1 Introduction

Let X be an m-dimensional smooth projective variety and \(\mathcal {F}\) be a rank n locally free sheaf over it. A way to study \(\mathcal {F}\) is to study its extensions \(0\rightarrow \mathcal {L}\rightarrow \mathcal {E}\rightarrow \mathcal {F}\rightarrow 0\) which, up to isomorphism, are parametrized by \(\text {Ext}^1(\mathcal {F},\mathcal {L})\). In [2, 3, 5, 6, 1013] and [1] the adjoint forms associated to \(\xi \in \mathrm{{Ext}}^{1}(\mathcal {O}_X, \mathcal {F})\) are deeply studied and many applications are given. Let us recall the notion of adjoint form in the case \(\mathcal {L}=\mathcal {O}_X\).

Given \(\xi \in \mathrm{{Ext}}^{1}(\mathcal {O}_X, \mathcal {F})\), take an \((n+1)\)-dimensional subspace W of the kernel of the cup-product homomorphism \(\partial _{\xi }:H^0(X,\mathcal {F})\rightarrow H^1(X,\mathcal {O}_X)\). Denote by \(\lambda ^{i}W\) the image of \(\bigwedge ^iW\) through the natural homomorphism \(\lambda ^i:\bigwedge ^i H^0(X,\mathcal {F})\rightarrow H^0(X,\bigwedge ^{i}\mathcal {F})\). If \(\mathcal {B}:=\langle \eta _{1},\ldots ,\eta _{n+1}\rangle \) is a basis of W and \(s_1,\ldots , s_{n+1}\in H^0(X,\mathcal {E})\) are liftings of \(\eta _{1},\ldots ,\eta _{n+1}\), respectively, then the map \(\Lambda ^{n+1}:\bigwedge ^{n+1} H^0(X,\mathcal {E})\rightarrow H^0(X,\bigwedge ^{n+1}\mathcal {E})\) gives the top form \( \Omega :=\Lambda ^{n+1}(s_1\wedge s_2 \wedge \ldots \wedge s_{n+1})\in H^0(X,\det \mathcal {E})\). The section \(\Omega \) corresponds to a top form \(\omega _{\xi ,W,{\widehat{\mathcal {B}}}}\in H^0(X,\det \mathcal {F})\) via the isomorphism \(\det \mathcal {F}\simeq \det \mathcal {E}\), where \({\widehat{\mathcal {B}}}=\langle s_1,\ldots , s_{n+1}\rangle \); the form \(\omega _{\xi ,W,{\widehat{\mathcal {B}}}}\) is called an adjoint form of W and \(\xi \). To the basis \(\mathcal {B}\) there are also naturally associated \(n+1\) elements \(\omega _i:=\lambda ^{n}(\eta _1\wedge \ldots \wedge \eta _{i-1}\wedge {\widehat{\eta _i}}\wedge \eta _{i+1}\wedge \ldots \wedge \eta _{n+1})\), \(i=1,\ldots ,n+1\), obtained by the basis \(\langle \eta _1\wedge \ldots \wedge \eta _{i-1}\wedge {\widehat{\eta _i}}\wedge \eta _{i+1}\wedge \ldots \wedge \eta _{n+1}\rangle _{i=1}^{n+1}\) of \(\bigwedge ^n W\). Note that if we change the liftings \(s_1,\ldots , s_{n+1}\in H^0(X,\mathcal {E})\) with other liftings \({\widetilde{s}_1},\ldots , {\widetilde{s}_{n+1}}\), then \(\omega _{\xi ,W,\widehat{\mathcal {B}}}\) is a linear combination of \(\omega _{\xi ,W,\widetilde{\mathcal {B}}}\) and \(\omega _1\),..., \(\omega _{n+1}\). The natural problem of this theory is to characterize the condition \(\omega _{\xi ,W,\widehat{\mathcal {B}}}\in \lambda ^nW\) in terms of the fixed divisor \(D_W\) of \(|\lambda ^nW|\subset \mathbb {P} H^0(X,\det \mathcal {F})\) and of the base locus \(Z_W\) of the moving part \(M_W\in \mathbb {P} H^0(X,\det \mathcal {F}\otimes _{\mathcal {O}_X}\mathcal {O}_{X}(-D_W))\), where \(|\lambda ^nW|=D_W+|M_W|\).

In this paper we consider the general case where \(\mathcal {L}\) is an invertible sheaf not necessarily equal to \(\mathcal {O}_X\). In this case \( \det \mathcal {E}=\mathcal {L}\otimes \det \mathcal {F}\) and liftings \(s_1,\ldots , s_{n+1}\in H^0(X,\mathcal {E})\) of \(\eta _{1},\ldots ,\eta _{n+1}\in H^0(X,\mathcal {F})\) determine \(\Omega :=\Lambda ^{n+1}(s_1\wedge s_2 \wedge \ldots \wedge s_{n+1})\in H^0(X,\det \mathcal {E})\) which is now called a generalized adjoint form. We define as before \(\omega _i:=\lambda ^{n}(\eta _1\wedge \ldots \wedge \eta _{i-1}\wedge {\widehat{\eta _i}}\wedge \eta _{i+1}\wedge \ldots \wedge \eta _{n+1})\), \(i=1,\ldots , n+1\) and we characterize the case where \(\Omega \) belongs to the image of \(H^0(X,\mathcal {L})\otimes \lambda ^nW\rightarrow H^0(X,\det \mathcal {E})\) by the natural tensor product map. The game is more complicated than in the above-mentioned papers because the linear system \(|\lambda ^nW|\) is inside \(\mathbb {P} H^0(X,\det \mathcal {F})\) and we have to relate the fixed divisor \(D_W\) of \(|\lambda ^nW|\) and the base locus \(Z_W\) of the moving part \(M_W\in \mathbb {P} H^0(X,\det \mathcal {F}\otimes _{\mathcal {O}_X}\mathcal {O}_{X}(-D_W))\) to forms which are not anymore inside \(H^0(X,\det \mathcal {F})\). Nevertheless the result is analogue to the one of [11, Theorem 1.5.1] and [13, Theorem 2.1.7]:

Theorem [A] Let X be an m-dimensional complex compact smooth variety. Let \(\mathcal {F}\) be a rank n locally free sheaf on X and \(\mathcal {L}\) an invertible sheaf. Consider an extension \(0\rightarrow \mathcal {L}\rightarrow \mathcal {E}\rightarrow \mathcal {F}\rightarrow 0\) corresponding to \(\xi \in \mathrm{{Ext}}^1(\mathcal {F},\mathcal {L})\). Let \(W=\langle \eta _1,\ldots ,\eta _{n+1}\rangle \) be an \(n+1\)-dimensional sublinear system of \(\ker ({\partial _\xi })\subset H^0(X,\mathcal {F})\). Let \(\Omega \in H^0(X,\det \mathcal {E})\) be a generalized adjoint form associated to W as above. It holds that if \(\Omega \in \mathrm {Im}\,(H^0(X,\mathcal {L})\otimes \lambda ^nW\rightarrow H^0(X,\det \mathcal {E}))\) then \(\xi \in \ker (H^1(X,\mathcal {F}^\vee \otimes \mathcal {L})\rightarrow H^1(X,\mathcal {F}^\vee \otimes \mathcal {L}(D_W)))\).

Theorem [A], called Adjoint Theorem, can be thought as a general version of the well-known Castelnuovo’s free pencil trick; c.f. see Theorem 2.8.

We have also a viceversa of the Adjoint Theorem; see: Theorem 2.9:

Theorem [B] Under the same hypothesis of Theorem [A], assume also that \(H^0(X,\mathcal {L})\cong H^0(X,\mathcal {L}(D_W))\). It holds that if \(\xi \in \ker (H^1(X,\mathcal {F}^\vee \otimes \mathcal {L})\rightarrow H^1(X,\mathcal {F}^\vee \otimes \mathcal {L}(D_W)))\), then \(\Omega \in \mathrm {Im}\,(H^0(X,\mathcal {L})\otimes \lambda ^nW\rightarrow H^0(X,\det \mathcal {E}))\).

In particular in the case \(D_W=0\) Theorem [B] is a full characterization of the condition \(\Omega \in \mathrm {Im}\,(H^0(X,\mathcal {L})\otimes \lambda ^nW\rightarrow H^0(X,\det \mathcal {E}))\).

