Introduction
We will consider in this paper smooth hypersurfaces $X_f\subset \mathbb {P}^{n}$ of degree
$d$ defined by a homogeneous polynomial equation
$f$. By the Lefschetz theorem on hyperplane sections, only the degree
$n-1$ cohomology group
$H^{n-1}(X_f,\mathbb {Z})$ carries a nontrivial Hodge structure, and its primitive part
$H^{n-1}(X_f,\mathbb {Z})_{\rm prim}:={\rm Ker}\,(H^{n-1}(X_f,\mathbb {Z})\rightarrow H^{n+1}(\mathbb {P}^{n},\mathbb {Z}))$ carries a Hodge structure polarized by the cup-product
$\langle \,,\,\rangle _X$ (the two groups agree when
$n-1$ is odd, otherwise they differ by
$\mathbb {Z}h^{{(n-1)}/{2}}$, where
$h=c_1(\mathcal {O}_X(1))$). The global Torelli problem for hypersurfaces thus asks whether the existence of an isomorphism of polarized Hodge structures
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU1.png?pub-status=live)
(extending to an isomorphism of Hodge structures $H^{n-1}(X_f,\mathbb {Z})\cong H^{n-1}(X_{f'},\mathbb {Z})$ preserving the classes
$h^{{(n-1)}/{2}}$ on both sides when
$n-1$ is even) implies that
$X_f\cong X_{f'}$. There are very few cases where this statement is known. For plane curves, we can apply the Torelli theorem for curves. For quartic surfaces, the global Torelli theorem is proved by Piateski-Shapiro and Shafarevich [Reference Piateski-Shapiro and ShafarevichPS71]. For cubic threefolds, the global Torelli theorem is proved by Clemens and Griffiths [Reference Clemens and GriffithsCG72] and Beauville [Reference BeauvilleBea82], and for cubic fourfolds it was first proved in [Reference VoisinVoi86] (alternative proofs are now available, see e.g. [Reference Huybrechts and RennemoHR19]).
The generic Torelli theorem for hypersurfaces of degree $d$ and dimension
$n-1$ is the following statement that we will study in this paper.
Let $X_f$ be a very general smooth hypersurface of degree
$d$ in
$\mathbb {P}^{n}$. Then any smooth hypersurface
$X_{f'}$ of degree
$d$ in
$\mathbb {P}^{n}$ such that there exists an isomorphism of Hodge structures
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn1.png?pub-status=live)
is isomorphic to $X_f$.
We will explain in § 1.1 why the ‘very general’ assumption is natural in this statement. This is related to the Cattani–Deligne–Kaplan theorem [Reference Cattani, Deligne and KaplanCDK95] which implies that the set of pairs $(f,\,{f'})$ such that an isomorphism as in (1) exists is a countable union of closed algebraic subsets in
$U_{d,n}\times U_{d,n}$, where
$U_{d,n}$ is the moduli space of smooth hypersurfaces of degree
$d$ in
$\mathbb {P}^{n}$.
Remark 0.1 The Torelli theorem is usually stated for the isomorphisms of polarized Hodge structures. However, using infinitesimal arguments (see [Reference VoisinVoi03, 6.3.1]) we can see that, except in the case of cubic surfaces, the Hodge structure on the primitive cohomology of $X_f$ has a unique polarization (up to a scalar) for very general
$X_f$. So the polarized and nonpolarized statements are equivalent.
In the case of cubic surfaces, the generic Torelli theorem is clearly wrong, since they have moduli, while their variation of Hodge structure is trivial. The case of plane quartics is also a counterexample to the generic Torelli theorem with rational coefficients, since in genus $3$, a general curve is not determined by the isogeny class of its Jacobian. Donagi proved in [Reference DonagiDon83] the following beautiful result.
Theorem 0.2 The generic Torelli theorem holds for smooth hypersurfaces of degree $d$ in
$\mathbb {P}^{n}$, with
$n\geq 3$,
$(d,n)\not =(3,3)$ and the following possible exceptions:
(1)
$d$ divides
$n+1$;
(2)
$d=4$,
$n=4m+1$, with
$m\geq 1$;
(3)
$d=6$,
$n=6m+2$, with
$m\geq 1$.
Remark 0.3 The Torelli theorem is usually stated for integral Hodge structures, and Donagi's original statement indeed concerned integral Hodge structures. In fact, his proof works as well for rational Hodge structures, since it relies on the study of the (complex!) variation of Hodge structure for hypersurfaces of given degree and dimension and its local invariants. Another instance where a generic Torelli theorem has been proved for rational Hodge structures is the case of curves of genus $g\geq 4$ which is treated in [Reference Bardelli and PirolaBP89]. In this case, Bardelli and Pirola prove that a very general curve of genus at least
$4$ is determined by the isogeny class of its Jacobian.
Remark 0.4 The Cattani–Deligne–Kaplan algebraicity theorem mentioned above appeared much later than [Reference DonagiDon83], so that Theorem 0.2 is in fact a slightly strengthened version of Donagi's theorem, taking into account [Reference Cattani, Deligne and KaplanCDK95].
Cox and Green solved in [Reference Cox and GreenCG90] the case (3), that is $d=6$, but the two infinite series (1) and (2) essentially remained open. The starting point of Donagi's proof is the description due to Griffiths [Reference GriffithsGri69] and Carlson and Griffiths [Reference Carlson, Griffiths and BeauvilleCG80] of the infinitesimal variation of Hodge structure of a smooth hypersurface. Denote by
$S^{*} =\mathbb {C}[X_0,\ldots,X_n]$ the graded polynomial ring of
$\mathbb {P}^{n}$ and by
$R_f^{*}=S^{*}/J_f^{*}$ the Jacobian ring of
$f$, where
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn2.png?pub-status=live)
is the Jacobian ideal of $f$, generated by the partial derivatives of
$f$. The infinitesimal variation of Hodge structure on the primitive cohomology of degree
$n-1$ of
$X_f$ is given, according to Griffiths [Reference GriffithsGri69], see also [Reference VoisinVoi03, 6.1.3], by linear maps
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn3.png?pub-status=live)
for $p+q=n-1$. Here, the space
$R^{d}_f$ is naturally identified with the first-order deformations of
$X_f$ in
$\mathbb {P}^{n}$ modulo the infinitesimal action of
${\rm PGL}(n+1)$. It also identifies via the Kodaira–Spencer map to the subspace
$H^{1}(X_f,T_{X_f})_0\subset H^{1}(X_f,T_{X_f})$ of deformations of
$X_f$ induced by a deformation of
$f$. Griffiths constructs residue isomorphisms
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn4.png?pub-status=live)
and the paper [Reference Carlson, Griffiths and BeauvilleCG80] in turn describes (3) using the isomorphisms (4) as follows.
Theorem 0.5 Via the isomorphisms (4), the maps (3) identify up to a scalar coefficient with the map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn5.png?pub-status=live)
induced by multiplication in $R_f^{*}$. In other words, the following diagram is commutative up to a coefficient.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn6.png?pub-status=live)
Furthermore, the Serre pairing between $H^{n-1-q,q}(X_f)_{{\rm prim}}$ and
$H^{q,n-1-q}(X_f)_{{\rm prim}}$ identifies with the Macaulay pairing
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn7.png?pub-status=live)
given by the product in $R_f^{*}$.
Donagi's proof starts with the observation that Theorem 0.2 is implied by the following result.
Theorem 0.6 Let $X$ be a smooth hypersurface of degree
$d$ in
$\mathbb {P}^{n}$, with
$n\geq 3$. Assume
$(d,n)\not =(3,3)$ and we are not in the cases (1), (2), (3) listed in Theorem 0.2. Then
$X$ is determined by the data (5) for all
$q$ and the Macaulay pairings (7), hence, using Theorem 0.5, by its polarized infinitesimal variation of Hodge structure.
Concretely, Theorem 0.6 says that if $X_f$ and
$X_{f'}$ are two smooth hypersurfaces of degree
$d$ and dimension
$n-1$ such that there exist isomorphisms,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU2.png?pub-status=live)
compatible with the Macaulay pairing (7) for $f$ and
$f'$, and such that the following diagram commutes,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn8.png?pub-status=live)
then $X_f$ is isomorphic to
$X_{f'}$.
Donagi's proof of Theorem 0.6 consists in recovering from the data (5) its polynomial structure (see § 1), and more precisely, reconstructing the whole Jacobian ring of $f$ from its partial data appearing in (5). He then applies the Mather–Yau theorem (see Proposition 3.2) which says that
$f$ is determined by
$J_f^{d-1}\subset S^{d-1}$.
Donagi's method does not work in the case where $(d,n)=(4,3)$, that is, quartic
$K3$ surfaces, because Theorem 0.6 is clearly wrong in this case. In fact, Theorem 0.2 is also wrong for quartic surfaces, due to the fact that it is stated for rational Hodge structures. More generally, Donagi's method to recover the polynomial structure, based on the use of the symmetrizer lemma (Proposition 1.1), gives nothing more, when
$d$ divides
$n+1$, than the subring
$R_f^{*d}\subset R^{*}_f$ defined as the sum of the graded pieces of
$R_f^{*}$ of degree divisible by
$d$. This is why Donagi's method fails to give the result in that case. The goal of this paper is to extend Theorem 0.2 to most families of hypersurfaces not covered by Donagi's theorem.
Theorem 0.7
(1) The generic Torelli theorem holds for smooth hypersurfaces of degree
$d$ in
$\mathbb {P}^{n}$ when
$d$ divides
$n+1$ and
$d$ is large enough. In particular, it holds for Calabi–Yau hypersurfaces of degree
$d$ large enough.
(2) The generic Torelli theorem holds for smooth hypersurfaces of degree
$4$ in
$\mathbb {P}^{4m+1}$ for
$m$ sufficiently large.
These results combined with Donagi's theorem (Theorem 0.2) and Cox and Green's result in [Reference Cox and GreenCG90] imply the following result.
Corollary 0.8 The generic Torelli theorem holds for hypersurfaces of degree $d$ in
$\mathbb {P}^{n}$ with finitely many exceptions.
The proof of Theorem 0.7(2) will be given in § 2. We will give there an effective estimate for $m$, which can probably be improved by refining the method. In that case, the method of proof follows closely Donagi's ideas, and in particular passes through a proof of Theorem 0.6, at least for
$X$ generic.
Theorem 0.7(1) had been also proved in [Reference Voisin, Hulek, Catanese, Peters and ReidVoi99] in the case of quintic threefolds, the first case which is not covered by Theorem 0.2, by extending Theorem 0.6 to that case. It is quite possible that Theorem 0.6 is true more generally when $d$ divides
$n+1$ and
$d$ is sufficiently large, but the proof given in [Reference Voisin, Hulek, Catanese, Peters and ReidVoi99] is very technical and ad hoc, hence not encouraging.
Our proof of Theorem 0.7(1) also rests on the algebraic analysis of the finite-order variation of Hodge structure, but it does not pass through a proof of Theorem 0.6. Theorem 0.6 tells that for the given pairs $(d,n)$, a hypersurface of degree
$d$ and dimension
$n-1$ can be reconstructed from its first order variation of Hodge structure. Instead, our proof will involve the higher-order variation of Hodge structure.
We introduce in this paper a main new ingredient, which is the notion of Schiffer variation of a hypersurface (see § 3). These Schiffer variations are of the form
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn9.png?pub-status=live)
(up to a change of variable $t$) and we believe they are interesting for their own. The chosen terminology comes from the notion of Schiffer variations for a smooth curve
$C$. They consist in deforming the complex structure of
$C$ in a way that is supported on a point
$p$ of
$C$. First-order Schiffer variations are in that case the elements
$u_p\in \mathbb {P}(H^{1}(C,T_C))$ given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU3.png?pub-status=live)
First-order Schiffer variations (9) of hypersurfaces $X_f$ are the tangent directions at
$0$ of Schiffer variations of
$f$. They are parameterized by the
$d$th Veronese embedding of
$\mathbb {P}(S^{1})$ in
$\mathbb {P}(S^{d})$ projected to
$\mathbb {P}(R^{d}_f)$ via the linear projection
$\mathbb {P}(S^{d})\dashrightarrow \mathbb {P}(R^{d}_f)$. Although the following result is easy to prove, it is crucial for our strategy.
Proposition 0.9 (cf. Proposition 3.6)
Let $X_f,\,X_g$ be two smooth hypersurfaces of degree
${d\geq 4}$ and dimension
$n-1\geq 3$, with
$f$ generic. If there exists a linear isomorphism
$R_f^{d}\cong R_g^{d}$ mapping the set of first order Schiffer variations of
$f$ to the set of first-order Schiffer variations of
$g$,
$X_f$ is isomorphic to
$X_g$.
Our strategy then consists in characterizing Schiffer variations by the formal properties of the variation of Hodge structure along them. An obvious but key point (see Lemma 3.8) is the fact that the structure of the Jacobian ring (hence of the infinitesimal variation of Hodge structure) does not change much along them. This follows from the fact that the Jacobian ideals of $f$ and
$f_t=f+tx^{d}$ agree modulo the ideal generated by
$x^{d-1}$. This is a higher-order property since it concerns the variation of the infinitesimal variation of Hodge structures. It would be nice to have a better understanding and a more Hodge-theoretic, less formal, characterization of Schiffer variations.
Remark 0.10 Schiffer variations have very special first-order properties mentioned above, as their tangent vector at any point lies in the Veronese variety. However, crucial to our argument is the fact that they also satisfy higher-order conditions, saying that the hypersurface in $\mathbb {P}^{n-1}$ defined by
$f_{\mid x=0}$ is constant, independent of
$t$.
Our main result can be rephrased as follows.
Theorem 0.11 (cf. Claim 4.18)
Let $d$ be large enough, and
$n\geq d$. Let
$f\in U_{d,n}$ be generic,
$U\subset U_{d,n}$,
$V\subset U_{d,n}$ be Euclidean open sets with
$f\in U$, and let
$i:U\cong V$ be a holomorphic diffeomorphism inducing an isomorphism of complex variations of Hodge structures
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU4.png?pub-status=live)
on $U$. Then for a Schiffer variation
$(f_t)_{t\in \Delta }$ of
$f$ contained in
$U$, where
$\Delta$ is a disc, with tangent vector
$\phi =({df_t}/{dt})_{\mid t=0}\in T_{U,f}$,
$i_*\phi$ is a first order Schiffer variation of
$f':=i(f)$.
This theorem easily implies Theorem 0.7(1) using Proposition 0.9. With more work, it could be improved in two ways.
(1) Under the same assumptions as above, for a general Schiffer variation
$(f_t)_{t\in \Delta }$ of
$f$,
$(i(f_t))_{t\in \Delta }$ is a Schiffer variation of
$f'=i(f)$.
(2) One should be able to replace the Schiffer variation
$(f_t)_{t\in \Delta }$ of
$f$ parameterized by a disc by a second-order Schiffer variation of
$f$.
The paper is organized as follows. In § 1, we first explain (see § 1.1) how Theorem 0.6 implies Theorem 0.2 and we next discuss the notion of polynomial structure on the data of an infinitesimal variation of Hodge structure of a hypersurface. We discuss various recipes toward proving uniqueness of the polynomial structure, including Donagi's method. For example, we exhibit a very simple recipe to show that the natural polynomial structure for most hypersurfaces of degree $d$ dividing
$n+1$ is rigid. In § 2, we prove Theorem 0.7(2), that is, degree 4. This proof follows Donagi's argument but provides a different recipe to prove the uniqueness of the polynomial structure of the infinitesimal variation of Hodge structure in these cases.
The main new ideas and results of the paper appear starting from § 3 where we introduce Schiffer variations of hypersurfaces and discuss their formal properties. The proof of Theorem 0.7(1) is given in § 4.2, where we give a characterization of Schiffer variations based on the local analysis of the variation of Hodge structure along the corresponding family of hypersurfaces and prove Theorem 0.11 (see Proposition 4.14 and Claim 4.18).
