1 Introduction
Let
$X=V(f)$
be a hypersurface of the projective space
${\mathbb {P}}^n$
over an algebraically closed field
${\mathbb {K}}$
of characteristic zero. In the case where the determinant
$h_f=\det (H_f)$
of the associated Hessian matrix
$H_f$
is not equivalently zero – for example, for X smooth – it is well known that the associated Hessian hypersurface
${\mathcal {H}}_f=V(h_f)\subset {\mathbb {P}}^n$
contains much information of X itself.
A sort of generic Torelli theorem for Hessian hypersurfaces is also supposed to be valid (see [Reference Ciliberto and Ottaviani8]), up to some known cases. In particular, in [Reference Ciliberto and Ottaviani8] the so-called Hessian map

where S denotes the ring
${\mathbb {K}}[x_0,\dots ,x_n]=\oplus _{d\geq 0} S^d$
, is studied. In the specific case of cubic hypersurfaces, namely when
$d=3$
, it is known that
$h_{3,1}$
has the generic fiber of dimension
$1$
,
$h_{3,2}$
has the generic fiber consisting of
$3$
points, and
$h_{3,3}$
is birational onto its image. As in this last case, the Ciliberto-Ottaviani conjecture states that the Hessian map
$h_{d,n}$
should be birational onto its image, for higher values of n and even for different values of d. Very recently, this conjecture has been analysed for the case of plane curves in [Reference Beorchia2] and [Reference Ciliberto, Ottaviani, Caro and Duque-Rosero9].
One can argue that also the singular locus
$\operatorname {\mathrm {Sing}}({\mathcal {H}}_f)$
of the Hessian hypersurface
${\mathcal {H}}_f$
must keep track of some crucial aspects of X. Along this line, the aim of this paper is a deep study of the singularities of the hypersurface
${\mathcal {H}}_f$
associated with a smooth cubic
$(n-1)$
-fold; in particular, we are interested in the dimension of
$\operatorname {\mathrm {Sing}}({\mathcal {H}}_f)$
and in the irreducibility of
${\mathcal {H}}_f$
itself. Notice that, for low dimensional hypersurfaces (i.e.,curves or surfaces), this analysis is classical (see, for instance, [Reference Hutchinson19] and [Reference Segre28]). More recently, an approach has been developed in [Reference Adler and Ramanan1] for the case of cubic threefolds, while in [Reference Bricalli, Favale and Pirola6], the authors have dealt with the higher-dimensional cases.
To explain our main result (Theorem A), let us recall that a polynomial f (or a hypersurface
$X=V(f)$
) is said to be of Thom-Sebastiani type, TS for brevity, if up to a change of coordinates it can be written using two disjoint sets of variables (see Definition 3.1). The name comes from the works [Reference Sebastiani and Thom27] of Sebastiani and Thom and [Reference Thom29] of Thom. These polynomials have been extensively studied in several contexts (for example, about their Jacobian ideals in some classical works of Bertini, Longo and Mammana – [Reference Bertini3], [Reference Longo22] and [Reference Mammana24]) and have appeared with other names too in the literature (for example, they are called direct sums in [Reference Buczy’nska, Buczy’nski, Kleppe and Teitler7] and [Reference Fedorchuk13]). The Hessian hypersurface associated with a TS polynomial is not irreducible, and its singular locus has dimension
$n-2$
. The interesting fact is that this is actually a characterization, as proved in the following.
Theorem A (Theorem 5.1).
Assume that
$n\leq 5$
and consider
$f\in {\mathbb {K}}[x_0,\dots ,x_n]$
defining a smooth cubic. Then the Hessian hypersurface
${\mathcal {H}}_f\subset {\mathbb {P}}^n$
is irreducible and normal if and only if f is not of Thom-Sebastiani type.
The problem of determining whether a polynomial is of TS type is interesting and investigated in the literature (see, for example, [Reference Buczy’nska, Buczy’nski, Kleppe and Teitler7] or [Reference Fedorchuk13]), also from an algorithmic point of view. As said above, one can apply a sort of ‘hessian test’: if the Hessian hypersurface associated to a polynomial f is normal, then f cannot be of TS type. Furthermore, with a strong geometric approach, Theorem A guarantees that this ‘hessian test’ is actually a complete test for (smooth) cubic forms with at most 6 variables.
Beyond the smoothness hypothesis, which is anyway necessary (see Remark 5.2 for details), one could conjecture that the same result still holds for higher dimensions or higher degrees. Even if we strongly believe that Theorem A is valid for smooth cubic hypersurface of any dimension, one can see that this is not the case – for example, in degree
$4$
(see again Remark 5.2).
Besides the fact that cubic hypersurfaces are classically endowed with interesting and particular properties in relation to their geometry and also, for example, to their associated Hessian variety (see, for example, [Reference Dolgachev12], [Reference Russo26], [Reference Gondim and Russo15], [Reference Huybrechts20],...), the peculiarity of the case of (smooth) cubics lies in the framework we want to deal with and in the techniques used. Indeed, Hessian loci of cubic hypersurfaces are equipped, among other things, with a special symmetry that will be a key ingredient in the whole article and which makes a crucial difference with the higher degree cases. Indeed, if
$X=V(f)$
is a general cubic hypersurface, then its Hessian
${\mathcal {H}}_f$
is a singular Calabi-Yau variety with a fixed point free rational involution. Indeed, for
$d=3$
, one can observe that for all
$x,y\in {\mathbb {K}}^{n+1}$
,

Such a relation can be easily translated in terms of the associated apolar ring
$A_f={\mathcal {L}}/\operatorname {\mathrm {Ann}}_{{\mathcal {L}}}(f)$
, where
${\mathcal {L}}={\mathbb {K}}[\frac {\partial }{\partial x_0},\dots ,\frac {\partial }{\partial x_0}]$
, as
$x\cdot y=y\cdot x$
. Under the natural identification
${\mathbb {P}}(A^1)\simeq {\mathbb {P}}^n$
(which comes from the Gorenstein duality of the apolar ring; see [Reference Macaulay23] or the comprehensive book [Reference Harima, Maeno, Morita, Numata, Wachi and Watanabe17]), one can define the natural incidence correspondence

Observe that it dominates, via the two projections, the Hessian hypersurfaces
${\mathcal {H}}_f$
. Moreover,
$\Gamma _f$
is also equipped with an involution
$\tau (([x],[y]))=([y],[x])$
which is fixed point free and descends to the above-mentioned fixed point free rational involution defined on the Hessian hypersurface
${\mathcal {H}}_f$
.
Another key point is the determinantal structure of the Hessian hypersurface. Indeed,
${\mathcal {H}}_f$
is equipped with a natural rank-decreasing filtration

where
${\mathcal {D}}_k(f)=\{[x]\in {\mathbb {P}}^n \ | \ \text {Rank}(H_f(x))\leq k\}$
. In the case where
$V(f)$
is any smooth cubic hypersurface, the second inclusion in the above filtration is actually an equality as proved in [Reference Adler and Ramanan1] for cubic threefolds and in [Reference Bricalli, Favale and Pirola6] for
$n\geq 5$
. Moreover, the authors proved that for X general, each of the
$D_i$
’s has the expected dimension, and we computed the basic invariant in some low dimensional cases. In particular, for a smooth cubic
$V(f)$
, we have

Indeed,
${\mathcal {H}}_f$
is reduced, and the expected value (namely,
$n-3$
) is achieved for f general. Observe that in the other case, we would have that the Hessian locus
${\mathcal {H}}_f$
is not normal (indeed, see [Reference Hartshorne18, Proposition 8.23]; recall that a hypersurface in
${\mathbb {P}}^n$
is normal if and only if it is regular in codimension
$1$
). Then, with Theorem A, we partially answer the following natural question:

When
$n=2$
, given a smooth cubic curve
$X=V(f)$
, the associated Hessian curve
${\mathcal {H}}_f$
is singular if and only if X is the Fermat curve. It is remarkable that Beniamino Segre in 1943 proved that a similar result also holds for cubic surfaces in a projective
$3$
-dimensional space. Indeed (see [Reference Segre28]), given a smooth cubic surface
$X=V(f)\subset {\mathbb {P}}^3$
,
${\mathcal {H}}_f$
is reducible if and only if X is cyclic; that is, up to a change of coordinate, we can write
$f=z_0^3+g(z_1,z_2,z_3)$
.
It is a real misfortune, due likely to the war and to racial issues, that the book of Beniamino Segre which focuses on non-singular cubic surfaces, is not easy to find. Its analysis is based on the use of Sylvester’s Pentahedral Theorem which, with a modern terminology, says that the general cubic surface has Waring rank equal to
$5$
(i.e., after a suitable change of coordinates, it can be written as the zero locus of
$\sum _{i=1}^{5}L_i^3$
where
$L_1,\dots ,L_5\in S^1$
). This description has also been recently used, for example, in [Reference Dardanelli and van Geemen10] and [Reference Ciliberto and Ottaviani8]. With Theorem A, we extend Segre’s result to cubic threefolds and fourfolds.
Let us now explain the main features of our proof. First of all, let us observe that, since no useful ‘Sylvester form tool’ seems to exist for cubic forms in
${\mathbb {P}}^n$
with
$n\geq 4$
, a completely new strategy must be used. In this environment, it is a great pleasure to acknowledge our main source of inspiration: Adler’s work. In a remarkable series of appendices to the book [Reference Adler and Ramanan1], among many other results, Adler set up a method to study the singular locus of the Hessian locus
${\mathcal {H}}_f$
associated with a cubic hypersurface
$V(f)$
. He considered the correspondence
$\Gamma _f$
introduced above, which can be seen as a partial desingularization of
${\mathcal {H}}_f$
, and moreover, he proved that the singular locus of
$\Gamma _f$
has a ‘triangle structure’. More precisely, a point
$([x],[y])$
is singular for
$\Gamma _f$
if and only if there exists
$[z]\in {\mathbb {P}}^n$
such that
$([x],[y]),([x],[z]),([y],[z])\in \Gamma _f$
.
Our crucial observation is that if
$\text {Sing}({\mathcal {H}}_f)$
contains a component of dimension
$n-2$
, then the same holds also for
$\text {Sing}(\Gamma _f)$
. We have then a large amount of triangles to deal with, and moreover, such a description is greatly enlightened by using the apolar-geometric method we already exploited in our proof of a Gordan-Noether theorem (see [Reference Gordan and Nöther16], [Reference Russo26], [Reference Bricalli, Favale and Pirola5]). The whole proof is then devoted to showing that ‘too many triangles’ for
${\mathcal {H}}_f$
force f to be of TS type.
Two results can be thought of as the main ingredients to this aim. First of all, we give a characterisation of the cubic polynomials of TS type in terms of the Hessian loci
${\mathcal {D}}_k$
appearing in the above-mentioned filtration:
Theorem B (Theorem 3.4).
A polynomial
$f\in {\mathbb {K}}[x_0,\dots ,x_n]$
defining a smooth cubic is of TS type of the form
$f(x_0,\dots ,x_n)=f_1(x_0,\dots ,x_k)+f_2(x_{k+1},\dots ,x_n)$
if and only if
${\mathcal {D}}_{k+1}(f)$
contains a
${\mathbb {P}}^{k}$
.
The second result allows us to make specific assumptions on the general triangle we will deal with. In particular, by considering an irreducible component
${\mathcal {F}}$
of the variety parametrizing these triangles for
${\mathcal {H}}_f$
and denoting by
$\pi _i$
its natural projections, we have the following:
Theorem C (Theorem 4.6).
Assume
$n\leq 5$
and let
$X=V(f)$
be a smooth cubic hypersurface in
${\mathbb {P}}^n$
not of TS type. If
${\mathcal {F}}$
is an irreducible family of triangles for
${\mathcal {H}}_f$
with
$\dim (\pi _1({\mathcal {F}}))=\dim ({\mathcal {F}})=n-2$
, then the general element in
${\mathcal {F}}$
is such that none of its vertices belongs to X.
Even if, a fortiori, the situation presented in the above theorem cannot be realised, let us stress that this result will allow us to set the right framework on which all the proof is based.
The problem (after some reduction preliminaries) becomes to compute the Zariski tangent space at the general point of
${\mathcal {F}}$
. This approach leads almost immediately to a conclusion in the case of cubic surfaces, and it is reasonably accessible for
$n=4$
. In the fourfold case, the computation becomes instead much more complicated: there are really many sub-cases to be considered (this is certainly due also to the fact that for
$n\leq 4$
all the cubics of Thom-Sebastiani type are indeed cyclic, which is not true anymore for
$n\geq 5$
).
It is interesting to notice that there is a family of cubic fourfolds (which is considered in Lemma 6.4), where the infinitesimal methods are not enough in order to conclude. For these hypersurfaces, in the spirit of the possible Torelli theorem, we recover the equation of the cubic fourfold
$V(f)$
, and then, with a direct computation, we show that the dimension of the singular locus of
${\mathcal {H}}_f$
is actually the expected one (i.e.
$2$
, unless f is of TS type).
Plan of the paper
After setting the notation and proving some preliminary results in Section 2, in Section 3 we deal with polynomials of Thom-Sebastiani type, and we prove Theorem B. In Section 4, we focus on the study of particular families of triangles, and we prove Theorem C. Finally, in Sections 5 and 6, we prove our main result, namely Theorem A, respectively for
$n\leq 4$
and for the case of cubic fourfolds.
2 Preliminaries and first results
In this first section, we set the notation and present some preliminary results, some of them proved in [Reference Bricalli, Favale and Pirola6]. For a complete comprehension of standard Artinian Gorenstein Algebras, which we are going to introduce, one can refer to [Reference Harima, Maeno, Morita, Numata, Wachi and Watanabe17]. Consider
${\mathbb {K}}$
an algebraically closed field of characteristic
$0$
and the projective space
${\mathbb {P}}^n$
for
$n\geq 2$
. Let us set

so that S is the homogeneous coordinate ring of
${\mathbb {P}}^n$
and
${\mathcal {L}}$
is the graded algebra of linear differential operators on S, where we define
$y_i$
as the first partial derivative with respect to
$x_i$
; that is,

If
$v\in {\mathbb {K}}^{n+1}$
, we will denote by
${\partial _{v}}$
the derivative in the direction of v (i.e.,
$\sum _{i=0}^n v_iy_i$
).
Let us now consider a homogeneous polynomial f of degree d (i.e., an element of
$S^d$
). Two objects can then be associated with f in a natural way:
-
• the Jacobian ring of f, defined as the quotient
$R_f=S/J_f$ , where
$J_f$ denotes the Jacobian ideal of f, spanned by the partial derivatives of f;
-
• the apolar ring of f, defined as the quotient
$A_f={\mathcal {L}}/\mathrm {Ann}_{{\mathcal {L}}}(f)$ , where
$\mathrm {Ann}_{\mathcal {L}}(f)$ is the annihilator ideal of f (i.e,. the ideal in
${\mathcal {L}}$ given by
$\{\delta \in {\mathcal {L}} \ | \ \delta (f)=0\}$ ).
Both the Jacobian and the apolar ring of f are graded Artinian algebras with socle in degree respectively
$(n+1)(d-2)$
and d (i.e.,
$R_f=R^0\oplus R^1\oplus \cdots \oplus R^{(n+1)(d-2)}$
and
$A_f=A^0\oplus A^1\oplus \cdots \oplus A^d$
). One can also see that they are standard (i.e., generated in degree
$1$
) and that they satisfy the so-called Poincaré (or Gorenstein duality) – for example, the multiplication map
$A^{d-k}\times A^k\rightarrow A^d$
is a perfect pairing for every suitable positive integer k. In other words, they are both examples of what we call SAGAs, an acronym for standard Artinian Gorenstein algebras.
Finally, given f as above, we can then define the associated Hessian matrix and the hessian polynomial, respectively the square symmetric matrix whose entries are the second partial derivatives of f with respect to the
$x_i$
’s and the determinant of such a matrix; that is,

Let us observe that if the zero locus of f,
$X=V(f)\subset {\mathbb {P}}^n$
is a smooth hypersurface, then the hessian determinant
$h_f$
belongs to
$S^{(n+1)(d-2)}\setminus \{0\}$
. In this case, one can define the Hessian hypersurface
${\mathcal {H}}_f$
associated with f (or with X) as the zero locus of such a polynomial; that is,

The smoothness of
$X=V(f)$
implies also that the associated apolar ring
$A_f$
is such that
$A^1$
has dimension
$n+1$
: indeed, such a dimension is strictly smaller than
$n+1$
if and only if
$V(f)$
is a cone. From the natural pairing
$S\times {\mathcal {L}}\rightarrow S$
, one can then deduce an isomorphism

From now on, let us focus on the case
$\mathbf {d=3}$
: the first result we need to recall is the following Proposition (see [Reference Bricalli, Favale and Pirola6, Proposition 1.2]), which allows us to interpret the cubic hypersurface
$X=V(f)$
, its singular locus and its associated Hessian variety in terms of the apolar ring
$A_f$
.
Proposition 2.1. Given a cubic hypersurface
$X=V(f)$
(not a cone) and the corresponding
$A_f$
, we have
-
1. Under the identification
${\mathbb {P}}^n\simeq {\mathbb {P}}(A^1)$ ,
$H_f(x)\cdot y=\nabla (xy(f))$ , for
$x,y\in {\mathbb {K}}^{n+1}$ ;
-
2.
${\mathcal {H}}_f=\{[y]\in {\mathbb {P}}(A^1)\,|\,\, \exists \, [x]\in {\mathbb {P}}(A^1) \mbox { with } xy=0 \ in \ A^2\}$ ;
-
3.
$X=\{[y]\in {\mathbb {P}}(A^1)\,|\, y^3=0\}$ ;
-
4.
$\operatorname {\mathrm {Sing}}(X)=\{[y]\in {\mathbb {P}}(A^1)\,|\, y^2=0\}$ .
In this paper, we deal with a homogeneous cubic polynomials whose zero locus is smooth: we will denote by
${\mathcal {U}}\subset {\mathbb {P}}(S^3)$
, the locus of such elements.
As done in [Reference Bricalli, Favale and Pirola6], given
$[f]\in {\mathcal {U}}$
, let us introduce some objects which will be used extensively in what follows. First of all, for
$[x]\in {\mathbb {P}}^n$
, we set