Now by the Adjoint Theorem and by Theorem [B] we can study extension classes of sheaves via adjoint forms. Indeed even if \(\mathcal {F}\) has no global sections we can always take the tensor product with a sufficiently ample linear system \(\mathcal {M}\) such that \(\mathcal {F}\otimes \mathcal {M}\) has enough global sections in order to apply the theory of adjoint forms. By applying the above idea to the case where \(n>2\), \(X\subset \mathbb {P}^n\) is an hypersurface of degree \(d>3\) and \(\mathcal {F}:=\Omega ^{1}_{X}\otimes _{\mathcal {O}_X}\mathcal {O}_X(2)\) we have a reformulation of the infinitesimal Torelli Theorem for X in the setting of generalized adjoint theory. In this paper we will not recall the theory concerning infinitesimal Torelli Theorems, for which a reference is [16], in any case a quick introduction to this topic is also given in [13]. Here we point out only that given a degree d form \(F\in \mathbb {C}[\xi _0,\ldots ,\xi _n]\) the Jacobian ideal of F is the ideal \(\mathcal {J}\) generated by the partial derivatives \(\frac{\partial {F}}{\partial {\xi _i}}\) for \(i=0,\ldots ,n\) and by [9][Theorem 9.8], any infinitesimal deformation \(\xi \in H^1(X,\Theta _X)\), where \(X=(F=0)\) and \(\Theta _X\) is the sheaf of tangent vectors on X, is given by a class [R] in the quotient \(\mathbb {C}[\xi _0,\ldots ,\xi _n]/\mathcal {J}\) where R is a homogeneous form of degree d.

Theorem [C] For a smooth hypersurface X of degree d in \(\mathbb {P}^n\) with \(n\ge 3\) and \(d>3\) the following are equivalent:

  • (1) the differential of the period map is zero on the infinitesimal deformation

    $$\begin{aligned}{}[R]\in ( \mathbb {C}[\xi _0,\ldots ,\xi _n]/\mathcal {J})_{d}\simeq H^1(X,\Theta _X) \end{aligned}$$
  • (2) R is an element of the Jacobian ideal \(\mathcal {J}\)

  • (3) \(\Omega \in \mathrm {Im}\,(H^0(X,\mathcal {O}_X(2))\otimes \lambda ^nW \rightarrow H^0(X,\mathcal {O}_X(n+d-1)))\) for the generic generalized adjoint \(\Omega \)

  • (4) The generic generalized adjoint \(\Omega \) lies in \(\mathcal {J}\).

Note that Theorem [C] has a different flavor with respect to the analogue [9, Theorem 9.8] since we essentially use meromorphic 1-forms over X; see Proposition 3.7. Finally we want to mention that in a forthcoming paper [14] we show how to recover also the Green’s infinitesimal Torelli Theorem for a sufficiently ample divisor of a smooth variety in terms of generalized adjoint theory.

2 The theory of generalized adjoint forms

2.1 Definition of generalized adjoint form

Let X be a smooth compact complex variety of dimension m and let \(\mathcal {F}\) and \(\mathcal {L}\) be two locally free sheaves on X of rank n and 1, respectively. Consider the exact sequence of locally free sheaves

$$\begin{aligned} 0\rightarrow \mathcal {L}\rightarrow \mathcal {E}\rightarrow \mathcal {F}\rightarrow 0 \end{aligned}$$
(2.1)

associated to an element \(\xi \in \text {Ext}^1(\mathcal {F},\mathcal {L})\cong H^1(X,\mathcal {F}^\vee \otimes \mathcal {L})\). Recall that the invertible sheaf \(\det \mathcal {F}:=\bigwedge ^n\mathcal {F}\) fits into the exact sequence

$$\begin{aligned} 0\rightarrow \bigwedge ^{n-1}\mathcal {F}\otimes \mathcal {L}\rightarrow \bigwedge ^n\mathcal {E}\rightarrow \det \mathcal {F}\rightarrow 0, \end{aligned}$$
(2.2)

which still corresponds to \(\xi \) under the isomorphism \(\text {Ext}^1(\mathcal {F},\mathcal {L})\cong \text {Ext}^1(\det \mathcal {F},\bigwedge ^{n-1}\mathcal {F}\otimes \mathcal {L})\cong H^1(X,\mathcal {F}^\vee \otimes \mathcal {L})\). Furthermore \(\det \mathcal {F}\) satisfies

$$\begin{aligned} \det \mathcal {F}\otimes \mathcal {L}\cong \det \mathcal {E}. \end{aligned}$$
(2.3)

Let \(\partial _\xi :H^0(X,\mathcal {F})\rightarrow H^1(X,\mathcal {L})\) be the connecting homomorphism related to (2.1), and let \(W\subset \ker (\partial _\xi )\) be a vector subspace of dimension \(n+1\). Choose a basis \(\mathcal {B}:=\{\eta _1,\ldots ,\eta _{n+1}\}\) of W. By definition we can take liftings \(s_1,\ldots ,s_{n+1}\in H^0(X,\mathcal {E})\) of the sections \(\eta _1,\ldots ,\eta _{n+1}\). If we consider the natural map

$$\begin{aligned} \Lambda ^n:\bigwedge ^{n}H^0(X,\mathcal {E})\rightarrow H^0(X,\bigwedge ^n\mathcal {E}) \end{aligned}$$

we can define the sections

$$\begin{aligned} \Omega _i:=\Lambda ^n(s_1\wedge \ldots \wedge \hat{s_i}\wedge \ldots \wedge s_{n+1}) \end{aligned}$$
(2.4)

for \(i=1,\ldots ,n+1\). Denote by \(\omega _i\), for \(i=1,\ldots ,n+1\), the corresponding sections in \(H^0(X,\det \mathcal {F})\). Obviously we have that \(\omega _i=\lambda ^n(\eta _1\wedge \ldots \wedge \hat{\eta _i}\wedge \ldots \wedge \eta _{n+1})\), where \(\lambda ^n\) is the natural morphism

$$\begin{aligned} \lambda ^n:\bigwedge ^{n}H^0(X,\mathcal {F})\rightarrow H^0(X,\det \mathcal {F}). \end{aligned}$$

The vector subspace of \(H^0(X,\det \mathcal {F})\) generated by \(\omega _1,\ldots ,\omega _{n+1}\) is denoted by \(\lambda ^nW\).

Definition 2.1

If \(\lambda ^nW\) is nontrivial, it induces a sublinear system \(|\lambda ^n W|\subset \mathbb {P}(H^0(X,\det \mathcal {F}))\) that we will call adjoint sublinear system. We call \(D_W\) its fixed divisor and \(Z_W\) the base locus of its moving part \(|M_W|\subset \mathbb {P}(H^0(X,\det \mathcal {F}(-D_W)))\).

Definition 2.2

The section \(\Omega \in H^0(X,\det \mathcal {E})\) corresponding to \(s_1\wedge \ldots \wedge s_{n+1}\) via

$$\begin{aligned} \Lambda ^{n+1}:\bigwedge ^{n+1}H^0(X,\mathcal {E})\rightarrow H^0(X,\det \mathcal {E}) \end{aligned}$$
(2.5)

is called generalized adjoint form.

Remark 2.3

It is easy to see by local computation that this section is in the image of the natural injection \(\det \mathcal {E}(-D_W)\otimes \mathcal {I}_{Z_W}\rightarrow \det \mathcal {E}\).

We want to study the condition

$$\begin{aligned} \Omega \in \mathrm {Im}\,(H^0(X,\mathcal {L})\otimes \left\langle \Omega _i\right\rangle \rightarrow H^0(X,\det \mathcal {E})) \end{aligned}$$
(2.6)

or, equivalently,

$$\begin{aligned} \Omega \in \mathrm {Im}\,(H^0(X,\mathcal {L})\otimes \lambda ^nW \rightarrow H^0(X,\det \mathcal {E})). \end{aligned}$$
(2.7)

The first map is given by the wedge product, the second one by (2.3). Note that if \(H^0(X,\mathcal {L})=0\) this condition is equivalent to \(\Omega =0\).