1. Polynomial structure and the Torelli theorem
1.1 Donagi's strategy and reduction to Theorem 0.6
For completeness, and because this argument will be also used in the last section, we explain in this section how Theorem 0.6 implies Theorem 0.2. Assume $f\in U_{d,n}$ is very general and
$X_{f'}$ is a smooth hypersurface of degree
$d$ and dimension
$n-1$ with polarized Hodge structure on
$H^{n-1}(X_{f'},\mathbb {Q})_{{\rm prim}}$, isomorphic to the Hodge structure on
$H^{n-1}(X_f,\mathbb {Q})_{{\rm prim}}$. We claim that this implies, except in the case where
$(d,n)=(3,3)$, that
$X_{f'}$ is also very general and there is an isomorphism of variations of Hodge structures on respective neighborhoods
$U$,
$V$ of
$f$ and
$f'$ in their moduli space
$U_{d,n}$. Indeed, the Hodge locus
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU5.png?pub-status=live)
defined as the set of points $t,\,t'\in U\times V$ such that the given isomorphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU6.png?pub-status=live)
induces an isomorphism of Hodge structures
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU7.png?pub-status=live)
is by the Cattani–Deligne–Kaplan theorem [Reference Cattani, Deligne and KaplanCDK95] the restriction to $U\times V$ of a closed algebraic subset (that we also denote
$\Gamma _\phi$) of
$U_{d,n}\times U_{d,n}$. As
$f$ is very general and in the image of
${\rm pr}_1: \Gamma _\phi \rightarrow U_{d,n}$,
${\rm pr}_1$ has to be dominant.
An important point is the fact that, as an easy consequence of Macaulay theorem [Reference VoisinVoi03, Theorem 6.19], the smooth hypersurfaces of degree $d$ in
$\mathbb {P}^{n}$, with
$(d,n)\not =(3,3)$ satisfy the infinitesimal Torelli theorem. This means that the period map is an immersion at the points
$f$ of the open set
$U_{d,n}^{0}\subset U_{d,n}$ parameterizing automorphisms free hypersurfaces, once
$(d,n)\not =(3,3)$. As
$f$ is very general, we can assume that
$f$ belongs to
$U_{d,n}^{0}$. It follows that the projection
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU8.png?pub-status=live)
must be locally finite since, by definition, the fiber of ${\rm pr}_2$ over any
$t'\in V$ parameterizes points
$t$ with isomorphic Hodge structures on
$H^{n-1}(X_t,\mathbb {Q})_{{\rm prim}}$. In particular,
$f'$ is also very general so we can assume that
$f'$ is also automorphism free.
The two projections ${\rm pr}_1$,
${\rm pr_2}$ are thus immersions and dominant morphisms, and hence they must be étale. It thus follows that
$\Gamma _\phi$ induces a local holomorphic diffeomorphism
$i$ between
$U$ and
$V$ which, by definition of
$\Gamma _\phi$, has the property that the isomorphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU9.png?pub-status=live)
of trivial local systems on $U$ induces an isomorphism of variations of Hodge structures. Here, if
$\pi :\mathcal {X}_{d,n}\rightarrow U_{d,n}^{0}$ is the universal hypersurface,
$H^{n-1}_\mathbb {C}$ is the local system
$R^{n-1}\pi _*\mathbb {C}_{{\rm prim}}$ on
$U_{d,n}^{0}$. Taking the differential of this isomorphism provides a commutative diagram where the vertical maps are isomorphisms,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn10.png?pub-status=live)
where the vertical map on the left is the differential $i_*$ at
$f\in U$ and the vertical map on the right is induced by the isomorphism of Hodge structures
$\phi : H^{n-1}(X_f,\mathbb {Q})_{{\rm prim}}\rightarrow H^{n-1}(X_{f'},\mathbb {Q})_{{\rm prim}}$. By Theorem 0.5, we then get a commutative diagram (8) to which Theorem 0.6 applies.
1.2 Polynomial structure and the symmetrizer lemma
The method used by Donagi to prove Theorem 0.6 consists in applying the ‘symmetrizer lemma’ (Proposition 1.1 below), in order to recover from the data (5) the whole Jacobian ring in degrees divisible by $l$, where
$l$ is the greatest common denominator of
$n+1$ and
$d$. This result proved first in [Reference DonagiDon83] for the Jacobian ring of generic hypersurfaces, and reproved in [Reference Donagi and GreenDG84] for any smooth hypersurface (and more generally quotients
$R_{f_{\bullet }}$ of the polynomial ring
$S=\mathbb {C}[X_0,\ldots,X_n]$ by a regular sequence
$f_{\bullet }=(f_0,\ldots,f_n)$ with
${\rm deg}\,f_i=d-1$), is the following statement. Consider the multiplication map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn11.png?pub-status=live)
Proposition 1.1 Let $N=(n+1)(d\,{-}\,2)$. Then, if
$\,{\rm Max}\,(k,N\,{-}\,k')\geq d\,{-}\,1$ and
${N-k-k'>0}$, the multiplication map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU10.png?pub-status=live)
is determined by the multiplication map (11) as follows
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn12.png?pub-status=live)
Coming back to the case of a Jacobian ring $R_f$, when
$d$ divides
$n+1$, the infinitesimal variation of Hodge structure (3) of
$X_f$, translated in the form (5), involves only pieces
$R^{k}_f$ of the Jacobian ring of degree
$k$ divisible by
$d$. Hence the symmetrizer lemma at best allows us, starting from the infinitesimal variation of Hodge structure (IVHS) of the hypersurface, to reconstruct the Jacobian ring in degrees divisible by
$d$. At the opposite, when
$d$ and
$n+1$ are coprime, repeated applications of the symmetrizer lemma allow us to reconstruct the whole Jacobian ring. In degree
$< d-1$, the Jacobian ring coincides with the polynomial ring, and hence we directly recover in that case the multiplication map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU11.png?pub-status=live)
and its kernel $J_f^{d}$. The proof of Donagi is then finished by applying Mather and Yau's theorem [Reference Mather and YauMY82] (see also Proposition 3.2).
This leads us to the following definition. Suppose that we have two integers $d,\ n$ and the partial data of a graded ring structure
$R^{*}$, namely finite dimensional vector spaces
$R^{d},\ R^{-(n+1)+id}$ for
$i$ such that
$-(n+1)+id\geq 0$ with multiplication maps
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn13.png?pub-status=live)
When $d$ divides
$n+1$, we get all the upper-indices divisible by
$d$, and an actual ring structure
$R^{d*}$, but in general (13) is the sort of data provided by the infinitesimal variation of Hodge structure of a hypersurface of degree
$d$ in
$\mathbb {P}^{n}$. Let
$S^{k}$ be the degree
$k$ part of the polynomial ring in
$n+1$ variables.
Definition 1.2 A polynomial structure in $n+1$ variables for the partial data of a graded ring structure
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU12.png?pub-status=live)
is the data of a rank $n+1$ base-point free linear subspace
$J\subset S^{d-1}$ generating a graded ideal
$J^{*}\subset S^{*}$, of a linear isomorphism
$S^{d}/J^{d}\cong R^{d}$ and, for all
$i$, of linear isomorphisms
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU13.png?pub-status=live)
compatible with the multiplication maps, i.e. making the following diagrams commutative.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn14.png?pub-status=live)
The group ${\rm GL}(n+1)$ acts in the obvious way on the set of polynomial structures. We will say that the polynomial structure of
$(R^{d},\,R^{-(n+1)+id},\,\mu _i)$ is unique if all its polynomial structures are conjugate under
${\rm GL}(n+1)$. As explained above, Donagi's Theorem 0.6 has the more precise form that, under some assumptions on
$(d,n)$, the polynomial structure of the infinitesimal variation of Hodge structure
$(R^{d}_f,\,R^{-(n+1)+id}_f,\,\mu _i)$ of a smooth hypersurface
$X_f$ is unique, and this is sufficient to imply the generic Torelli theorem for hypersurfaces of these degree and dimension. We will prove a similar statement in the case (2) (that is degree
$d=4$) of Theorem 0.7, at least for generic
$f$ and
$n$ large enough.
For the main series of cases not covered by Donagi's theorem, namely when $d$ divides
$n+1$, we have not been able to prove the uniqueness of the polynomial structure of
$R^{d*}_f$ (even for generic
$f$), although it is likely to be true (and it is proved in [Reference Voisin, Hulek, Catanese, Peters and ReidVoi99] for
$d=5$,
$n=4$). We finish this section by the proof of a weaker statement that provides evidence for the uniqueness. We will say that a polynomial structure is rigid if its small deformations are given by its orbit under
${\rm GL}(n+1)$. We have the following.
Proposition 1.3 Assume $n+1\geq 8$ and
$d\geq 6$. Let
$f\in S^{d}$ be a generic homogeneous polynomial of degree
$d$ in
$n+1$ variables and
$R_{f}^{d*}$ be its Jacobian ring in degrees divisible by
$d$. Then the natural polynomial structure
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU14.png?pub-status=live)
given by the quotient map is rigid.
Remark 1.4 The case where $d=4$ and
$n+1\cong 2$ mod
$4$ will be studied in next section. We will prove there, using a different recipe, that the polynomial structure on
$R^{2*}_f$ is unique for
$n$ large enough.
Remark 1.5 Proposition 1.3 implies that the natural polynomial structure of $R_{f_{\bullet }}^{d*}$ for a generic rank
$n+1$ regular sequence
$f_{\bullet }$ of degree
$d-1$ homogeneous polynomials is rigid.
We will use in fact only the multiplication map in degree $d$
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU15.png?pub-status=live)
Proposition 1.3 will be implied by Proposition 1.8 below. For our original polynomial structure on $R_f^{d*}$, and for each
$x\in S^{1}$, we get a pair of vector subspaces
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn15.png?pub-status=live)
which form an ideal in the sense that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn16.png?pub-status=live)
It is not hard to see that the multiplication map by $x$, from
$R^{d-1}_{f}$ to
$R^{d}_{f}$, is injective for a generic
$x\in S^{1}$ when
$f$ is generic with
$d\geq 4$ and
$n\geq 3$ (or
$d\geq 3$ and
$n\geq 5$). In fact, we even have the following (statement (ii) will be used only later on).
Lemma 1.6
(i) The multiplication map by
$x$ is injective on
$R^{2d-1}_{f}$ when
$f$ is generic,
$x\in S^{1}$ is generic and
(17)(for example,\begin{equation} 2(2d-1)<(d-2)(n+1) \end{equation}
$n+1\geq 5$ and
$d>8$, or
$n+1\geq 6$ and
$d>4$ work).
(ii) The multiplication map by
$x$ is injective on
$R^{3d-1}_{f}$ when
$f$ is generic,
$x\in S^{1}$ is generic and
(for example,\[ 2(3d-1)<(d-2)(n+1) \]
$n+1\geq 5$ and
$d>8$, or
$n+1\geq 6$ and
$d>4$ work).
(iii) The multiplication map by
$x^{l}$ is injective on
$R^{k}_{f}$ when
$f$ is generic,
$x\in S^{1}$ is generic and
\[ 2k+l\leq (d-2)(n+1). \]
Proof. As the dimensions of the vector spaces $R^{k}_f$ are independent of
$f$ (assumed to define a smooth hypersurface), the conclusions are open properties of
$f$, and hence it suffices to check them for a particular
$f$. Take for
$f$ the Fermat polynomial
$f_{\rm Fermat}=\sum _{i=0}^{n}X_i^{d}$. Then
$R_{f_{\rm Fermat}}^{*}$ identifies with the cohomology ring
$H^{2*}((\mathbb {P}^{d-2})^{n+1},\mathbb {C})$ (indeed, it has generators
$X_i$ and relations
$X_i^{d-1}=0$) and
$x=\sum _iX_i$ corresponds to an ample class in
$H^{2}((\mathbb {P}^{d-2})^{n+1},\mathbb {C})$. By the hard Leschetz theorem for
$(\mathbb {P}^{d-2})^{n+1}$, the multiplication by
$x$ is thus injective on
$R^{2d-1}_{f_{\rm Fermat}}$ if
$2(2d-1)<(d-2)(n+1)$, and injective on
$R^{3d-1}_{f_{\rm Fermat}}$ if
$2(3d-1)<(d-2)(n+1)$. More generally, the Lefschetz isomorphism for the power
$x^{l}$ gives the injectivity of
$x^{l}$ on
$R^{k}_{f_{\rm Fermat}}$ when
$2k+l\leq (d-2)(n+1)$.
Remark 1.7 The estimate in (i) is optimal for dimension reasons. Indeed, the dimensions of the graded pieces $R^{k}_{f}$ are increasing in the interval
$k\leq {((d-2)(n+1))}/{2}$, and decreasing in the interval
${((d-2)(n+1))}/{2}\leq k\leq (d-2)(n+1)$.
It follows from Lemma 1.6 that, assuming inequality (17), the space $I_x^{d}$ defined in (15) has generic dimension
$r_{d-1}:={\rm dim}\, R^{d-1}_{f}$, while
$I_x^{2d}$ has generic dimension
$r_{2d-1}:={\rm dim}\, R^{2d-1}_{f}$.
Proposition 1.8 If $f$ is generic of degree
$d$ in
$n+1$ variables, and
$d\geq 6,\ n\geq 9$, then the subset
$Z_{\rm ideal}=\{[I^{d}_x]\in G(r_{d-1},R^{d}_{f}),\,x\in S^{1}\}$ of the Grassmannian
$G(r_{d-1},R^{d}_{f})$ is a reduced component of the closed algebraic subset
$Z\subset G(r_{d-1},R^{d}_{f})$ defined as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn18.png?pub-status=live)
Proof. The tangent space to $Z_{\rm ideal}$ at the point
$[I^{d}_x]\in G(r_{d-1},R^{d}_{f})$ is the image of
$S^{1}/\langle x\rangle$ in
${\rm Hom}(R^{d-1}_{f},R^{d}_{f}/xR^{d-1}_{f})=T_{G(r_{d-1},R^{d}_{f}),[I_x^{d}]}$ given by multiplication by
$y\in S^{1}/\langle x\rangle$, where we identify
$I_x^{d}$ with
$R^{d-1}_{f}$ via multiplication by
$x$. Let us now compute the Zariski tangent space to
$Z$ at
$[I^{d}_x]$ for
$f$ and
$x$ generic. As
${\rm dim}\,I_x^{2d}=r_{2d-1}$ is maximal by the claim above, the condition (18) provides the following infinitesimal conditions:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn19.png?pub-status=live)
Equation (19) says that $h: R^{d-1}_{f}\rightarrow R^{d}_{f}/xR^{d-1}_{f}$ is a ‘morphism of
$R^{d}_{f}$-modules’, the set of which we will denote by
${\rm Mor}_{R^{d}_f}(R^{d-1}_{f},R^{d}_{f}/xR^{d-1}_{f})$, in the sense that we have a commutative diagram for some
$h'\in {\rm Hom}(R^{2d-1}_{f},R^{2d}_{f}/\langle x\rangle )$,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn20.png?pub-status=live)
where the horizontal maps are given by multiplication. The equality $T_{Z_{\rm ideal}}=T_Z$ at the point
$[I_x^{d}]$ is thus equivalent to the fact that all the ‘
$R^{d}_{f}$-modules morphisms’
$h: R^{d-1}_{f}\rightarrow R^{d}_{f}/xR^{d-1}_{f}$, are given by multiplication by some
$y\in S^{1}$, followed by reduction modulo
$x$. This is the statement of the following
Lemma 1.9 Let ${f}$ be a generic homogeneous degree
$d$ polynomial in
$n+1$ variables with
$d \geq 5,\ n\geq 9$ (or
$d\geq 6$ and
$n\geq 7$), and let
$x\in S^{1}$ be generic. Then the natural map
$S^{1}/\langle x\rangle \rightarrow {\rm Mor}_{R_{f}^{d}}(R^{d-1}_{f},R^{d}_{f}/\langle x\rangle )$ is surjective.