This is either empty (exactly when
$[x]\not \in {\mathcal {H}}_f$
) or a projective linear space of dimension
$n-\operatorname {\mathrm {Rank}}(H_f(x))$
. It is then natural to consider the Hessian loci

which give a stratification of the projective space
${\mathbb {P}}^n$
and in particular of the Hessian hypersurface
${\mathcal {H}}_f$
(for example, we have
${\mathcal {D}}_{n+1}(f)={\mathbb {P}}^n$
and
${\mathcal {D}}_n(f)={\mathcal {H}}_f$
). Moreover, in general, for
$k\leq n-1$
,
${\mathcal {D}}_{k-1}(f)\subseteq {\mathcal {D}}_k(f)\subset {\mathcal {H}}_f$
. We will simply write
${\mathcal {D}}_k$
, when it is clear which polynomial we are referring to in the following. In [Reference Bricalli, Favale and Pirola6], the authors actually proved that for every integer
$k\in \{2,\dots ,n\}$
,
${\mathcal {D}}_{k-1}(f)\subseteq \operatorname {\mathrm {Sing}}({\mathcal {D}}_k(f))$
and that equality holds for
$[f]\in {\mathcal {U}}$
general. For any
$[f]\in {\mathcal {U}}$
, let us introduce a useful incidence correspondence:

and let us denote by
$\operatorname {\mathrm {pr}}_i$
the two natural projections.
Remark 2.2. By the relation
$H_f(x)\cdot y=H_f(y)\cdot x$
(which is equivalent to the relation
$xy=yx$
in
$A_f$
, by Proposition 2.1), the standard involution

sends
$\Gamma _f$
to itself. As a consequence,
$\Gamma _f$
dominates
${\mathcal {H}}_f$
via both
$\operatorname {\mathrm {pr}}_1$
and
$\operatorname {\mathrm {pr}}_2$
.
By Proposition 2.1, the loci just introduced can be described also in terms of the apolar ring as follows:

Through the article, we will use one description or the other according to the convenience.
We summarise here the main results of [Reference Bricalli, Favale and Pirola6]:
Theorem 2.3. The following hold:
-
1. for any
$[f]\in {\mathcal {U}}$ , we have
$\operatorname {\mathrm {Sing}}({\mathcal {H}}_f)={\mathcal {D}}_{n-1}(f)$ ;
-
2. if
$[f]\in {\mathcal {U}}$ is general, then
$\Gamma _f$ is smooth and
$\operatorname {\mathrm {pr}}_i:\Gamma _f\rightarrow {\mathcal {H}}_f$ is a desingularization;
-
3. the expected codimension of
${\mathcal {D}}_k(f)$ is
$\binom {n-k+2}{2}$ .
In particular, the expected dimension of
$\operatorname {\mathrm {Sing}}({\mathcal {H}}_f)$
equals
$n-3$
.
Hence, by Theorem 2.3, the Hessian variety associated with any smooth cubic hypersurface in
${\mathbb {P}}^n$
for
$n\geq 3$
is always singular, and in the general case,
$\Gamma _f$
is a desingularization for it and we have a lower bound for the dimension of
$\operatorname {\mathrm {Sing}}({\mathcal {H}}_f)$
for any
$[f]\in {\mathcal {U}}$
. We are now interested in giving an upper bound for this dimension and more generally for
$\dim ({\mathcal {D}}_k(f))$
.
Remark 2.4. It is well known that the diagonal
$\Delta _{{\mathbb {P}}^n}\subseteq {\mathbb {P}}^n\times {\mathbb {P}}^n$
has decomposition in the Chow group of
${\mathbb {P}}^n\times {\mathbb {P}}^n$
given by

where H is a hyperplane in
${\mathbb {P}}^n$
and
$\pi _i$
are the standard projections. Hence, every effective cycle of dimension at least n intersects
$\Delta _{{\mathbb {P}}^n}$
.
Proposition 2.5. Consider any
$[f]\in {\mathcal {U}}$
. Then the following hold:
-
1. the variety
$\Gamma _f$ is a connected complete intersection in
${\mathbb {P}}^n\times {\mathbb {P}}^n$ of pure dimension
$n-1$ ;
-
2. for each k, one has
$\dim ({\mathcal {D}}_k(f))\leq k-1$ ;
-
3. there is a bijective correspondence between irreducible components of
$\Gamma _f$ and the irreducible components of the various loci
${\mathcal {D}}_{k}(f)$ for which the bound in
$(b)$ is sharp.
Proof. For
$(a)$
, first of all, observe that
$\Gamma _f\cap \Delta _{{\mathbb {P}}^n}=\emptyset $
since, otherwise,
$V(f)$
would be singular by Proposition 2.1. Hence, by Remark 2.4, we have that
$\Gamma _f$
has dimension at most
$n-1$
. However, by definition, we have that
$\Gamma _f$
is cut by
$n+1$
divisors of
${\mathbb {P}}^n\times {\mathbb {P}}^n$
of bidegree
$(1,1)$
so each component of
$\Gamma _f$
has dimension at least
$n-1$
.
Since
$\Gamma _f$
is a complete intersection, its connectedness follows by the Fulton-Hansen-type theorem (see [Reference Fulton and Hansen14], [Reference Lazarsfeld21, Ch.3] or [Reference Martinelli, Naranjo and Pirola25]).
For
$(b)$
, let us assume by contradiction that there exists an irreducible component W of
${\mathcal {D}}_k(f)$
of dimension
$d\geq k$
. Over the general point
$[w]$
of W, the fiber of the projection
$\operatorname {\mathrm {pr}}_1$
from
$\Gamma _f$
is a projective space
$\iota ([x])\simeq {\mathbb {P}}^s$
with
$s\geq n-k$
. Therefore, there exists a component of
$\Gamma _f$
of dimension at least
$d+s\geq n$
. This would mean, by Remark 2.4, that
$\Gamma _f\cap \Delta _{{\mathbb {P}}^n}$
is not empty, giving a contradiction.
For
$(c)$
, assume that W is as in
$(b)$
and of dimension
$k-1$
. Then the same reasoning as above implies the existence of an irreducible component of
$\Gamma _f$
dominating W. For the converse, let G be an irreducible component of
$\Gamma _f$
, set
$G'=\operatorname {\mathrm {pr}}_1(G)$
and let m be the dimension of
$G'$
. If
$m=n-1$
, then
$G'$
is a component of
${\mathcal {H}}_f={\mathcal {D}}_n(f)$
, so we are done. Otherwise,
$m=n-1-a$
with
$a>0$
so that
$\operatorname {\mathrm {pr}}_1|_G$
has the general fiber F of dimension a. Since the whole fiber of
$\operatorname {\mathrm {pr}}_1$
over a general point of
$[x]\in G'$
is a projective space containing a fiber of dimension a, we have that
$\operatorname {\mathrm {Rank}}(H_f(x))\leq n-a$
(i.e., the general point of
$G'$
lies in
${\mathcal {D}}_{n-a}(f)$
), as claimed.
As a consequence of Proposition 2.5, recalling that
$\operatorname {\mathrm {Sing}}({\mathcal {H}}_f)={\mathcal {D}}_{n-1}(f)$
, we have in particular that

and for f general, such dimension coincides with the expected one (i.e.,
$\dim (\operatorname {\mathrm {Sing}}({\mathcal {H}}_f))=n-3$
). In this paper, we are interested in answering the following:
Question 2.6. For which
$[f]\in {\mathcal {U}}$
does
$\operatorname {\mathrm {Sing}}({\mathcal {H}}_f)$
have dimension
$n-2$
?
In this study, the singularities of
$\Gamma _f$
will play a central role.
Definition 2.7. Given
$f\in S^3$
, we set

Elements in
${\mathcal {T}}$
are called triangles for
${\mathcal {H}}_f$
.
Remark 2.8. Recall that if f is a cubic polynomial, then the incidence variety
$\Gamma _f$
is symmetric with respect to the involution
$\tau $
. This implies that

for all
$[x],[y]\in {\mathbb {P}}^n$
, so a permutation of the vertices of a triangle yields again a triangle. Furthermore, if
$[f]\in {\mathcal {U}}$
, two vertices of the same triangle cannot be equal, since we have
$\Delta _{{\mathbb {P}}^n}\cap \Gamma _f=\emptyset $
, and so by construction, each vertex of a triangle lies necessarily in
${\mathcal {D}}_{n-1}(f)$
.
The following result links triangles for
${\mathcal {H}}_f$
and singularities of
$\Gamma _f$
. It has been proved for
$n=4$
in [Reference Adler and Ramanan1] and in [Reference Bricalli, Favale and Pirola6] for the general case.
Lemma 2.9. For
$[f]\in {\mathbb {P}}(S^3)$
, a point
$([x],[y])$
is singular for
$\Gamma _f$
if and only if there exists a third point
$[z]\in {\mathcal {H}}_f$
such that the triple
$([x],[y],[z])$
is a triangle for
${\mathcal {H}}_f$
.
To conclude this first section, let us present a couple of technical results which will be useful in what follows.
Lemma 2.10. Assume that
$[f]\in {\mathcal {U}}$
. Consider a point
$P\in \Gamma _f$
. Then
-
1. the squares of its coordinates are independent;
Moreover, if
$T\in {\mathcal {T}}$
is a triangle, then
-
2. the squares of the vertices of T are independent;
-
3. the vertices of T span a
${\mathbb {P}}^2$ .
Proof. In order to prove the claims, we will use extensively that
$\operatorname {\mathrm {Sing}}(V(f))$
can be identified with
$\{[x]\in {\mathbb {P}}(A^1)\,|\, x^2=0\}$
(see Proposition 2.1). More precisely, we will proceed by contradiction by proving that if the conclusion of
$(a),(b)$
or
$(c)$
is false, then there exists an element whose square is
$0$
(i.e., a singular point for f), which is impossible by assumption.
Let us start by proving
$(a)$
. For
$P=([y_1],[y_2])\in \Gamma _f$
, assume, by contradiction, that
$y_1^2$
and
$y_2^2$
are linearly dependent. Then, there exists
$\lambda \in {\mathbb {K}}$
such that
$y_1^2=-\lambda ^2y_2^2$
. As
$y_1y_2=0$
vanishes, we would have that
$(y_1+\lambda y_2)^2=0$
, which is impossible.
By
$(a)$
, in order to prove
$(b)$
, it is enough to show that
$z^2\not \in \left \langle x^2,y^2\right \rangle $
in
$A^2_f$
: let us assume again by contradiction that it is the case (i.e., there exist
$\alpha $
and
$\beta $
in
${\mathbb {K}}^*$
such that
$z^2=-\alpha ^2 x^2-\beta ^2 y^2$
). In the same way as before, we can consider the square
$(z+\alpha x+\beta y)^2$
, which is zero since
$xy=xz=yz=0$
, leading a contradiction.
For
$(c)$
, assume, by contradiction, that we could write
$z=\alpha x+\beta y$
for some
$\alpha ,\beta \in {\mathbb {K}}^*$
. Then, by definition of triangle, we would have that
$x^2,y^2$
and
$z^2$
are linearly dependent, which is impossible by
$(b)$
.
Assume that
$[f]\in {\mathcal {U}}$
. If
${\mathcal {F}}$
is variety of
$({\mathbb {P}}^n)^3\simeq {\mathbb {P}}(A^1)^3$
whose points are triangles for
${\mathcal {H}}_f$
(i.e.,
${\mathcal {F}}\subseteq {\mathcal {T}}$
), we will refer to
${\mathcal {F}}$
as a family of triangles (for
${\mathcal {H}}_f$
). Recall that the tangent space to
$({\mathbb {P}}^n)^3\simeq {\mathbb {P}}(A^1)^3$
at
$T=([x_1],[x_2],[x_3])$
is given by

Lemma 2.11. Assume that
$[f]\in {\mathcal {U}}$
. Let
$T=([x_1],[x_2],[x_3])$
be a triangle for
${\mathcal {H}}_f$
, and consider a (Zariski) tangent vector
$\underline {v}=(v_1,v_2,v_3)\in T_{{\mathcal {T}},T}$
. Then, for
$l\in \{1,2,3\}$
and for any representatives
$x_l'\in A^1$
of the class
$v_l\in A^1/\langle x_l\rangle $
, one has

whenever
$\{i,j,k\}=\{1,2,3\}$
. In particular,
$v_i\in \operatorname {\mathrm {Ann}}_{A^1}(x_j^2,x_k^2)/\langle x_i\rangle $
.
Proof. The variety
${\mathcal {T}}$
can be described in
${\mathbb {P}}(A^1)^3$
as

so that the forms
$xy,xz$
and
$yz$
vanish identically on
${\mathcal {T}}$
. If
$\underline {v}=(v_1,v_2,v_3)\in T_{{\mathcal {T}},T}$
is a tangent vector and
$v_l=[x_l']$
as in the statement, then
$(x_1+tx_1',x_2+tx_2',x_2+tx_3')$
satisfies the equations
$xy=xz=yz=0$
at first order:

so
$x_ix_j'+x_jx_i'=0$
, as claimed. Multiplying by
$x_j$
, we get
$x_j^2x_i'=0$
(i.e., we have
$x_i'\in \mathrm {Ann}_{A^1}(x_j^2,x_k^2)/\left \langle x_i\right \rangle $
).
Notice that in the proof of Lemma 2.11, all the computations do not depend on the choice made for
$x_l'$
in
$v_l$
. Hence, for brevity, we will use indifferently
$x_l'$
and
$v_l$
in similar situations.
As a consequence of Lemma 2.10, we can then define a morphism

Proposition 2.12. The morphism
$\psi $
has everywhere injective differential, and it is injective modulo permutations of the vertices.
Proof. Let us start by proving that the map
$\psi $
is injective (up to the permutation of such vertices). Let
$T=([x_1],[x_2],[x_3])$
and
$T'=([y_1],[y_2],[y_3])$
be two triangles which are not equivalent via permutation of the vertices. Assume, by contradiction, that
$\psi (T)=\psi (T')$
(i.e., the two triples of vertices span the same projective plane). We can then write
$y_i=\sum _{j=1}^3a_{ij}x_j$
. Recall that
$x_kx_l=\delta _{kl}x_k^2$
, since
$x_kx_l=0$
for every
$k\neq l$
, while
$x_i^2\neq 0, y_i^2\neq 0$
for every i, by the smoothness of
$V(f)$
. Hence, for
$i\neq j$
, we get

From Lemma 2.10, we have then that
$a_{ik}a_{jk}=0$
for every
$k=1,2,3$
and
$i\neq j$
. From this, one can easily see that, for every
$i=1,2,3$
, at least (and at most, by construction) two coefficients among
$a_{i1},a_{i2},a_{i3}$
are zero. Hence, the vertices of
$T'$
and of T are the same up to a permutation.
Fix a triangle
$T=([x_1],[x_2],[x_3])$
in
${\mathcal {T}}$
. We claim now that the differential

of
$\psi $
at T is injective. Let us consider a nontrivial vector

Via the isomorphism
$T_{\operatorname {\mathrm {Gr}}(k,V),W}\simeq \operatorname {\mathrm {Hom}}(W,V/W)$
, we have that
$d_T\psi (\underline {v})$
is the homomorphism

Hence, if we assume that
$d_T\psi (\underline {v})\equiv 0$
, we have
$x_i'\in \left \langle x_1,x_2,x_3\right \rangle $
, so we can write

for suitable
$a_{im}\in {\mathbb {K}}$
. By Lemma 2.11, we have
$x_ix_j'+x_i'x_j=0$
for
$i\neq j$
, so using the relations in Equation (2.6), we obtain