Remark 2.4

The choice of the liftings is not relevant for this purpose. Take different liftings \(s_1',\ldots ,s_{n+1}'\in H^0(X,\mathcal {E})\) of \(\eta _1,\ldots ,\eta _{n+1}\) and call \(\Omega _i'\in H^0(X,\bigwedge ^n\mathcal {E})\) and \(\Omega '\in H^0(X,\det \mathcal {E})\) the corresponding sections constructed as above. Obviously

$$\begin{aligned} \mathrm {Im}\,(H^0(X,\mathcal {L})\otimes \left\langle \Omega _i\right\rangle \rightarrow H^0(X,\det \mathcal {E}))=\mathrm {Im}\,(H^0(X,\mathcal {L})\otimes \left\langle \Omega _i'\right\rangle \rightarrow H^0(X,\det \mathcal {E})), \end{aligned}$$
(2.8)

since they are both equal to \(\mathrm {Im}\,(H^0(X,\mathcal {L})\otimes \lambda ^nW \rightarrow H^0(X,\det \mathcal {E}))\). It is also easy to see that \(\Omega \in \mathrm {Im}\,(H^0(X,\mathcal {L})\otimes \left\langle \Omega _i\right\rangle \rightarrow H^0(X,\det \mathcal {E}))\) iff \(\Omega '\in \mathrm {Im}\,(H^0(X,\mathcal {L})\otimes \left\langle \Omega _i'\right\rangle \rightarrow H^0(X,\det \mathcal {E}))\).

Remark 2.5

Consider another basis \(\mathcal {B}':=\{\eta _1',\ldots ,\eta _{n+1}'\}\) of W and let A be the matrix of the basis change. The sections \(s_1',\ldots ,s_{n+1}'\) obtained from \(s_1,\ldots ,s_{n+1}\) through the matrix A are liftings of \(\eta _1',\ldots ,\eta _{n+1}'\). The section \(\Omega ':=\Lambda ^{n+1}(s_1'\wedge \ldots \wedge s_{n+1}')\) satisfies \(\Omega '=\det A\cdot \Omega \). Moreover \(\Omega \in \mathrm {Im}\,(H^0(X,\mathcal {L})\otimes \left\langle \Omega _i\right\rangle \rightarrow H^0(X,\det \mathcal {E}))\) iff \(\Omega '\in \mathrm {Im}\,(H^0(X,\mathcal {L})\otimes \left\langle \Omega _i'\right\rangle \rightarrow H^0(X,\det \mathcal {E}))\).

Lemma 2.6

If \(\Omega \in \mathrm {Im}\,(H^0(X,\mathcal {L})\otimes \left\langle \Omega _i\right\rangle \rightarrow H^0(X,\det \mathcal {E}))\), then we can find liftings \(\tilde{s_i}\in H^0(X,\mathcal {E})\), \(i=1,\ldots ,n+1\), such that \(\tilde{\Omega }:=\Lambda ^{n+1}(\tilde{s_1}\wedge \ldots \wedge \tilde{s}_{n+1})=0\).

Proof

By hypothesis there exist \(\sigma _i\in H^0(X,\mathcal {L})\) such that

$$\begin{aligned} \Omega =\sum ^{n+1}_{i=1} \sigma _i\wedge \Omega _i \end{aligned}$$
(2.9)

We can define new liftings for the element \(\eta _i\):

$$\begin{aligned} \tilde{s_i}:=s_i+(-1)^{n-i}\sigma _i. \end{aligned}$$

Now, since

$$\begin{aligned} \tilde{s_1}\wedge \ldots \wedge \tilde{s}_{n+1}=s_1\wedge \ldots \wedge s_{n+1}-\sum ^{n+1}_{i=1} s_1\wedge \ldots \wedge \hat{s_i}\wedge \ldots \wedge s_{n+1}\wedge \sigma _i, \end{aligned}$$
(2.10)

we immediately deduce \(\tilde{\Omega }=0\). \(\square \)

From the natural map

$$\begin{aligned} \mathcal {F}^\vee \otimes \mathcal {L}\rightarrow \mathcal {F}^\vee \otimes \mathcal {L}(D_W) \end{aligned}$$

we have a homomorphism

$$\begin{aligned} H^1(X,\mathcal {F}^\vee \otimes \mathcal {L})\mathop {\rightarrow }\limits ^{\rho } H^1(X,\mathcal {F}^\vee \otimes \mathcal {L}(D_W)); \end{aligned}$$

we call \(\xi _{D_W}=\rho (\xi )\).

2.2 Castelnuovo’s free pencil trick

Consider the case where both \(\mathcal {L}\) and \(\mathcal {F}\) are of rank one, while X has dimension m. In this case \(W=\left\langle \eta _1,\eta _2\right\rangle \subset H^0(X,\mathcal {F})\) has dimension two; as usual we choose liftings \(s_1,s_2\in H^0(X,\mathcal {E})\) of \(\eta _1,\eta _2\). Note also that \(\omega _1=\eta _2\) and \(\omega _2=\eta _1\), in particular \(W=\lambda ^1W\) so \(D_W\) is the fixed part of W and \(Z_W\) is the base locus of its moving part. Call \(\tilde{\eta }_i \in H^0(X,\mathcal {F}(-D_W))\) the sections corresponding to the \(\eta _i\)’s via \(H^0(X,\mathcal {F}(-D_W))\rightarrow H^0(X,\mathcal {F})\). The following lemma is well known and it is the core of the Castelnuovo base point free pencil trick.

Lemma 2.7

We have an exact sequence

$$\begin{aligned} 0\rightarrow \mathcal {F}^\vee (D_W)\mathop {\rightarrow }\limits ^{i}\mathcal {O}_X\oplus \mathcal {O}_X\mathop {\rightarrow }\limits ^{\nu }\mathcal {F}(-D_W)\otimes \mathcal {I}_{Z_W}\rightarrow 0 \end{aligned}$$
(2.11)

where the morphism i is given by contraction with \(-\tilde{\eta }_1\) and \(\tilde{\eta }_2\), while \(\nu \) is given by evaluation with \(\tilde{\eta }_2\) on the first component and \(\tilde{\eta }_1\) on the second one.

It is easy to see by local computation that sequence (2.11) fits into the following commutative diagram

(2.12)

The morphism \(\mathcal {E}^\vee \rightarrow \mathcal {O}_X\oplus \mathcal {O}_X\) is given by contraction with the sections \(-s_1\) and \(s_2\), the morphism \(\mathcal {L}^\vee \rightarrow \mathcal {F}(-D_W)\otimes \mathcal {I}_{Z_W}\) by contraction with the adjoint \(\Omega \). We can prove now the following

Theorem 2.8

Let X be an m-dimensional complex compact smooth variety. Let \(\mathcal {F}\), \(\mathcal {L}\) be invertible sheaves on X. Consider \(\xi \in H^1(X,\mathcal {F}^\vee \otimes \mathcal {L})\) associated to the extension (2.1). Define \(W=\left\langle \eta _1,\eta _{2}\right\rangle \subset \ker ({\partial _\xi })\subset H^0(X,\mathcal {F})\) and \(\Omega \) as above. We have that \(\Omega \in \mathrm {Im}\,(H^0(X,\mathcal {L})\otimes W\rightarrow H^0(X,\det \mathcal {E}))\) if and only if \(\xi _{D_W}=0\).

Proof

Tensoring (2.12) by \(\mathcal {L}\) and passing to cohomology we have the following diagram

(2.13)

Obviously \(\beta (1)=\Omega \) and, by commutativity, \(\delta (\beta (1))=\xi _{D_W}\). We have then \(\xi _{D_W}=0\) if and only if \(\Omega \in \mathrm {Im}\,(H^0(\mathcal {L}\oplus \mathcal {L})\mathop {\rightarrow }\limits ^{\nu } H^0(\mathcal {F}(-D_W)\otimes \mathcal {I}_{Z_W}\otimes \mathcal {L}))\). Since \(\nu \) is given by the sections \(\tilde{\eta }_2\) and \(\tilde{\eta }_1\), this condition is equivalent to \(\Omega \in \mathrm {Im}\,(H^0(X,\mathcal {L})\otimes W\rightarrow H^0(X,\det \mathcal {E}))\), since \(\det \mathcal {E}=\mathcal {F}\otimes \mathcal {L}\). \(\square \)

2.3 The Adjoint Theorem

We go back now to the general case with \(\mathcal {F}\) locally free of rank n. By obvious identifications the natural map

$$\begin{aligned} \text {Ext}^1(\det \mathcal {F},\bigwedge ^{n-1}\mathcal {F}\otimes \mathcal {L})\rightarrow \text {Ext}^1(\det \mathcal {F}(-D_W),\bigwedge ^{n-1}\mathcal {F}\otimes \mathcal {L}) \end{aligned}$$

gives an extension \(\mathcal {E}^{(n)}\) and a commutative diagram:

(2.14)

2.3.1 The proof of the Adjoint Theorem

By the hypothesis \(\Omega \in \mathrm {Im}\,(H^0(X,\mathcal {L})\otimes \lambda ^nW\rightarrow H^0(X,\det \mathcal {E}))\) and by lemma (2.6), we can choose liftings \(s_i\in H^0(X,\mathcal {E})\) of \(\eta _i\) with \(\Omega =0\).