Proof. The existence of $h'$ as in (20) says that for any tensor
$\sum _i A_i\otimes B_i\in R^{d}_{f}\otimes R^{d-1}_{f}$ such that
$\sum _iA_iB_i=0$ in
$R^{2d-1}_{f}$,
$\sum _i A_ih(B_i)=0$ in
$R^{2d}_{f}/\langle x\rangle$.
Claim 1.10 Under the same assumptions as in Lemma 1.9, for a generic $q\in R^{d-1}_{f}$, the multiplication map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU18.png?pub-status=live)
is injective.
Proof. This is proved again by looking at the Fermat polynomial $f_{\rm Fermat}=\sum _iX_i^{d}$ and choosing carefully
$x$ so that multiplication by
$x$ is injective on
$R^{2d-1}_{f_{\rm Fermat}}$, and multiplication by
$q$ is injective on
$R^{d+1}_{f_{\rm Fermat}}/\langle x\rangle$. We write
$f_{\rm Fermat}=f'_{\rm Fermat}+f''_{\rm Fermat}$, where
$f'_{\rm Fermat}=\sum _{i=0}^{4}X_i^{d}$ and
$f''_{\rm Fermat}=\sum _{i=5}^{n}X_i^{d}$. We take
$x=\sum _{i=0}^{4}X_i$ and
$q=\big (\sum _{i=5}^{N}X_i\big )^{d-1}$. We observe that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU19.png?pub-status=live)
as graded rings, and that $x$ acts by multiplication on the left term
$R_{f'_{\rm Fermat}}^{*}$, while
$q$ acts by multiplication on the right term
$R_{f''_{\rm Fermat}}^{*}$. So it suffices to show that multiplication by
$x$ is injective on
$R_{f'_{\rm Fermat}}^{k}$ for
$k\leq 2d-1$ and multiplication by
$q$ is injective on
$R_{f''_{\rm Fermat}}^{k}$ for
$k\leq d+1$. The first statement follows from Lemma 1.6(i) when
$2(d+1)<5(d-2)$, hence when
$d\geq 5$. The second statement holds by Lemma 1.6(iii) when
$2(d+1)+d-1\leq (n-4)(d-2)$, and in particular if
$d\geq 6$ and
$n\geq 9$.
We deduce from Claim 1.10 that for any tensor $\sum _i A_i\otimes B_i\in S^{1}\otimes R^{d-1}_{f}$ such that
${\sum _iA_iB_i=0}$ in
$R^{d}_{f}$, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn21.png?pub-status=live)
since this becomes true after multiplication by $q$. It follows now that
$h$ vanishes on
$\langle x\rangle$. Indeed, let
$b=xb'$. Then for any
$y\in S^{1}$, we have
$yb=xb''$ with
$b''=yb'$. Hence by (21), we get
$yh(b)=xh(b'')=0$. Hence
$yh(b)=0$ in
$R^{d+1}_f/\langle x\rangle$ for any
$y\in S^{1}$, and it follows, by choosing
$y$ such that multiplication by
$y$ is injective on
$R^{d}_{f}/\langle x\rangle$, that
$h(b)=0$ in
$R^{d}_{f}/\langle x\rangle$. Thus
$h$ induces a morphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU20.png?pub-status=live)
which also satisfies (21). Assuming $d\geq 6,\,n\geq 9$, we now show by similar arguments as above that for generic
$z,\,y\in S^{1}/\langle x\rangle$, the following holds. For any
$p,\,q\in R^{d}_{f}/\langle x\rangle$,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn22.png?pub-status=live)
for some $r\in R^{d-1}_f/\langle x\rangle$. Furthermore we already know that the multiplication map by
$z$ from
$R^{d}_{f}/\langle x\rangle$ to
$R^{d+1}_{f}/\langle x\rangle$ is injective. It follows that there exists
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU21.png?pub-status=live)
inducing $\bar {h}$, that is,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn23.png?pub-status=live)
for any $p\in R^{d-2}_{f}/ \langle x\rangle$, and any
$a\in S^{1}/\langle x\rangle$. Indeed,
$y$ and
$z$ being as above, we have for any
$p\in R^{d-2}_{f}/ \langle x\rangle$
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU22.png?pub-status=live)
hence by (21), we get that $y\bar {h}(zp)-z \bar {h}(yp)=0$ in
$R^{d+1}_{f}/\langle x\rangle$, and by (22), this gives
$\bar {h}(zp)=z\bar {h}''(p)$, which defines
$\bar {h}''$. One then shows that the map
$\bar {h}''$ so defined does not depend on
$z$ and satisfies (23), which is easy. To finish the proof, we construct similarly
$\bar {h}''':R^{d-3}_{f}/\langle x\rangle \rightarrow R^{d-2}_{f}/\langle x\rangle$ inducing
$\bar {h}''$ and
$\bar {h}^{iv}:R^{d-4}_{f}/\langle x\rangle \rightarrow R^{d-3}_{f}/\langle x\rangle$ inducing
$\bar {h}'''$. As
$R^{i}_{f}/\langle x\rangle =S^{i}/\langle x\rangle$ for
$i\leq d-2$, it is immediate to show that
$\bar {h}^{iv}$ is multiplication by some element of
$S^{1}$, hence also
$\bar {h}$.
The proof of Proposition 1.8 is thus complete.
Proof Proof of Proposition 1.3
Let $f$ be generic of degree
$d\geq 6$ in
$n+1\geq 10$ variables. We first claim that for any
$x\in S^{1}$, the multiplication map by
$x:R^{d-1}_{f}\rightarrow R^{d}_{f}$ is injective, and that the morphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn24.png?pub-status=live)
so constructed is an embedding. None of these statements is difficult to prove. The first statement says that if ${f}$ is a generic homogeneous degree
$d$ polynomials in
$n+1$ variables,
${f}$ does not satisfy an equation
$\partial _u(f)_{\mid H}=0$ for some hyperplane
$H \subset \mathbb {P}^{n}$ and vector field
$u$ on
$\mathbb {P}^{n}$. The obvious dimension count shows that this holds if
$h^{0}(\mathbb {P}^{n-1},\mathcal {O}(d))>n-1+{(n+1)^{2}}/{2}$, which holds if
$d\geq 4,\ n\geq 3$. As for the second statement, suppose that
$xR^{d-1}_f=yR^{d-1}_f$ for some nonproportional
$x,\,y\in S^{1}$. Then there is a subspace of dimension
$\geq {\rm dim}\,S^{d-1}$ of pairs
$(p,q)\in S^{d-1}\times S^{d-1}$ such that
$xp=yq$ in
$R^{d}_f$, that is
$xp-yq\in J^{d}_f$. As the kernel of the map
$x-y: S^{d-1}\times S^{d-1}\rightarrow S^{d}$ is of dimension
${\rm dim}\,S^{d-2}$, this would imply that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn25.png?pub-status=live)
As ${\rm dim}\,J_f^{d}=(n+1)^{2}$, (25) is impossible if
$h^{0}(\mathbb {P}^{n-1},\mathcal {O}(d-1))>(n+1)^{2}$, which holds if
$n\geq 5, d\geq 4$. We have thus proved that the map
$\Phi$ of (24) is injective. That it is an immersion follows in the same way because the differential at
$x$ is given by the multiplication map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU23.png?pub-status=live)
and $\mu _y$ is zero if and only if
$y R^{d-1}_f\subset xR^{d-1}_f$, which has just been excluded. The claim is thus proved.
It follows from the claim and from Proposition 1.8 that, if we have a family of polynomial structures
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU24.png?pub-status=live)
with $\phi _0$ the natural one, then there is an isomorphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU25.png?pub-status=live)
such that for any $x\in S^{1}$,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU26.png?pub-status=live)
Such a projective isomorphism is induced by a linear isomorphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU27.png?pub-status=live)
and composing $\phi _t$ with the automorphism of
$S^{d*}$ induced by
$\tilde {\psi }_t^{-1}$, we conclude that we may assume that for any
$x\in S^{1}$,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn26.png?pub-status=live)
We claim that this implies $\phi _t: S^{d}\rightarrow R^{d}_f$ is the natural map of reduction mod
$J_f$. To see this, choose a general
$x$, so that the multiplication map by
$x$ is injective on
$R^{2d-1}_f$. The polynomial structure given by
$\phi _t$ and satisfying (26) provides two linear maps
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU28.png?pub-status=live)
such that $x\phi '_t=\phi _t\circ x: S^{d-1}\rightarrow R^{d}_f$,
$x\phi ''_t=\phi _t\circ x:S^{2d-1}\rightarrow R^{2d}_f$, and the injectivity of the map of multiplication by
$x$ on
$R^{2d-1}_f$ implies that the following diagram commutes, since it commutes after multiplying the maps by
$x$.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn27.png?pub-status=live)
The horizontal maps in the diagram above are the multiplication maps. Following Donagi [Reference DonagiDon83], the multiplication map on the bottom line determines the polynomial structure of $R_f^{*}$, because it determines (for
$d\geq 3$)
$S^{1}$ and the multiplication map
$S^{1}\otimes R^{d-1}_f\rightarrow R^{d}_f$ by the symmetrizer lemma (Proposition 1.1). The diagram (
) then says that up to the action of an automorphism $g$ of
$S^{*}$, the polynomial structure given by
$(\phi '_t,\,\phi _t)$ is the standard one. Finally, as
$g$ must act trivially on the space
$Z_{\rm ideal}$ of ideals by (26),
$g$ is proportional to the identity.
2. The case of degree
$4$
We explain in this section how to recover the polynomial structure of the infinitesimal variation of Hodge structure of a generic hypersurface of degree $4$ so as to prove Theorem 0.7(2), namely the cases where
$d=4$,
$n=4m+1$, with
$m$ large. Note that the methods of Schiffer variations that we will develop later would presumably also apply to this case, but it is much more difficult and does not prove Theorem 0.6 (saying that one can recover a hypersurface from its IVHS).
The congruence conditions is equivalent to the fact that we have ${\rm gcd}(4,n+1)=2$. The infinitesimal variation of Hodge structure (as translated in (5) using Theorem 0.5)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn28.png?pub-status=live)
has for smallest degree term the multiplication map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU29.png?pub-status=live)
and the symmetrizer lemma (see Proposition 1.1) allows us to reconstruct in these cases the whole ring $R^{2*}_f$, and in particular the multiplication map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn29.png?pub-status=live)
(Note that $R^{2}_f=S^{2}$.) We thus only have to explain in both cases how to recover the polynomial structure of (28) from (29), at least for a generic polynomial
$f$. We use the notation
$\mathcal {S}q^{2l}_f\subset R^{2l}_f$ for the set of squares
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU30.png?pub-status=live)
This is a closed algebraic subset which is a cone in $R_f^{2l}$ and we will denote by
$\mathbb {P}(\mathcal {S}q^{2l})$ the corresponding closed algebraic subset of
$\mathbb {P}(R^{2l}_f)$. When
$d=4,\ d'=2$, (29) determines
$\mathcal {S}q^{4}_f$. Our proof of Theorem 0.7(2) will be based (following Donagi's strategy described in the previous section) on the following.
Claim 2.1 The algebraic subset $\mathcal {S}q^{2}_f\subset R^{2}_f=S^{2}$ determines the polynomial structure of the even degree Jacobian ring
$R^{2*}_f$ of
$f$.
Proof. Indeed, passing to the projectivization of these affine cones, $\mathbb {P}(\mathcal {S}q^{2}_f)$ is the second Veronese embedding of
$\mathbb {P}(S^{1})$ in
$\mathbb {P}(S^{2})$. Thus the positive generator
$H$ of
${\rm Pic}(\mathbb {P}(\mathcal {S}q^{2}_f))$ satisfies the property that
$H^{0}(\mathbb {P}(\mathcal {S}q^{2}_f), H)=:V$ has dimension
$n+1$ and the restriction map
$(S^{2})^{*}\rightarrow {\rm Sym}^{2}V$ is an isomorphism. The dual isomorphism gives the desired isomorphism
${\rm Sym}^{2}S^{1}\cong S^{2}$, with
${S^{1}:=V^{*}}$.
We observe now that the closed algebraic subset $\mathcal {S}q^{2}_f\subset R^{2}_f$ has the following property:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn30.png?pub-status=live)
We prove now the following result, which by Claim 2.1 completes the proof of Theorem 0.7(2).
Proposition 2.2 Let $f$ be a generic homogeneous polynomial of degree
$4$ in
$n+1$ variables, with
$n\geq 599$. Then the only subvariety
$T\subset R^{2}_f=S^{2}$ of dimension
$\geq n+1$ satisfying the condition
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU31.png?pub-status=live)
is $\mathcal {S}q^{2}_f$.
Note that in this statement, we can clearly assume that $T$ is a cone, since the conditions are homogeneous.
Proof Proof of Proposition 2.2
We observe that by a proper specialization argument, the schematic version of the statement, saying that, furthermore, (30) define, at least generically, the reduced structure of $\mathcal {S}q^{2}_f$, is an open condition on the set of polynomials
$f$ for which
$R_f^{4}$, or equivalently
$J^{4}_f$, has the right dimension. We will thus prove this schematic version for one specific
$f$, for which
$J^{4}_f$ has the right dimension.
Let us first explain the specific polynomials we will use. We will first choose general linear sections $\mathbb {P}\cap Pf_4$ of the Pfaffian quartic
$\mathcal {P}f_4\subset \mathbb {P}\big (\bigwedge ^{2} V_8\big )$, where
$\mathbb {P}\subset \mathbb {P}\big (\bigwedge ^{2} V_8\big )$ is a linear subspace of dimension
$23$ or
$24$. We get this way polynomials
$f_i$ of degree
$4$ in
$24$ or
$25$ variables. In higher dimension, we will then consider polynomials of the form
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU32.png?pub-status=live)
with $i_1,\ldots,i_l\in \{24,\,25\}$, which allows us to construct degree
$4$ polynomials with any number
$n+1$ of variables starting from
$600$.
If a degree $d$ polynomial
$f$ defines a smooth hypersurface
$X_f$, the Jacobian ideal
$J_f$ is generated by the regular sequence
$J_f^{d-1}$ of degree
$d-1$ polynomials, and hence, assuming
$d\geq 3$, we find that the multiplication map
$S^{1}\otimes J_f^{d-1}\rightarrow J_f^{d}$ is an isomorphism, so that
${{\rm dim}\,J_f^{d}=(n+1)^{2}}$. In general, assuming
$X_f$ reduced with regular locus
$X_{f,{\rm reg}}$ and using the normal bundle sequence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn31.png?pub-status=live)
we see that the kernel of the map $S^{1}\otimes J_f^{d-1}\rightarrow J_f^{d}$ identifies naturally with the set of infinitesimal automorphisms of
$X_f$ induced by an infinitesimal automorphism of
$\mathbb {P}^{n}$. The hypersurfaces
$X_f$ discussed above are singular. We nevertheless have the following.
Lemma 2.3 For a polynomial $f$ of the form above,
$J_f^{4}$ has the right dimension, i.e.
$(n+1)^{2}$.
