(\alpha \) and \(N-\alpha \) are effective.

Proof

  1. 1.

    From eq. (2) we see that the sequence

    $$\begin{aligned} 0 \rightarrow I^N \rightarrow S^N \xrightarrow {\Lambda } {\mathbb {C}}\rightarrow 0 \end{aligned}$$

    is exact.

  2. 2.

    Define \(\Phi :R^\alpha \times R^{N-\alpha } \rightarrow {\mathbb {C}}\) as \(\Phi (x,y) = \Lambda ({{\bar{x}}}{{\bar{y}}})\), where \({{\bar{x}}}\), \({{\bar{y}}}\) are pre-images of x, y in S. One easily checks that this is well defined and that via the isomorphism \(R^N\simeq k\) it coincides with the pairing (3). Now if \(x\in R^\alpha \) and \(\Phi (x,y) = 0\) for all \(y\in R^{N-\alpha }\) then \(\Lambda ({{\bar{x}}}{{\bar{y}}}) =0\) for all \({{\bar{y}}}\in S^{N-\alpha }\) so that \({{\bar{x}}} \in I^\alpha \), i.e., \(x=0\). \(\square \)

Let \(f_0,\dots ,f_d\) be homogeneous polynomials, \(f_i\in S^{\alpha _i}\), where \(d=\dim {\mathbb {P}}_\Sigma \) and each \(\alpha _i\) is ample, and let \( N = \sum _i\alpha _i-\beta _0\), where \(\beta _0\) is the anticanonical class of \({\mathbb {P}}_\Sigma \). Assume that the \(f_i\) have no common zeroes in \({\mathbb {P}}_\Sigma \), i.e., \(V_{{\mathbb {T}}}(I)=\emptyset \) if \(I=(f_0,\dots ,f_d)\).

In [1, 7, 9] it is shown that for each \(G \in S^N\) one can define a meromorphic d-form \(\xi _G\) on \({\mathbb {P}}_\Sigma \) by letting

$$\begin{aligned} \xi _G = \frac{G\,\Omega }{f_0\cdots f_d} \end{aligned}$$

where \(\Omega \) is a Euler form on \({\mathbb {P}}_\Sigma \). The form \(\xi _G\) determines a class in \(H^d({\mathbb {P}}_\Sigma ,\omega )\), where \(\omega \) is the canonical sheaf of \({\mathbb {P}}_\Sigma \) (the sheaf of Zariski d-forms on \({\mathbb {P}}_\Sigma \)), and in turn the trace morphism \({\text {Tr}}_{{\mathbb {P}}_\Sigma }:H^d({\mathbb {P}}_\Sigma ,\omega )\rightarrow {\mathbb {C}}\) associates a complex number to G, so we can define \(\Lambda \in (S^N)^\vee \) as

$$\begin{aligned} \Lambda (G) = {\text {Tr}}_{{\mathbb {P}}_\Sigma }([\xi _G])\in {\mathbb {C}}. \end{aligned}$$
(4)

Finally, we can prove a toric version of Macaulay’s theorem.

Theorem 3.5

The linear map defined in Eq. (4) satisfies the condition in Definition 3.2. Therefore, the ideal \(I=(f_0,\dots ,f_d)\) is a Cox-Gorenstein ideal of socle degree N.

Proof

By Theorem 6 in [7] the map \(\Lambda \) establishes an isomorphism \(R^N \simeq {\mathbb {C}}\). Hence, if \(f\in S^\alpha \) is such that \(\Lambda (fg)=0\) for all \(g\in S^{N-\alpha }\), then \(fg\in I^N\), which implies \(f\in I^\alpha \). On the other hand, it is clear that \(\Lambda (fg)=0\) if \(f\in I^\alpha \) and \(g\in S^{N-\alpha }\). \(\square \)

Another example is given in terms of toric Jacobian ideals. For every ray \(\rho \in \Sigma (1)\) we shall denote by \(v_\rho \) its rational generator, and by \(x_\rho \) the corresponding variable in the Cox ring. Recall that d is the dimension of the toric variety \({\mathbb {P}}_\Sigma \), while we denote by \(r=\#\Sigma (1)\) the number of rays. Given \(f\in S^\beta \) one defines its toric Jacobian ideal as

$$\begin{aligned} J_0(f) = \left( x_{\rho _1} \frac{\partial f }{\partial x_{\rho _1}}, \dots , x_{\rho _r} \frac{\partial f }{\partial x_{\rho _r}} \right) . \end{aligned}$$

We recall from [1] the definition of nondegenerate hypersurface and some properties (Def. 4.13 and Prop. 4.15).