Since \(D_W\) is the fixed divisor of the linear system \(|\lambda ^n W|\) and the sections \(\omega _i\) generate this linear system, then the \(\omega _i\) are in the image of

$$\begin{aligned} \det \mathcal {F}(-D_W)\rightarrow \det \mathcal {F}, \end{aligned}$$

so we can find sections \(\tilde{\omega }_i\in H^0(X,\det \mathcal {F}(-D_W))\) such that

$$\begin{aligned} \tilde{\omega }_i\cdot d=\omega _i, \end{aligned}$$
(2.15)

where d is a global section of \(\mathcal {O}_X(D_W)\) with \((d)=D_W\). Hence, using the commutativity of (2.14), we can find liftings \(\tilde{\Omega }_i\in H^0(X,\mathcal {E}^{(n)})\) of the sections \(\Omega _i\). The evaluation map

$$\begin{aligned} \bigoplus _{i=1}^{n+1}\mathcal {O}_X\mathop {\rightarrow }\limits ^{\tilde{\mu }}\mathcal {E}^{(n)} \end{aligned}$$

given by the global sections \(\tilde{\Omega }_i\), composed with the map \(\alpha \) of (2.14), induces a map \(\mu \) which fits into the following diagram

We point out that the morphism \(\mu \) is given by multiplication by \(\tilde{\omega }_i\) on the i-th component. The sheaf \(\mathrm {Im}\,\tilde{\mu }\) is torsion free since it is a subsheaf of the locally free sheaf \(\mathcal {E}^{(n)}\). Moreover, since \(\Omega =0\), a local computation shows that \(\mathrm {Im}\,\tilde{\mu }\) has rank one outside \(Z_W\). On the other hand the sheaf \(\mathrm {Im}\,\mu \) is by definition

$$\begin{aligned} \mathrm {Im}\,\mu =\det \mathcal {F}(-D_W)\otimes \mathcal {I}_{Z_W}. \end{aligned}$$

The morphism

$$\begin{aligned} \alpha :\mathcal {E}^{(n)}\rightarrow \det \mathcal {F}(-D_W) \end{aligned}$$

induces a surjective morphism, that we continue to call \(\alpha \),

$$\begin{aligned} \mathrm {Im}\,\tilde{\mu }\mathop {\rightarrow }\limits ^{\alpha }\mathrm {Im}\,\mu \end{aligned}$$

between two sheaves that are locally free of rank one outside \(Z_W\). This morphism is also injective, because its kernel is a torsion subsheaf of the torsion free sheaf \(\mathrm {Im}\,\tilde{\mu }\), hence it is trivial.

We have proved that

$$\begin{aligned} \mathrm {Im}\,\tilde{\mu }\cong \det \mathcal {F}(-D_W)\otimes \mathcal {I}_{Z_W}, \end{aligned}$$

so

$$\begin{aligned} \mathcal {E}^{(n)}\supset (\mathrm {Im}\,\tilde{\mu })^{\vee \vee }\cong \det \mathcal {F}(-D_W). \end{aligned}$$

This isomorphism gives the splitting

Since \(\xi _{D_W}\) is the element of \(H^1(X,\mathcal {F}^\vee \otimes \mathcal {L}(D_W))\) associated to this extension, we conclude that \(\xi _{D_W}=0\).

We have proved the Adjoint Theorem.

2.3.2 An inverse of the Adjoint Theorem

We prove now an inverse of the Adjoint Theorem.

Theorem 2.9

Let X be an m-dimensional complex compact smooth variety. Let \(\mathcal {F}\) be a rank n locally free sheaf on X and \(\mathcal {L}\) an invertible sheaf. Consider an extension \(0\rightarrow \mathcal {L}\rightarrow \mathcal {E}\rightarrow \mathcal {F}\rightarrow 0\) corresponding to \(\xi \in {\mathrm{Ext}}^1(\mathcal {F},\mathcal {L})\). Let \(W=\langle \eta _1,\ldots ,\eta _{n+1}\rangle \) be a \(n+1\)-dimensional sublinear system of \(\ker ({\partial _\xi })\subset H^0(X,\mathcal {F})\). Let \(\Omega \in H^0(X,\det \mathcal {E})\) be an adjoint form associated to W as above. Assume that \(H^0(X,\mathcal {L})\cong H^0(X,\mathcal {L}(D_W))\). If \(\xi \in \ker (H^1(X,\mathcal {F}^\vee \otimes \mathcal {L})\rightarrow H^1(X,\mathcal {F}^\vee \otimes \mathcal {L}(D_W)))\), then \(\Omega \in \mathrm {Im}\,(H^0(X,\mathcal {L})\otimes \lambda ^nW\rightarrow H^0(X,\det \mathcal {E}))\).

Proof

If \(\mathcal {F}\) is a rank one sheaf, then (2.8) gives the thesis without the extra assumption \(H^0(X,\mathcal {L})\cong H^0(X,\mathcal {L}(D_W))\). We assume then rank \(\mathcal {F}\) \(\ge 2\).

By (2.3), we can write \((\Omega )=D_W+F\) with F effective. In the first step of the proof we want to find a global section

$$\begin{aligned} \Omega '\in H^0\left( X,\bigwedge ^n\mathcal {E}\otimes \mathcal {L}(-F)\right) \end{aligned}$$

which restricts, through the natural map

$$\begin{aligned} \bigwedge ^n\mathcal {E}\otimes \mathcal {L}(-F)\rightarrow \det \mathcal {E}(-F), \end{aligned}$$

to the section \(d\in H^0(\det \mathcal {E}(-F))\), where \((d)=D_W\).

Consider the commutative diagram:

By the hypothesis \(\xi _{D_W}=0\) it follows easily that there exists a lifting \(\tilde{\Omega }\in H^0(X,\bigwedge ^{n}\mathcal {E}\otimes \mathcal {L})\) of \(\Omega \). Indeed, tensor (2.14) by \(\mathcal {L}\) and take a global lifting \(f\in H^0(X, \det \mathcal {E}(-D_W))\) of \(\Omega \). Since \(\xi _{D_W}=0\), f can be lifted to a section \(e\in H^0(X, \mathcal {E}^{(n)}\otimes \mathcal {L})\). Define \(\tilde{\Omega }:=\psi (e)\). By commutativity, \(H_3(\tilde{\Omega }|_F)=0\) hence we call \(\bar{\mu }\in H^0(X,\bigwedge ^{n-1}\mathcal {F}\otimes \mathcal {L}^{\otimes 2}|_F)\) the lifting of \(\tilde{\Omega }|_F\). A local computation shows that the connecting homomorphism

$$\begin{aligned} \delta :H^0(X,\bigwedge ^{n-1}\mathcal {F}\otimes \mathcal {L}^{\otimes 2}|_F)\rightarrow H^1(X,\bigwedge ^{n-1}\mathcal {F}\otimes \mathcal {L}^{\otimes 2}(-F)) \end{aligned}$$

maps \(\bar{\mu }\) to \(\xi _{D_W}\), which is zero by hypothesis. Then there exists a global section

$$\begin{aligned} \mu \in H^0(X,\bigwedge ^{n-1}\mathcal {F}\otimes \mathcal {L}^{\otimes 2}) \end{aligned}$$

which is a lifting of \(\bar{\mu }\). The section

$$\begin{aligned} \hat{\Omega }:=\Omega -\tau (\mu )\in H^0(X,\bigwedge ^{n}\mathcal {E}\otimes \mathcal {L}) \end{aligned}$$

is a new lifting of \(\Omega \) that, by construction, vanishes when restricted to F. We call

$$\begin{aligned} \Omega '\in H^0(X,\bigwedge ^{n}\mathcal {E}\otimes \mathcal {L}(-F)) \end{aligned}$$

the global section which lifts \(\hat{\Omega }\). It is easy to see that \(G_2(\Omega ')=d\) so \(\Omega '\) is the section we wanted.