Proof. One has $f=\sum _j f_{j}$, where each
$f_{j}$ involves variables
$X_{j,1},\ldots, X^{j,i_j}$. It is immediate to check that the statement for each
$f_{j}$ implies the statement for
$f$. Turning to the
$f_{j}$, they are either general linear sections of the quartic Pfaffian hypersurface in
$\mathbb {P}^{27}$ by a
$\mathbb {P}^{n}$,
$n=23$ or
$24$. Let us show that each of them has no infinitesimal automorphism. The automorphism group of the general Pfaffian hypersurface
$\mathcal {P}f_k\subset \mathbb {P}\big (\bigwedge ^{2}V_{2k}\big )$ is the group
${\rm {P}GL}(2k)$. We claim that the automorphism group of a general linear section of dimension
$>2(2k-2)={\rm dim}\,G(2,V_{2k})$ is also contained in
${\rm {P}GL}(2k)$. This follows from the fact that after blowing-up
$\mathcal {P}f_{k}$ along its singular locus, which parameterizes forms of rank
$<2k-2$, we get a dominant morphism
$\widetilde {\mathcal {P}f_{k}}\rightarrow G(2,V_{2k})$, which to a degenerate form associates its kernel. If we consider a general linear section
$X_l$ of
$\mathcal {P}f_{2k}$ of dimension
$>{\rm dim}\,G(2,V_{2k})$ defined by a
$r$-dimensional vector subspace
$W\subset \bigwedge ^{2}V_{2k}^{*}$, the same remains true and we get a morphism
$\tilde {X}_l\rightarrow G(2,V_{2k})$ which is dominant with connected fiber of positive dimension. Thus the automorphism group of
$X_l$ has to act on
$G(2,V_{2k})$ and it has to identify with the group of automorphisms of
$G(2,V_{2k})$, or automorphisms of
$\mathbb {P}(V_{2k})$ preserving the space
$W\subset \bigwedge ^{2}V_{2k}^{*}$. It is easy to check that this space is zero once
$r\geq 3$. Coming back to our situation where
$k=d=4$, our choices of
$r$ are
$r=3$ or
$r=4$ for
$k=4$. In all cases, the variety
$X_l$ has dimension
$>2(2k-2)$ so the analysis above applies.
We now prove Proposition 2.2 for $f$ as above.
Lemma 2.4 Let $f=\sum _jf_j$ be a polynomial of degree
$4$ in
$n+1$ variables as constructed above. Then if
$T\subset S^{2}$ is an algebraic subvariety of dimension
$\geq n+1$ such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU33.png?pub-status=live)
for any $A,B\in T$,
$T= \mathcal {S}q^{2}$.
Proof. Let as above $f=\sum _l f_{j}$. The singular locus
$Z_f$ of
$X_f=V(f)$ is the join of the singular loci
$Z_{j}$ of
$V(f_j)$ in
$\mathbb {P}^{i_j}$. This means that, introducing the natural rational projection map
$\pi :\mathbb {P}^{\sum _j i_j-1}\dashrightarrow \prod _l\mathbb {P}^{i_j-1}$, one has
$Z_f=\pi ^{-1}\big (\prod _jZ_{j}\big )$.
Claim 2.5 The varieties $Z_{f}$ are not contained in any quadric.
Proof. Consider first the case of the Pfaffian linear sections $Z_j$. The claim follows in this case because they are general linear sections of the singular locus
$Z$ of the quartic universal Pfaffian
$\mathcal {P}f_4$ in
$\mathbb {P}\big (\bigwedge ^{2} V_8\big )$, which is defined by the equations
$\omega ^{3}=0$, that is, by cubics, and is not contained in any quadric. The last point can be seen by looking at the singular locus of
$Z$, which consists of forms of rank
$2$, that is the Grassmannian
$G(2,V_8^{*})$. Along this locus, the Zariski tangent space of
$Z$ is the full Zariski tangent space of
$\mathbb {P}\big (\bigwedge ^{2}V_8\big )$. A quadric containing
$Z$ should thus be singular along
${\rm Sing}\,Z$. But
${\rm Sing}\,Z=G(2,V_8^{*})$ is not contained in any proper linear subspace of
$\mathbb {P}\big (\bigwedge ^{2} V_8\big )$. It follows that
$Z$ is not contained in any quadric. It remains to conclude that the same statement is true for the general linear section
$\mathbb {P}^{i_l-1}\cap \mathcal {P}f_4$, with
$i_l=24,\ 25$. Its singular locus
$Z_l$ is the general linear section
$\mathbb {P}^{i_l-1}\cap Z$, and we show inductively that any quadric containing
$Z_l$ is the restriction of a quadric containing
$Z$. This statement only needs that
$Z_l$ is nonempty (it has dimension
$\geq 18$ in our case) and that all the successive linear sections
$\mathbb {P}^{j}\cap Z$, with
$j\geq i_l$, are linearly normal in
$\mathbb {P}^{j}$, which is not hard to prove. Finally we have to show that the same is true for a general
$f=\sum f_j$. As already mentioned,
$Z_f$ is then a join
$Z_1*\ldots *Z_l$ and a join of varieties not contained in any quadric is not contained in any quadric.
We also prove the following.
Claim 2.6
(a) The restriction map
$S^{1}\rightarrow H^{0}(Z_f,\mathcal {O}_{Z_f}(1))$ is an isomorphism.
(b) The only
$n$-dimensional family
$\{D_A\}$ of divisors on
$Z_f$ such that for some fixed effective divisor
$D_0$,
is the family of hyperplane sections of\[ 2D_A+ D_0\in |\mathcal{O}_{Z_f}(2)| \]
$Z_f$.
Proof. We know that $Z_f$ is the join of the
$Z_j\subset \mathbb {P}^{i_j}$, where each
$Z_j$ is a smooth linear section of the singular locus
$Z$ of
$\mathcal {P}f_4$ by either a
$\mathbb {P}^{24}$ or a
$\mathbb {P}^{23}$. Let us first conclude the proof when f is one of the
$f_j$ (and thus Z is one of the
$Z_j$). We observe that
$Z\subset \mathbb {P}\big (\bigwedge ^{2}V_8\big )$ is the set of
$2$-forms of rank
$\leq 4$ (the generic element of
$Z$ being of rank exactly
$4$), and has a natural birational model
$\tilde {Z}\rightarrow Z$, where
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU35.png?pub-status=live)
The statement (a) easily follows from the above description of $\tilde {Z}$ and the fact that the
$Z_j$ are general linear sections of codimension
$3$ or
$4$ of
$Z$. Let us prove (b). The variety
$\tilde {Z}$ is smooth and, being a projective bundle fibration over
$G(4,V_8^{*})$, has Picard rank
$2$. Its effective cone is very easy to compute: indeed, the line bundle
$l$ which is pulled-back from the Plücker line bundle on the Grassmannian via the first projection
$pr_1$ is clearly one extremal ray of the effective cone since the corresponding morphism has positive dimensional fibers. There is a second extremal ray of the effective cone, which is the class of the divisor
$D$ contracted by the birational map
$\tilde {Z}\rightarrow Z$ (induced by the second projection
$pr_2$). One easily computes that this class is
$2h-l$, where
$h$ is the pull-back of hyperplane class on
$\mathbb {P}\big (\bigwedge ^{2}V_8\big )$ by
$pr_2$. We observe that the fibers of
$pr_1$ are of dimension
$5$, so that, when we take a general linear section of
$Z$ by a codimension
$3$ or
$4$ projective subspace, getting the singular locus
$Z_j$ of
$X_j$, all the properties above remain satisfied, and thus
${\rm Pic}\,\tilde {Z}_j=\mathbb {Z} h_j+\mathbb {Z}l_j$, with effective cone generated by
$l_j$ and
$2h_j-l_j$. We now finish the argument for
$Z_j$: we lift our data
$\{D_A\},\ D_0$ to
$\tilde {Z}_j$. We then have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU36.png?pub-status=live)
in ${\rm Pic}\,\tilde {Z}_j$, with
$\tilde {D}_0$ effective and
${\rm dim}|\tilde {D}_A|\geq n$, where
$n$ is
${\rm dim}|h_{\mid Z_j}|$ by (a). Let us write
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU37.png?pub-status=live)
with $\alpha,\beta \in 2\mathbb {Z}$. Then the analysis above shows that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU38.png?pub-status=live)
As $2h_j-\tilde {D}_0$ is effective, we also have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU39.png?pub-status=live)
As $\alpha$ is even, we only get the possibilities
$\alpha =0,\ 2$. If
$\alpha =2$, we get
$\beta =0$ and thus
$2h_j-\tilde {D}_0=0$, contradicting the assumption
${\rm dim}|2h-\tilde {D}_0|\geq n$. If
$\alpha =0$, we get from the inequalities above
$\beta =0$ since
$\beta$ has to be even, and
$2h=2\tilde {D}_A$, which proves statement (b) (using the fact that
${\rm Pic}\,Z_j$ has no
$2$-torsion and statement (a)).
We now have to prove the same result for the join $Z_f=Z_1*Z_2*\ldots * Z_l$, which is done inductively on the number
$l$, assuming, as this is satisfied in our situation, that the
$Z_i$ are simply connected. This way we are reduced to consider only the join
$Z_1*Z_2\subset \mathbb {P}^{n}$, with
$n=n_1+n_2+1$, of two linearly normal varieties
$Z_1\subset \mathbb {P}^{n_1},\ Z_2\subset \mathbb {P}^{n_2}$, which satisfy the properties (a) and (b). We observe that
$Z_1*Z_2$ is dominated by a
$\mathbb {P}^{1}$-bundle over
$Z_1\times Z_2$, namely
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn32.png?pub-status=live)
the two sections being contracted to $Z_1$, respectively
$Z_2$, by the natural morphism to
$Z_1*Z_2\subset \mathbb {P}^{n}$. The description (32) of the join immediately proves (a) for
$Z_1*Z_2$ once we have it for
$Z_1$ and
$Z_2$. We now turn to (b). Let
$h=\mathcal {O}_{\mathbb {P}(\mathcal {O}_{Z_1}(1)\oplus \mathcal {O}_{Z_2}(1)})$ on
$\widetilde {Z_1*Z_2}$ and let
$D_0$ be a fixed effective divisor and
$\{D_A\}$ be a mobile family of divisors on
$\widetilde {Z_1*Z_2}$ such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn33.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn34.png?pub-status=live)
Then either $D_0$ or
$D_A$ is vertical for
$\pi$. Indeed, they both restrict otherwise to a divisor of degree
$\geq 1$ on the fibers of
$\pi$, contradicting (33). Assume
$D_0$ is vertical for
$\pi$, that is,
$D_0=\pi ^{-1}(D'_0)$. The equality
$2h-D'_0=2D_A$ says that
$D'_0=2D''_0$ as divisors on
$Z_1\times Z_2$ and, as
$Z_1$ and
$Z_2$ are simply connected,
$D''_0\in |pr_1^{*}D''_{0,1}+pr_2^{*}D''_{0,2}|$ and both
$2D''_{0,1}, \ 2D''_{0,2}$ are effective. The divisors
$D''_{0,i}$ on
$Z_i$ have the property that the linear system
$|h-pr_1^{*}D''_{0,1}-pr_2^{*}D''_{0,2}|$ on
$\mathbb {P}(\mathcal {O}_{Z_1}(1)\oplus \mathcal {O}_{Z_2}(1))$ has dimension
$\geq n$, which says that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU40.png?pub-status=live)
As $2D''_{0,1}$ is effective on
$Z_1$ and
$2D''_{0,2}$ is effective on
$Z_2$, we conclude that
$D''_{0,i}=0$ and that the
$D_A$ belong to
$|\mathcal {O}_{Z_1*Z_2}(1)|$, so (b) is proved in this case.
In the case where $D_0$ is not vertical, then restricting again to the fibers of
$\pi$,
$D_0=2h-D'_0$ where
$D'_0$ is effective and comes from
$Z_1\times Z_2$, and
$D_A$ is vertical,
$D_A=\pi ^{-1}(D'_A)$. Hence we have again
$D'_0=2D''_0$, and
${\rm dim}\,|h-D'_0-D'_A|\geq n$, so the proof finishes as before that
$D'_0=0$ and
$D_A=0$ which contradicts (34).
We now complete the proof of Lemma 2.4. We have $J_f\subset I_{Z_f}$, since
$Z_f$ is contained in
${\rm Sing}\,X_f$. Let
$T\subset S^{2}$ be a closed algebraic subset satisfying the assumptions of Lemma 2.4 for
$f$. Then for any
$A,B\in T$,
$AB_{|Z_f}=M^{2}_{|Z_f}$, for some
$M\in S^{2}$. This implies that the moving part of
${\rm div}\,A_{\mid Z_f}$ appears with multiplicity
$2$. Hence
${\rm div}\,A_{\mid Z_f}=D_0+2D_A$. We now use Claim 2.5 which implies that the family of divisors
${\rm div}\,A_{\mid Z_f}$ is of dimension
$\geq n$, hence also the family
$\{D_A\}$ of divisors. We then conclude from Claim 2.6 that
$T\subset \mathcal {S}q^{2}$.
In order to complete the proof of Proposition 2.2, it suffices now to make Lemma 2.4 more precise by analyzing the schematic structure of a closed algebraic subset $T\subset S^{2}$ satisfying the assumptions of this lemma. We first observe the following.
Lemma 2.7 Let $f$ be as in Lemma 2.4 and let
$A,B\in S^{1}$ be general. Let
$M:=AB$ and consider the subspace
$MS^{2}=MR^{2}_f\subset R^{4}_f$. Then
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn35.png?pub-status=live)
where as usual, the notation $[ MS^{2}:A^{2}]$ is used for
$\{S\in S^{2},\,SA^{2}\in \langle M\rangle \}$.
The lemma is obvious, using restriction to $Z_f$ and using Claim 2.6.
We now complete the proof of Proposition 2.2. We observe that $MS^{2}$ is the tangent space to
$\mathcal {S}q^{4}$ at
$M^{2}$, while
$BS^{1}\subset S^{2}$ is the tangent space to
$\mathcal {S}q^{2}$ at
$B^{2}$. Equation (35) thus says that a space
$T\subset S^{2}$ satisfying the assumptions of Lemma 2.4 must be generically the reduced
$\mathcal {S}q^{2}$. The completion of the proof then follows by a specialization and cycle-theoretic argument, using the fact that these sets
$T$ above are cones, hence come from closed algebraic subsets
$\mathbb {P}(T)$ of
$\mathbb {P}(S^{2})$.
3. Schiffer variations and Jacobian ideals
Definition 3.1 A Schiffer variation of a homogeneous polynomial $f$ of degree
$d$ in
$n+1$ variables
$X_0,\ldots, X_n$ is a one-parameter family
$f+t x^{d},\ t\in \mathbb {C}$, where
$x\in S^{1}$ is a linear form of the variables
$X_0,\ldots, X_n$.
In the definition above, we are not interested in the linear character of the parameterization, as this does not make sense anymore after projection of this line to the moduli space. We should thus consider more generally one-parameter families of polynomials supported (up to the action of ${\rm GL}(n+1)$) on a line as above. We can also speak of finite-order Schiffer variations, which consist in looking at a finite-order arc in an affine line as above passing through
$f$. Observe that if
$g=f+ x^{d}$, then for any
$u\in H^{0}(\mathbb {P}^{n},T_{\mathbb {P}^{n}}(-1))=H^{0}(\mathbb {P}^{n}, \mathcal {O}_{\mathbb {P}^{n}}(1))^{*}$ such that
$\partial _u(x)=0$, one has
$\partial _u(f)=\partial _u(g)$. It follows that the Jacobian ideals
$J_f$, respectively
$J_g$ generated by the partial derivatives of
$f$, respectively
$g$, satisfy the condition
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn36.png?pub-status=live)
It turns out that (36) is in most cases a characterization of Schiffer variations, as shows Proposition 3.4 below. A well-known result due (in various forms) to Carlson and Griffiths [Reference Carlson, Griffiths and BeauvilleCG80], Donagi [Reference DonagiDon83], and Mather and Yau [Reference Mather and YauMY82] says the following.