By Lemma 2.10, squares of vertices of a triangle are independent, so we obtain
$a_{ij}=0$
for
$i\neq j$
and
$x^{\prime }_i=a_{ii}x_i$
. This is impossible since
$x_i'\in A^1/\langle x_i\rangle $
, and we would have
$\underline {v}=0$
, whereas
$\underline {v}$
is assumed to be nontrivial.
3 Characterisation of TS Polynomials
In Section 2, we posed a question about a possible description of cubic forms
$[f]\in {\mathcal {U}}$
whose Hessian locus has singularities in codimension
$1$
(see Question 2.6). First of all, let us notice that the locus in
${\mathbb {P}}(S^3)$
we are interested in is not empty. Indeed, one can easily exhibit polynomials whose Hessian locus is reducible.
Definition 3.1. Given
$f\in S^d\setminus \{0\}$
, we say that f is a Thom-Sebastiani Polynomial (TS, for brevity) if

for suitable
$0\leq k\leq n-1$
,
$\{l_0,\dots ,l_n\}$
independent linear forms and
$f_1,f_2$
polynomials of degree d in
$k+1$
and
$n-k$
variables, respectively.
We will denote by
${\mathcal {V}}$
the set of smooth hypersurfaces which are not of Thom-Sebastiani type.
Remark 3.2. Notice that the set of Thom-Sebastiani polynomials of degree d is not closed in
${\mathbb {P}}(S^d)$
. For instance, one can see Examples 1.3 and 1.4 of [Reference Buczy’nska, Buczy’nski, Kleppe and Teitler7]. Nevertheless, in Theorem 4.5 of [Reference Buczy’nska, Buczy’nski, Kleppe and Teitler7], the authors obtain a normal form for polynomials which are limits of TS polynomials and are not themselves of TS type (which are called direct sums with their language). All these limits correspond to singular varieties: this proves that
${\mathcal {V}}$
is an open set in
${\mathcal {U}}$
(i.e., the open set of polynomials corresponding to smooth varieties) and consequently in
${\mathbb {P}}(S^d)$
.
Examples of TS polynomials are the ones whose zero locus is a cone. These are all singular, clearly. It is easy to see that if f is a TS polynomial as in Equation (3.1),
$X=V(f)$
is smooth if and only if
$V(f_1(l_0,\dots , l_k),l_{k+1},\dots , l_n)$
and
$V(f_2(l_{k+1},\dots , l_n),l_0,\dots , l_k)$
are smooth. This is also equivalent to ask that both
$V(f_1)\subset {\mathbb {P}}^k$
and
$V(f_2)\subset {\mathbb {P}}^{n-k-1}$
are smooth.
For brevity, if
$\{x_0,\dots , x_n\}$
are linear forms in
${\mathbb {P}}^n$
and if
$f_1$
and
$f_2$
are polynomials in
$k+1$
and
$n-k$
variables, respectively, let us define

Remark 3.3. Let f be a TS polynomial. Then, we can choose coordinates
$\{x_0,\dots , x_n\}$
for
${\mathbb {P}}^n$
and write
$f=f_1(\underline {x})+f_2(\underline {x}')$
for suitable polynomials in
$k+1$
and
$n-k$
variables of degree d. Set
$g_1:=f_1(\underline {x})$
and
$g_2:=f_2(\underline {x}')$
so that
$g_1,g_2\in S^d$
with
$g_1$
that depends only on the variables
$x_0,\dots , x_k$
and
$g_2$
that depends only on the other variables. Then, it is clear that

so
$h_f(\underline {x},\underline {x}')=h_{f_1}(\underline {x})h_{f_2}(\underline {x}')$
. In particular, the Hessian variety
${\mathcal {H}}_f$
associated to a TS polynomial is reducible, and it is the union of the two cones
$W_1=V(g_1)$
and
$W_2=V(g_2)$
. Moreover,
$\operatorname {\mathrm {Sing}}({\mathcal {H}}_f)$
has dimension
$n-2$
since it contains the intersection
$W_1\cap W_2$
.
When
$f=x_0^d+f_2(x_1,\dots ,x_{n})$
(in other words, in Definition 3.1, we are taking
$k=0$
), one talks about cyclic polynomials (see also Example 3.8 at the end of this section). The name comes from the fact that the projection of
$X=V(f)$
from the point
$P_0=(1:0:\dots :0)$
to
$V(x_0)\simeq {\mathbb {P}}^{n-1}$
gives a natural structure of cyclic cover of
${\mathbb {P}}^{n-1}$
branched along the hypersurface
$V(f_2)$
. For a smooth cubic
$X=V(f)$
, being cyclic gives a strong condition both on the associated Hessian locus and on the Jacobian ideal of f. Indeed, in [Reference Bricalli and Favale4], it has been proved that being cyclic is equivalent to having a linear component in the Hessian variety and a point in
${\mathcal {D}}_1(f)$
. This point corresponds to an element in
$J_f$
which is a square of some linear form in
$S^1$
(i.e., it gives a nilpotent element of order
$2$
in the Jacobian ring of f).
The main purpose of this section is to give a characterization of these Thom-Sebastiani polynomials in terms of the existence of suitable linear projective spaces in some Hessian loci. In particular, we will prove Theorem B:
Theorem 3.4. A polynomial
$f\in {\mathcal {U}}$
is of Thom-Sebastiani type of the form
$f(x_0,\dots ,x_n)=f_1(x_0,\dots ,x_k)+f_2(x_{k+1},\dots ,x_n)$
if and only if
${\mathcal {D}}_{k+1}(f)$
contains a
${\mathbb {P}}^{k}$
.
First of all, if we assume that
${\mathcal {D}}_k(f)\neq {\mathcal {D}}_{k-1}(f)$
, we can define the map

whose indeterminacy locus is
${\mathcal {D}}_{k-1}(f)$
.
Proposition 3.5. Assume that
${\mathcal {D}}_k(f)\neq {\mathcal {D}}_{k-1}(f)$
. The injectivity of
$\varphi $
can only fail on points along a line contained in
${\mathcal {D}}_{k}(f)$
and cutting
${\mathcal {D}}_{k-1}(f)$
. In particular, if
${\mathcal {D}}_{k-1}(f)=\emptyset $
or if
${\mathcal {D}}_{k}(f)$
does not contain lines,
$\varphi $
is injective.
Proof. If
${\mathcal {D}}_{k}(f)\setminus {\mathcal {D}}_{k-1}(f)$
is a single point,
$\varphi $
is clearly injective. Assume then that
$z_1,z_2\in {\mathcal {D}}_{k}(f)\setminus {\mathcal {D}}_{k-1}(f)$
are distinct and that
$\varphi ([z_1])=\varphi ([z_2])$
. Then
$\iota ([z_1])=\iota ([z_2])$
, so
$H_f(z_1)$
and
$H_f(z_2)$
have the same kernel. Up to a change of coordinates, we can assume that
$\ker (H_f(z_i))=\langle e_0,\dots , e_{n-k}\rangle $
, where
$\{e_0,\dots ,e_n\}$
is the basis corresponding to the basis
$\{y_0,\dots ,y_n\}$
under the identification
${\mathbb {P}}^n\simeq {\mathbb {P}}(A^1)$
. Hence, there exist two square matrices
$A_1$
and
$A_2$
of order k with coefficients in
${\mathbb {K}}$
and maximal rank such that

Being
$x\mapsto H_f(x)$
linear,
${\mathbb {P}}(\langle z_1,z_2\rangle )\simeq {\mathbb {P}}^{1}$
is clearly contained in
${\mathcal {D}}_k(f)$
.
Set
$p(\lambda ,\mu )$
to be the polynomial
$\det (\lambda A_1+\mu A_2)$
. Since
$\det (A_i)\neq 0$
by assumption, we have that p is homogeneous of degree k and nontrivial. Hence, there exists
$[\lambda _0:\mu _0]$
such that
$p(\lambda _0,\mu _0)=0$
(i.e.,
$H_f(\lambda _0 z_1+\mu _0 z_2)$
has rank at most
$k-1$
). Thus,
${\mathbb {P}}(\langle z_1,z_2\rangle )$
cuts
${\mathcal {D}}_{k-1}(f)$
, as claimed.
Proposition 3.6. Let
$f\in {\mathcal {U}}$
and assume that the
$(n-k-1)$
-plane
$\Pi ={\mathbb {P}}(V)$
is contained in
${\mathcal {D}}_{n-k}(f)$
. Then there exists
${\mathbb {P}}(U)\simeq {\mathbb {P}}^k$
in
${\mathcal {D}}_{1+k}(f)$
. Moreover, for all
$[u]\in {\mathbb {P}}(U)$
, one has that
$\Pi \subseteq \iota ([u])$
with equality holding for
$[u]$
general.
Proof. By assumption, one has
$\operatorname {\mathrm {Rank}}(H_f(v))\leq n-k$
for all
$v\in V\setminus \{0\}$
. One can see ([Reference Bricalli, Favale and Pirola6]) that the quadric of
${\mathbb {P}}^n$
given by the vanishing of the polynomial
${\partial _{v}}(f)$
is represented by the square symmetric matrix
$H_f(v)$
. Then the singular locus of the quadric
$V({\partial _{v}}(f))$
contains a
${\mathbb {P}}^{k}$
. By setting
$W=\{{\partial _{v}}(f)\}_{v\in V}$
, we can then observe that
$|W|$
is a linear subsystem of dimension
$n-k-1$
, since the map
$v\mapsto {\partial _{v}}(f)$
is injective as
$V(f)$
is smooth (it would have been enough to ask that
$V(f)$
is not a cone).
Let
$J=J_f$
be the Jacobian ideal of f. Since
$|W|\subset |J^2|$
and
$J^2$
is spanned by a regular sequence, we have that
$B:=\operatorname {\mathrm {BL}}(|W|)$
has pure dimension k. Indeed, being B cut by
$n-k$
quadrics, we have that
$\dim (B)\geq k$
. However, if there were a component of B with dimension at least
$k+1$
, then we would be able to complete a basis of W in such a way that
$J^2$
is not spanned by a regular sequence, against our assumptions. Indeed, if
$q_0,\dots ,q_{n-k-1}$
is a basis of W, then for any choice of elements
$q_{n-k},\dots ,q_n$
in
$J^2$
, we would have
$\bigcap _iV(q_i)\neq \emptyset $
. This cannot happen since f is smooth by assumption.
By Bertini’s theorem, the general element of
$|W|$
is smooth away from B, which has pure dimension k. However, as observed before, all quadrics of
$|W|$
have a
${\mathbb {P}}^k$
contained in the singular locus. Therefore, there exists a component of B which is a
${\mathbb {P}}^k$
. Since B has dimension k, it contains at most a finite number of
${\mathbb {P}}^k$
: there exists a component of B which is a
${\mathbb {P}}^k$
contained in the singular locus of all the elements of
$|W|$
. Let
${\mathbb {P}}(U)\simeq {\mathbb {P}}^k$
be this linear space.
Consider
$[u]\in {\mathbb {P}}(U)$
. Since all the quadrics
$V({\partial _{v}}(f))$
parametrized by W are singular along
${\mathbb {P}}(U)$
, we have

This implies that
${\partial _{v}}({\partial _{u}}(f))=0$
for all
$[v]\in {\mathbb {P}}(V)$
:
$V({\partial _{u}}(f))$
is a quadric whose singular locus contains the
$(n-k-1)$
-plane
$\Pi ={\mathbb {P}}(V)$
. Therefore,
$H_f(u)$
has rank at most
$k+1$
, and thus,
${\mathbb {P}}(U)\subseteq {\mathcal {D}}_{k+1}(f)$
.
Finally, notice that
$\dim ({\mathcal {D}}_{k}(f))\leq k-1$
by Proposition 2.5, so
${\mathbb {P}}(U)\simeq {\mathbb {P}}^k$
cannot be contained in
${\mathcal {D}}_{k}(f)$
. Hence, for the general point
$[u]\in {\mathbb {P}}(U)$
, the singular locus of
$V({\partial _{u}}(f))$
is exactly the
$(n-k-1)$
-plane
$\Pi $
. In other terms, we have
$\iota ([u])={\mathbb {P}}(V)$
for
$[u]\in {\mathbb {P}}(U)$
general.
Corollary 3.7. Let
$f\in {\mathcal {U}}$
and assume that there exists
$k\geq 1$
such that
${\mathcal {D}}_{n-k}(f)$
contains a
$(n-k-1)$
-plane. Then
${\mathcal {D}}_{k}(f)\neq \emptyset $
.
Proof. Assume that
${\mathbb {P}}(V)$
is a
$(n-k-1)$
-plane in
${\mathcal {D}}_{n-k}(f)$
. By Proposition 3.6, we have that there exist
${\mathbb {P}}^k\simeq {\mathbb {P}}(U)\subseteq {\mathcal {D}}_{k+1}(f)$
such that
${\mathbb {P}}(V)\subseteq \iota ([u])$
for
$[u]\in {\mathbb {P}}(U)$
with equality holding for
$[u]$
general. Since, in this case,
${\mathcal {D}}_{k+1}(f)$
and
${\mathcal {D}}_k(f)$
do not coincide, for dimensional reason (by Proposition 2.5), we can define the map
$\varphi :{\mathcal {D}}_{k+1}(f)\setminus {\mathcal {D}}_{k}(f)\to G(n-k-1,{\mathbb {P}}^n)$
. Then, the injectivity of
$\varphi $
fails on two general points of
${\mathbb {P}}(U)$
. Finally, by Proposition 3.5, we have that
${\mathbb {P}}(U)\cap {\mathcal {D}}_{k}(f)\neq \emptyset $
, as claimed.
We can now prove Theorem 3.4:
Proof. First of all, let us assume that for a fixed
$k\geq 0$
, there exists
${\mathbb {P}}^k\simeq {\mathbb {P}}(V)\subseteq {\mathcal {D}}_{k+1}(f)$
. Then, by Proposition 3.6, there also exists
${\mathbb {P}}(U)\simeq {\mathbb {P}}^{n-k-1}$
contained in the locus
${\mathcal {D}}_{n-k}(f)$
. Moreover, we know that for all
$[u]\in {\mathbb {P}}(U)$
, the projective space
${\mathbb {P}}(V)$
is contained in
$\iota ([u])$
. This means that for every
$[u]\in {\mathbb {P}}(U)$
and
$[v]\in {\mathbb {P}}(V)$
, we have
$uv=0$
, with the identification
${\mathbb {P}}^n={\mathbb {P}}(A^1)$
. Let us notice that the spaces
${\mathbb {P}}(U)$
and
${\mathbb {P}}(V)$
are skew in
${\mathbb {P}}^n$
and of complementary dimension. Indeed, if their intersection was nontrivial, we could find a point
$[x]\in {\mathbb {P}}(V)\cap {\mathbb {P}}(U)$
: from above, we would obtain
$x^2=0$
, against the smoothness of
$V(f)$
. We can then consider for
${\mathbb {P}}^n$
a coordinate system
$x_0,\dots ,x_n$
where
${\mathbb {P}}(V)=V(x_0,\dots ,x_{k})$
and
${\mathbb {P}}(U)=V(x_{k+1},\dots ,x_n)$
. Up to a change of coordinates, we can then write the polynomial f with respect to these variables: since, by construction,
$x_ix_j=0$
in
$A^2$
for every
$i=0,\dots ,k$
and
$j=k+1,\dots ,n$
, we get the claim.
Let us now assume that f is TS: as in Remark 3.3, we can write it (up to a possible change of coordinates) as
$f=f_1(\underline {x})+f_2(\underline {x}')$
. As already observed, the Hessian matrix of f is of the form

Hence, by defining
${\mathbb {P}}(V):=V(x_0,\dots ,x_k)\simeq {\mathbb {P}}^k$
, one easily sees that
$\operatorname {\mathrm {Rank}}(H_f(v))\leq k+1$
for every
$[v]\in {\mathbb {P}}(V)$
(i.e.,
${\mathbb {P}}(V)\subseteq {\mathcal {D}}_{k+1}(f)$
, as claimed).
To end this section, let us present some key examples of TS polynomials.
Example 3.8 (Cyclic cubics).
The simplest examples of TS polynomials are the cyclic polynomials. We recall that a polynomial
$f\in S^d$
, where
$S={\mathbb {K}}[x_0,\cdots ,x_{n}]$
is cyclic if, up to a change of coordinates, it can be written as
$f=x_0^d+g(x_1,\dots ,x_{n})$
, where
$g\in {\mathbb {K}}[x_1,\dots ,x_{n}]_d$
.
As observed before,
$X=V(f)$
is smooth exactly when
$V(g)\subset {\mathbb {P}}^{n-1}$
is smooth and
$h_f= d(d-1)x_0^{d-2}\cdot h_g(x_1,\dots ,x_n)$
, so the Hessian variety splits as the union of a hyperplane and a hypersurface of degree
$n(d-2)$
, namely

Notice that W does not need to be irreducible, but this is the case if g is general (and
$n\geq 3$
). Under the identification
${\mathbb {P}}^{n-1}\simeq V(x_0)$
, we can say that the Hessian loci
${\mathcal {D}}_k(g)$
live in
$H=V(x_0)$
. We denote by
$\hat {{\mathcal {D}}}_k(g)$
the cone over
${\mathcal {D}}_k(g)\subseteq V(x_0)$
with vertex the coordinate point
$P_0=(1:0:\cdots :0)$
. For example, one has
$W=\hat {{\mathcal {D}}}_{n-1}(g)$
. Then, using the explicit description of the Hessian matrix of f as block matrix such as in Equation (3.2), one can easily prove that

It is well known that the general cubic surface
$S=V(g)\subseteq {\mathbb {P}}^3$
has an irreducible Hessian variety which is a quartic with
$10$
nodes as the only singularities. This was known already by B. Segre (see [Reference Segre28]), but one can also refer to the more recent [Reference Dardanelli and van Geemen10]. In particular,

Using this observation and Equation (3.3), one can describe the stratification given by the Hessian loci of a general cyclic cubic threefold
$X=V(f)$
:


Among these, only
$H,W\subset {\mathcal {D}}_4(f)$
,
$H\cap W={\mathcal {H}}_g\subset {\mathcal {D}}_{3}(f)$
and
$\{P_0\}\subset {\mathcal {D}}_1(f)$
give irreducible components of
$\Gamma _f$
(this will be clear from Lemma 4.1).
Since smooth binary cubic forms can be written as sum of
$2$
cubes, every TS polynomial in
$n+1$
variables, with
$n\leq 4$
, is necessarily cyclic (see [Reference Bricalli and Favale4] for details). Let us now describe a new phenomenon arising in
${\mathbb {P}}^5$
.
Example 3.9 (A TS cubic which is not cyclic).
Let
$g_1,g_2\in S_w={\mathbb {K}}[w_0,w_1,w_2]$
be such that
$V(g_1)$
and
$V(g_2)$
are smooth cubic curves in
${\mathbb {P}}^2$
which are not projectively equivalent to the Fermat curve. A classical result implies that this is equivalent to ask that
$V(g_i)$
is a cubic whose Hessian
$V(h_{g_i})$
is irreducible. This equivalence also follows easily from the main result of this paper (see Theorem 5.1 or, more specifically, Proposition 5.6) and can be found in [Reference Ciliberto and Ottaviani8].
A smooth cubic fourfold X of TS type is not cyclic if and only if, up to a change of coordinates, it is defined by a polynomial
$f=g_1(x_0,x_1,x_2)+g_2(x_3,x_4,x_5)$
. From the point of view of moduli, such fourfolds form a dimension
$2$
variety in the moduli space of smooth cubic fourfolds.
Consider the subvarieties of
${\mathbb {P}}^5$
defined by


where
$J(C_1,C_2)$
is the joint variety of
$C_1$
and
$C_2$
(namely, the union of all lines joining a point of
$C_1$
and a point of
$C_2$
). Notice that, by construction, for
$\{i,j\}=\{1,2\}$
, the variety
$W_i$
is a cone over
$C_j$
with vertex
$\Pi _i$
and
$C_i$
is isomorphic to the curve
$V(h_{g_i})$
. Moreover, by the assumptions on the curves
$V(g_1)$
and
$V(g_2)$
,
$W_i$
is irreducible. Being f a TS polynomial, one has that
${\mathcal {H}}_f$
is indeed reducible. More precisely, since
$h_f=h_{g_1}(x_0,x_1,x_2)h_{g_2}(x_3,x_4,x_5)$
, one has that

The other strata of the stratification induced by f are

whereas
${\mathcal {D}}_1(f)=\emptyset $
as X is not cyclic (by the results in [Reference Bricalli and Favale4]).
It is worth highlighting two facts. First of all,
$\Pi _1$
and
$\Pi _2$
are two
$2$
-planes contained in
${\mathcal {D}}_3(f)$
. These are exactly the k-planes contained in
${\mathcal {D}}_{k+1}(f)$
whose existence is guaranteed by Theorem 3.4 since f is a TS polynomial. Moreover, note that for all
$k\in \{2,3,4,5\}$
, the dimension of
${\mathcal {D}}_k(f)$
equals
$k-1$
(i.e., the maximum predicted by Proposition 2.5). In particular,
$\Gamma _f$
splits as the union of
$7$
irreducible fourfolds (this follows from Lemma 4.1).
4 Families of triangles of high dimension
In this section, we focus on the study of suitable families of triangles for
${\mathcal {H}}_f$
arising naturally, as we will see in a moment, when
$\operatorname {\mathrm {Sing}}({\mathcal {H}}_f)$
exceeds the expected dimension. Moreover, we prove Theorem C.
Let us now set some notations and prove some technical results.
We recall that, given a smooth cubic
$V(f)\subseteq {\mathbb {P}}^n\simeq {\mathbb {P}}(A^1)$
, a family of triangles for
${\mathcal {H}}_f$
is a subvariety
${\mathcal {F}}$
of

We will denote by
$\pi _i$
the natural projections from
${\mathcal {F}}$
on the factors. For simplicity, if
${\mathcal {F}}$
is a family of triangles for
${\mathcal {H}}_f$
, we will set
$Y_i=\pi _i({\mathcal {F}})$
for
$i\in \{1,2,3\}$
. Notice that
$\dim (Y_i)\leq n-2$
since
$Y_i\subseteq {\mathcal {D}}_{n-1}(f)$
which has dimension at most
$n-2$
by Proposition 2.5.
Moreover, recall that if
$[f]\in {\mathcal {U}}$
, by Proposition 2.5, all components of
$\Gamma _f$
come from the Hessian loci of f. More precisely, if Z is an irreducible component of
${\mathcal {D}}_{k}(f)$
of dimension
$k-1$
, there exists a unique irreducible component of
$\Gamma _f$
which dominates Z by first projection. We will denote by
$\tilde {Z}$
such component.
Lemma 4.1. Let
$[f]\in {\mathcal {U}}$
and assume
$k\in \{1,\dots , n-1\}$
. Consider an irreducible component W of
${\mathcal {H}}_f$
and an irreducible component Z of
${\mathcal {D}}_{k}(f)$
of dimension
$k-1$
which is contained in W. Then
$\tilde {Z}\cap \tilde {W}$
dominates Z via the first projection
$\operatorname {\mathrm {pr}}_1$
. In particular, there exists a family of triangles
${\mathcal {F}}$
for
${\mathcal {H}}_f$
of dimension at least
$k-1$
. Moreover, if
$k=n-1$
, then every family of triangles as above has dimension exactly
$n-2$
.
Proof. Notice that Z is not contained in
${\mathcal {D}}_{k-1}(f)$
by Proposition 2.5, as we are assuming
$\dim (Z)=k-1$
. Hence, the general point
$z\in Z$
lies in
${\mathcal {D}}_k(f)\setminus {\mathcal {D}}_{k-1}(f)$
, and thus,
$\iota ([z])\simeq {\mathbb {P}}^{n-k}$
. Since the general fiber of
$\operatorname {\mathrm {pr}}_1|_{\tilde {Z}}:\tilde {Z}\to Z$
has dimension
$n-k$
by construction, one has that the whole fiber
$\operatorname {\mathrm {pr}}_1^{-1}([z])=\{[z]\}\times \iota ([z])$
is contained in
$\tilde {Z}$
. However,
$\operatorname {\mathrm {pr}}_1|_{\tilde {W}}:\tilde {W}\to W$
is surjective and
$Z\subseteq W$
by assumption, so there exists at least a point
$p=([z],[y])$
of the whole fiber
$\pi _1^{-1}([z])$
with
$p\in \tilde {W}$
. Then
$p \in U=\tilde {W}\cap \tilde {Z}$
and
$\operatorname {\mathrm {pr}}_1|_{U}:U\to Z$
is such that
$\operatorname {\mathrm {pr}}_1|_{U}(p)=z$
. In particular,
$\operatorname {\mathrm {pr}}_1|_{U}$
dominates Z.
By the above argument, we have that
$\tilde {W}$
and
$\tilde {Z}$
are irreducible components of
$\Gamma _f$
which meet in a variety U of dimension at least
$k-1$
. Then, we have a family of dimension
$k-1$
since this variety is contained in
$\operatorname {\mathrm {Sing}}(\Gamma _f)$
by construction and each point yields (at least) a triangle by Lemma 2.9.
Let
$Z\subseteq {\mathcal {D}}_{n-1}(f)$
be an irreducible component of dimension
$n-2$
and let
${\mathcal {F}}$
be a family of triangles dominating Z via
$\pi _1$
, so that
$\dim ({\mathcal {F}})\geq n-2$
. By construction, the general point
$[x]$
of Z is such that
$\iota ([x])\simeq {\mathbb {P}}^1$
so
$\pi _1^{-1}([x])\subset \{[x]\}\times {\mathbb {P}}^1\times {\mathbb {P}}^1$
. If the general fiber
$\pi _1^{-1}([x])$
has positive dimension, we would have that
$\pi _1^{-1}([x])\cap \{[x]\}\times \Delta _{{\mathbb {P}}^1}$
is not empty, thus giving rise to a singular point of
$V(f)$
. Then the general fiber of
$\pi _1$
has dimension
$0$
, and thus,
$\dim ({\mathcal {F}})=n-2$
.
Remark 4.2. If
${\mathcal {H}}_f$
is not normal, we have that there exists at least a family of triangles of dimension
$n-2$
. Indeed, we have that the singular locus of
${\mathcal {H}}_f$
has dimension
$n-2$
and equals
${\mathcal {D}}_{n-1}(f)$
by Theorem 2.3. Hence, given an irreducible component Z of
$\operatorname {\mathrm {Sing}}({\mathcal {H}}_f)$
of dimension
$n-2$
, we have that Z yields a family of triangles of dimension
$n-2$
as a consequence of Lemma 4.1.
Lemma 4.3. Let
$X=V(f)\subset {\mathbb {P}}^n$
be a smooth cubic hypersurface and let
${\mathcal {F}}$
be an irreducible family of triangles. Then the following hold:
-
1. If
$\dim (Y_i)\geq 1$ and
$Y_i\subset X$ , then
$\dim (Y_i)\leq n-3$ ;
-
2. If
$n\geq 3$ and
$\dim (Y_i)=n-2$ for some i, then no projection has dimension
$0$ unless
$V(f)$ is of Thom-Sebastiani type.
Proof. For
$(a)$
, w.l.o.g. we can assume
$\dim (Y_1)\geq 1$
and
$Y_1\subset V(f)$
. If
$T=([x],[y],[z])\in {\mathcal {F}}$
is a general triangle, then the differential
$d\pi _{1,T}:T_{{\mathcal {F}},T}\rightarrow T_{Y_1,[x]}$
is surjective and it sends a tangent vector
$(x',y',z')$
to
$x'$
. By Lemma 2.11,
$x'\in \operatorname {\mathrm {Ann}}_{A^1}(y^2,z^2)/\langle x\rangle $
. Moreover, since
$Y_1\subset X$
, by Proposition 2.1, we have
$X=\{[x]\,|\, x^3=0\}$
so
$T_{X,[x]}=\operatorname {\mathrm {Ann}}_{A^1}(x^2)/\langle x\rangle $
. Hence,

Since T is a triangle, by Lemma 2.10, one has that
$\langle x^2,y^2,z^2\rangle $
has dimension
$3$
, and thus, by the Gorenstein duality,
$\dim (\operatorname {\mathrm {Ann}}_{A^1}(x^2,y^2,z^2))=n+1-3=n-2$
. Moreover, being
$[x]\in X$
, one has
$x\in \operatorname {\mathrm {Ann}}(x^2,y^2,z^2)$
, so
$\operatorname {\mathrm {Ann}}_{A^1}(x^2,y^2,z^2)/\langle x\rangle $
has dimension
$n-3$
.
For
$(b)$
, w.l.o.g. assume that
$Y_3$
is of dimension
$n-2$
. By contradiction, let us assume that
$Y_1=\{[x]\}$
and that f is not of TS type. Hence,
$Y_3\subseteq \iota ([x])$
, and this implies that
$\iota (x)={\mathbb {P}}^s$
with
$s\in \{n-2,n-1\}$
.
Then, we would get
$\iota ([x])={\mathbb {P}}^{n-2}=Y_3\subseteq {\mathcal {D}}_{n-1}(f)$
and
$[x]\in {\mathcal {D}}_1(f)$
, respectively. Both cases yield a contradiction by Theorem 3.4.
Before proving Theorem C (Theorem 4.6), let us focus on (families of) triangles with all vertices on the cubic X. These are linked to families of
$2$
-planes in the cubic hypersurface:
Remark 4.4. First of all, recall that if
$T=([x],[y],[z])$
is a triangle for
${\mathcal {H}}_f$
, then
$\left \langle [x],[y],[z]\right \rangle ={\mathbb {P}}^2$
, by Lemma 2.10. If we assume, moreover, that all the vertices of T belong to the cubic hypersurface X, we have
$x^3=y^3=z^3=0$
and also
$xy=yz=zx=0$
; this implies that the
$2$
-plane is actually contained in X. Hence, a triangle with three vertices on X cannot exist if X is a smooth cubic hypersurface of dimension at most
$3$
. Furthermore, since on smooth cubic fourfolds one has at most a finite number of
$2$
-planes (see, for example, [Reference Degtyarev, Itenberg and Ottem11]), by Proposition 2.12, we can have at most a finite number of triangles with all the vertices on the cubic X.
Lemma 4.5. Assume that
${\mathcal {F}}$
is a family of triangles for
${\mathcal {H}}_f$
with
$\dim (Y_1)=\dim ({\mathcal {F}})>\dim (Y_3)$
. Then, none of the fibers of the projection
$\pi _3$
can be contracted to points via
$\pi _1$
.
Proof. This follows from a more general fact: if
$g:X\to Z$
is a surjective morphism between irreducible varieties and
$f:X\to Y$
is a morphism, the locus

is an open Zariski set of Z. In order to prove this, let us consider the map
$F=(f,g):X\rightarrow Y\times Z$
and denote by
$X'\subseteq Y\times Z$
the image of X under the morphism F. Moreover, let
$p_1$
and
$p_2$
be the two projections from
$X'$
to Y and Z, respectively. It is then clear that