Definition 3.6

Let \(f\in S(\Sigma )^\beta \), with \(\beta \) an ample Cartier class. The associated hypersurface \(X_f\) is nondegenerate if for all \(\sigma \in \Sigma \) the affine hypersurface \(X_f\cap O(\sigma )\) is a smooth codimension one subvariety of the orbit \(O(\sigma )\) of the action of the torus \({\mathbb {T}}^d\).

Proposition 3.7

  1. 1.

    Every nondegenerate hypersurface is quasi-smooth.

  2. 2.

    If f is generic then \(X_f\) is nondegenerate.

The following is part of Prop. 5.3 in [9], with some changes in the terminology.

Proposition 3.8

Let \(f\in S(\Sigma )^\beta \), and let \(\{\rho _1,\dots ,\rho _d\}\subset \Sigma (1)\) be such that \(v_{\rho _1},\dots ,v_{\rho _d}\) are linearly independent.

  1. 1.

    The toric Jacobian ideal of f coincides with the ideal

    $$\begin{aligned} \left( f, x_{\rho _1} \frac{\partial f }{\partial x_{\rho _1}}, \dots , x_{\rho _d} \frac{\partial f }{\partial x_{\rho _d}} \right) . \end{aligned}$$
  2. 2.

    The following conditions are equivalent:

    1. (a)

      f is nondegenerate;

    2. (b)

      the polynomials \(x_{\rho _i} \frac{\partial f }{\partial x_{\rho _i}}\), \(i=1,\dots ,r\), do not vanish simultaneously on \(X_f\);

    3. (c)

      the polynomials f and \(x_{\rho _i} \frac{\partial f }{\partial x_{\rho _i}}\), \(i=1,\dots ,d\),do not vanish simultaneously on \(X_f\).

  3. 3.

    If moreover \(\beta \)is ample and f is nondegenerate, then \(J_0(f)\) is a Cox-Gorenstein ideal of socle degree \(N = (d+1){\beta} -{\beta _0}\), where \(\beta _0\)is the anticanonical class of \({\mathbb {P}}_\Sigma ^d\).

4 Asymptotic Hodge conjecture

In this section we prove Theorem 4.2. Let us recall part of the notation and assumptions of [4]. Let \({{\mathbb {P}}}_\Sigma ^{2k+1}\) be an Oda variety with an ample Cartier class \(\beta \) such that \(k\beta -\beta _0=n\eta \), where \(\beta _0\) is the anticanonical class, \(\eta \) is a primitive ample class and \(n\in {\mathbb {N}}\).

We need to define a pre-order in the group

$$\begin{aligned} N^1({{\mathbb {P}}}_\Sigma ^{2k+1}) = {\text {Pic}}({{\mathbb {P}}}_\Sigma ^{2k+1})\otimes {\mathbb {Q}}/ \text {numerical equivalence}, \end{aligned}$$

by letting \(\alpha < \alpha '\) if \(\alpha '-\alpha \) is an effective class.

Let \(X_f\in \vert \beta \vert \) be a quasi-smooth hypersurface in the Noether-Lefschetz locus associated to a nontrival Hodge class \(\lambda \in F^k{\mathcal {H}}^{2k}(U)\). Again, its degree is computed by intersecting with the ample class \(\eta \), i.e., \(\deg X_f = [X_f]\cdot \eta \). Let r be number of rays of \(\Sigma \), so that \(r\ge 2(k+1)\). Assuming that n is big enough, it follows from Proposition 4.7 or Theorem 6.1 in [4] that there exists a k-dimensional subvariety V of \(X_f\) satisfying the following conditions:

  • \(\deg V\le 2\delta m_\beta \) with \(0< \delta < \frac{1}{4(r-(k+1))}\) (the number \(m_\beta \) was defined in Eq. (1));