Proposition 3.2 Let $f,\ g$ be two homogeneous polynomials in
$n+1$ variables, defining smooth hypersurfaces in
$\mathbb {P}^{n}$. If the Jacobian ideals
$J_f^{d-1}$ and
$J_g^{d-1}$ coincide, then
$f$ and
$g$ are in the same orbit under the group
${\rm PGL}(n+1)$.
A nice proof of this statement is given in [Reference DonagiDon83]. The example of the Fermat equation $f=\sum _iX_i^{d}$ and its variations
$g=\sum _i \alpha _iX_i^{d}$ shows that one does not always have
$f= \mu g$ for some coefficient
$\mu$, under the assumptions of Proposition 3.2. The Mather–Yau theorem is the following variant which is more precise but works only for
$d$ large enough and
$f$ generic.
Proposition 3.3 Let $f,\ g$ be two homogeneous polynomials of degree
$d$ in
$n+1$ variables, defining smooth hypersurfaces in
$\mathbb {P}^{n}$. Assume
$d\geq 4,\ n\geq 4$. If
$f$ is generic and the Jacobian ideals
$J_f$ and
$J_g$ coincide, then
$f=\mu g$ for some coefficient
$\mu$.
Let us prove a closely related statement concerning the case where $J_f^{d-1}$ and
$J_g^{d-1}$ are not equal but almost equal, that is, satisfy (36).
Proposition 3.4 Let $d,\ n$ be such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn37.png?pub-status=live)
Then for a generic polynomial $f\in H^{0}(\mathbb {P}^{n},\ \mathcal {O}_{\mathbb {P}^{n}}(d))$, (36) holds if and only if
$g$ belongs to a Schiffer variation of
$\lambda f$ for some coefficient
$\lambda$.
Note that (37) holds if $d\geq 4$ and
$n\geq 4$.
Proof Proof of Proposition 3.4
As $f$ is generic, its Jacobian ideal
$J_f^{d-1}$ has dimension
$n+1$. Equation (37) can thus be written in the following form, for adequate choices of linear coordinates
$X_0,\ldots,X_n$ and
$Y_0,\ldots,Y_n$ on
$\mathbb {P}^{n}$
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn38.png?pub-status=live)
Let us now use the symmetry of partial derivatives:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU41.png?pub-status=live)
Combined with (38), it provides, for any $i,\,j$ between
$0$ and
$n-1$, the following second-order equations:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn39.png?pub-status=live)
Lemma 3.5 Under the numerical assumption (37), a generic polynomial $f$ of degree
$d$ in
$n+1$ variables does not satisfy a nontrivial second-order partial differential equation of the type (39).
Proof. This is a dimension count. The differential equations appearing in (39) are linear second-order equations determined by elements $U$ of
${\rm rank}\leq 4$ in
${\rm Sym}^{2}H^{0}(\mathbb {P}^{n},\mathcal {O}_{\mathbb {P}^{n}}(1))^{*}$. The nonzero elements
$U$ of
${\rm rank}\leq 4$ are parameterized by a variety of dimension
$4(n-3)+9$. Given a nonzero
$U$, the differential equation
$\partial ^{2}_U \phi =0$ determines a linear subspace
$H_U$ of
$H^{0}(\mathbb {P}^{n},\mathcal {O}_{\mathbb {P}^{n}}(d))$ of codimension
$h^{0}(\mathbb {P}^{n},\mathcal {O}_{\mathbb {P}^{n}}(d-2))$ since
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU42.png?pub-status=live)
is surjective. If (37) holds, the union of the spaces $H_U$ does not fill-in a Zariski open set of
$H^{0}(\mathbb {P}^{n},\mathcal {O}_{\mathbb {P}^{n}}(d))$.
It follows that all the equations appearing in (39) are trivial, which says equivalently that for any $i,\ j$
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU43.png?pub-status=live)
These equations exactly say that for some $\lambda \in \mathbb {C}$,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU44.png?pub-status=live)
for $i=0,\ldots,n-1$. Thus
$g-\lambda f$ satisfies
$({\partial }/{\partial X_i})(g-\lambda f)=0$ for
$i=0,\ldots,n-1$, and hence
$g=\lambda f +\alpha X_n^{d}$ for some coefficient
$\alpha$, which completes the proof.
We complete this section with the following proposition which shows the relevance of first-order Schiffer variations to our subject.
Proposition 3.6 Let $f,\ g$ be two degree
$d$ homogeneous polynomials in
$n+1$ variables defining smooth hypersurfaces
$X_f$,
$X_g$. Assume
$d\geq 4,\ n\geq 4$ and
$f$ is generic. Then if there exists a linear isomorphism
$i: R^{d}_f\cong R^{d}_g$ mapping the set of first-order Schiffer variations of
$f$ to the set of first-order Schiffer variations of
$g$,
$X_f$ is isomorphic to
$X_g$.
Proof. The conditions on $d,\ n$ are used in the following.
Lemma 3.7 For $f$ generic with
$d\geq 4,\ n\geq 2$, if
$x\in S^{1}$ is nonzero, then
$x^{d}\not =0$ in
$R^{d}_f$. Furthermore the map
$x\mapsto x^{d}$ induces a (incomplete) Veronese imbedding
$v_f: \mathbb {P}(S^{1})\hookrightarrow \mathbb {P}(R_f^{d})$.
Proof. The statement is equivalent to proving that, if $f$ is a generic homogeneous polynomial of degree
$d$ in
$n+1$ variables,
$J_f^{d}\subset S^{d}$ does not contain any power
$x^{d}$ of a linear form, or any sum
$x^{d}-y^{d}$ of two such powers. In the first case, we get that
$x\cdot x^{d-1}=0$ in
$R^{d}_f$ and in the second case, we get that
$(x-y)(x^{d-1}+x^{d-2}y+\cdots +y^{d-1})=0$ in
$R^{d}_f$. By an easy dimension count, one sees that for generic
$f$, the multiplication map
$x:R_f^{d-1}\rightarrow R_f^{d}$ by any nonzero linear form
$x\in S^{1}$ is injective, so in the first case we conclude that
$x^{d-1}=0$ in
$R_f^{d-1}$ and in the second case,
$x^{d-1}+x^{d-2}y+\cdots +y^{d-1}=0$ in
$R_f^{d-1}$, or equivalently
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn40.png?pub-status=live)
Using the fact that $n\geq 2$, and choosing a coordinate system such that
$x=X_0,\,y=X_1$, we can write (40) as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn41.png?pub-status=live)
for some nonzero coefficients $\alpha _i$. We thus get in both cases a nontrivial second-order equation
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU45.png?pub-status=live)
which is excluded by Lemma 3.5.
We now complete the proof. Using the lemma, the projectivized isomorphism $i$ induces an isomorphism
$i_1:\mathbb {P}(S^{1})\cong \mathbb {P}(S^{1})$ between the two projected Veronese
$v_f(\mathbb {P}(S^{1})) \subset \mathbb {P}(R^{d}_f)$ and
$v_g(\mathbb {P}(S^{1})) \subset \mathbb {P}(R^{d}_g)$, that is,
$i_1$ satisfies
$i\circ v_f=v_g\circ i_1$. The projective isomorphism
$i_1$ lifts to a linear isomorphism
$\tilde {i}_1:S^{1}\cong S^{1}$. The incomplete Veronese embeddings
$v_f$, respectively
$v_g$ factor canonically through the complete Veronese embeddings
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU46.png?pub-status=live)
(respectively $\mathbb {P}^{n}(S^{1})\stackrel {V_d}{\rightarrow } \mathbb {P}(S^{d})\dashrightarrow \mathbb {P}(R^{d}_f)$,) which implies that the following diagram,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn42.png?pub-status=live)
where $\widetilde {i_d}:S^{d}\cong S^{d}$ is induced by
$\widetilde {i_1}$, is commutative up to a scalar.
The vertical quotient maps have for respective kernels $J_f^{d}$,
$J_g^{d}$. We thus conclude that
${\widetilde {i_d}(J_f^{d})=J_g^{d}}$, and thus, by Proposition 3.2,
$X_f$ is isomorphic to
$X_g$.
3.1 Formal properties of Schiffer variations
Our strategy for the proof of Theorem 0.7 when $d$ divides
$n+1$ consists in finding a characterization of the set of Schiffer variations of a hypersurface
$X_f$ that can be read from its local variation of Hodge structure. In fact we will need not only the IVHS of
$X_f$ but also the ‘deformation of the IVHS’ along the Schiffer variation, which is a higher-order argument. The IVHS itself provides the first-order invariants of the variation of Hodge structure of
$X_f$ at the point
$[f]$, hence part of the multiplicative structure of the Jacobian ring
$R_f$, by (6). We wish in this section to analyze the specifics of the first-order Schiffer variations
$\phi \in R^{d}_f$ as elements of the Jacobian ring
$R_f^{*}$ and also analyze, using (36), the way the Jacobian ring deforms along them.
Recall that a first-order Schiffer variation of $f$ is an element
$\phi =x^{d}\in R^{d}_f$, where
$x\in S^{1}$. We will consider only the subring
$R_f^{d*}$ of
$R_f^{*}$, because, by Theorem 0.5, this is, when
$d$ divides
$n+1$, the data that we get from the IVHS of
$f$. As
$R_f^{d*}$ contains only the graded pieces of degree divisible by
$d$, it does not contain the linear form
$x$ and by Proposition 3.6, recovering the polynomial structure of
$R_f^{d}$ precisely means recognizing the set of powers
$\phi =x^{d}$. However, the ideal of
$R_f^{*d}$ generated by such a
$\phi \in R_f^{d}$ has some special properties that we can describe using only the multiplication map
$R^{id}_f\otimes R^{jd}_f\rightarrow R^{(i+j)d}_f$ for
$i+j\leq 3$.
Given $\phi \in R^{d}_f$ and vector subspaces
$I_k^{*d}\subset R^{*d}$ for
$*=1,2,3$, and
$1\leq k\leq d-1$, consider the following condition
$(*)$:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn43.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn44.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn45.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn46.png?pub-status=live)
Then condition $(*)$ is satisfied by
$I_{x,k}^{*d}:=x^{k}R^{*d-k}_f\subset R^{*d}_f$ for
$x\in S^{1}$ generic, with
$\phi =x^{d}$, at least if
$d$ is large enough. Indeed, condition (43) follows in that case from the fact that the multiplication by
$x$ is injective in the relevant degrees (see Lemma 1.6), at least if
$d$ or
$n$ are large enough, and the other conditions are obvious.
The second obvious property of a Schiffer variation is described in the following lemma.
Lemma 3.8 Let $x\in S^{1}$ determine a first-order Schiffer variation
$f_t=f+tx^{d}$ of
$f$ with tangent vector
$\phi =x^{d}\in R^{d}_f$, and let
$I_{x,d-1,t}^{*d}:=x^{d-1}R^{*d-d+1}_{f_t}\subset R^{*d}_{f_t}$ be defined as above. Then the quotient ring
$R^{*d}_{f_t}/I_{x,d-1,t}^{*d}$ does not deform (as a ring) along the Schiffer variation
$(f_t)$.
Proof. Indeed, if $f_t=f+tx^{d}$, then
$J_{f_t}=J_{f}$ modulo
$x^{d-1}$, and hence the quotient
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU47.png?pub-status=live)
is constant. A fortiori, its isomorphism class as a graded ring does not depend on $t$.
4. Proof of Theorem 0.7 when
$d$ divides
$n+1$
4.1 Specialization and Schiffer variations
We will consider in this section singular hypersurfaces $X_f$ of degree
$d$ in
$\mathbb {P}^{n}$ defined by a polynomial of the form
$f=\sum _{i=1}^{m}f_ig_i$, with
$n-2m\geq 0$. If
$d=2d'$ is even, we will choose the
$f_i$ and
$g_i$ to be of degree
$d'$ and if
$d=2d'+1$ we will choose the
$f_i$ of degree
$d'$ and the
$g_i$ of degree
$d'+1$. The hypersurface
$X_f$ is then singular along the variety
$Z$ defined by the polynomials
$f_i$ and
$g_i$ which are all of degree
$\leq {(d+1)}/{2}$, and when they are generically chosen, it is of dimension
$n-2m$. Let us start with the following result of independent interest, which will be important below and in the next section.
Proposition 4.1 Let $f$ be a homogeneous polynomial of degree
$d$ in
$n+1$ variables, defining a hypersurface
$X_f$ singular along a smooth subvariety
$Z$ defined by homogeneous polynomial equations of degree
$\leq {(d+1)}/{2}$. Then the dimension of the space
$R^{k}_f$ is equal to the dimension of the space
$R^{k}_{f_{\rm gen}}$ for a generic polynomial
$f_{\rm gen}$, assuming
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn47.png?pub-status=live)
Proof Proof of Proposition 4.1
Recall that $J_f$ is generated by the partial derivatives
${\partial f}/{\partial X_i}$ for
$i=0,\ldots, n$. For the generic polynomial
$f_{\rm gen}$, these partial derivatives form a linear system
$W$ of degree
$d-1$ polynomials with no base-point on
$\mathbb {P}^{n}$, and the associated Koszul resolution
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn48.png?pub-status=live)
twisted by $\mathcal {O}_{\mathbb {P}^{n}}(k)$, allows us to compute the dimension of
$R^{k}_{f_{\rm gen}}$ or, equivalently, of
$J^{k}_{f_{\rm gen}}={\rm Im}\,H^{0}(\alpha (k))$, using the fact that this twisted Koszul complex, that we will denote by
$\mathcal {K}_{\mathbb {P}^{n},k}^{*}$, remains exact at the level of global sections, at least in strictly negative degrees, where we put the last term
$\mathcal {O}_{\mathbb {P}^{n}}(k)$ in degree
$0$, which is what we need to compute
${\rm dim}\,J^{k}_{f_{\rm gen}}$. Indeed, we then get an equality
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn49.png?pub-status=live)
In our special case, the partial derivatives ${\partial f}/{\partial X_i}$ form a linear system
$W$ of degree
$d-1$ polynomials on
$\mathbb {P}^{n}$ with base-locus
$Z$ and we have to understand how this affects the computation. Clearly, the Koszul complex (48) is no more exact. Let
$\tau :Y\rightarrow \mathbb {P}^{n}$ be the blow-up of
$\mathbb {P}^{n}$ along
$Z$, and let
$E$ be the exceptional divisor of
$\tau$. Then
$W$ provides a base-point free linear system, that we also denote by
$W$, of sections of the line bundle
$L:=\tau ^{*}\mathcal {O}_{\mathbb {P}^{n}}(d-1)(-E)$ on
$Y$. Then we have an exact Koszul complex on
$Y$ associated with
$W$, which has the following form:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn50.png?pub-status=live)
We now twist by $\tau ^{*}\mathcal {O}_{\mathbb {P}^{n}}(k)$ so that
${\rm Im}\,H^{0}(\alpha '(k))= {\rm dim}\,J_f^{k}$. Let
$\mathcal {K}_{Y,k}^{*}$ be this twisted Koszul complex. We observe that, as only nonnegative twists of
$E$ appear in
$\mathcal {K}_{Y,k}^{*}$, the global sections of
$\mathcal {K}_{Y,k}^{*}$ are
$\bigwedge ^{i} W\otimes S^{k-i(d-1)}$ in degree
$-i$. So our problem is actually to prove that, if the inequality (47) holds, the complex
$K_{Y,k}^{*}$ of global sections of
$\mathcal {K}_{Y,k}^{*}$ is as before exact in strictly negative degrees. To prove this, we have to analyze the hypercohomology spectral sequence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn51.png?pub-status=live)
of $\mathcal {K}_{Y,k}^{*}$. As
$\mathcal {K}_{Y,k}^{*}$ is exact, one has
$\mathbb {H}^{p+q}(Y, \mathcal {K}_{Y,k}^{*})=0$, and hence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn52.png?pub-status=live)
We have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU48.png?pub-status=live)
where
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn53.png?pub-status=live)
We now observe that, when $-p< n-{\rm dim}\,Z$, we have
$H^{q}(Y,\mathcal {K}_{Y,k}^{p})=0$ for
$q\not = 0,\, n$. This is because one then has
$R^{i}\pi _* \mathcal {K}_{Y,k}^{p}=0$ for
$i>0$, and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU49.png?pub-status=live)
When $-p\geq n-{\rm dim}\,Z$, we have
$H^{q}(Y,\mathcal {K}_{Y,k}^{p})=0$ for
$q< n-{\rm dim}\,Z-1$. Indeed, this follows again from the Leray spectral sequence of
$pL(k)$ with respect to the map
$\tau$ and the fact that
$R^{0}\tau _*(pL(k))$ has nonzero cohomology only in degree
$n$, and the only nonzero other higher direct image is
$R^{n-{\rm dim}\,Z-1}\tau _*(pL(k))$, which contributes only to cohomology of degree
$\geq n-{\rm dim}\,Z-1$.