and thus, it is an open subset of Z. In our situation, if we assume that a fiber of
$\pi _3$
is contracted to points via
$\pi _1$
, then the same is true for the general one, contradicting the assumption
$\dim (Y_1)=\dim ({\mathcal {F}})$
.
We are now ready to prove Theorem C, one of the main ingredients in the proof of Theorem A.
Theorem 4.6. Assume
$n\leq 5$
and consider a smooth cubic
$V(f)$
not of TS type. If
${\mathcal {F}}$
is an irreducible family of triangles for
${\mathcal {H}}_f$
with
$\dim ({\mathcal {F}})=\dim (\pi _1({\mathcal {F}}))=n-2$
, then the general element in
${\mathcal {F}}$
is such that none of its vertices belongs to
$X=V(f)$
.
Proof. Notice that the proof follows easily if none of the three projections of
${\mathcal {F}}$
is contained in X. Indeed, in this case,
$\pi _i^{-1}(Y_i\cap X)$
, the locus where the triangles of
${\mathcal {F}}$
have the i-th vertex on the cubic X, is a proper closed subset of
${\mathcal {F}}$
. We would like to show that also for
$n\leq 5$
, this is indeed the only possible case (i.e., no projection of
${\mathcal {F}}$
can be contained in X). For
$n=3$
, This is an easy consequence of Lemma 4.3, so we can assume
$n\geq 4$
.
By hypothesis,
$\pi _1$
is generically finite (and thus, by Lemma 4.3,
$Y_1$
is not contained in X) and, by contradiction,
$Y_3\subset X$
. Notice that, under these assumptions,
$Y_2$
is not contained in X. Otherwise,
$\pi _1^{-1}(Y_1\cap X)$
would be an
$(n-3)$
-dimensional family of triangles with all the vertices contained in X. Then, we would have a contradiction as observed in Remark 4.4. For brevity, set
${\mathcal {F}}_c={\mathcal {F}}\cap (X\times X\times X)$
(i.e.,
${\mathcal {F}}_c$
is the locus of the triangles of
${\mathcal {F}}$
with all
$3$
vertices on the cubic hypersurface X).
Assume that
$n=4$
. By Lemma 4.3, since
$Y_3\subset X$
and
$Y_1$
has dimension
$2=n-2=\dim ({\mathcal {F}})$
, we have that
$\dim (Y_3)=1$
. Then, since
${\mathcal {F}}$
is irreducible, all the fibers of
$\pi _3$
have pure dimension
$1$
. The dimension of
$Y_2$
is either
$1$
or
$2$
. If the dimension of
$Y_2$
is
$1$
, all the fibers of
$\pi _2$
are curves too and
$Y_2\cap X$
is not empty. Consider
$[y_0]\in Y_2\cap X$
and its fiber
$C=\pi _2^{-1}([y_0])$
. By Lemma 4.5, C cannot be contracted by
$\pi _1$
, so
$\pi _1(C)$
is a curve. Then,
$\pi _1(C)\cap X$
is not empty, and we can consider a point
$[x_0]$
in this intersection. Hence, any element in
$\pi _1^{-1}([x_0])\cap C\neq \emptyset $
is a triangle in
${\mathcal {F}}_c$
. This is impossible by Remark 4.4. Then we have necessarily
$\dim (Y_2)=2$
.
Let Y be an irreducible component of
$Y_2\cap X$
. Being
$Y_2\not \subseteq X$
and of dimension
$2$
, Y is a curve, and there exists an irreducible component C in
$\pi _2^{-1}(Y)$
of dimension
$1$
(since
${\mathcal {F}}$
is irreducible) dominating Y via the second projection. If either
$\pi _1(C)$
is a curve or
$\pi _1(C)=[x_0]$
with
$[x_0]\in X$
, we have an element in
${\mathcal {F}}_c$
, so the only possible case is
$\pi _1(C)=[x_0]$
with
$[x_0]\not \in X$
.
Looking at the third projection, we have that
$\pi _3(C)$
either is a point
$[z]$
or it coincides with
$Y_3$
. Moreover, let us observe that
$\iota ([x_0])\simeq {\mathbb {P}}^s$
with
$s\in \{1,2\}$
such that Y and
$\pi _3(C)$
are contained in
$\iota ([x_0])$
. To rule out the first case, namely
$\pi _3(C)=[z]$
, first of all, observe that
$[z]\not \in Y$
; otherwise, we would have a singular point for the cubic X. Then s is forced to be
$2$
. Moreover, by construction, we have
$yz=0$
and
$y^3=z^3=0$
for any
$[y]\in Y$
: by reasoning as in Remark 4.4, all the lines
$\langle [y],[z]\rangle $
lie in X. This implies that the whole
$\iota ([x_0]) \simeq {\mathbb {P}}^2$
is contained in the smooth threefold X, but this is not possible, as observed in Remark 4.4.
For the remaining case, we have that
$\pi _3(C)=Y_3$
, and by construction, both the curves Y and
$Y_3$
are contained in
$\iota ([x_0])$
. Then, if
$s=1$
, we necessarily have
$Y=Y_3\simeq {\mathbb {P}}^1$
and C is a family of triangles of dimension
$1$
in
$\{[x_0]\}\times Y\times Y$
. This yields a contradiction since C has to meet
$\{[x_0]\}\times \Delta _{Y}$
, thus giving a singular point for X. Hence, we have necessarily
$s=2$
, and we can assume
$Y\neq Y_3$
or
$Y=Y_3\neq {\mathbb {P}}^1$
. In both cases, as done above, considering the lines
$\langle [y],[z]\rangle $
with
$[y]\in Y, \ [z]\in Y_3$
and
$yz=0$
, we get that the
$2$
-plane
$\iota ([x_0])$
is contained in the cubic threefold X, which is not possible.
Assume now that
$n=5$
. We are working in the following framework:
${\mathcal {F}}$
is an irreducible
$3$
-dimensional family of triangles with
$\pi _1$
generically finite,
$Y_1,Y_2$
not contained in X and
$Y_3\subseteq X$
(this will lead to a contradiction). Being
$Y_3\subseteq X$
by assumption,
${\mathcal {F}}_c$
is cut out from
${\mathcal {F}}$
by two divisors, so its expected dimension is
$n-4=1$
. Then, either
${\mathcal {F}}_c$
is empty or
$\dim ({\mathcal {F}}_c)\geq 1$
. However, as observed in Remark 4.4, under the above-mentioned hypotheses, we have that
$\dim ({\mathcal {F}}_c)\leq 0$
, so
${\mathcal {F}}_c$
is necessarily empty. We will now prove that
${\mathcal {F}}_c$
is not empty, thus leading to a contradiction.
First of all, notice that
$\dim (Y_2)\in \{1,2,3\}$
by Lemma 4.3. If
$\dim (Y_2)\leq 2$
, the general fiber of
$\pi _2$
has positive dimension and cannot be contracted to a point by
$\pi _1$
by Lemma 4.5. Then, its image meets X, and thus, we produce a triangle in
${\mathcal {F}}_c$
as the analogous case for the threefold. We can then assume
$\dim (Y_2)=3$
so that
$\pi _2$
is generically finite as
$\pi _1$
.
Denote by Y an irreducible component of
$Y_2\cap X$
. The preimage
$\pi _2^{-1}(Y)$
has dimension
$2$
, and we can consider an irreducible component
$S\subset \pi _2^{-1}(Y)$
dominating Y. If
$\pi _1(S)$
is not a point or is a point on the cubic fourfold, as in the threefold case, one can easily construct an element in
${\mathcal {F}}_c$
: we can assume
$\pi _1(S)=[x_0]\not \in X$
. As done in the previous case, we have that
$\iota ([x_0])\simeq {\mathbb {P}}^s$
containing Y and
$\pi _3(S)$
(thus,
$s\in \{2,3\}$
since X is not of TS type), where
$\pi _3(S)\subseteq Y_3$
can be either a point
$[z]$
, a curve
$C\not \subset Y_3$
(with
$\dim (Y_3)=2$
) or the whole
$Y_3$
(with
$\dim (Y_3)\in \{1,2\})$
. To conclude the proof, let us study these distinguished cases.
-
•
$\mathbf {\pi _3(S)=[z_0]}$ : since
$[z_0]\not \in Y$ (otherwise we would have a singular point in X), s is forced to be equal to
$3$ . Considering again the lines
$\langle [y],[z_0]\rangle $ with
$[y]$ varying in Y, we have that the whole
$3$ -space
$\iota ([x_0])$ is contained in the smooth cubic fourfold X, which is clearly not possible.
-
•
$\mathbf {\pi _3(S)=C}$ : In this case, one has
$\iota ([x_0])\simeq {\mathbb {P}}^3$ since, otherwise, we would have
$C\subseteq Y={\mathbb {P}}^2$ and
$S\subseteq \{[x_0]\}\times Y\times Y$ and then a singular point for X as in a previous case. If we assume
$C\not \subset Y\subset \iota ([x_0])={\mathbb {P}}^3$ , then one easily sees that the lines
$\langle [y],[z]\rangle $ with
$[y]\in Y, [z]\in C$ and
$yz=0$ cover
$\iota ([x_0])$ : we have a contradiction since we would have a projective
$3-$ space in X. The only remaining case to analyse is then the one where
$C\subset Y\subset {\mathbb {P}}^3$ with Y surface which is not a
${\mathbb {P}}^2$ . In this case, S is a surface in
$\{[x_0]\}\times Y\times C$ with the projections
$p_2=\pi _2|_S$ and
$p_3=\pi _3|_S$ which are surjective. Then, for all
$[z]\in C$ ,
$p_3^{-1}([z])$ has pure dimension
$1$ . Let
$[z]$ be a point in C and let D be an irreducible component of one of those fibers. For all
$[y]\in p_2(D)$ , we have
$yz=y^3=z^3=0$ and
$[z]\not \in p_2(D)$ , so the joint variety
$J(p_2(D),[z])$ has dimension
$2$ , is a cone with vertex
$[z]$ and is completely contained in
$\iota ([x_0])\cap X$ . Since
$\iota ([x_0])\simeq {\mathbb {P}}^3$ cannot be contained in X, these cones have to vary at most discretely, when
$[z]$ moves in C. Notice that Y lies, by construction, in the union of these cones, so
$[z]$ is in the vertex
$\operatorname {\mathrm {Vert}}(Y)$ of Y. Then
$C\subseteq \operatorname {\mathrm {Vert}}(Y)$ , and this forces Y to be a
${\mathbb {P}}^2$ , which is against our assumptions.
-
•
$\mathbf {\pi _3(S)=Y_3}$ : As in the previous case, we necessarily have
$\iota ([x_0])\simeq {\mathbb {P}}^3$ containing both the surface Y and
$\pi _3(S)$ , which can be either a curve or a surface. We can assume, moreover
$Y_3=\pi _3(S)\subseteq Y$ , since otherwise, proceeding as above, we would have
$\iota ([x_0])\subset X$ . In particular, Y is not a
$2$ -plane. If
$Y_3$ is a curve, we can obtain a contradiction as in the previous case by considering the cones with vertex
$[z]\in Y_3$ spanned by the curves in Y whose elements annihilate
$[z]$ . Hence, we can assume
$Y=Y_3$ (and thus,
$\pi _3$ is generically finite). By construction, for any element
$[y]\in Y$ , there exists at least one element
$[z]\in Y$ such that
$yz=0$ ; hence, again, the line
$\left \langle [y],[z]\right \rangle $ is contained in
$\iota ([x_0])\cap X$ . If, for
$[y]$ general, at least one of these lines is not contained in Y, one can see that the whole
$3$ -space
$\iota ([x_0])$ is contained in X, yielding a contradiction. We can then assume that for
$[y]\in Y$ general, the above-mentioned lines are contained in Y. Our aim is now to show that
$Y\simeq {\mathbb {P}}^2$ , against our assumptions. First of all, let us show that the union
$\Lambda $ of these lines as subset in the Grassmannian
$\operatorname {\mathrm {Gr}}(1,{\mathbb {P}}^3)$ has dimension
$2$ . If, by contradiction,
$\dim (\Lambda )=1$ , this would mean that for all
$\ell \in \Lambda $ and for all
$[y]\in \ell $ , there exists
$[z]\in \ell $ with
$yz=0$ . In other words, this would yield a correspondence in
${\mathbb {P}}^1\times {\mathbb {P}}^1$ , which intersects the diagonal
$\Delta _{{\mathbb {P}}^1}$ nontrivially: the cubic X would be singular, which is not possible. Let us finally consider the incidence variety
$$ \begin{align*}\Psi:=\{(y,\ell) \ | \ y\in\ell\in\Lambda\}\subset Y\times \operatorname{\mathrm{Gr}}(1,{\mathbb{P}}^3)\end{align*} $$
$\psi _1$ and
$\psi _2$ the two projections. We have just shown that
$\operatorname {\mathrm {Im}}(\psi _2)=\Lambda $ , and moreover, it is clear that if
$\ell \in \Lambda $ , then
$\psi _2^{-1}(\ell )$ is described by
$\ell $ itself; hence, such a fiber has dimension
$1$ . Then
$\dim (\Psi )=3$ , and looking at the first projection
$\psi _1$ , we have that there exist infinitely many lines in
$\Lambda $ contained in Y and passing through the general point
$[y]\in Y$ . Hence, Y has to be a cone with
$[y]\in \operatorname {\mathrm {Vert}}(Y)$ : from the generality of
$[y]$ , it follows that
$Y\simeq {\mathbb {P}}^2$ , as claimed.
Remark 4.7. Observe that if
$n=2$
, the hypotheses of Theorem cannot be realized since the existence of a triangle implies that the locus
${\mathcal {D}}_1(f)$
is nonempty. Hence, by Theorem 3.4, the cubic f is of TS type, against our assumption.
5 Proof of main theorem: the cubic threefold case
In this section, we state and begin to prove the main result of this article – namely, the following.
Theorem 5.1 (Theorem A).
Assume that
$2\leq n\leq 5$
and consider
$f\in {\mathbb {K}}[x_0,\dots ,x_n]$
defining a smooth cubic. Then, the singular locus of the Hessian hypersurface
${\mathcal {H}}_f\subset {\mathbb {P}}^n$
has the expected dimension if and only if f is not of TS type. In particular,
$f\in {\mathcal {V}}$
if and only if
${\mathcal {H}}_f$
is irreducible and normal.
As we observe now, the assumptions on the degree and the smoothness of the hypersurface
$X=V(f)$
are essential.
Remark 5.2. Let us stress that the result stated in Theorem 5.1 is false for smooth hypersurfaces of degree
$d\geq 4$
and for non-smooth cubics. We provide here two simple examples proving these claims.
-
• Let
$f(x,y,z)=x^4+y^4+z^4+x(y^3+z^3)$ and consider
$C=V(f)$ . Then one easily sees that C is a smooth quartic plane curve and that
$$ \begin{align*}h_f=54\cdot yz\left( 8x^4 + 16x^3(y + z) + 32x^2yz - x(y^3+z^3) - 2yz(y^2 + z^2)\right).\end{align*} $$
$h_f$ yields a smooth quartic by the Jacobian criterion and thus an irreducible one. This also implies that f is not of TS type since, otherwise, we would have the Hessian polynomial which is product of linear factors: there are smooth hypersurfaces of degree
$d\geq 4$ , which are not of TS type, with reducible Hessian variety.
-
• Let
$f(x_0,x_1,x_2,x_3)=x_0x_1^2+x_1x_2^2+x_2x_3^2$ and consider the cubic surface
$S=V(f)$ . One can see that S is an irreducible cubic surface whose singular locus coincides with the point
$p_0=(1:0:0:0)$ , which is a singularity of type
$D_5$ . Its associated Hessian variety is a reducible and non-reduced quartic surface
$V(x_1^2(x_1x_2-x_3^2))$ . Notice that the quadratic factor of the Hessian polynomial
$h_f$ is irreducible so, reasoning as in the previous example, one can see that
$f\neq f_1(z_0,z_1)+f_2(z_2,z_3)$ for suitable coordinates
$\{z_0,\dots ,z_3\}$ . With a direct and easy computation, also the cyclic case is ruled out as follows. If we assume that f is cyclic, then
$V(f)$ would be projectively equivalent to
$V(g)$ where
$g=x_0^3+m(x_1,x_2,x_3)$ . Notice that
$h_g=x_0\cdot h_{m}(x_1,x_2,x_3)$ . Since a linear form has to appear in the factorization of
$h_g$ with multiplicity
$2$ (since
$x_1^2$ divides
$h_f$ ) we have that
$h_m$ is the product of
$3$ linear forms, which is incompatible with the above description. Hence, f is not of TS type although its Hessian is reducible (and thus non-normal) and non-reduced. The same phenomenon happens for the cuspidal cubic curve (see, for example, [Reference Ciliberto and Ottaviani8]).
Nevertheless, not every type of singularity gives the same behaviour as in the last example above. Indeed, the nodal cubic curve and the
$1$
-nodal cubic surface
$V(x_0(x_1^2+x_2^2+x_3^2)+x_1^2x_3+x_2x_3^2)$
have irreducible and normal associated Hessian variety (and thus, they are not of TS type). One can easily construct examples of
$1$
-nodal cubic threefolds and fourfolds with the same property.
Going back to the case of smooth cubic hypersurfaces, we make the following:
Conjecture 5.3. The same result stated in Theorem 5.1 holds for smooth cubic hypersurfaces in
${\mathbb {P}}^n$
for every
$n\geq 2$
.
The techniques used in the proof of Theorem 5.1 do not seem to adapt to an argument that could be valid in any dimension: already for cubic fourfolds, one can see the large amount of cases one has to consider. For this reason, we will give the proof of the Theorem for
$n\leq 4$
at the end of this section, while the case of cubic fourfolds is treated in the subsequent one, since it is more involved, even if the techniques are similar.
We stress that, for a smooth cubic
$V(f)$
, the implication

is always true for all
$n\geq 2$
as we have seen in Remark 3.3: the hard part of the conjecture is to prove the other implication.
Let us now explain the strategy that will be used for the proof of the Theorem for the various values of
$n\leq 5$
.
Framework 5.4. Assume that f defines a smooth cubic
$X=V(f)$
which is not of TS type (i.e.,
$[f]\in {\mathcal {V}}$
) such that
${\mathcal {H}}_f$
is not normal. Then by Lemma 4.1 and Remark 4.2, we have a family
${\mathcal {F}}$
of dimension
$n-2$
of triangles for
${\mathcal {H}}_f$
dominating, via the first projection, a component
$Y_1$
of
${\mathcal {D}}_{n-1}(f)$
of the same dimension. As we have done in the previous sections, we denote by
$Y_i$
the images of the projections
$\pi _i$
. As just observed,
$\pi _1:{\mathcal {F}}\to Y_1$
is generically finite.
If
$T=([x],[y],[z])=([x_1],[x_2],[x_3])$
is a general point of
${\mathcal {F}}$
, by generic smoothness, we can assume that the differentials
$d_T\pi _i:T_{{\mathcal {F}},T}\rightarrow T_{Y_i,[x_i]}$
are surjective; in particular,
$d_T\pi _1$
is an isomorphism. Moreover, since we are assuming
$n\leq 5$
, by Theorem 4.6, we have that none of the vertices of T belongs to the cubic
$V(f)$
(i.e.,
$x_i^3\neq 0$
for
$i\in \{1,2,3\}$
). If we set

by Lemma 2.10 and by Gorenstein duality, we have that
$\dim _{{\mathbb {K}}}(V_1)=3$
and
$\dim _{{\mathbb {K}}}(V_2)=n-2$
. Moreover, since
$x^3,y^3,z^3\neq 0$
, we also have
$V_1\cap V_2=\{0\}$
. Hence, by dimension reason, one has

If
$\{i,j,k\}=\{1,2,3\}$
, by Lemma 2.10 and Lemma 2.11, one has also

By dimension reason and since
$x_i\not \in V_2$
, we have a canonical isomorphism
$V_2\simeq \operatorname {\mathrm {Ann}}_{A^1}(x_j^2,x_k^2)/\langle x_i\rangle $
induced by the inclusion
$V_2\hookrightarrow \operatorname {\mathrm {Ann}}_{A^1}(x_j^2,x_k^2)$
followed by the quotient by
$\langle x_i\rangle $
. By Lemma 2.11 and since we have that
$d_{T}\pi _i$
is surjective, we have

so we can interpret
$d_T{\pi _i}$
as maps
$T_{{\mathcal {T}},T}\to V_2$
. Being
$d_T\pi _1$
an isomorphism, we have then the endomorphisms

Our approach is to analyse obstructions for such a configuration, by studying the Zariski tangent spaces. These information can be naturally codified by looking at the operators
$\psi _m$
; this strategy can be followed a priori in every dimension and in the cases of our interest yield to a direct and conclusive computation. In the proof of Theorem 5.1, we will start by analyzing these specific maps, ruling out both the cases where
$\psi _2$
(or
$\psi _3$
) can be diagonalized or not and ultimately proving that a family of triangles of dimension
$n-2$
cannot exist.
As we have done in the proof of Lemma 2.11, to a tangent vector
$\underline {v}=(x',y',z')\in T_{{\mathcal {T}},T}$
, we can associate the ‘first-order deformation’ of T in the direction of
$\underline {v}$
(for brevity,
$\underline {v}$
-deformation of T), which we write in a compact way as

Lemma 5.5. Let
${\mathcal {F}}$
be a family of triangles. Assume furthermore that both
${\mathcal {F}}$
and
$Y_1$
have dimension
$n-2$
. If the general element
$T=([x],[y],[z])\in {\mathcal {F}}$
has no vertices on
$X=V(f)$
, then
$x\cdot :V_2\to A^2$
is injective.
Proof. Since
$Y_1$
has dimension
$n-2$
, its general point
$[x]$
is in
${\mathcal {D}}_{n-1}(f)\setminus {\mathcal {D}}_{n-2}(f)$
(i.e., the kernel of the multiplication map
$x\cdot :A^1\to A^2$
has dimension
$2$
). The point
$[x]$
is also the vertex of an element
$T=([x],[y],[z])$
of
${\mathcal {F}}$
, and thus,
$\ker (x\cdot )=\langle y,z\rangle \subseteq V_1$
. However, by the assumptions, one has
$V_1\cap V_2={0}$
, as observed above.
Let us now show the Theorem 5.1 in the first cases.
Proposition 5.6. Theorem 5.1 is true for
$n\in \{2,3\}$
.
Proof. By contradiction, let us assume that
$[f]\in {\mathcal {V}}$
and
$\dim (\operatorname {\mathrm {Sing}}({\mathcal {H}}_f))=n-2$
. Notice that if
$n=2$
, since
$\operatorname {\mathrm {Sing}}({\mathcal {H}}_f)={\mathcal {D}}_1(f)$
, we have a contradiction by Theorem 3.4. Hence, we can assume
$n=3$
.
Fix the notation explained in the framework (see 5.4). Since
$n=3$
, we have that

for suitable
$u\in A^1\setminus \{0\}$
. All projections of
${\mathcal {F}}$
have dimension exactly
$1$
by Lemma 4.3, so
$T_{{\mathcal {F}},T}=\langle (Au,Bu,Cu)\rangle $
with
$A,B,C\in {\mathbb {K}}^*$
. By Lemma 2.11, the associated first-order deformation
$T+u(A,B,C)$
has to satisfy