  • the graded ideals \(I_V\) and

    $$\begin{aligned} E=\{g\in S^{\bullet } \mid \sum _{i=1}^b \lambda _i \int _{{\text {Tub}}\gamma _i} \frac{gh\Omega _0}{f^{k+1}}=0 \ \text{ for } \text{ all }\ h\in S^{N-\bullet } \} , \end{aligned}$$
    (5)

    coincide in degree less than or equal to \(\left( m_{\beta }-2-(r-j)\deg V\right) \eta \) for some j, with \(0<j<r\). Here \({\text {Tub}}(-)\) is the adjoint of the residue map, and \(N=(k+1)\beta -\beta _0\) is the socle degre of the Cox-Gorenstein ideal E, while

    $$\begin{aligned} \lambda _f =\left( \sum _{i=1}^b\,\lambda _i\gamma _i\right) ^{pd} \end{aligned}$$

    is the Poincaré dual of some rational combination of the homology cycles \(\gamma _i\) generating \(H_{2k}(X_f,{\mathbb {Q}})\). Moreover, via the isomorphism \(T_fU\simeq S^{\beta }\), the degree \(\beta \) summand \(E^\beta \) of E is identified with the tangent space \(T_fN^{k,\beta }_{\lambda ,U}\) to the Noether-Lefschetz locus, so that \(E^\beta \) contains the degree \(\beta \) part \(J(f)^\beta \) the Jacobian ideal of f.

Lemma 4.1

The toric Jacobian ideal \(J_0(f)\) is contained in E.

Proof

\(J_0(f)\subset J(f)\), so that \(J_0(f)^\beta \subset J(f)^\beta \subset E^\beta \), and since \(J_0(f)\) is generated in degree \(\beta \), one has \(J_0(f)\subset E\). \(\square \)

We denote by \(\lambda _V\) the class of V in \(H^{k,k}_{\mathrm{prim}}(X_f,{\mathbb {Q}})\). In the following theorem we assume that \({\text {Pic}}({{\mathbb {P}}}_\Sigma ^{2k+1})=1\), i.e., that \({{\mathbb {P}}}_\Sigma ^{2k+1}\) is a (possibly fake) weighted projective space [6, 13] (cf. [10] Exer. 5.1.13). This implies that \(h^{p,p}({{\mathbb {P}}}_\Sigma ^{2k+1})=1\) for all p.

Theorem 4.2

If V is a quasi-smooth intersection subvariety, there exists \(c\in {\mathbb {Q}}^*\) such that \(\lambda _f=c\lambda _V\).

Proof

We divide the proof in three steps.

Step I: \(\lambda _V \ne 0\). For clarity, for every cohomology class of a subvariety we denote in the cohomology of which ambient variety we consider it (so we write \([V]_{X_f}\) and \([V]_{{{\mathbb {P}}}_\Sigma ^{2k+1}}\)). Since \(V\subset X_f\) is a regular embedding we have

$$\begin{aligned} \begin{array}{ccl} [V]^2_{X_f}&{}=&{}\int _{V} c_k(N_{V/X_f})=\int _V \left[ c(N_{V/{{\mathbb {P}}}_\Sigma ^{2k+1}}) / c(N_{X_f/{{\mathbb {P}}}_\Sigma ^{2k+1}|V})\right] _k\\[8pt] &{}=&{} \int _{{{\mathbb {P}}}_\Sigma ^{2k+1}} [V]_{{{\mathbb {P}}}_\Sigma ^{2k+1}} \cup \Xi _k \end{array} \end{aligned}$$
(6)

where \(\Xi _k\) is the component in \(H^{k,k}({{\mathbb {P}}}_\Sigma ^{2k+1})\) of

$$\begin{aligned} \Xi = \frac{\prod _i (1+A_i)}{1 + [X_f]_{{{\mathbb {P}}}_\Sigma ^{2k+1}}}; \end{aligned}$$

here \(A_1,\dots ,A_{k+1}\) are the classes in \({\text {Cl}}({{\mathbb {P}}}_\Sigma ^{2k+1})\) of the hypersurfaces that cut the quasi-smooth intersection subvariety V. The claim is proved by contradiction: if \([V]_{X_f}\) is the restriction of a class in \(H^{k,k}({{\mathbb {P}}}_\Sigma ^{2k+1})\), i.e.,

$$\begin{aligned} {[}V]_{X_f} =b \, \eta ^k_{\vert X_f} \end{aligned}$$

for some b, then comparing this with (6) we obtain

$$\begin{aligned} \deg V = m_k \,\deg X_f, \end{aligned}$$
(7)

where \(m_k\) is defined by \( \Xi _k =m_k\,\eta ^k\). But (7) cannot be true when \(\deg X_f\) is big enough.