A second source of vanishing for the $E_1^{p,q}$ of this spectral sequence comes from Kodaira vanishing applied to the line bundle
$pL(k)$ on
$Y$. We observe that
$Z$ is by assumption defined by equations of degree
$\leq {(d+1)}/{2}$, so that the line bundle
$\tau ^{*}\mathcal {O}_{\mathbb {P}^{n}}(l)(-mE)$ on
$Y$ is nef and big when
$m\geq 0$ and
$l>m {(d+1)}/{2}$. Using (53), Kodaira vanishing thus tells us that
$H^{q}(Y,\mathcal {K}_{Y,k}^{p})=0$ for
$q< n$ if
$-p(d-1)-k>-p{(d+1)}/{2}$.
Summarizing, we proved that the spectral sequence (51) has vanishing as follows:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn54.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn55.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn56.png?pub-status=live)
We now complete the proof. The complex $K_{Y,k}^{*}$ of global sections of the complex
$\mathcal {K}_{Y,k}^{*}$ is the complex
$E_1^{*,0}$ of our spectral sequence, and hence its cohomology is the complex
$E_2^{*,0}$. Recall that we have to prove the vanishing of the cohomology of
$K_{Y,k}^{*}$ in strictly negative degrees. Let
$a<0$ be a fixed negative integer. There is no nonzero differential
$d_r$ with
$r\geq 2$ starting from
$E_2^{a,0}$ since
$E_r^{p,q}=0$ for
$q<0$. As
$E_\infty ^{a,0}=0$ (see 52), it follows that, if
$E_2^{a,0}$ is nonzero, there must be a nonzero differential
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU50.png?pub-status=live)
Let $p=a-r$. By the vanishing statements (54), (55), we must have
$r-1=n$ if
$-p< n-{\rm dim}\,Z$ and
$r-1\geq n-{\rm dim}\,Z-1$ if
$-p\geq n-{\rm dim}\,Z$. If
$r-1=n$, then, as
$a<0$,
${p=a-r<-n-1}$. The term
$\mathcal {K}_{Y,k}^{p}$ is then
$0$. Hence the only nontrivial differential appears when
$r-1\not =n$. But then,
$r\geq n-{\rm dim}\,Z$ and thus
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn57.png?pub-status=live)
Furthermore, using (56), $-p(d-1)-k\leq -p{(d+1)}/{2}$, that is,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn58.png?pub-status=live)
Combining (57) and (58), we proved that the existence of a nonzero $E_2^{a,0}$ for some
$a<0$ implies
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn59.png?pub-status=live)
which contradicts inequality (47). Proposition 4.1 is thus proved.
Imposing the dimension of $Z$ to be at most
$4$ (we will later choose dimension of
$Z$ to be equal to
$3$ if
$n$ is even and
$4$ if
$n$ is odd), we get the following.
Corollary 4.2 For a polynomial $f$ as in Proposition 4.1 with
$d$ dividing
$n+1$ and
${\rm dim}\,Z\leq 4$, the dimensions of the spaces
$R^{d}_f$,
$R^{2d}_f$ and
$R^{3d}_f$ are respectively equal to the dimensions of the spaces
$R^{d}_{f_{\rm gen}}$,
$R^{2d}_{f_{\rm gen}}$ and
$R^{3d}_{f_{\rm gen}}$ for generic
$f_{\rm gen}$, assuming
$d\geq 13$.
Proof. Indeed, if ${\rm dim}\,Z\leq 4$ and
$k\leq 3d$, (47) is satisfied if
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn60.png?pub-status=live)
As $n\geq d-1$ and
$d\geq 3$, (60) is satisfied if
$3d< (d-4)({(d-3)}/{2})$, hence if
$d\geq 13$.
Remark 4.3 The estimate of Corollary 4.2 is sharp only when $d=n+1$.
We now assume $f=\sum _{i=1}^{m}f_ig_i$ with
$f_i,\ g_i$ generic and
$d$,
$n$ are such that the conclusion of Corollary 4.2 holds. We observe that, with the same notation as above, as
$f$ is singular along
$Z$, one has
$J_f\subset I_Z$, and hence the Jacobian ring
$R_f^{d*}$ has
$H^{0}(Z,\mathcal {O}_Z(d*))$ as a quotient. We will use the notation
$g_{\mid Z}$ for the image of an element
$g$ in this quotient. For subspaces
$I_k^{id}\subset R^{id}_f$, let us denote
$\bar {I}_k^{id}:=I_{k\mid Z}\subset H^{0}(Z,\mathcal {O}_Z(id))$. Let us prove the following.
Lemma 4.4 Let $\phi \in R_f^{d}$ and
$I_k^{*d}\subset R_f^{*d}$ satisfy condition
$(*)$ (see (43)–(46)). Then either
${\bar {I}_{d-1}^{d}=0}$ and
$\bar {I}_{d-1}^{2d}=0$ or there exists an element
$g$ of
$R^{k}_f$ with
$k\geq d-1$, such that
$g_{\mid Z}\not =0$ and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn61.png?pub-status=live)
Proof. For a nonzero linear system $W$ on
$Z$, let us denote by
${\rm FL}(W)$ (for the ‘fixed locus’) the divisorial part of the base-locus of
$W$. We now observe that, as
$Z$ is a smooth complete intersection of dimension at least
$3$, one has
${\rm Pic}\,Z=\mathbb {Z}\mathcal {O}_Z(1)$ by the Grothendieck–Lefschetz theorem. In particular, if
$\bar {I}_l^{k}\not =0$, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU51.png?pub-status=live)
for some nonnegative integers $d_{l,k}$.
We first make the following claim.
Claim 4.5 For $d$ large enough, one has
$\bar {I}_1^{d}\not =0$, and
$d_{1,d}\leq 1$.
Proof. This is proved by a dimension argument. Indeed, it suffices to prove that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn62.png?pub-status=live)
As ${\rm dim}\,I_1^{d}={\rm dim}\,R^{d-1}_f$ by (43), one has
${\rm dim}\,\bar {I}_1^{d}\geq {\rm dim}\,S^{d-1}-{\rm dim}\,I_Z(d)$ and it thus suffices to prove that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn63.png?pub-status=live)
Recalling that $Z$ is a complete intersection of
$2m< n$ hypersurfaces defined by equations
$f_i$ of degree
$d'$ and
$g_i$ of degree
$d''$, with
$d''-1\leq d'\leq d''$, and
$d'+d''=d$, we conclude that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU52.png?pub-status=live)
and that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU53.png?pub-status=live)
Inequality (63) will thus be a consequence of
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn64.png?pub-status=live)
Inequality (64) easily follows (at least for $d$ large enough) from our conditions
$n> 2m$ and
$d'=d'' =d/2$ if
$d$ is even,
$d'=d''_1 =(d-1)/2$ if
$d$ is odd.
We now use the fact that $\bar {I}_1^{d}\bar {I}_{d-1}^{d}\subset \phi _{\mid Z}\cdot H^{0}(Z,\mathcal {O}_Z(d))$ (see (44)). Combined with Claim 4.5, this implies that, either
$\bar {I}_{d-1}=0$, or
$d_{d-1,d}\geq d-1$. Similarly, as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU54.png?pub-status=live)
we conclude that $d_{d-1,2d}=:k\geq d-1$. Finally we use the fact that
$H^{0}(Z,\mathcal {O}_Z(d))\cdot \bar {I}_{d-1}^{d}\subset \bar {I}_{d-1}^{2d}$ (see (45)) to deduce that the same
$g$ of degree
$k\geq d-1$ works for both
$\bar {I}_{d-1}^{d}$ and
$\bar {I}_{d-1}^{2d}$. Lemma 4.4 is now proved.
We now want to study, for a generic polynomial $f$ of the form
$\sum _{i=1}^{m}f_ig_i$ as above, the elements
$\phi \in R^{d}_f$ which both satisfy condition
$(*)$ and the property described in the assertion of Lemma 3.8. As Lemma 4.1 and Corollary 4.2 hold only for
$R^{d}_f$,
$R^{2d}_f$ and
$R^{3d}_f$ and not for the whole
$R^{*d}_f$ when
$f$ is singular, we are going to use only the data of the multiplication map of
$R^{*}_f$ in degree
$d$, which is described by a triple
$(R^{d}_f,R^{2d}_f,\mu )$ consisting of (isomorphism class of) two vector spaces of dimensions
${\rm dim}\,R^{d}_{f_{\rm gen}}$, respectively
${\rm dim}\,R^{2d}_{f_{\rm gen}}$, and a symmetric linear map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU55.png?pub-status=live)
We will also consider similar data $\bar {\mu }_f: \bar {R}^{d}\otimes \bar {R}^{d}\rightarrow \bar {R}^{2d}$ for quotients of
$R_f^{d*}$ and
$\mu :S^{d}\otimes S^{d}\rightarrow S^{2d}$ for the multiplication in the polynomial ring itself. We will call such data a ‘partial ring.’
We now study elements $\phi \in R^{d}_f$ satisfying the following condition
$(**)$ (satisfied by Schiffer variations, see § 3.1).
$(**)$ (i) For
$1\leq k\leq d-1$, there exist vector subspaces
$I_{k}^d\subset R^d_f$,
$I_{k}^{2d}\subset R^{2d}_f$,
$I_{k}^{3d}\subset R^{3d}_f$ satisfying condition
$(*)$ (see (43)–(46)).
(ii) Along a one-parameter family $f_t$, with
$f_0=f$ and
$({d}/{dt})(f_t)_{|t=0}=\phi$, there exist data
$I_{k,t}^{*d}\subset R_{f_t}^{*d}$,
$*=1,2,3$,
$k=1,\ldots, d-1$, associated to
$\phi _t=({d}/{dt})(f_t)\in R^{d}_{f_t}$, and also satisfying condition
$(*)$.
(iii) The (isomorphism class of the) partial ring $(R^{d}_{f_t}/I_{d-1,t}^{d}, R^{2d}_{f_t}/I_{d-1,t}^{2d},\bar {\mu })$ does not deform with
$t$.
Proposition 4.6 For $d$ sufficiently large and for a generic
$f=\sum _{i=1}^{m}f_ig_i$ as above, any
$\phi \in R^{d}_f$ satisfying
$(**)$ is a first-order Schiffer variations of
$f$.
Remark 4.7 We will also prove later on (see Lemma 4.15) that, in the situation of Proposition 4.6, for a generic first-order Schiffer variation $\phi =x^{d}$, the only spaces
$I_k^{id}$ satisfying Condition
$(*)$ with the given
$\phi$ are the spaces
$I_{x,k}^{id}=x^{k}R^{id-k}_f$, and hence are determined by
$\phi$.
The proof of Proposition 4.6 will use several preliminary lemmas.
Lemma 4.8 The assumptions being the same as in Proposition 4.6, then we have the following.
(a) If
$\bar {I}_{d-1}^{2d}=0$,
$f_t$ remains singular along
$Z$ (or rather, a subvariety deduced from
$Z$ by the action of an automorphism of
${\rm PGL}(n+1)$). In particular
$\phi _{\mid Z}=0$.
(b) If
${\rm FL}(\bar {I}_{d-1}^{2d})$ is defined by
$g\in H^{0}(Z,\mathcal {O}_Z(k))$,
$f_t$ remains (modulo the action of
${\rm PGL}(n+1)$) singular along the locus
$Z_g:=\{g=0\}\subset Z$.
(c) If
$k\geq d$ in (b),
$f_t$ remains (modulo the action of
${\rm PGL}(n+1)$) singular along
$Z$.
Proof. (a) If $\bar {I}_{d-1}^{2d}=0$, the partial ring
$(R^{d}_{f}/I_{d-1}^{d}, R^{2d}_{f}/I_{d-1}^{2d},\bar {\mu }_f)$ admits the partial ring
$(H^{0}(Z,\mathcal {O}_Z(d)), H^{0}(Z,\mathcal {O}_Z(2d)),\mu _Z)$ as a quotient. As by assumption, the quotient
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU56.png?pub-status=live)
of $(R^{d}_{f_t}, R^{2d}_{f_t},{\mu }_{f_t})$ is isomorphic to
$(R^{d}_f/I_{d-1}^{d}, R^{2d}_f/I_{d-1}^{2d},\bar {\mu })$, we conclude that the partial ring
$(R^{d}_{f_t},R^{2d}_{f_t},\mu _{f_t})$ also admits the partial ring
$(H^{0}(Z,\mathcal {O}_Z(d)), H^{0}(Z,\mathcal {O}_Z(2d)),\mu _Z)$ as a quotient. Denoting by
$\alpha _t: S^{*d}\rightarrow H^{0}(Z,\mathcal {O}_Z(*d))$ the quotient map for
$*=1,\,2$, this means that we have a commutative diagram,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn65.png?pub-status=live)
where $\alpha _t$ is surjective with kernel containing
$J_{f_t}$, since it factors through
$R_{f_t}$. The map
$\alpha _t$ gives an embedding
$j_t$ of
$Z$ in
$\mathbb {P}((S^{d})^{*})$. As the quadrics in
${\rm Ker}\,\mu$ are the defining equations for the
$d$th Veronese embedding
$V_d(\mathbb {P}^{n})$ in
$\mathbb {P}((S^{d})^{*})$, one concludes that
$j_t$ factors through an embedding
$j'_t$ of
$Z$ in
$\mathbb {P}^{n}=\mathbb {P}((S^{1})^{*})$, that is
$j_t=V_d\circ j'_t$. As
$Z\stackrel {j'_0}{\hookrightarrow }\mathbb {P}^{n}$ is the natural embedding of a complete intersection in
$\mathbb {P}^{n}$ of dimension
$>0$, the small deformations of the morphism
$j'_0: Z\rightarrow \mathbb {P}^{n}$ are induced by the action of
${\rm PGL}(n+1)$. Hence for
$t$ close to
$0$,
$j'_t$ is, up to the action of
${\rm PGL}(n+1)$, the original embedding. Finally, as the map
$\alpha _t=(j'_t)^{*}$ contains
$J_{f_t}$ in its kernel,
$J_{f_t}$ vanishes on
$j'_t(Z)$, which means that
$f_t$ is singular along
$j'_t(Z)$.