We can then observe that the three independent points
$Bx+Ay, \ Cx+Az, \ Cy+Bz$
belong to the kernel of the multiplication by u. In other words,
$\iota (u)\supseteq {\mathbb {P}}^2$
, so that
$[u]\in {\mathcal {D}}_1(f)$
. Hence, as before, we have a contradiction.
Proposition 5.7. Theorem 5.1 is true for
$n=4$
: for a smooth cubic threefold
$X=V(f)$
, the Hessian quintic threefold
${\mathcal {H}}_f$
is normal if and only if f is not of TS type.
Proof. By contradiction, let us assume that
$[f]\in {\mathcal {V}}$
and
$\dim (\operatorname {\mathrm {Sing}}({\mathcal {H}}_f))=2$
. Then we are in the situation described in 5.4:
$T=([x],[y],[z])$
will denote a general triangle in
${\mathcal {F}}$
(recall that by Theorem 4.6 none of its vertices belongs to the cubic
$X=V(f)$
). Since
$n=4$
, we have that
$\dim (V_2)=2$
. Recall that we have the endomorphisms

which have image
$T_{Y_2,[y]}$
and
$T_{Y_3,[z]}$
, respectively. We have that either one of the two is diagonalizable or that none is. We treat differently the two cases.
Case (I): Let us suppose that at least one of the two above endomorphisms is diagonalizable. W.l.o.g. we can assume that
$\{u,w\}$
is a basis of
$V_2$
whose elements are eigenvectors for
$\psi _2$
. Then two independent tangent vectors to
${\mathcal {F}}$
in T are given as

for suitable
$A,B,C,D,E,F$
scalars depending on the triangle, where, in particular, A and B are the eigenvalues corresponding respectively to u and w.
We recall that, for a given tangent vector
$\underline {v}=(x',y',z')$
at
$T=([x],[y],[z])$
, we have the relation
$xy'+yx'=0$
by Lemma 2.11. For brevity, we refer to this relation with the notation
$(\underline {v})_{xy}$
. We denote by
$(\underline {v})_{xz}$
and
$(\underline {v})_{yz}$
the analogous relations.
For example, using the vectors in (5.6), we have the corresponding first-order deformations

which yield

By elementary operations, one gets two new relations:

For example, the former relation is obtained by substituting in
$(\underline {v})_{yz}$
of Equation (5.7), the products
$uy$
,
$wy$
, and
$uz$
obtained respectively from
$(\underline {v})_{xy}$
,
$(\underline {v}')_{xy}$
and
$(\underline {v})_{xz}$
of Equation (5.7).
By Lemma 5.5, the multiplication map
$x\cdot :V_2\to A_2$
is injective, so

Let us now observe that A and B cannot be simultaneously zero when evaluated in a general triangle T; otherwise, the second projection of
${\mathcal {F}}$
would be zero-dimensional, contradicting Lemma 4.3.
Lemma 5.8. In this situation, for T general, none of the two eigenvalues of the endomorphism
$\psi _2$
can be zero.
Proof. Let
$T=([x],[y],[z])$
be a general triangle of
${\mathcal {F}}$
. W.l.o.g. we can assume by contradiction that
$A\neq 0$
and
$B=0$
. We claim that
$Y_2\subseteq \iota (z)$
and
$Y_3\subseteq {\mathcal {D}}_2(f)$
.
Since
$A\neq 0$
, we also get that
$C=D=E=0$
by Equation (5.8). Hence, the first-order deformations of T are given by

Since the differential maps of the projections from
${\mathcal {F}}$
are surjective, we have
$\dim (Y_1)=2$
and
$\dim (Y_2)=\dim (Y_3)=1$
.
Consider the curve
$C_T=\pi _3^{-1}([z])$
. By construction, the tangent to
$C_T$
in T is spanned by
$\underline {v}$
which is projected to u and
$Au$
via
$d_T(\pi _1|_{C_T})$
and
$d_T(\pi _2|_{C_T})$
, respectively. Hence,
$\pi _1(C_T)$
is a curve in
$Y_1$
and
$\pi _2(C_T)=Y_2$
. In particular,
$\pi _1(C_T)\cup Y_2\subseteq \iota ([z])$
, as claimed. Moreover, since
${\mathcal {D}}_1(f)=\emptyset $
, by assumption (by Theorem 3.4), we get that
$\iota ([z])\simeq {\mathbb {P}}^2$
up to the case where it is a projective line coinciding both with
$Y_2$
and
$\pi _1(C_T)$
. But in this last case, we would have an involution on
$Y_2\simeq {\mathbb {P}}^1$
, which yields a fixed point and then a singular point for
$V(f)$
. Hence,
$Y_3\subseteq {\mathcal {D}}_2(f)$
, as claimed.
Notice that the same argument can be used to prove that
$Y_3\subseteq \iota ([y])$
and thus that
$Y_2\subseteq {\mathcal {D}}_2(f)$
. By Proposition 3.5, one can see that for two general points
$[z]$
and
$[z']$
in
$Y_3$
, we have that
$\iota ([z])\neq \iota ([z'])$
. Hence,
$Y_2\subseteq \iota ([z])\cap \iota ([z'])={\mathbb {P}}^1$
, and thus,
$Y_2={\mathbb {P}}^1\subseteq {\mathcal {D}}_2(f)$
. This is impossible by Theorem 3.4 since f is not of TS type.
From the above Lemma 5.8, since T is general, we have that both A and B are not zero; hence, by Equation 5.8, we also get
$B=-A$
. Indeed, if
$A+B\neq 0$
, we would obtain
$C=D=E=F=0$
, which is not possible by Lemma 4.3. Since
$A=-B\neq 0$
, from Equations (5.8), we have
$C=F=0$
. Then, the first-order deformation, in this case, can be written as

and the conditions (5.7) are equivalent to

Moreover, we cannot have
$D=E=0$
as observed above, so we can assume
$D\neq 0$
.
Since these equations hold, by assumption, for the general point
$T\in {\mathcal {F}}$
, by deforming T at the first order in the direction of
$\underline {v}$
, that is, by considering a curve

also the corresponding eigenvectors of
$\psi _2$
‘move’. More precisely, we have two curves

defined in a neighbourhood of
$0$
such that
$\gamma _u(0)=u$
,
$\gamma _w(0)=w$
and
$\{\gamma _u(t),\gamma _w(t)\}$
is a basis of eigenvectors of
$d_{T(t)}\pi _2\circ d_{T(t)}\pi _1^{-1}$
. These eigenvectors satisfy equations analogous to the ones in (5.7) where the coefficients depend on t. As observed above, the sum of the two eigenvalues of
$\psi _2$
is
$0$
also in a neighbourhood of T, so that the Equations (5.11) hold also locally.
We can then consider an expansion of the curves

and substitute them in the Equations (5.11) in order to get new relations. We write
$A(t)=A+A't+t^2(\cdots )$
for the curve following the eigenvalue relative to
$\gamma _u(t)$
with an analogous notation for the coefficients that appear in Equations (5.11).
For example, from the condition
$u(y+Ax)=0$
, one has

so we get
$2Au^2+A'xu+u'(Ax+y)=0$
. One can do the same reasoning for the
$\underline {v}'$
-deformation, and we also can use two ‘parameters’ to take into account in a compact description the deformation of T in the direction of
$t\underline {v}+s\underline {v}'$
. In this way, u and w ‘deform’ at first order as

respectively. Moreover, we can assume that
$u',u"$
and
$w',w"$
do not depend on u and w, respectively. This argument yields the relations




First of all, observe that multiplying by x Equation (5.14)
${}_I$
, since
$xy=0=x^2w$
, we get

Hence, multiplying by x Equation (5.15)
${}_I$
, we obtain

Since
$xuw=0$
, from Equation (5.11), one can easily see that also

We claim now that
$A'=A"=0$
and
$xu^2,xw^2\neq 0$
. Indeed, let us observe that the product
$xu$
vanishes if multiplied by
$x, \ y, \ z, \ w$
. If
$xu^2=0$
too, then by the Gorenstein duality in the apolar ring
$A_f$
, we would get that
$xu=0$
, which is not possible as observed with Lemma 5.5.
In the same way, one sees that
$xw^2\neq 0$
. Then, recalling that
$u(Ax+y)=0=w(y-Ax)$
, we can multiply by u and by w, respectively, Equations (5.12)
${}_{II}$
and (5.14)
${}_I$
, getting
$A"xu^2=0$
and
$A'xw^2=0$
, and so the claim:

Lemma 5.9. The tangent vectors
$w'$
and
$u"$
are trivial.
Proof. Let us prove it for
$w'$
. Since we can assume that
$w'$
does not depend on w, we can write it as
$w'=\alpha x+\beta y+\gamma z+\delta u$
. Multiplying Equation (5.14)
${}_I$
by x and y, we get
$x^2w'=0$
and
$y^2w'=0$
respectively. Moreover, multiplying by z Equation (5.13)
${}_I$
, we have also
$z^2w'=0$
. These last conditions yield

but since no vertex for the general triangle T belongs to
$V(f)$
, we have that
$\alpha =\beta =\gamma =0$
. Finally, since we have just shown that
$A'=0$
, from Equation (5.14)
${}_I$
, we get

Since, from Equations (5.11), we get
$uy=-Aux$
, we would have
$-2\delta Axu=0$
, which implies that
$\delta =0$
, by Lemma 5.5.
Then
$w'=0$
, as claimed. The same reasoning can also be used to prove that
$u"=0$
.
Remark 5.10. From the above lemma, one can see that we can assume that also
$E\neq 0$
. Indeed, if
$E\equiv 0$
locally (and thus, we can simply set
$E'=E"=0$
in the above equations), from Equation (5.13)
${}_I$
we would have
$w^2=0$
. Hence,
$[w]$
would be a singularity for
$V(f)$
, which is not possible.
We claim now that
$wu^2=0$
and
$w^2u=0$
. Multiplying by u Equation (5.15)
${}_I$
, one gets

Since we have just shown that for T general, also the condition
$zu^2=0$
is satisfied (see Equation (5.17)), we can deform it at the first order:

Putting together Equations (5.19) and (5.20), one gets

Let us now do the same for the second claim. Multiplying by w the Equation (5.13)
${}_{II}$
, we get
$2Euw^2+zww"=0$
. Moreover, by deforming the condition
$zw^2=0$
(see Equation (5.17)), we get
${Euw^2+2zww"=0}$
. As before, putting together these last conditions, one gets

by using
$E\neq 0$
(see Remark 5.10).
We claim now that
$zuw=0$
. In order to show this last claim, let us deform the condition just obtained (i.e.,
$u^2w=0$
):

Write
$u'$
as
$\alpha x+\beta y+\gamma z+\delta w$
for simplicity. Since
$xuw=yuw=uw^2=0$
and since
$z^2w=0$
by definition of w, one has

If
$\gamma \neq 0$
we have done, let us then assume
$\gamma =0$
: we get
$z^2u'=0$
, and multiplying by z Equation (5.15)
${}_I$
we get
$2Dzuw=0$
, as desired.
Finally, having
$zuw=0$
yields a contradiction: in this case, from Equations (5.11), we would have
$xw^2=0$
, which is not possible by Equation (5.18). Hence, neither
$\psi _2$
nor
$\psi _3$
can be diagonalizable for T general.
Case(II): For T general, the map
$\psi _2$
is not diagonalizable. We can choose a basis
$\{u,w\}$
of
$V_2$
in such a way that
$\psi _2$
is written in its Jordan normal form

As done before, the corresponding first-order deformations are

with
$C,D,E,F$
not all simultaneously zero (by Lemma 4.3).
These, using Lemma 2.11, yield

Again, by elementary operations, one gets:

By Lemma 5.5, the multiplication map
$x\cdot :V_2\to A_2$
is injective so

Notice that in the case where
$A\neq 0$
, from the above relations, one easily sees that also
$D=C=F=E=0$
, which is not possible, as we have stressed before, so we can assume

One can then observe that the matrix associated to the endomorphism
$\psi _3$
with respect to the basis
${u,w}$
is of the form

If
$C\neq 0$
, the map
$\psi _3$
would be diagonalizable: this is not possible for T general as proved in Case (I). We can then assume that

We are then considering the first-order deformations

with
$E\neq 0$
, and the relations (5.21) are then equivalent to

By considering the
$\underline {v}$
-deformation and the
$\underline {v}'$
-deformation of the first two equations, we obtain

We claim now that

Conditions
$yu^2=zu^2=0$
and
$xu^2=0$
follow easily by multiplying by u or by x the equations in (5.23). One obtains
$u^3=0$
from Equation (5.24)
${}_{II}$
after multiplying by u and by remembering that
$yu=0$
. Since
$u^2$
annihilates
$x,y,z$
and u, it cannot annihilate w too, since, otherwise, from the perfect pairing induced by the Gorenstein duality in
$A_f$
,
$[u]$
would give a singular point for
$V(f)$
.
As a consequence of the above relation, notice that
$([y],[z],[u])$
is a triangle for
${\mathcal {H}}_f$
since
$yz=yu=zu=0$
. Hence, by Lemma 2.10, we have
$\dim (\langle y^2,z^2,u^2\rangle )=3$
. Being
$([x],[y],[z])$
a triangle and by Equation (5.25), we can conclude

We claim now that
$u"\in \langle x,u\rangle $
. By the above relation, it is enough to show that
$y^2u"=z^2u"=u^2u"=0$
. The first relation comes from (5.24)
${}_{II}$
if we multiply both terms by y and use (5.26). One gets the second relation working on the Equation (5.24)
${}_{IV}$
and using
$E\neq 0$
. To get the third and last relation, let us simply observe that we have shown that the equation
$u^3=0$
holds for the general triangle T in
${\mathcal {F}}$
and so, we can write its
$\underline {v}'$
-deformation:

as claimed.
As a consequence of the last claim, we can write
$u"=\alpha x+\beta u$
for suitable
$\alpha ,\beta \in {\mathbb {K}}$
. Now consider Equation (5.24)
${}_{II}$
and recall that
$yu=0$
by Equation (5.23). By substituting, one obtains

which is impossible by Proposition 2.1, since
$V(f)$
is smooth. This concludes the analysis of Case (II) and, consequently, the proof of the main theorem for the case of cubic threefolds.
6 Proof of main theorem: the cubic fourfold case
In this section, we prove Theorem A in the last remaining case: given any smooth cubic fourfold
$X=V(f)$
, the Hessian variety
${\mathcal {H}}_f$
is normal and irreducible if and only if f is not of TS type.
We set ourselves in the framework described in 5.4. We assume by contradiction that given
$X=V(f)$
, a smooth cubic fourfold, with f which is not of TS type the associated variety
${\mathcal {H}}_f$
is not normal. Then there exists an irreducible
$3$
-dimensional family
${\mathcal {F}}$
of triangles for
${\mathcal {H}}_f$
with the first projection dominating a
$3$
-dimensional component of
$\operatorname {\mathrm {Sing}}({\mathcal {H}}_f)$
. Fixing a general triangle
$T=([x],[y],[z])\in {\mathcal {F}}$
for
${\mathcal {H}}_f$
, let us now study the behaviour of
$\psi _2$
and
$\psi _3$
as endomorphisms of
$V_2$
. We distinguish the following mutually exclusive cases:
-
1. for the general T,
$\psi _2$ (or
$\psi _3$ ) has a Jordan decomposition with one Jordan block;
-
2. for the general T,
$\psi _2$ (or
$\psi _3$ ) has a Jordan decomposition with two Jordan blocks;
-
3. for the general T,
$\psi _2$ and
$\psi _3$ are diagonalizable.
We will rule out all the possibilities, by proving the following Lemmas 6.1 (for the case
$(a)$
), 6.2 (for
$(b)$
), 6.3 and 6.4 (both of them dealing with the case
$(c)$
). The last one, concerning a particular subcase of
$(c)$
, is proved in the dedicated subsection 6.1.
Let us start by ruling out case
$(a)$
.
Lemma 6.1. For T general, neither the map
$\psi _2$
nor
$\psi _3$
can have a Jordan decomposition with only one block.
Proof. Let us suppose, w.l.o.g, that
$\psi _2$
has a Jordan decomposition with one Jordan block. We can choose a basis
$\{u,v,w\}$
of
$V_2=\operatorname {\mathrm {Ann}}_{A^1}(x^2,y^2,z^2)$
in such a way that
$\psi _2$
is written in its Jordan normal form with w as eigenvector. Then we have a basis
$\{x,y,z,u,v,w\}$
of
$A^1$
with the first three vectors such that
$x^3,y^3,z^3\neq 0$
. Then three independent tangent vectors to
${\mathcal {F}}$
in T are given as

for suitable scalars depending on the triangle.
As done in the case of threefolds, one uses Lemma 2.11 in order to obtain conditions from the first-order deformations associated to
$\underline {v},\underline {v}'$
and
$\underline {v}"$
. By elementary operations between these equations and by using Lemma 5.5, one gets the following relations on the coefficients appearing in the above description of the tangent vectors:

Note that if
$A\neq 0$
, then one has that all the other coefficients have to be
$0$
, which is not possible since the second and third projections cannot send
${\mathcal {F}}$
to a point by Lemma 4.3. Hence,
$A=0$
and then
$E=H=I=C+G=B+F=F+L=0$
. We can then write the above tangent vectors as

Then, the above-mentioned equations can be reduced to the following system of equations:

As usual, if T is deformed in the direction of
$t\underline {v}+s\underline {v}'+r\underline {v}"$
, we have the corresponding deformation
$u+tu'+su"+ru"'$
of u (and analogously the ones for v and w).
Claim:
$uw^2\neq 0$
,
$L=0$
and
$zw=0$
.
First of all, let us study
$\ker (w^2\cdot : A^1\to A^3)$
. Clearly,
$yw^2=0$
by Equation
$(\underline {v}")_{xy}$
. Moreover, if we multiply by w Equation
$(\underline {v}')_{xy}$
and use Equation
$(\underline {v}")_{xy}$
, we get
$xw^2=0$
. Similarly, one gets
$zw^2=0$
upon multiplying by w Equation
$(\underline {v}")_{xz}$
. Since Equation
$(\underline {v}")_{xy}$
holds for T general, one can deform it in the direction of
$t\underline {v}+s\underline {v}'+r\underline {v}"$
and obtain

This yields

If we multiply by w these relations, we get
$w^2v=0$
and
$w^3=0$
. Hence, we have

Observe now that
$uw^2\neq 0$
. Indeed, if
$uw^2=0$
, then we would also have
$w^2\cdot A^1=\{0\}$
so, by Gorenstein duality, this would imply
$w^2=0$
, which contradicts the smoothness of the cubic fourfold
$V(f)$
.
Finally, by deforming Equation
$(\underline {v}")_{xz}$
and multiplying by w, using the various vanishings obtained before, we get
$2Luw^2=0$
, and thus,
$L=0$
. Then one has the claim by Equation
$(\underline {v}")_{xz}$
.
Claim:
$zv=0$
,
$G=0$
and
$D\neq 0$
.
By deforming equations
$(\underline {v})_{xy}$
in the direction of
$t\underline {v}+s\underline {v}'$
we get, respectively,

If one multiplies these by z, one obtains
$zuv=zv^2=0$
. One can now observe that
$\ker (zv\cdot :A^1\to A^3)=A^1$
, so by Gorenstein duality, one has
$zv=0$
, as claimed. Since
$xw\neq 0$
(by Lemma 5.5), from Equation
$(\underline {v}')_{xz}$
one obtains
$G=0$
and, consequently, by Lemma 4.3, also
$D\neq 0$
.
Claim:
$uw^2=0$
.
Since
$zv=0$
for T general, we can deform this equation in the direction of
$t\underline {v}$
. We get
$zv'+Dvw=0$
, and so
$Dv^2w=0$
, if we multiply by v. Being
$D\neq 0$
, one has also
$v^2w=0$
. Let us now deform
$(\underline {v})_{xz}$
and
$(\underline {v}')_{xy}$
in the direction of
$t\underline {v}$
in order to get

Upon multiplying by w, one gets
$xww'+2uw^2=xww'+uw^2=0$
, which yields
$xww'=uw^2=0$
. This is impossible as observed in the first claim above.
Let us now prove that case
$(b)$
cannot be realised.
Lemma 6.2. For T general, neither the map
$\psi _2$
nor
$\psi _3$
can have a Jordan decomposition with two blocks.
Proof. Let us suppose, w.l.o.g, that
$\psi _2$
has a Jordan decomposition with two Jordan blocks. As done in Lemma 6.1, we can choose a basis
$\{u,v,w\}$
of
$V_2=\operatorname {\mathrm {Ann}}_{A^1}(x^2,y^2,z^2)$
in such a way that
$\psi _2$
is written in its Jordan normal form with v and w as eigenvectors. Then, in this case, we can write three independent tangent vectors to
${\mathcal {F}}$
in T as

for suitable scalars depending on the triangle.
As done in the previous cases, one gets the following relations involving the coefficients appearing in the above description of the tangent vectors:

We distinguish four cases, depending on the vanishing of the two eigenvalues.
Case (I):
$\mathbf {A=B=0}$
.
Since
$A=B=0$
, we also have
$F=H=I=C+G=0$
. Among the various equations obtained by deforming at first order the general triangle T, one gets

We claim now that
$uv^2\neq 0$
. By deforming at first order the equation
$yv=0$
in the direction of
$t\underline {v}$
, one gets
$yv'+v^2=0$
, which implies that

However, this has to be an equality; otherwise, we would have
$v^2=0$
by Gorenstein duality. In particular,
$uv^2\neq 0$
.
From the Equation
$yv'+v^2=0$
, one can also see that
$yv'\neq 0$
; otherwise, we would contradict the smoothness of
$V(f)$
. We claim now that
$yv'=0$
, so we conclude Case (I).
Since T is a triangle and by Equations (6.2), we have
$\langle x,z,v,w\rangle \subseteq \ker (y\cdot :A^1\to A^2)$
, so in order to prove
$yv'=0$
, it is enough to show that
$v'$
does not depend on y and u. One easily sees that
$0=y(yv'+v^2)=y^2v'$
. Since, by assumption,
$\langle x,z,u,v,w\rangle =\ker (y^2\cdot :A^1\to A^3)$
, we get that
$v'$
does not depend on y. Moreover, since we have just shown in Equation (6.3) that for T general, also the equation
$v^3=0$
holds, we can deform it and in the same way, one proves that
$v'$
does not depend on u.
Case (II):
$\mathbf {A=0, B\neq 0}$
.
In this case, we also have
$E=F=C+G=H=I=L=M=0$
so that

Among the equations deduced by deforming the general triangle, one gets the conditions

We claim now that
$C=0$
. First of all, notice that

Indeed, we have
$v\in V_2$
by assumption so
$x^2v$
, and since T is a triangle, we also have
$xyv=xzv=0$
. The last two vanishing can be easily obtained form Equations (6.4) upon a multiplication by v and w (and by recalling that we are assuming
$B\neq 0$
). In particular, we have
$xuv\neq 0$
, by Lemma (5.5).
Then by taking the fifth and sixth equations in (6.4) multiplied by u and v, respectively, one has

so
$Cxuv=0$
. Then, since
$xuv\neq 0$
, we have necessarily
$C=0$
for the general triangle, and so
$D\neq 0$
by Lemma 4.3. In particular, we have
$\psi _3(u)=Dv\neq 0$
and
$\psi _3(v)=\psi _3(w)=0$
. Hence, for the general triangle T,
$\psi _3$
is not diagonalizable and has two Jordan blocks with eigenvalues both equal to
$0$
. This is impossible, as seen in Case (I).
Case (III):
$\mathbf {A\neq 0, B=0}$
.
This case can be treated in a ‘geometric’ way as done in Lemma 5.8. Indeed, since
$A\neq 0$
and
$B=0$
, we have also all the other variables, besides M, are zero. Moreover, by Lemma 4.3,
$M\neq 0$
. In particular, the tangent vectors to
${\mathcal {F}}$
in T are spanned by

and the varieties
$\pi _i({\mathcal {F}})=Y_i$
have dimension
$3,2$
and
$1$
, respectively. As done in the other cases, by studying the relations coming from the deformation at the first order, one can easily see that
$Y_3$
is contained in
${\mathcal {D}}_2(f)$
and
$Y_2$
is a surface living in
${\mathcal {D}}_3(f)$
. Since
$Y_2$
cannot be contained in
${\mathcal {D}}_2(f)$
(otherwise, we would have singular points for
$V(f)$
by Proposition 2.5), for the general
$[y]\in Y_2$
, we have that
$\iota ([y])\simeq {\mathbb {P}}^2$
.
For a general triangle
$T=([x],[y],[z])$
, consider the curve
$C_T=\pi _2^{-1}([y])$
. The tangent to
$C_T$
in T is generated by
$\underline {v}"$
which is projected to w and
$Mw$
via
$d_T(\pi _1|_{C_T})$
and
$d_T(\pi _3|_{C_T})$
, respectively. Hence,
$\pi _1(C_T)$
is a curve in
$Y_1$
and
$\pi _3(C_T)=Y_3$
.
In particular,
$\pi _1(C_T)\cup Y_3\subseteq \iota ([y])\simeq {\mathbb {P}}^2$
. By varying the point
$[y]$
, the kernel has to move, since the curve
$\pi _1(C_T)$
has to cover the threefold
$Y_1$
. Hence,
$Y_3$
lies in the intersection of distinct projective planes: we have
${\mathbb {P}}^1\simeq Y_3$
and thus a line in
${\mathcal {D}}_2(f)$
. This implies, by Theorem 3.4, that f is of TS type, against our assumptions.
Case (IV):
$\mathbf {A,B\neq 0}$
.
First of all, notice that assuming
$A,B\neq 0$
implies that
$A+B=0$
. Indeed, if we assume also
$A+B\neq 0$
, we would obtain that all the other coefficients are equal to
$0$
. This is impossible by Lemma 4.3. Then

with
$E,L$
not both zero. In particular,
$\psi _3$
is not diagonalizable for the general triangle T and all its eigenvalues are zero. This is impossible as seen in the previous cases.
As a consequence of Lemmas 6.1 and 6.2, the maps
$\psi _2$
and
$\psi _3$
have to be diagonalizable. In what follows, we rule out this remaining case, splitting it up into two lemmas, the second of which is postponed in the following subsection.
Lemma 6.3. For T general, neither the map
$\psi _2$
nor
$\psi _3$
can be diagonalizable.
Proof. As a consequence of Lemma 6.1 and 6.2, we have that
$\psi _2$
and
$\psi _3$
are both diagonalizable for the general triangle. We can choose a basis
$\{u,v,w\}$
of
$V_2=\operatorname {\mathrm {Ann}}_{A^1}(x^2,y^2,z^2)$
in such a way that
$\psi _2$
is in diagonal form. Thus, three independent tangent vectors to
${\mathcal {F}}$
in T are

for suitable coefficients depending on the triangle.
As done so far, one gets the following relations:

Notice that
$A,B$
and C cannot be all equal to zero by Lemma 4.3. Hence, we distinguish three cases, depending on the vanishing of the three eigenvalues.
Case (I):
$\mathbf {A=B=0}$
and
$\mathbf {C\neq 0}$
.
In this case, one can easily see that three tangent vectors to
${\mathcal {F}}$
at T can be written as

for suitable coefficients so that
$\dim (Y_1)=3$
,
$\dim (Y_2)=1$
and
$\dim (Y_3)\in \{1,2\}$
.
-
• Claim:
$\dim (Y_3)=1$ and
$Y_2\subseteq {\mathcal {D}}_2(f)$ .
Among the equations obtained by deforming the general triangle T, one has
$yu=yv=0$ . Moreover, by definition of triangle, we clearly have also
$yx=yz=0$ : the general
$[p]\in Y_2$ lives in
${\mathcal {D}}_2(f)$ ; that is, for
$[p]\in Y_2$ , one has
$\iota ([p])\simeq {\mathbb {P}}^3$ (since
${\mathcal {D}}_1(f)$ is empty by hypothesis). In the same way, since one gets
$zw=0$ , one also has
$Y_3\subseteq {\mathcal {D}}_3(f)$ . Notice that the surface
$\pi _2^{-1}([p])=S_T$ projects onto a surface in
$Y_1$ via
$\pi _1$ and dominates
$Y_3$ via
$\pi _3$ . This means that
$Y_3\subset \iota ([p])\simeq {\mathbb {P}}^3$ . Observe that if
$[q_1]$ and
$[q_2]$ are distinct points of
$Y_2$ , then their kernels have to be distinct: indeed, if
$\Lambda :=\iota ([q_1])=\iota ([q_2])$ , then by symmetry,
$\iota (\Lambda )\supset {\mathbb {P}}(\left \langle [q_1],[q_2]\right \rangle $ (i.e.,
$\Lambda \subseteq {\mathcal {D}}_4(f)$ ), which is not possible by Theorem 3.4. Then, for
$[q_1]$ and
$[q_2]$ two distinct general points of
$Y_2$ , one gets that
$Y_3\subseteq \iota ([q_1])\cap \iota ([q_2])\simeq {\mathbb {P}}^s$ with
$s\in \{1,2\}$ . From this, one can see that
$\dim (Y_3)=1$ : indeed, if
$Y_3$ is a surface, then we necessarily have
$s=2$ and
$Y_3\simeq {\mathbb {P}}^2$ , but this means there exists a projective plane in
${\mathcal {D}}_3(f)$ , which is impossible by Theorem 3.4
-
• Claim:
$Y_3\subseteq {\mathcal {D}}_2(f)$ .
Since
$Y_3$ is a curve, the endomorphism
$\psi _3|_{\langle u, v\rangle }$ has necessarily rank
$1$ (i.e., there exists a vector
$au+bv$ that is sent to
$0$ by
$\psi _3$ ). Among the first-order conditions given by the tangent vectors above, one has
$$ \begin{align*}(Du+Ev)x+zu=(Gu+Hv)x+zv=0.\end{align*} $$
$\psi _3(au+bv)=a(Du+Ev)+b(Gu+Hv)=0$ , one has
$z(au+bv)=0$ . Hence,
$\iota ([z])\simeq {\mathbb {P}}^3$ , and we have that
$Y_3$ is contained in
${\mathcal {D}}_2(f)$ too.
Being f not of TS type, and being
$Y_3$ a curve in
${\mathcal {D}}_2(f)$ , we have that
$Y_3$ is not a line. However, for
$[q_1]$ and
$[q_2]$ distinct general points in
$Y_2$ as above,
$Y_3\subseteq \iota ([q_1])\cap \iota ([q_2])\simeq {\mathbb {P}}^2$ , so the general triangle
$T=([x],[y],[z])$ is such that
$\iota ([y])\simeq {\mathbb {P}}^3$ contains a fixed
${\mathbb {P}}^2$ , denoted by
$\Pi $ , which coincides with the projective plane spanned by
$Y_3$ . To conclude, let us take a general point
$[\eta ]\in \Pi $ : by symmetry, the general
$[p]\in Y_2$ is such that
$[p]\in \iota ([\eta ])$ , and so
$Y_2\subset \iota ([\eta ])\simeq {\mathbb {P}}^r$ . Since
$Y_2\not \simeq {\mathbb {P}}^1$ , we have that
$r\geq 2$ : this means that
$\Pi \simeq {\mathbb {P}}^2\subseteq {\mathcal {D}}_3(f)$ , which yields a contradiction as above.
Let us stress that having a
${\mathbb {P}}^2$
contained in
${\mathcal {D}}_3(f)$
is a phenomenon that happens exactly when
$V(f)$
is a smooth, non-cyclic cubic of
$TS$
type as described in the specific Example 3.9.
Case (II):
$\mathbf {A=0}$
and
$\mathbf {B,C\neq 0}$
.
This case cannot occur. It will be treated in Lemma 6.4.
Case (III):
$\mathbf {A,B,C\neq 0}$
.
Being
$A,B,C\neq 0$
, one has
$D=H=N=0$
. Notice that the three values
$A+B,A+C$
and
$B+C$
cannot be simultaneously zero; moreover, at least one of them has to be
$0$
, since otherwise we would get
$D=E=F=G=H=I=L=M=N=0$
, which is impossible by Lemma 4.3. W.l.o.g, we distinguish
$2$
cases: either
$A+C=0$
and
$A+B,B+C\neq 0$
or
$A+B=A+C=0$
and
$B+C\neq 0$
. In the first case,
$\psi _3$
is diagonalizable with one zero eigenvalue, whereas in the second case, one has the same conclusion or that
$\psi _3$
is not diagonalizable and its Jordan normal form has
$1$
Jordan block with
$0$
as the only eigenvalue. Both conclusions yield a contradiction as observed in the previous cases or in Lemma 6.1.
6.1 The end of the proof
To end the proof of Theorem 5.1, we have to rule out a last remaining possibility which could arise in the case where the both the maps
$\psi _2$
and
$\psi _3$
are diagonalizable (case
$(c)$
, as stated at the beginning of Section 6). This subsection is devoted to this subcase, which is ruled out with the following Lemma 6.4, which yields also the end of the proof of the main theorem. To prove this last Lemma, we start by analyzing the usual framework 5.4 and obtaining different relation that both the vertices of the general triangle and the tangent vectors to it have to satisfy. After that, we will use these conditions to reconstruct the cubic fourfolds which the framework is, in this case, associated with, showing that these do not actually satisfy the hypotheses we are setting.
Lemma 6.4. For T general, neither the map
$\psi _2$
nor
$\psi _3$
can be diagonalizable with dimension of the kernel equal to
$1$
.
Proof. We refer to the notations introduced at the beginning of Lemma 6.3. W.l.o.g, we can set
$A=0$
so that
$B,C\neq 0$
by hypothesis. Then, one has
$E=F=G=H=L=N=0$
. First of all, notice that
$B+C=0$
. Indeed, otherwise, we would get
$I=M=0$
, so v and w would be two eigenvectors for
$\psi _3$
with associated eigenvalue
$0$
. This is impossible as observed in Case (I).
Then, the tangent vectors to
${\mathcal {F}}$
at T can be written as

with
$B\neq 0$
and
$(D,I,M)\neq (0,0,0)$
by Lemma 4.3. Moreover, notice that
$I,M\neq 0$
since, otherwise, we would have that
$\psi _3$
is not diagonalizable (and this cannot happen for T general by Lemma 6.2).
The first-order conditions obtained as a consequence of Lemma 2.11 are