Step II. Let \(E_{\mathrm{alg}}\) and E be the Cox-Gorenstein ideals associated to \(\lambda _V\) and \(\lambda _f\), respectively, as in Eq. (5). To prove the theorem it is enough to show that \(E=E_{\mathrm{alg}}\). Note that \(I_V+J_0(f)\) is contained in E and \(E_{\mathrm{alg}}\). Moreover, since \(V\subset X_f\), and f is quasi-smooth, there exist \(K_1,\dots K_{k+1}\in B\) such that \(f=A_1K_1+\dots A_{k+1}K_{k+1}\) and \((A_1,\dots ,A_{k+1}, K_1,\dots K_{k+1})\) is a Cox-Gorenstein ideal with socle degree N; this will follow from the next step, which concludes the proof.

Step III. It is enough to show that every Cox-Gorenstein ideal \({\mathcal {I}}\) of socle degree N containing \(I_V+J_0(f)\) also contains \((A_1,\dots ,A_{k+1}, K_1,\dots K_{k+1})\). By assumption

$$\begin{aligned} \left( A_j,j\in \{1,\dots , k+1\}, \sum _{j=1}^{k+1}x_i\frac{\partial A_j}{\partial x_i}K_j, i\in {1,\dots , r} \right) \subset {\mathcal {I}}. \end{aligned}$$

Let us see that \(K_j\in {\mathcal {I}}\) for every \(j\in \{1,\dots ,k+1\}\). Let \(M\in {\text {Mat}}(r\times (k+1))\) be the matrix \([x_i\frac{\partial A_j}{\partial x_i}]\) and K the column \((K_j)_{j\in \{1,\dots ,k+1\}}\). Let \(I\subset \{1,\dots r\}\) with cardinality \(k+1\) and let \(M_I\) be the matrix obtained extracting the \(i\in I\)-arrows of M. We have that \(\sum _{j=1}^{k+1}x_i\frac{\partial A_j}{\partial x_i}K_j=(MK)_i=(M_IK)_i\); multiplying by the adjoint of \(M_I\) we get that \(\det (M_I)K_j\in {\mathcal {I}}\) for all \(j\in \{1,\dots k+1\}\). On one hand the ideal \(({\mathcal {I}}:K_j)\) contains the ideal

$$\begin{aligned} {\mathcal {J}}=I_V+\left\langle \det M_I\,\vert \, I\subset \{1,\dots ,r\},\ \# I = k+1\right\rangle . \end{aligned}$$

Since V is a smooth complete intersection subvariety, it follows that \({\mathcal {J}}\) is base point free, and therefore it contains a complete intersection Cox-Gorenstein ideal \({\mathcal {J}}'\) by the toric Macaulay theorem, Theorem 3.5. Since \({\mathcal {J}}\) is generated in degree less than or equal to \((\deg V) \eta \), we can take \({\mathcal {J}}'\) with the same property. It follows that

$$\begin{aligned} soc({\mathcal {J}}') \le 2(k+1)(\deg V)\eta - \beta _0 \le 2rm_\beta \delta \eta -\beta _0. \end{aligned}$$

On the other hand if \(K_j\notin {\mathcal {I}}\) then \(({\mathcal {I}}: K_j)\) contains a Cox-Gorenstein ideal with socle degree

$$\begin{aligned} N-\deg K_j\ge N-\beta = k\beta -\beta _0; \end{aligned}$$

then comparing the above two inequalities and keeping in mind that \(r\ge 2(k+1)\), we get

$$\begin{aligned} \delta \ge \frac{1}{2r} \ge \frac{1}{4(r-(k+1))}, \end{aligned}$$

which is absurd. \(\square \)

On behalf of all authors, the corresponding author states that there is no conflict of interest.