(b) We know that for $t=0$, and
$*=1,\,2$,
$(I_{d-1}^{*d})_{\mid Z}$ is contained in the ideal generated by
$g$. It follows that the partial ring
$(R^{d}_{f}/I_{d-1}^{d}, R^{2d}_{f}/I_{d-1}^{2d},\bar {\mu }_f)$ has the partial ring
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU57.png?pub-status=live)
as a quotient, where $Z_g:=\{g=0\}\subset Z$. We can then argue exactly as before, using the fact that
$Z_g$ is a complete intersection of strictly positive dimension in
$\mathbb {P}^{n}$. We then conclude that
$f_t$ is singular along
$j'_t(Z_g)\subset \mathbb {P}^{n}$ and that the embedding
$j'_t$ of
$Z_g$ in
$\mathbb {P}^{n}$ is deduced from
$j'_0$ by the action of an element of
${\rm PGL}(n+1)$.
(c) By (b), we know that, modulo the action of ${\rm PGL}(n+1)$,
$f_t$ remains singular along the hypersurface
$\{g=0\}$ in
$Z$. As the singular locus of
$f_t$ is defined by the partial derivatives of
$f_t$ which are degree
$d-1$ polynomials, and
$Z$ is smooth connected, we conclude, when
${k={\rm deg}\,g\geq d}$, that the partial derivatives of
$f_t$ vanish along
$Z$, which proves (c).
We next make the following observation.
Lemma 4.9 Let $Z$ be a smooth complete intersection of dimension
$\geq 3$ of hypersurfaces
$X_{h_j}$ of degrees
$d_j\geq 2$ and let
$f_t$ be a polynomial of degree
$d$ such that
$f_t$ is singular along
$Z_g$ for some
$0\not =g\in H^{0}(\mathcal {O}_Z(k))$ with
$k\geq d-1$. Then either
$f_t$ is singular along
$Z$ or there exists an element
$x\in S^{1}$ such that
$g=x^{d-1}_{\mid Z}$ and
$f_t-\alpha _t x^{d}$ is singular along
$Z$ for some scalar
$\alpha _t$.
Proof. We first claim that if $f_{t\mid Z}=0$, then
$f_t$ is singular along
$Z$. This is proved as follows. As
$f_{t\mid Z}=0$, we can write
$f_t=\sum _ja_j h_j$, with
${\rm deg}\,a_j=d-d_j$. As
$Z$ is smooth, the differential of
$f_t$ vanishes at a point
$z\in Z$ if and only if all
$a_j$ vanish at
$z$. As the
$a_j$ are of degree
$< d-1$ and
${\rm deg}\,g\geq d-1$, the vanishing of
$df_t$ along
$Z_g$ implies the vanishing of
$df_t$ along
$Z$.
Next, if $k\geq d$, we conclude that the partial derivatives of
$f$ vanish identically along
$Z$, since they vanish along
$Z_g$, so
$f$ is singular along
$Z$. We can thus assume that
$f_{\mid Z}\not =0$ and
$k=d-1$.
We then claim that there exists an $x\in S^{1}$ and a scalar
$\alpha _t$ such that
$(f_t-\alpha _tx^{d})_{\mid Z}=0$ and
$g= x^{d-1}_{\mid Z}$. We use here the fact that
${\rm dim}\,Z\geq 3$ so that
${\rm Pic}\,Z=\mathbb {Z}\mathcal {O}_Z(1)$. We decompose
$g\in H^{0}(Z,\mathcal {O}_Z(d-1))$ into irreducible factors as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU58.png?pub-status=live)
where $\gamma _j\in H^{0}(Z,\mathcal {O}_Z(d_j))$ and
$\sum _ja_jd_j=d-1$. Now if
$f_{t\mid Z}$ vanishes to order
$b_j$ along
$\{\gamma _j=0\}$,
$df_{t}$ vanishes to order
$\leq b_j-1$ along
$\{\gamma _j=0\}$. We thus conclude that
$b_j\geq a_j+1$. As
$\sum _jb_j\leq d$ and
$\sum _ja_j=d-1$, we conclude that there is a single
$j$ and the corresponding
$a_j$ equals
$d-1$, which proves that
$g= x^{d-1}_{\mid Z}$ for some
$x\in S^{1}$. It follows that
$f_{\mid Z}$ vanishes along
$x^{d-1}_{\mid Z}=0$ and the fact that the derivatives of
$f$ also vanish along
$x^{d-1}_{\mid Z}=0$ implies that
$f_{\mid Z}$ is proportional to
$x^{d}_{\mid Z}$, proving the second claim.
The second claim finally implies Lemma 4.9 since $f_t-\alpha _t x^{d}$ vanishes along
$Z$ and is also singular along
$Z_g$, with
$g=x^{d-1}_{\mid Z}$, so that the first claim applies to show that
$f_t-\alpha _tx^{d}$ is singular along
$Z$.
Proof Proof of Proposition 4.6
With the notation and assumptions of Proposition 4.6, Lemma 4.8 tells us that, modulo the action of ${\rm GL}(n+1)$, we can assume
$f_t$ is singular along
$Z$ or
$f_t$ is singular along
$Z_g$. Lemma 4.9 then says that, for some
$x\in S^{1}$,
$g=x^{d-1}_{\mid Z}$ and
$f_t-\alpha _t x^{d}$ is singular along
$Z$ for any
$t$, and the same is true for
$\phi =({\partial f_t}/{\partial t})_{\mid t=0}$. It follows that either (i)
$\phi \in I_Z^{2}(d)$ or (ii)
$\phi -x^{d}\in I_Z^{2}(d)$.
We use now the fact (this is (44) in condition $(*)$) that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn66.png?pub-status=live)
for $1\leq i\leq d$, with
$I_k^{d}\subset R^{d}_f$ of dimension equal to
${\rm dim}\,R^{d-k}_f$ (this is (43) in condition
$(*)$). We previously used this condition only for
$k=1$. We are going to use it for
$k=3$ to prove the following claim which excludes case (i).
Claim 4.10 For $d$ large enough and
$f,\,Z$ generic, a nonzero element
$\phi \in R^{d}_f$ satisfying condition
$(*)$ for adequate spaces
$I_k^{*d}\subset R^{*d}_f$ cannot belong to
$I_Z^{2}(d)$.
Proof. We argue by contradiction and assume that $\phi \in I_Z^{2}(d)$ and its image in
$R^{d}_f$ satis- fies condition
$(*)$. First of all, we use the same dimension arguments as in the proof of Claim 4.5 to show that
$\bar {I}^{d}_3:=({I}_3^{d})_{\mid Z}\not =0$. More precisely, we can show that it is of dimension
$>h^{0}(Z,\mathcal {O}_Z(d-4))$, at least if
$d$ is large enough. As
$\phi \in I_Z^{2}(d)$, we have in particular
$\phi _{\mid Z}=0$, and thus
$I_{d-3}^{d}\subset I_Z(d)$ since
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn67.png?pub-status=live)
On the one hand, as ${\rm dim}\,I_{d-3}^{d}={\rm dim}\,S^{3}$ for
$d>4$, and
${\rm dim}\,I_Z^{2}(d)<{\rm dim}\,S^{3}$,
$I_{d-3}^{d}$ is not contained in
$I_Z^{2}(d)$. On the other hand, if we look at the image of
$I_{d-3}^{d}$ in
$I_Z(d)/(I_Z^{2}(d)+J_f^{d})$, it is annihilated by multiplication by elements of
$\bar {I}_3^{d}$ acting by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU59.png?pub-status=live)
This follows indeed from the condition that $\phi \in I_Z^{2}(d)$ and (67). Now, writing
$f=\sum _jf_jg_j$ with
${\rm deg}\,f_j=d'$ and
${\rm deg}\,g_j=d''$ , we have a graded isomorphism (given by differentiation along
$Z$)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU60.png?pub-status=live)
which to $\sum _ja_jf_j+b_jg_j$ associates
$(a_{j|Z},b_{j|Z})_{j=1,\ldots,m}$. By the Leibniz rule, this isomorphism maps
${\partial f}/{\partial X_i}\in J_f$ to the
$2m$-uple
$({\partial g_j}/{\partial X_i},{\partial f_j}/{\partial X_i})_{j=1,\ldots,m}$. In other words, observing that we have a natural isomorphism
$N_{Z/\mathbb {P}^{n}}\cong N_{Z/\mathbb {P}^{n}}^{*}(d)$ given by the quadratic form defined as the Hessian of
$f$ along
$Z$, we have on the one hand the composite morphism
$\mathcal {I}_Z\rightarrow N_Z^{*}\cong N_Z(-d)$ and on the other hand the normal bundle sequence of
$Z$
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn68.png?pub-status=live)
Then the computation above shows that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn69.png?pub-status=live)
and these isomorphisms are compatible with the multiplication map by $b\in H^{0}(Z,\mathcal {O}_Z(d))$. Finally, the exact sequence (68) together with the fact that
${\rm dim}\,Z\geq 3$ show that the right hand side in (69) is isomorphic to
$H^{1}(Z,T_Z)$ for
$*=d$. Let now
$w\in I_{d-3}^{d}\subset I_Z(d)$ such that
$w\not =0$ in
$I_Z(d)/(J_f^{d}+I_Z^{2}(d))$. Then
$w$ has a nonzero image
$\bar {w}\in H^{1}(Z,T_Z)$ and
$\bar {w}$ is annihilated by multiplication by any
$b\in \bar {I}_3^{d}\subset H^{0}(Z,\mathcal {O}_Z(d))$, that is,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn70.png?pub-status=live)
for any $b\in \bar {I}_3^{d}$. The extension class
$\bar {w}\in H^{1}(Z,T_{Z})$ determines a vector bundle
$F$ on
$Z$ which fits in an exact sequence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn71.png?pub-status=live)
and the condition (70) says equivalently that $\bar {I}_3^{d}\subset H^{0}(Z,\mathcal {O}_Z(d))$ lifts to sections of
$F(d)$. Let
$\mathcal {G}\subset F(d)$ be the coherent subsheaf generated by the global sections of
$F(d)$. Observe that
${\rm det}\,\mathcal {G}=\mathcal {O}_Z(k)$ with
$k\geq 0$ since
${\rm Pic}\,Z=\mathbb {Z}\mathcal {O}_Z(1)$ and
$\mathcal {G}$ is generated by its sections. Assume first that
$\mathcal {G}$ has rank
$1$. Then we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU61.png?pub-status=live)
and we already noted that the right hand side is $>h^{0}(Z,\mathcal {O}_Z(d-4))$. It follows that
$k\geq d-3$, and that for some
$0\not =\sigma \in h^{0}(Z,\mathcal {O}_Z(3))$, one has
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn72.png?pub-status=live)
Equation (72) says that $\bar {w}$ is coming from a section of
$H^{0}(Z_\sigma,T_{Z\mid Z_\sigma }(3))$, where
${Z_\sigma :=\{\sigma =0\}\subset Z}$. A dimension count shows that for
$d$ large enough and
$Z$ generic as above, there does not exist a cubic section
$Z_\sigma$ of
$Z$ and a nonzero section of
$T_{Z\mid Z_\sigma }(3)$. This case is thus ruled out. We thus conclude that the rank of
$\mathcal {G}$ is at least
$2$. We then get a contradiction as follows. Let now
$\mathcal {G}':={\rm Ker}\,\alpha : (\mathcal {G}\rightarrow \mathcal {O}_Z(d))$, where the morphism
$\alpha$ is the restriction to
$\mathcal {G}$ of the morphism
$F(d)\rightarrow \mathcal {O}_Z(d)$ deduced from the exact sequence (71). By this exact sequence,
$\mathcal {G'}$ is a subsheaf of
$T_Z(d)$ and we have
${\rm det}\,\mathcal {G}'=\mathcal {O}_Z(k')$ with
$k'\geq -d$. Thus the slope of
$\mathcal {G}'$ is at least
$-d$. Recall that
$Z$ is a complete intersection of
$m$ hypersurfaces of degree
$d'$ and
$m$ hypersurfaces of degree
$d''$ with
$d'+d''=d$ and that
$s:={\rm dim}\,Z$ is equal to
$3$ or
$4$. It follows that
$n=2m+s$ and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU62.png?pub-status=live)
It follows that ${\rm det}\,T_Z(d)=\mathcal {O}_Z(-m(d-2)+s+1+sd)$ and for
$-m(d-2)+s+1+sd<0$, the slope of
$T_Z(d)$ is thus at most
${(-m(d-2)+s+1+sd)}/{4}\leq {(-m(d-2)+5+4d)}/{4}$. Hence we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU63.png?pub-status=live)
which holds if $m\geq 10,\ d\geq 13$. This gives a contradiction for
$d$ large enough since
$Z$ is a variety with ample canonical bundle, and hence has stable tangent bundle by [Reference BogomolovBog78], [Reference YauYau78] or [Reference TsujiTsu88]. The claim is thus proved.
We are thus in case (ii), that is,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn73.png?pub-status=live)
for some $\alpha \in I_Z^{2}(d)$, and we need to show that, in fact,
$\phi = x^{d}$ mod
$J_f^{d}$. We start with the following lemma, where we use again the notation
$\bar {I}_{k}^{d}:=(I_{k}^{d})_{\mid Z}$.
Lemma 4.11 One has $\bar {I}_{1}^{d}\subset xH^{0}(Z,\mathcal {O}_Z(d-1))$.
Proof. As $\phi _{\mid Z}=x^{d}_{\mid Z}$, we have, by (44) and (46) of condition
$(*)$ followed by restriction to
$Z$,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn74.png?pub-status=live)
If $\bar {I}_{1}^{d}\not \subset xH^{0}(Z,\mathcal {O}_Z(d-1))$, then (74) imply that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn75.png?pub-status=live)
By Lemma 4.8(c), this implies that $f_t$ remains singular along
$Z$. Thus
$f_t\in I_Z^{2}(d)$ and
$\phi \in I_Z^{2}(d)$, contradicting (73).
Corollary 4.12 Let $I'_1\subset I_1^{d}$ be defined by
$I'_1=I_1^{d}\cap xR^{d-1}_f$, and let
$\bar {I}'_1:=(I'_1)_{\mid Z}\subset xH^{0}(Z,\mathcal {O}_Z(d-1))$. Then for
$d$ (hence also
$n$) large enough
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU64.png?pub-status=live)
Proof. Indeed, as $\bar {I}^{d}_1\subset xH^{0}(Z,\mathcal {O}_Z(d-1))$, we have
$I^{d}_1\subset xR^{d-1}_f+I_Z(d)$, hence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU65.png?pub-status=live)
which implies a fortiori
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU66.png?pub-status=live)
Using the fact that ${\rm dim}\,I_1^{d}={\rm dim}\,R^{d-1}_f$ (see (43) in condition
$(*)$), the inequality
${\rm dim}\,\bar {I}'_1>h^{0}(Z,\mathcal {O}_Z(d-2))$ is then proved for
$d$ large enough in the same way as the inequality (62) proved in Claim 4.5.
We come back to our $\phi =x^{d}+\alpha$ satisfying condition
$(**)$, with
$0\not =x\in S^{1}$, and
$\alpha \in I_Z^{2}(d)$. By (74) and using the fact that
${\rm dim}\,\bar {I}_1^{d}> h^{0}(Z,\mathcal {O}_Z(d-2))$, we conclude that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU67.png?pub-status=live)
so that we can write, for any $w\in I_{d-1}^{d}$,
$w=x^{d-1}y+k_y$, where
$y\in S^{1}$ and
$k_y\in I_Z(d)$ mod
$J_f^{d}$. For
$a=xa'\in I_1'\subset I_{1}^{d}\subset R^{d}_f$, and
$w\in I_{d-1}^{d}$, we then have by (44) and (46) in condition
$(*)$ and recalling that
$\phi =x^{d}+\alpha$,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn76.png?pub-status=live)
Restricting to $Z$, and using the fact that
$k_y\in I_Z(d),\,\alpha \in I_Z^{2}(d)$, we get
$(\gamma _{a',w})_{\mid Z}=a'_{\mid Z}y_{\mid Z}$, which we write
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU68.png?pub-status=live)
for some $\gamma '_{a',w}\in I_Z(d)$ which depends linearly on
$a'$, for
$w$ fixed.