In the second part of the proof we prove that \(\Omega \in \mathrm {Im}\,(H^0(X,\mathcal {L})\otimes \lambda ^nW\rightarrow H^0(X,\det \mathcal {E}))\). The global sections

$$\begin{aligned} \omega _i:=\lambda ^n(\eta _1\wedge \ldots \wedge \hat{\eta _i}\wedge \ldots \wedge \eta _{n+1})\in H^0(X,\det \mathcal {F}) \end{aligned}$$

generate \(\lambda ^nW\) and by definition they vanish on \(D_W\), that is there exist global sections \(\tilde{\omega }_i\in H^0(X,\det \mathcal {F}(-D_W))\) such that

$$\begin{aligned} \omega _i=\tilde{\omega }_i\cdot d. \end{aligned}$$

We consider the commutative diagram

(2.16)

The map \(\beta \) is locally defined by

$$\begin{aligned} (f_1,\ldots ,f_{n+1})\mapsto (-1)^nf_1\cdot s_1+\cdots +f_{n+1}\cdot s_{n+1}. \end{aligned}$$

The map \(\alpha \) is defined in the following way: if \(f\in \mathcal {L}(-F)(U)\) is a local section, then, locally on U, \(\alpha \) is given by

$$\begin{aligned} f\mapsto (\tilde{\omega }_1(f),\ldots ,\tilde{\omega }_{n+1}(f)), \end{aligned}$$

where we observe that the sections \(\tilde{\omega }_i\) are global sections of the dual sheaf of \(\mathcal {L}(-F)\). The sheaf \(\bar{\mathcal {F}}\) is by definition the cokernel of the first row. Now, tensoring by \(\mathcal {L}^\vee \), we have

(2.17)

Dualizing and tensoring again by \(\mathcal {O}_X(D_W)\), we obtain the commutative square

where we have used the isomorphism of vector spaces \(W^\vee \cong \bigwedge ^{n}W\), given by

$$\begin{aligned} \eta ^i\mapsto \eta _1\wedge \ldots \wedge \hat{\eta _i}\wedge \ldots \wedge \eta _{n+1}=:e_i \end{aligned}$$

where \(\eta ^1,\ldots ,\eta ^{n+1}\) is the basis of \(W^\vee \) dual to the basis \(\eta _1,\ldots ,\eta _{n+1}\) of W. By definition of \(\alpha \) we have that \(\alpha ^\vee \) is the evaluation map given by the global sections \(\tilde{\omega }_i\). Note that \(\mathcal {E}^\vee \otimes \mathcal {L}(D_W)\cong \bigwedge ^{n}\mathcal {E}\otimes \mathcal {L}(-F)\). Taking global sections we have

The section \(\Omega '\in H^0(X,\mathcal {E}^\vee \otimes \mathcal {L}(D_W))\) produces in \(H^0(X,\det \mathcal {E})\) the adjoint \(\Omega \), so by commutativity

$$\begin{aligned} \Omega =\overline{\alpha ^\vee }(\overline{\beta ^\vee }(\Omega ')). \end{aligned}$$

We have

$$\begin{aligned} \overline{\beta ^\vee }(\Omega ')=\sum _{i=1}^{n+1}c_i\cdot e_i\otimes \sigma _i, \end{aligned}$$

where \(c_i\in \mathbb {C}\) and \(\sigma _i\in H^0(X,\mathcal {L}(D_W))\). By our hypothesis \(H^0(X,\mathcal {L})\cong H^0(X,\mathcal {L}(D_W))\), there exists sections \(\tilde{\sigma _i}\in H^0(X, \mathcal {L})\) with \(\sigma _i=\tilde{\sigma _i}\cdot d\). So

$$\begin{aligned} \Omega =\overline{\alpha ^\vee }(\overline{\beta ^\vee }(\Omega '))=\overline{\alpha ^\vee }(\sum _{i=1}^{n+1}c_i\cdot e_i\otimes \sigma _i)=\sum _{i=1}^{n+1}c_i\cdot \tilde{\omega }_i\cdot \sigma _i=\sum _{i=1}^{n+1} c_i\cdot \tilde{\omega }_i\cdot d\cdot \tilde{\sigma }_i=\sum _{i=1}^{n+1} c_i\cdot \omega _i\cdot \tilde{\sigma }_i. \end{aligned}$$

This is exactly our thesis. \(\square \)

By the Adjoint Theorem and (2.9) we deduce the following

Corollary 2.10

If \(D_W=0\), then \(\xi =0\) iff \(\Omega \in \mathrm {Im}\,(H^0(X,\mathcal {L})\otimes \lambda ^nW\rightarrow H^0(X,\det \mathcal {E}))\).

3 Infinitesimal Torelli Theorem for projective hypersurfaces

In this section we want to study adjoint images in the case of smooth hypersurfaces of the projective space \(\mathbb {P}^n\).

3.1 Meromorphic 1-forms on a smooth projective hypersurface

Let \(V\subset \mathbb {P}^n\) be a smooth hypersurface defined by a homogeneous polynomial \(F\in \mathbb {C}[\xi _0,\ldots ,\xi _n]\) of degree \(\deg F=d\). An infinitesimal deformation \(\xi \in \text {Ext}^1(\Omega ^1_V,\mathcal {O}_V)\) of V gives an exact sequence for the sheaf of differential forms \(\Omega _V^1\):

$$\begin{aligned} 0\rightarrow \mathcal {O}_V\rightarrow \Omega ^1_{\mathcal {V}|V}\rightarrow \Omega ^1_V\rightarrow 0. \end{aligned}$$
(3.1)

We assume that \(n\ge 3\), hence \(H^0(V,\Omega ^1_V)=0\) and we can not construct the adjoint of this sequence directly, so we twist (3.1) by a suitable integer a such that \(\Omega ^1_V(a)\) has at least \(n=\mathrm {rank}\,(\Omega ^1_V)+1\) global sections. A standard computation shows that \(a=2\) is enough for this purpose, so from now on we will consider the sequence

$$\begin{aligned} 0\rightarrow \mathcal {O}_V(2)\rightarrow \Omega ^1_{\mathcal {V}|V}(2)\rightarrow \Omega ^1_V(2)\rightarrow 0 \end{aligned}$$
(3.2)

which again corresponds to \(\xi \in \text {Ext}^1(\Omega ^1_V(2),\mathcal {O}_V(2))\cong \text {Ext}^1(\Omega ^1_V,\mathcal {O}_V)\cong H^1(V,\Theta _V)\), where \(\Theta _V\) denotes the sheaf of vector fields on V. Denote by \(\mathcal {J}\) the Jacobian ideal of F, that is the ideal of \(\mathbb {C}[\xi _0,\ldots ,\xi _n]\) generated by the partial derivatives \(\frac{\partial {F}}{\partial {\xi _i}}\) for \(i=0,\ldots ,n\). Following [9][Theorem 9.8], the deformation \(\xi \) is given by a class [R] of degree d in the quotient \(\mathbb {C}[\xi _0,\ldots ,\xi _n]/\mathcal {J}\). If we choose a representative R of degree d for this class, then \(F~+~t R=0\), for small t, is the equation of the hypersurface that is the associated deformation of V.