Consider the following subsets of
${\mathcal {L}}={\mathbb {K}}[x,y,z,u,v,w]$
:



and the elements

Notice that all the monomials in
$M_0$
are
$0$
in
$A_f={\mathcal {L}}/\operatorname {\mathrm {Ann}}_{\mathcal {L}}(f)$
by the conditions imposed by our framework. We want to prove that the same holds for all the elements in
$M_i$
for
$i\in \{1,\dots , 5\}$
and for
$r_1,r_2$
and
$r_3$
, whereas all the monomials in
${\mathcal {M}}^{nv}$
are not
$0$
(in
$A_f$
).
If T is deformed in the direction of
$t\underline {v}+s\underline {v}'+r\underline {v}"$
, the corresponding first-order deformation of u is written as
$u+tu'+su"+ru"'$
(and analogously the ones for
$v,w, B,D, I$
and M).
Claim: All the monomials in
$M_1$
and in
$M_2$
are
$0$
.
One has
$yu=0$
from Equation
$(\underline {v})_{xy}$
. Upon multiplying by u, the other equations in (6.5), one gets the vanishing for the monomials in
$M_1$
. By deforming Equation
$(\underline {v})_{xy}$
in the direction of
$s\underline {v}'+r\underline {v}"$
, we get

Multiplying by u these relations and by using the vanishing
$yu=0$
, one gets the claim.
Claim: All the monomials in
$M_3$
are
$0$
.
Observe that it is enough to show hat
$xvw=0$
: all the other vanishings come from Equations (6.5) after multiplication by v or w and vanishing in
$M_1$
or
$M_2$
.
Let us deform Equation
$(\underline {v}')_{xy}$
in the direction of
$r\underline {v}"$
:

Multiplying by x, we get the relation
$x^2v"'=0.$
Recalling that
$V_2=\langle u,v,w\rangle =\operatorname {\mathrm {Ann}}_{A_1}(x^2,y^2,z^2)$
, let us consider the vanishing
$x^2v=0$
: its deformation in the direction of
$r\underline {v}"$
yields
$x^2v"'+2xvw=0$
. Since, as just shown,
$x^2v"'=0$
, we get
$xvw=0$
, as claimed.
Claim: All the monomials in
${\mathcal {M}}^{nv}$
besides
$uv^2$
are not
$0$
and
$B"=B"'=0$
.
We have
$x^3,y^3,z^3\neq 0$
by assumption, since the general triangle of
${\mathcal {F}}$
cannot have a vertex on the cubic fourfold. As a consequence of the framework and since the monomials in
$M_1$
and
$M_3$
are
$0$
, we get that
$\langle x,y,z,u,w\rangle \subseteq \ker (xv\cdot : A^1\rightarrow A^3)$
. Hence,
$xv^2$
cannot be zero; otherwise, by Gorenstein duality, also
$xv$
would be
$0$
, which is not possible as observed in Lemma 5.5. In the same way, one gets that also
$xu^2$
and
$xw^2$
are not
$0$
. For the remaining monomials, they have to be different from
$0$
; otherwise, one would get a contradiction multiplying the equations in (6.5) by
$u, v$
or w.
For the second claim, observe that multiplying by v Equation (6.6), since
$(Bx+y)v=0$
by Equation
$(\underline {v}')_{xy}$
, one gets
$B"'xv^2=0$
. Being
$xv^2\neq 0$
, as just shown, we have also that
$B"'=0$
, as claimed. In order to show that
$B"=0$
, one proceeds in an analogous way by deforming Equation
$(\underline {v}")_{xy}$
in the direction of
$s\underline {v}'$
:

One gets the claim by multiplying by w.
Claim: The monomials in
$M_4$
are
$0$
.
Let us consider the first-order deformation of
$(\underline {v}')_{xy}$
and
$(\underline {v}")_{xy}$
in the direction of
$s\underline {v}'$
and
$r\underline {v}"$
, respectively:

Since
$B"=B"'=0$
as shown in the previous claim, if one multiplies the above Equations (6.8) by v and w, respectively, one gets the claim.
Claim: One has
$w"=v"'=0$
as tangent vectors.
Let us start by proving that
$w"=0$
. Since, by construction, we have that
$w\in V_2$
, one can deform in the direction of
$s\underline {v}'$
the relations
$x^2w=y^2w=z^2w=0$
. Recalling that
$xvw=yvw=zw^2=0$
by the previous claims, one obtains that
$w"\in V_2$
, so we can write
$w"=\alpha u+\beta v+\gamma w$
. By substituting this expression in Equation (6.7), one has

which, if multiplied by u, gives
$\alpha B xu^2=0$
. Since
$Bxu^2\neq 0$
, one has
$\alpha =0$
.
Being
$\alpha =0$
, it follows
$\beta (-Bx+y)v=0$
from the above equation. However, one has
$(Bx+y)v=0$
(see Equation
$(\underline {v}^{\prime }_{xy})$
), so
$\beta =0$
. Indeed, otherwise, we would get
$xv=0$
, which is not possible by Lemma 5.5. This means that
$w"=0$
in
$A^1/{\langle w\rangle }$
.
In order to get
$v"'=0$
, one proceeds in a similar way: first of all, one proves
$v"'\in V_2$
starting from
$v\in V_2$
and by using previous vanishings. Then, by substituting in Equation (6.6) and by using Equation
$(\underline {v}')_{xy}$
, one concludes as above.
Claim: The monomials in
$M_5$
are
$0$
.
We have shown that
$zv^2=0$
and
$xvw=0$
for the general triangle T, so we can deform these equations in the direction of
$s\underline {v}'$
. By using Equation
$(\underline {v})_{xz}$
and
$w"=0$
, one can write these relations as

As observed above, I is not
$0$
; thus, we deduce
$v^2w=0$
.
For the vanishing
$vw^2=0$
, one works in a similar way by deforming
$zw^2=0$
and
$xvw=0$
in the direction of
$r\underline {v}"$
, and by using Equation
$(\underline {v}")_{xz}, v"'=0$
and
$M\neq 0$
.
Claim:
$v'\in \langle v,w\rangle $
and
$u'$
does not depend on y.
Proceeding as we have done above for proving
$v"',w"\in V_2$
, one can also obtain that
$v'\in V_2=\langle u,v,w\rangle $
. Consider the first-order deformation of the Equation
$(\underline {v}')_{xy}$
in the direction of
$t\underline {v}$
, namely

Upon multiplication by u, using
$B\neq 0$
and the various vanishing shown above, one gets
$xuv'=0$
. Since
$xu^2\neq 0$
and
$xu\cdot \langle x,y,z,v,w\rangle =0$
, one has that
$v'\in \langle v,w\rangle $
.
For the second claim, by deforming
$(\underline {v})_{xy}$
in the direction of
$t\underline {v}$
, one gets
$yu'=0$
so
$y^2u'=0$
, and this implies that
$u'$
does not depend on y, since
$y^3\neq 0$
and
$y^2\cdot \langle x,z,u,v,w\rangle =0$
.
Claim:
$uv^2\neq 0$
in
$A_f$
.
Assume, by contradiction, that
$uv^2=0$
. We claim that
$v'=0$
as tangent vector. Consider the first-order deformation of the Equation
$zv^2=0$
in the direction of
$t\underline {v}$
– that is,

Since
$v'\in \langle v,w\rangle $
(by the previous claim),
$zv^2=0$
and
$zvw\neq 0$
, one has that
$v'=0$
as tangent vector.
Since we are assuming that
$uv^2=0$
for the general triangle in
${\mathcal {F}}$
, we can deform this equation in the direction of
$t\underline {v}$
. This operation yields the relation
$0=v^2u'+2uvv'=v^2u'$
. Now recall that
$u'$
does not depend on y,
$v^2\cdot \langle z,v,w\rangle $
by previous vanishings and
$uv^2=0$
, by assumption. Since
$xv^2\neq 0$
, from
$v^2u'=0$
, one has that
$u'$
does not depend on x.
This yields a contradiction by deforming
$x^2u=0$
in the direction of
$t\underline {v}$
. Indeed, one has
$0=x^2u'+2xu^2=2xu^2$
but
$xu^2\neq 0$
.
Claim: Elements
$r_1, r_2$
and
$r_3$
are
$0$
.
The relation
$r_1=x(Iw^2-Mv^2)=0$
is easily obtained from Equations
$(\underline {v}')_{xz}$
and
$(\underline {v}")_{xz}$
upon multiplication by w and v, respectively.
We prove now that
$r_2=u(Mv^2-Iw^2)=0$
. Consider the first-order deformation in the direction of
$s\underline {v}'$
of the Equations
$zuw=0$
and
$xuv=0$
, together with Equation (6.8)
${}_I$
multiplied by u, namely

Now, since
$w"=zw+Mxv=0$
, and
$B"=yu=0\neq B$
, we have

which give the desired relation, if substituted into Equation (6.9)
${}_{II}$
.
The other relation, namely
$r_3=u(Dw^2+Mvw)=0$
, is obtained in a similar way from the first-order deformation in the direction of
$r\underline {v}"$
of the Equations
$(\underline {v})_{xy}$
and
$(\underline {v})_{xz}$
upon multiplication by suitable elements (more precisely, the first one by w and
$Mv$
and the second one by w, respectively).
To sum up, we have proved that if we define

then

Now we would like to partially reconstruct the cubic fourfold f from the information about its apolar ring
$A_f$
obtained so far. For simplicity, we are using the same symbols for the indeterminates in
$S={\mathbb {K}}[x_0,\dots ,x_5]$
and in
${\mathcal {L}}={\mathbb {K}}[y_0,\dots , y_5]={\mathbb {K}}[x,y,z,u,v,w]$
. Consider the following cubics in
$S^3$
:


It is easy to see that

This can be checked directly by hand by writing
$f=\sum \alpha _m\cdot m$
, where m runs over the set of monomials of degree
$3$
in S. Each element in
${\mathcal {R}}$
is a linear differential equation satisfied by f and thus gives a linear closed condition on the vector space
$S^3$
. For example, since
$r_2=u(Mv^2-Iw^2)\in {\mathcal {R}}$
, we have the corresponding condition
$2M\alpha _{uv^2}-2I\alpha _{uw^2}=0$
on the coefficients of f.
Hence, any cubic polynomial that we are analysing in this case can be written as
$f=\sum _{i=0}^{6}p_i s_i$
for suitable
$p_i\in {\mathbb {K}}$
. Having proved that
${\mathcal {M}}^{nv}\cap \operatorname {\mathrm {Ann}}_{{\mathcal {L}}}(f)=\emptyset $
gives nontrivial open conditions, indeed, it is translated into

so all the cubic fourfolds satisfying Conditions (6.10) live in a dense open subset of
$|W|$
. Notice that the base locus of
$|W|$
is the line
$L=V(x,y,z,u)$
. Moreover, as
$p_4,B,I,M\neq 0$
and since

we have that the general cubic in
$|W|$
is indeed smooth on the points of L and thus smooth everywhere by Bertini.
Claim: One has
$D\neq 0$
.
Consider the first-order deformation of
$v^2w$
and of
$z^2v$
in the direction of
$s\underline {v}'$
, namely

where we also used
$w"=0$
. Assume, by contradiction, that
$D=0$
. Then, from the expression of f and as
$p_5\neq 0$
, we have that
$uvw=0$
in
$A_f$
. Hence,
$vw\cdot \langle x,y,u,v,w\rangle =0$
. Then, by previous vanishings and since
$zvw\neq 0$
, from Equation (6.12)
${}_I$
, we get that
$v"$
does not depend on z. This implies that
$z^2v"=0$
, so Equation (6.12)
${}_{II}$
yields
$2Izvw=0$
, and thus,
$I=0$
, which is impossible.
Claim: If f satisfies the Conditions in (6.10), then
$\operatorname {\mathrm {Sing}}({\mathcal {H}}_f)$
is of dimension
$2$
near
$[x]$
.
By changing coordinates, we can simplify a little the expression of f. Indeed, as B, D, I, M,
$p_4$
and
$p_5$
are not
$0$
, by an easy change of coordinates, and by redefining the
$p_i$
s, one can write

with
$\lambda ,p_0,p_1,p_2,p_3\neq 0$
.
By construction,
$[x]\in {\mathbb {P}}(A^1)\leftrightsquigarrow (1:0:0:0:0:0)\in {\mathbb {P}}^n$
is a vertex of a triangle for
${\mathcal {H}}_f$
so
$[x]\in \operatorname {\mathrm {Sing}}({\mathcal {H}}_f)$
. We are assuming also that there exists a family of dimension
$5-2=3$
whose general element is a triangle dominating via the first projection a component of dimension
$3$
of
$\operatorname {\mathrm {Sing}}({\mathcal {H}}_f)$
. Then, in order to conclude the proof of the lemma, it is enough to show that the local dimension of
$\operatorname {\mathrm {Sing}}({\mathcal {H}}_f)$
near
$[x]$
is actually
$2$
.
The Hessian matrix of f is

Since f is smooth, we have that
$\operatorname {\mathrm {Sing}}({\mathcal {H}}_f)={\mathcal {D}}_4(f)$
. In particular,
$\operatorname {\mathrm {Sing}}({\mathcal {H}}_f)$
is cut out by
$21$
quintic equations corresponding to the minors of order
$5$
of the Hessian matrix (there are
$36$
minors but
$15$
appear twice since
$H_f$
is symmetric). Let
$m_{ij}$
be the minor obtained by removing the i-th row and the j-th column. We are interested in the local expression of
$\operatorname {\mathrm {Sing}}({\mathcal {H}}_f)$
near
$[x]$
. Notice that the linear form
$(y_iy_j)(f)$
(defined as the second partial derivative of f as in Equation (2.1)), depends on x if and only if
$(i,j)\in \{(0,0),(3,3),(4,4),(5,5)\}$
, so no term of
$h_f$
can have as exponent of x an integer greater than
$4$
: this is a confirmation of the fact that
$[x]\in \operatorname {\mathrm {Sing}}({\mathcal {H}}_f)$
. By differentiating
$m_{ij}$
, it is easy to see that
$[x]$
is singular for
$V(m_{ij})$
if
$(i,j)\not \in \{(1,1),(1,2),(2,1),(2,2)\}$
. Consider the variety
$Z=V(m_{11},m_{12},m_{22})$
and notice that
$\operatorname {\mathrm {Sing}}({\mathcal {H}}_f)\subseteq Z$
by construction. Being defined by
$3$
equations, one has that
$\dim (Z)\geq 2$
. We claim that Z has dimension
$2$
near
$[x]$
; to do that, we will show that
$(1:0:0:0:0:0)$
is isolated in
$Z\cap V(u,v)=V(m_{11},m_{12},m_{22},u,v)$
.
We compute now the local expression of
$m_{11},m_{12}$
and
$m_{22}$
modulo
$(u,v)$
in the local ring
$A_m$
, where
$A={\mathbb {K}}[y,z,u,v,w]$
and m is the maximal ideal of the origin in
${\mathbb {A}}^5$
. By the explicit expression of
$H_f$
in Equation (6.14), one can easily see that

so one between w and
$w^2-p_3\lambda ^2 z(1-z)$
is zero.
Assume first that
$w=0$
. By a direct computation, one can see that


since
$p_0,p_1,p_3\neq 0$
by assumption and since both
$z-1$
and
$y^2+\lambda ^2 z^2-1$
are invertible in
$A_m$
. This shows that
$[x]$
is isolated in
$Z\cap V(u,v,w)$
.
Assume now that
$w^2=p_3\lambda ^2 z(1-z)$
. Again, by a direct computation, one can show that w appears only with even powers in
$m_{ij}(1,y,z,0,0,w)$
for
$i,j\in \{1,2\}$
, so one can substitute
$p_3\lambda ^2 z(1-z)$
to
$w^2$
in order to obtain the two expressions


In
$A_m$
, one has

with
$g(0,0)\neq 0$
. Since
$w^2=p_3\lambda ^2 z(1-z)$
, if we assume
$z=0$
, we also have that
$w=0$
, so we can conclude by the previous case. We can then suppose that
$3p_2 y + (3p_2 - p_3\lambda ^4)(z-1)=0$
in the local ring. This can happen if and only if
$3p_2=\lambda ^4p_3$
and
$y=0$
. However, if
$y=0$
, from the expression of
$r_{11}$
, one has that
$z(z-1)=0$
, and thus again,
$w^2=0$
. This shows that
$[x]$
is isolated in
$Z\cap V(u,v)$
too and thus that the local dimension of
${\mathcal {D}}_4(f)=\operatorname {\mathrm {Sing}}({\mathcal {H}}_f)$
near
$[x]$
is
$2$
.
With this lemma, we also conclude the proof of Theorem 5.1 in the case of cubic fourfolds.
Acknowledgements
The authors want to express their gratitude to Carlos D’Andrea and Giorgio Ottaviani for stimulating discussions and for pointing out some interesting papers related to some of the topics treated here. The authors also thank the anonymous referee for the useful observations.
Competing interest
The authors have no competing interests to declare.
Funding statement
The authors are partially supported by INdAM-GNSAGA and by PRIN 2022 ‘Moduli spaces and special varieties’. The first and second authors are partially supported by the INdAM – GNSAGA Project, ‘Classification Problems in Algebraic Geometry: Lefschetz Properties and Moduli Spaces’ (CUP
$\_$
E55F22000270001). The first author is holder of a research grant from Istituto Nazionale di Alta Matematica (INdAM).