We now use again the observation that ${\rm dim}\,I_Z(d)$ is (asymptotically) small compared to
$h^{0}(Z,\mathcal {O}_Z(d-2))$ and conclude that for
$a'$ in a subspace
$I_1''\subset I'_1$ such that
${\rm dim}(I''_1)_{\mid Z}> h^{0}(Z,\mathcal {O}_Z(d-2))$, one can take
$\gamma '_{a',w}=0$ in
$R^{d}_f$, so that (76) becomes
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU69.png?pub-status=live)
that is,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn77.png?pub-status=live)
The right hand side belongs to $(I_{Z}^{2}(2d)+J_f^{2d})/J_f^{2d}$. We argue now as in the proof of Claim 4.10 to deduce that
$k_y\in (I_{Z}^{2}(d)+J_f^{d})/J_f^{d}$. Indeed, we consider the image
$\bar {k}_y$ of
$k_y$ in
$I_Z(d)/(I_Z^{2}+J_f^{d})$ and (77) says that it is annihilated by multiplication by
$xa'$ for
$xa'\in I_1'$, which is of large dimension. Then we conclude that
$\bar {k}_y=0$.
Equations (77) are thus relations in $(I_Z^{2}+J_f)/J_f$. We claim that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn78.png?pub-status=live)
Indeed, recall from the proof of Claim 4.10 that the image of $J_f^{*}$ in
$I_Z(*)/I_Z^{2}$ identifies naturally with the image of
$H^{0}(\mathbb {P}^{n}, T_{\mathbb {P}^{n}}(*-d+1))$ in
$H^{0}(Z,N_Z(*-d+1))$. We have the exact sequence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU70.png?pub-status=live)
and we observe as in the proof of Claim 4.10 that the stability of the tangent bundle of $Z$ implies that
$h^{0}(Z,T_Z(d))=0$ for
$d$ (hence
$n$) large enough. It follows that the map
$H^{0}(Z,T_{\mathbb {P}^{n}\mid Z}(d))\rightarrow H^{0}(Z,N_Z(d))$ is injective, and thus
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU71.png?pub-status=live)
comes from $H^{0}(\mathbb {P}^{n},T_{\mathbb {P}^{n}}\otimes \mathcal {I}_{Z}(d))$, which proves (78). We thus conclude that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU72.png?pub-status=live)
which is, for $d$ (hence
$n$) large enough, much smaller than
${\rm dim}\,I''_1$. It follows that, taking representatives of
$k_y,\,\alpha$ in
$I_Z^{2}(d)$, (77) provides an actual vanishing
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn79.png?pub-status=live)
for some nonzero $a''\in S^{d}$, which implies that
$xk_y=\alpha y$ in
$I_Z^{2}(d+1)$. Using the fact that the space of
$(y,k_y)$ satisfying this property has dimension
$\geq n+1$, we conclude that
$k_y=0$ for generic
$(y,k_y)$ and thus
$\alpha =0$. Proposition 4.6 is now proved.
Remark 4.13 Note that, in turn, $\alpha =0$ and (79) imply that
$k_y=0$, so that we also proved that
$I_{d-1}^{d}=x^{d-1}S^{1}$ mod
$J_f^{d}$. This will be used below.
4.2 Proof of Theorem 0.7
We complete in this section the proof of Theorem 0.7. We start by establishing the following.
Proposition 4.14 Let $f$ be a generic homogeneous polynomial of degree
$d$ in
$n+1$ variables with
$d$ dividing
$n+1$ and
$d$ large enough. Let
$\phi \in R^{d}_f$,
$I_k^{*d}\subset R^{*d}_f$, for
$*\leq 3$ and
$1\leq k\leq d-1$ satisfy condition
$(**)$ of § 4.1. Then
$\phi$ is a (first-order) Schiffer variation of
$f$.
Proof. Proposition 4.6 proves Proposition 4.14 when $f=\sum _{j=1}^{m}f_jg_j$ is the singular polynomial used in previous section. It thus remains to see that this implies the same result for the generic
$f$. This almost follows because the condition
$(**)$ is closed on
$(f,\,\phi )$ once the dimensions of the spaces
$R^{d}_f$,
$R^{2d}_f$,
$R^{3d}_f$ remain respectively equal to the dimensions of the spaces
$R^{d}_{f_{\rm gen}}$,
$R^{2d}_{f_{\rm gen}}$,
$R^{3d}_{f_{\rm gen}}$ for the generic
$f_{\rm gen}$, which is guaranteed for
$d$ large enough by Lemma 4.1. This is not completely true because we did not prove the statement of Proposition 4.6 schematically for the special
$f$. In fact, what we have to do in order to finish is to prove the following complement to Proposition 4.6.
Lemma 4.15 Let the notation and assumption on $f,\,\phi$ be as in Proposition 4.6. Assume moreover that
$\phi =x^{d}$ in
$R^{d}_f$, with
$x$ generic in
$S^{1}$. Then
$I_k^{d}=x^{k}R^{d-k}_f$ for
$1\leq k\leq d-1$.
Furthermore, inside $\mathbb {P}(R^{d}_f)\times \prod _{k=0}^{d-1}{\rm Grass}(r_k,R^{d}_f)$, where
$r_k:={\rm dim}\,R^{d-k}_f$, the set of points
$(\phi =x^{d},I_{k}^{d}=x^{k}R^{d-k}_f)$, is schematically defined, at least at its generic point, by the condition
$(*)$.
Remark 4.16 The spaces $I_k^{2d},\ I_k^{3d}$ are defined by the spaces
$I_k^{d}$ using (45), so we can consider condition
$(*)$ as a condition on
$(\phi,\,I_{k}^{d})$ only.
Remark 4.17 We did not use up to now (46) of Condition $(*)$. We will need it for the proof of this lemma.
Proof Proof of Lemma 4.15
We already noted in Remark 4.13 that $I_{d-1}^{d}=x^{d-1}R^{d-1}_f$. We study again the equations
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn80.png?pub-status=live)
for $a$ in a subspace
$I_1^{d}\subset R^{d}_f$ of dimension
${\rm dim}\,R^{d-1}_f$, and
$w$ in
$I_{d-1}^{d}=x^{d-1} S^{1}$. We already proved in Lemma 4.11 that
$\bar {I}_1^{d}=I^{d}_{1\mid Z}\subset xH^{0}(Z,\mathcal {O}_Z(d-1))$. We thus conclude that elements
$a\in I_1^{d}$ can be written as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU73.png?pub-status=live)
with $a'\in S^{d-1},\ k_a\in I_Z(d)$. Restricting (80) to
$Z$, we also get
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU74.png?pub-status=live)
for some $\gamma '\in I_Z(d)$. Equation (80) then becomes
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn81.png?pub-status=live)
where $w'$ is generic in
$S^{1}$. One then easily concludes that
$k_a=0$ mod
$\langle xS^{d-1},J_f\rangle$, that is,
$I_1^{d}\subset xR^{d-1}_f$. Hence we proved (by dimension reasons) that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn82.png?pub-status=live)
We now use (46). We get
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU75.png?pub-status=live)
which provides, using (82) $x^{2}R^{2d-2}_f\subset I_2^{2d}$. Using (43), this inclusion gives in turn, by dimension reasons,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn83.png?pub-status=live)
Here, in order to apply the dimension argument, we need to know that multiplication by $x^{2}$ is injective on
$R^{2d-2}_f$. More generally we will need to know that multiplication by
$x^{i}$ is injective on
$R^{d+i}_f$ for
$1\leq i\leq d$, which is not hard to prove since
$x$ is generic. We next use (45)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU76.png?pub-status=live)
that is,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU77.png?pub-status=live)
and easily conclude that $I_2^{d}\subset x^{2}R^{d-2}_f$, hence
$I_2^{d}= x^{2}R^{d-2}_f$ by dimension reasons. We continue this way and prove that
$I_{k}^{d}=x^{k}R^{d-k}_f$ for all
$1\leq k\leq d-1$. Thus the first statement is proved.
In order to prove the schematic statement, we consider a first-order variation $(h,h_1,\ldots,h_{d-1})$ of
$(x^{d},I^{d}_1,\ldots,I^{d}_{d-1})$ satisfying conditions
$(*)$ at first order. We thus have a first-order deformation
$x^{d}+\epsilon h\in R^{d}_f$ of
$x^{d}$ and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU78.png?pub-status=live)
satisfying the infinitesimal version of (44)–(46). We have to prove that there is a $y\in S^{1}/x$ such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU79.png?pub-status=live)
is given by multiplication by $lyx^{l-1}$. We first observe that it suffices to prove the result for
$l=1$, because the reasoning above, which deduces the equality
$I_{k}^{d}=x^{k}R^{d-k}_f$ for all
$1\leq k\leq d-1$ from the equality (82) using (45) and (46) work as well schematically.
We thus have a first-order deformation $x^{d}+\epsilon h\in R^{d}_f$ of
$x^{d}$ and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU80.png?pub-status=live)
satisfying the equations
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn84.png?pub-status=live)
for any $a=x a'\in xR^{d-1}_f$,
$w=x^{d-1}w'\in x^{d-1}S^{1}$, and for
$\gamma _\epsilon =\gamma +\epsilon \gamma _1$, where
$\gamma$ is as in (80). We want to prove that there exists
$y\in S^{1}/\langle x\rangle$, such that for any
$a=xa'$, the following holds in
$R^{d}_f$
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn85.png?pub-status=live)
Looking at the previous proof, we deduce from (81) with $k_a=0$ that
$\gamma '=0$ (using injectivity of the multiplication by
$x^{d}$), so
$\gamma =a'w'$ in
$R^{d}_f$. We thus have
$\gamma _\epsilon =a'w'+\epsilon \gamma _1$. Equation (84) then gives
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn86.png?pub-status=live)
for any $a=x a'\in xR^{d-1}_f$,
$w=x^{d-1}w'\in x^{d-1}S^{1}$. We claim that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn87.png?pub-status=live)
Indeed, (86) first implies that $h=xh'$ since it becomes divisible by
$x$ after multiplication by any element of
$R^{d}_f$, and then, after simplification by
$x$, that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn88.png?pub-status=live)
We rewrite (88) in the form
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn89.png?pub-status=live)
We now restrict (89) to $Z$. As
$x^{d-2}$ and
$a'$ have no common divisor on
$Z$ for
$a'$ generic, it follows that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU81.png?pub-status=live)
which proves (87) since $w'\in S^{1}$ is generic.
We can even conclude by similar arguments that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU82.png?pub-status=live)
for some $m_1\in H^{0}(Z,\mathcal {O}_Z(1))$. We can see
$m_1$ as an element
$y\in S^{1}$ because the map of restriction to
$Z$ is an isomorphism in degree
$1$, and we can thus write in
$R_f^{d}$
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn90.png?pub-status=live)
where $k_1(a')\in I_Z(d)$ for any
$a'\in R^{d-1}_f$. Equation (89) then gives
$x^{d-2}((y a'+k_1(a'))w'-x\gamma _1)+a'(h_{d-1}(w)-h' w')=0$ in
$R^{2d-1}_f$, that is
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqn91.png?pub-status=live)
The term $x^{d-2}(k_1(a')w'-x\gamma _1)$ belongs by (91) to
$x^{d-2}I_Z(d+1)\cap \langle a'\rangle$. For
$a'$ generic, it is easy to show that it implies that it belongs to
$a'x^{d-2}I_Z(2)=0$. Thus
$x^{d-2}(k_1(a')w'-x\gamma _1)=0\ {\rm in}\ R^{2d-1}_f$, hence
$k_1(a')w'-x\gamma _1=0$, and
$k_1(a')=0$ mod
$\langle x\rangle$. This is true for
$a'$ generic in
$R^{d-1}_f$, hence for all
$a'$. Thus (85) is proved.
Lemma 4.15 is a schematic version of Proposition 4.6 that guarantees that the Veronese image $v_f(\mathbb {P}(S^{1}))\subset \mathbb {P}(R^{d}_f)$ is characterized not only set theoretically but also schematically (at the generic point) by condition
$(**)$ (in fact, we can see from the proof above that condition
$(*)$ even suffices for the scheme-theoretic statement, but condition
$(**)$ was needed to prove the set-theoretic statement for the special
$f$). It follows that for generic
$f_{\rm gen}$, the Veronese image
$v_f(\mathbb {P}(S^{1}))\subset \mathbb {P}(R^{d}_f)$ is also characterized by condition
$(**)$.
Proof Proof of Theorem 0.7(1)
Fix integers $d,\,n$ with
$d$ dividing
$n+1$, and for which the conclusion of Proposition 4.14 holds. We want to show that if
$X_f$ is a very general hypersurface of degree
$d$ in
$\mathbb {P}^{n}$, then any smooth hypersurface
$X_g$ of degree
$d$ in
$\mathbb {P}^{n}$ such that there exists an isomorphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU83.png?pub-status=live)
of rational Hodge structures, is isomorphic to $X_f$.
We first argue as in § 1.1. Denote by $U_{d,n}^{0}\subset U_{d,n}$ the Zariski open set parameterizing automorphisms free smooth hypersurfaces. As
$f$ is very general,
$f\in U_{d,n}^{0}$ and our assumption provides simply connected Euclidean open neighborhoods
$U\subset U_{d,n}^{0},\,V\subset U_{d,n}^{0}$ of
$f$,
$g$ respectively, a holomorphic diffeomorphism
$i:U\cong V$ with
$i(f)=g$, and an isomorphism of complex variations of Hodge structures
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU84.png?pub-status=live)
on $U$. Here, if
$\pi :\mathcal {X}_{d,n}\rightarrow U_{d,n}^{0}$ is the universal hypersurface,
$H^{n-1}_\mathbb {C}$ is the local system
$R^{n-1}\pi _*\mathbb {C}_{{\rm prim}}$ on
$U_{d,n}^{0}$, and
$F^{\cdot }\mathcal {H}^{n-1}$ is the Hodge filtration on the associated flat holomorphic vector bundle
$\mathcal {H}^{n-1}=H^{n-1}_\mathbb {C}\otimes \mathcal {O}_{U_{d,n}^{0}}$.
The differential $i_*:T_{U,f}\rightarrow T_{V,g}$ is a linear isomorphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20230406195754967-0150:S0010437X21007727:S0010437X21007727_eqnU85.png?pub-status=live)
Claim 4.18 In the situation described above, the differential $i_*$ sends the set of first-order Schiffer variations of
$f$ to the set of first-order Schiffer variations of
$g$.
Proof. Indeed, the local diffeomorphism $i$ induces an isomorphism of variations of Hodge structures. It thus sends a one-parameter Schiffer variation
$(f_t)_{t\in \Delta }$ of
$f$ to a one-parameter variation
$(g_t)_{t\in \Delta }$,
$g_t:=i(f_t)$, of
$g$, which satisfies the assumptions of Proposition 4.14. Proposition 4.14 then tells us that
$\psi :=({\partial g_t}/{\partial t})_{\mid t=0}$ is a first-order Schiffer variation of
$g$. But
$\phi :=({\partial f_t}/{\partial t})_{\mid t=0}$ is an arbitrary first-order Schiffer variation of
$f$ and we have
$\psi =i_*(\phi )$.
Having the claim, the proof of the theorem is finished using Proposition 3.6.
Acknowledgements
I thank Nick Shepherd-Barron for reminding me that the Donagi method a priori works only starting from $n=3$ and that Theorem 0.2 is actually wrong for
$(d,n)=(4,2)$. I also thank the referees for their very careful reading and constructive criticism. This work was started at MSRI during the program ‘Birational Geometry and Moduli Spaces’ in the Spring 2019. I thank the organizers for inviting me to stay there and the Clay Institute for its generous support.