Together with (3.2), we have the conormal exact sequence

$$\begin{aligned} 0\rightarrow \mathcal {O}_V(-d)\rightarrow \Omega ^1_{\mathbb {P}^n|V}\rightarrow \Omega ^1_V\rightarrow 0. \end{aligned}$$
(3.3)

If we put these sequences together we obtain the diagram

which can be completed as follows

(3.4)

By [9] the deformation \(\xi \) of (3.2) comes from \(R\in H^0(\mathbb {P}^n, \mathcal {O}_\mathbb {P}^n(d))\), then it gives the zero element in \(H^0(V,\Theta _{ \mathbb {P}^n |V} )\), hence we have that the sheaf \(\mathcal {G}\) in (3.4) is a direct sum \(\mathcal {G}=\mathcal {O}_V(2)\oplus \Omega ^1_{\mathbb {P}^n|V}(2)\) and we have a natural morphism \(\phi :\Omega ^1_{\mathbb {P}^n|V}(2)\rightarrow \Omega ^1_{\mathcal {V}|V}(2)\) which fits in the diagram

(3.5)

The morphism \(\phi \) gives in a natural way a morphism

$$\begin{aligned}&\phi ^n:H^0(V,\det (\Omega ^1_{\mathbb {P}^n|V}(2)))\cong H^0(V,\mathcal {O}_V(n-1))\rightarrow H^0(V,\det (\Omega ^1_{\mathcal {V}|V}(2)))\\&\quad \cong H^0(V,\mathcal {O}_V(n+d-1)). \end{aligned}$$

We can write explicitly the isomorphism between \(H^0(V,\det (\Omega ^1_{\mathbb {P}^n|V}(2)))=H^0(V,\Omega ^n_{\mathbb {P}^n|V}(2n))\) and \(H^0(V,\mathcal {O}_V(n-1))\). Note that \(H^0(\mathbb {P}^n,\Omega ^n_{\mathbb {P}^n}(2n))\rightarrow H^0(V,\Omega ^n_{\mathbb {P}^n|V}(2n))\) is surjective, so we will focus on the rational n-forms on \(\mathbb {P}^n\). By [9][Corollary 2.11] this forms may be written as \(\omega =\frac{P\Psi }{Q}\) where \(\Psi =\sum _{i=0}^n(-1)^i\xi _i(d\xi _0\wedge \ldots \wedge d\widehat{\xi _i}\wedge \ldots \wedge d\xi _n)\) gives a generator of \(H^0(\mathbb {P}^n,\Omega ^n_{\mathbb {P}^n}(n+1))\) and \(\deg Q=\deg P +(n+1)\). In our case Q is a polynomial of degree 2n, hence P has degree \(n-1\). This identification depends on the (noncanonical) choice of the polynomial Q and gives an isomorphism \(H^0(V,\Omega ^n_{\mathbb {P}^n|V}(2n))\rightarrow H^0(V,\mathcal {O}_V(n-1))\) defined by \(\omega |_V\mapsto P\).

Proposition 3.1

\(\phi ^n\) is given via the multiplication by the polynomial R (modulo F).

Proof

Locally we can see \(\mathcal {V}\) in the product \(\Delta \times \mathbb {P}^n\) of the projective space with a disk; here \(\mathcal {V}\) is defined by the equation \(F+tR=0\). Hence \(d(F+tR)=0\) in \(\Omega ^1_{\mathcal {V}}\), that is \(dF+dt\cdot R+dR\cdot t=0\).

Call \(F_i:=\frac{\partial F}{\partial \xi _i}\). Since V is smooth, there exist i such that \(U_i=(F_i\ne 0)\) is a nontrivial open subset; let for example \(U_1\) be nontrivial. Take local coordinates \(z_i=\frac{\xi _i}{\xi _0}\) in the open set \((\xi _0\ne 0)\cap U_1\). Then we have

$$\begin{aligned} d z_1=-\frac{Rdt}{F_1}-\frac{tdR}{F_1}-\sum _{i>1}\frac{F_i}{F_1}d z_i \end{aligned}$$
(3.6)

which gives in V (that is for \(t=0\))

$$\begin{aligned} d z_1=-\frac{Rdt}{F_1}-\sum _{i>1}\frac{F_i}{F_1}d z_i \end{aligned}$$
(3.7)

The image \(\phi ^n(\omega |_V)\) is then obtained by the substitution of (3.7) in \(\frac{P(z)}{Q(z)}dz_1\wedge \ldots \wedge dz_n\), which is the local form of \(\frac{P(\xi )\Psi }{Q(\xi )}\). Hence

$$\begin{aligned} \frac{P(z)}{Q(z)}dz_1\wedge \ldots \wedge dz_n=-\frac{P(z) R(z)}{Q(z) F_1(z)}dt\wedge dz_2\wedge \ldots \wedge dz_n. \end{aligned}$$
(3.8)

If we homogenize we obtain on \(U_1\)

$$\begin{aligned} \frac{P\Psi }{Q}=-\frac{PR}{QF_1}\sum _{i\ne 1}(-1)^{i-1}\text {sgn}(i-1)\xi _idt\wedge d\xi _0\wedge \widehat{d\xi _1}\ldots \wedge \widehat{d\xi _i}\wedge \ldots \wedge d\xi _n \end{aligned}$$

Hence

$$\begin{aligned} \phi ^n(\omega |_V)=-\frac{PR}{QF_1}\sum _{i\ne 1}(-1)^{i-1}\text {sgn}(i-1)\xi _idt\wedge d\xi _0\wedge \widehat{d\xi _1}\ldots \wedge \widehat{d\xi _i}\wedge \ldots \wedge d\xi _n \end{aligned}$$
(3.9)

and it is clear that \(\phi ^n\) is given by multiplication with R. \(\square \)

3.2 A canonical choice of adjoints on a hypersurface of degree \(d>2\)

We want now to construct adjoint forms associated to the sequence (3.2).

Assume that \(n\ge 3\), so that \(H^1(V,\mathcal {O}_V(2))=H^1(V,\mathcal {O}_V(2-d))=0\), and we can lift all the global sections of \(H^0(V,\Omega ^1_V(2))\) both in the horizontal and in the vertical sequence of (3.5).

We take \(\eta _1,\ldots ,\eta _n\in H^0(V,\Omega ^1_V(2))\) global forms and we want to find liftings \(s_1,\ldots ,s_n\in H^0(V,\Omega ^1_{\mathcal {V}|V})\). This can be done since \(H^1(V,\mathcal {O}_V(2))\) is zero. A generalized adjoint is then the global section of the sheaf \(\det (\Omega ^1_{\mathcal {V}|V}(2))=\mathcal {O}_V(n+d-1)\) given by \(\Omega :=\Lambda ^n(s_1\wedge \ldots \wedge s_n)\in H^0(V,\det (\Omega ^1_{\mathcal {V}|V}(2)))\).

We point out another interesting way to compute this generalized adjoint form using Proposition (3.1).

Consider the sequence (3.3), that is the vertical sequence in (3.5). Since \(H^1(V,\mathcal {O}_V(2-d))=0\), we can find liftings \(\tilde{s_1},\ldots ,\tilde{s_n}\in H^0(V, \Omega ^1_{\mathbb {P}^n|V}(2))\) of the sections \(\eta _1,\ldots ,\eta _n\). Furthermore they are unique if \(d>2\). We can thus consider the adjoint form associated to (3.3) given by \(\widetilde{\Omega }:=\Lambda ^n (\tilde{s_1}\wedge \ldots \wedge \tilde{s_n})\). This adjoint is independent from the deformation \(\xi \); it depends only on V and its embedding in \(\mathbb {P}^n\). If \(d>2\), then \(\widetilde{\Omega }\) is unique.

To describe \(\widetilde{\Omega }\) explicitly we first consider the exact sequence

$$\begin{aligned} 0\rightarrow \Omega ^1_{\mathbb {P}^n}(2-d)\rightarrow \Omega ^1_{\mathbb {P}^n}(2)\rightarrow \Omega ^1_{\mathbb {P}^n|V}(2)\rightarrow 0. \end{aligned}$$
(3.10)

If \(d>2\), the vanishing of \(H^0(\mathbb {P}^n,\Omega ^1_{\mathbb {P}^n}(2-d))\) and \(H^1(\mathbb {P}^n,\Omega ^1_{\mathbb {P}^n}(2-d))\) (c.f. Bott Formulas) gives the isomorphism \(H^0(\mathbb {P}^n,\Omega ^1_{\mathbb {P}^n}(2))=H^0(V,\Omega ^1_{\mathbb {P}^n|V}(2))\). Hence, the forms \(\tilde{s_i}\) are the restriction on V of global rational 1-forms. By [9][Theorem 2.9] we can write

$$\begin{aligned} \tilde{s_i}=\frac{1}{Q}\sum _{j=0}^n L^i_j d\xi _j \end{aligned}$$
(3.11)

where \(\deg Q=2\) and \(L^i_j\) is a homogeneous polynomial of degree 1 which does not contain \(\xi _j\) in its expression. Hence

$$\begin{aligned} \widetilde{\Omega }= \Lambda ^n(\tilde{s_1}\wedge \ldots \wedge \tilde{s_n})=\frac{1}{Q^n}\sum _{i=0}^n M_i d\xi _0\wedge \ldots \wedge \widehat{d\xi _i}\wedge \ldots \wedge d\xi _n \end{aligned}$$
(3.12)

where \(M_i\) is the determinant of the matrix obtained by

$$\begin{aligned} \begin{pmatrix} L^1_0&{}\ldots &{}L^n_0\\ \vdots &{}&{}\vdots \\ L^1_n&{}\ldots &{}L^n_n \end{pmatrix} \end{aligned}$$
(3.13)

removing the i-th row. Since \(\widetilde{\Omega }\) is a rational n-form on \(\mathbb {P}^n\), following [9][Corollary 2.11] it can be written as \(\frac{P\Psi }{Q^n}\), and we deduce that

$$\begin{aligned} \frac{M_i}{(-1)^i\xi _i}=P \end{aligned}$$
(3.14)

for all \(i=0,\ldots ,n\). P is a polynomial of degree \(n-1\) and it corresponds to \(\widetilde{\Omega }\) via the isomorphism \(H^0(V,\Omega ^n_{\mathbb {P}^n|V}(2n))\cong H^0(V,\mathcal {O}_V(n-1))\). Hence by (3.1) we have that the form \(\Omega \in H^0(V,\mathcal {O}_V(n+d-1))\) given by PR is a canonical choice of adjoint form for \(W=\langle \eta _1,\ldots , \eta _n\rangle \) and \(\xi \).

Remark 3.2

Alternatively this can be seen using the Euler sequence on V:

$$\begin{aligned} 0\rightarrow \mathcal {O}_V\rightarrow \bigoplus ^{n+1}\mathcal {O}_V(1)\rightarrow \Theta _{\mathbb {P}^n|V}\rightarrow 0. \end{aligned}$$
(3.15)

This sequence, dualized and conveniently tensorized gives

$$\begin{aligned} 0\rightarrow \Omega _{\mathbb {P}^n|V}^1(2)\rightarrow \bigoplus _{i=1}^{n+1}\mathcal {O}_V(1)\rightarrow \mathcal {O}_{V}(2)\rightarrow 0. \end{aligned}$$
(3.16)

The sections \(\tilde{s}_i\) are associated via the first morphism to an \(n+1\)-uple of linear polynomials \((L_i^0,\ldots ,L_i^n)\). Then, taking the wedge product of (3.16) we obtain an exact sequence

$$\begin{aligned} 0\rightarrow \Omega ^n_{\mathbb {P}^n|V}(2n)\cong \mathcal {O}_V(n-1)\rightarrow \bigwedge ^n\mathcal {O}_V(1)=\bigoplus ^{n+1}\mathcal {O}_V(n)\rightarrow \Omega ^{n-1}_{\mathbb {P}^n|V}(2n)\rightarrow 0 \end{aligned}$$
(3.17)

where the morphism \(\mathcal {O}_V(n-1)\rightarrow \bigoplus ^{n+1}\mathcal {O}_V(n)\) is given by

$$\begin{aligned} G\mapsto (G\xi _0,\ldots ,(-1)^nG\xi _n). \end{aligned}$$
(3.18)

Since \(\widetilde{\Omega }=\Lambda ^n (\tilde{s_1}\wedge \ldots \wedge \tilde{s_n})\in H^0(V,\Omega ^n_{\mathbb {P}^n|V}(2n))\) is sent exactly to \((L_0^0,\ldots ,L_0^n)\wedge \ldots \wedge (L_n^0,\ldots ,L_n^n)=(M_0,\ldots ,M_n)\) (using the same notation as above), then we conclude that \(\widetilde{\Omega }\) corresponds in \(H^0(V,\mathcal {O}_V(n-1))\) to a polynomial P which satisfies

$$\begin{aligned} \frac{M_i}{(-1)^i\xi _i}=P. \end{aligned}$$
(3.19)

3.3 The adjoint sublinear systems obtained by meromorphic 1-forms

To study the conditions given in (2.6) and (2.7), we need to describe the sections

$$\begin{aligned} \widetilde{\Omega _i}:=\Lambda ^{n-1}(\tilde{s_1}\wedge \ldots \wedge \hat{\tilde{s_i}}\wedge \ldots \wedge \tilde{s_n})\in H^0(V,\Omega ^{n-1}_{\mathbb {P}^n|V}(2n-2)) \end{aligned}$$

(c.f. (2.4)) and their images in \(H^0(V,\Omega ^{n-1}_{V}(2(n-1)))=H^0(V,\mathcal {O}_V(n+d-3))\) that we have denoted by \(\omega _i\).

A computation similar to the above shows that

$$\begin{aligned} \widetilde{\Omega _i}= \Lambda ^{n-1}(\tilde{s_1}\wedge \ldots \wedge \hat{\tilde{s_i}}\wedge \ldots \wedge \tilde{s_n})=\frac{1}{Q^{n-1}}\sum _{j<k}M^i_{jk}d\xi _0\wedge \ldots \wedge \hat{d\xi _j}\wedge \ldots \wedge \hat{d\xi _k}\wedge \ldots \wedge d\xi _n \end{aligned}$$
(3.20)

where \(M^i_{jk}\) is the determinant of the matrix obtained by (3.13) removing the i-th column and the j-th and k-th rows. On the other hand, rearranging the expression of [9][Theorem 2.9] we can write

$$\begin{aligned} \widetilde{\Omega _i}=\frac{1}{Q^{n-1}}\sum _{j} A^i_j(\sum _{k\ne j} (-1)^{k+j}\text {sgn}(k-j)\xi _k d\xi _0\wedge \ldots \wedge \hat{d\xi _j}\wedge \ldots \wedge \hat{d\xi _k}\wedge \ldots \wedge d\xi _n) \end{aligned}$$
(3.21)

with \(\deg A^i_j=n-2\).

Comparing (3.20) and (3.21) gives

$$\begin{aligned} M^i_{jk}=(-1)^{j+k}(A^i_j\xi _k-\xi _j A_k^i). \end{aligned}$$
(3.22)

As before this can be computed also via the Euler sequence.

We call \(\Xi _j:=\sum _{k\ne j} (-1)^{k+j}\text {sgn}(k-j)\xi _k d\xi _0\wedge \ldots \wedge \hat{d\xi _j}\wedge \ldots \wedge \hat{d\xi _k}\wedge \ldots \wedge d\xi _n\). Note that the sections \(\Xi _j\), for \(j=0,\ldots ,n\) give a basis of \(H^0(V,\Omega ^{n-1}_{\mathbb {P}^n|V}(n))\).

Proposition 3.3

\(\omega _i=\sum _{j} A^i_j\cdot F_j\) in \(H^0(V,\mathcal {O}_V(n+d-3))\)

Proof

It is enough to show that the image of \(\Xi _j\) through the morphism \(\Omega ^{n-1}_{\mathbb {P}^n|V}(n)\rightarrow \mathcal {O}_V(d-1)\) is \(F_j\). Consider the exact sequence of the tangent sheaf of V:

$$\begin{aligned} 0\rightarrow \Theta _V\rightarrow \Theta _{\mathbb {P}^n|V}\rightarrow \mathcal {O}_V(d)\rightarrow 0. \end{aligned}$$
(3.23)

The beginning of the Koszul complex is

$$\begin{aligned} \bigwedge ^n\Theta _{\mathbb {P}^n|V}\otimes \mathcal {O}_V(-d)\rightarrow \bigwedge ^{n-1}\Theta _{\mathbb {P}^n|V} \end{aligned}$$
(3.24)

which, tensored by \(\mathcal {O}_V(-n)\), gives

$$\begin{aligned} \bigwedge ^n\Theta _{\mathbb {P}^n|V}\otimes \mathcal {O}_V(-n-d)\rightarrow \bigwedge ^{n-1}\Theta _{\mathbb {P}^n|V}\otimes \mathcal {O}_V(-n). \end{aligned}$$
(3.25)

This is exactly the dual of \(\Omega ^{n-1}_{\mathbb {P}^n|V}(n)\rightarrow \mathcal {O}_V(d-1)\). Hence we only need to show that the morphism (3.25) composed with the contraction by \(\Xi _i\)

$$\begin{aligned} \bigwedge ^{n-1}\Theta _{\mathbb {P}^n|V}\otimes \mathcal {O}_V(-n)\mathop {\rightarrow }\limits ^{\Xi _i}\mathcal {O}_V \end{aligned}$$
(3.26)

is the multiplication by \(F_i\). This is easy to see by a standard local computation. \(\square \)

Remark 3.4

We immediately have that the polynomials associated to the sections \(\omega _i\) are in the Jacobian ideal of V.

Condition (2.7), that is

$$\begin{aligned} \Omega \in \mathrm {Im}\,(H^0(V,\mathcal {O}_V(2))\otimes \lambda ^nW \rightarrow H^0(V,\mathcal {O}_V(n+d-1))), \end{aligned}$$
(3.27)

can be written, modulo F, as

$$\begin{aligned} RP=\sum \omega _i\cdot S_i=\sum _{i,j} A^i_j\cdot F_j\cdot S_i, \end{aligned}$$
(3.28)

where \(\deg S_i=2\). In particular this implies that RP is in the Jacobian ideal of V.

Proposition 3.5

The base locus \(D_W\) of the linear system \(|\lambda ^nW|\) is zero for the generic W.

Proof

By [11][Proposition 3.1.6] it is enough to prove that \(H^0(V,\Omega ^1_V(2))\) generically generates the sheaf \(\Omega ^1_V(2)\) and that \(D_{H^0(V,\Omega ^1_V(2))}=0\). We have an explicit basis for \(H^0(V,\Omega ^1_V(2))\) given by

$$\begin{aligned} \frac{\xi _id\xi _j-\xi _jd\xi _i}{Q} \end{aligned}$$
(3.29)

where \(i<j\) and \(\deg Q=2\). The vector space \(\lambda ^nH^0(V,\Omega ^1_V(2))\subset H^0(V,\mathcal {O}_V(n+d-3))\) is obviously nonzero, hence \(H^0(V,\Omega ^1_V(2))\) generically generates the sheaf \(\Omega ^1_V(2)\).

It remains to prove that \(D_{H^0(V,\Omega ^1_V(2))}=0\). An easy computation (for example by induction) shows that \(\lambda ^nH^0(V,\Omega ^1_V(2))\) contains all the polynomials of the form

$$\begin{aligned} \xi _{i_{1}}\xi _{i_{2}}\ldots \xi _{i_{n-2}} \frac{\partial F}{\partial {\xi _j}} \end{aligned}$$
(3.30)

where \(\{i_1,\ldots ,i_{n-2}\}\subset \{1,\ldots ,n+1\}\) and \(j\notin \{i_1,\ldots ,i_{n-2}\}\). Since V is smooth, these polynomials do not vanish simultaneously on a divisor, hence \(D_{H^0(V,\Omega ^1_V(2))}=0\), and we are done. \(\square \)

3.4 On Griffiths’s proof of infinitesimal Torelli Theorem

In this section we will prove Theorem [C] of the Introduction.

It is well known by [9] that the deformation \(\xi \) is trivial if and only if R lies in the Jacobian ideal \(\mathcal {J}\) of the variety V. The following lemma gives a translation of this condition in the setting of adjoint forms.

Lemma 3.6

R is in the Jacobian ideal \(\mathcal {J}\) if and only if \(\Omega \in \mathrm {Im}\,(H^0(V,\mathcal {O}_V(2))\otimes \lambda ^nW \rightarrow H^0(V,\mathcal {O}_V(n+d-1)))\) for the generic \(\Omega \).

Proof

If \(\Omega \in \mathrm {Im}\,(H^0(V,\mathcal {O}_V(2))\otimes \lambda ^nW \rightarrow H^0(V,\mathcal {O}_V(n+d-1)))\), then by the Adjoint Theorem, \(\xi _{D_W}=0\). Since \(D_W=0\), the deformation is trivial, hence R lies in the Jacobian ideal.

Viceversa if \(R\in \mathcal {J}\), the deformation is trivial and by theorem (2.9), we have that \(\Omega \in \mathrm {Im}\,(H^0(V,\mathcal {O}_V(2))\otimes \lambda ^nW \rightarrow H^0(V,\mathcal {O}_V(n+d-1)))\). \(\square \)

Our theory gives another characterization for \([R]\in ( \mathbb {C}[\xi _0,\ldots ,\xi _n]/\mathcal {J})_{d}\simeq H^1(V,\Theta _V)\) to be trivial.

Proposition 3.7

Assume that \(\deg R=d>3\). Then R is in the Jacobian ideal \(\mathcal {J}\) if and only if \(RP\in \mathcal {J}\) for every polynomial \(P\in H^0(V,\mathcal {O}_V(n-1))\) corresponding to a generalized adjoint \(\widetilde{\Omega }\in H^0(V,\Omega ^n_{\mathbb {P}^n|V}(2n))\).

Proof

One implication is trivial.

To prove the other one the idea is to show that every monomial of \(H^0(V,\mathcal {O}_V(n-1))\) corresponds to a suitable generalized adjoint. Hence, if \(RP\in \mathcal {J}\) for every polynomial \(P\in H^0(V,\mathcal {O}_V(n-1))\) corresponding to a generalized adjoint, we have that \(R\cdot H^0(V,\mathcal {O}_V(n-1))\subset \mathcal {J}\) and we are done by Macaulay Theorem (c.f. [16] Theorem 6.19 and Corollary 6.20).

We work by induction at the level of \(\mathbb {P}^n\), since \(H^0(\mathbb {P}^n,\mathcal {O}_{\mathbb {P}^n}(n-1))\rightarrow H^0(V,\mathcal {O}_V(n-1))\) is surjective. The base of the induction is for \(n=2\). A simple computation shows that the map

$$\begin{aligned} \bigwedge ^2H^0(\mathbb {P}^2,\Omega ^1_{\mathbb {P}^2}(2))\rightarrow H^0(\mathbb {P}^2,\mathcal {O}_{\mathbb {P}^2}(1)) \end{aligned}$$
(3.31)

is surjective because its image contains the canonical basis of degree one monomials.

For the general case we show that every monomial of degree \(n-1\) is given by a generalized adjoint. Consider the natural homomorphism:

$$\begin{aligned} \bigwedge ^nH^0(\mathbb {P}^n,\Omega ^1_{\mathbb {P}^n}(2))\rightarrow H^0(\mathbb {P}^n,\mathcal {O}_{\mathbb {P}^n}(n-1)) \end{aligned}$$
(3.32)

and take a monomial M with \(\deg M=n-1\). There is a variable \(\xi _i\) which does not appear in M. We restrict to the hyperplane \(\xi _i=0\) and we use induction on \(\frac{M}{\xi _j}\), where \(\xi _j\) appears in M. There exist \(s_1,\ldots ,s_{n-1}\in H^0(\mathbb {P}^{n-1},\Omega ^1_{\mathbb {P}^{n-1}}(2))\) with \(s_1\wedge \ldots \wedge s_{n-1}\) which corresponds to \(\frac{M}{\xi _j}\), that is

$$\begin{aligned} s_1\wedge \ldots \wedge s_{n-1}=\frac{M\Psi '}{\xi _j\cdot Q^{n-1}} \end{aligned}$$
(3.33)

where \(\Psi '=\sum _{k=0,k\ne i}^n(-1)^k\xi _k(d\xi _0\wedge \ldots \wedge \hat{d\xi _i}\ldots \wedge \hat{d\xi _k}\ldots \wedge d\xi _n)\) gives a basis of \(H^0(\mathbb {P}^{n-1},\Omega _{\mathbb {P}^{n-1}}^{n-1}(n))\). It is easy to see that

$$\begin{aligned} s_1\wedge \ldots \wedge s_{n-1}\wedge \frac{(\xi _jd\xi _i-\xi _id\xi _j)}{Q}=\frac{M\Psi }{Q^{n}}, \end{aligned}$$
(3.34)

i.e. M corresponds to a generalized adjoint, which is exactly our thesis. \(\square \)

From the previous results we deduce immediately Theorem [C] of the Introduction.