72
Views
0
CrossRef citations to date
0
Altmetric
Research Article

The Variety of Polar Simplices II

, &

Abstract

We discuss the space VPS(Q, H) of ideals with Hilbert function H=(1,n,n,) that are apolar to a full rank quadric Q. We prove that its components of saturated ideals are closely related to the locus of Gorenstein algebras and to the Slip component in border apolarity. We also point out an important error in [Citation20] and provide the necessary corrections.

2020 MATHEMATICS SUBJECT CLASSIFICATION:

1 Introduction

For a fixed nondegenerate quadric q in n variables, a polar simplex is a tuple [l1],,[ln]Pn1 of points, where li are such that q=l12++ln2. The locus of all polar simplices is a principal homogeneous space for the group SO(q)SOn. Informally speaking, the variety of polar simplices is a compactification of this locus. As we discuss below, there is actually more than one possible compactification and choosing the correct one is subtle. The variety is important for two main reasons. First, it is a “simplest” example of a variety of sums of powers [Citation19], which are useful as examples of special projective varieties and for applications in tensors [Citation3–5, Citation10, Citation14, Citation21]. Second, as explained below, it serves as a mean to investigate the Gorenstein locus of the Hilbert scheme of points: it has smaller dimension, which is important for computations such as [Citation16] and at the same time contains each abstract Gorenstein subscheme of degree n.

Let S=k[x1,,xn] be a polynomial ring which we view as a homogeneous coordinate ring of the Pn1=P(S1*). Fix a number d and a nonzero fSd*, then F={f=0}Pˇn1. A finite subscheme ΓPn1 is apolar to F if [f] lies in the span of the d-uple reembedding of Γ in P(Sd*). Apolarity may also be formulated algebraically, see Section 2, by requiring that IΓf, the ideal of forms in S that annihilates f by differentiation. This condition may be formulated for ideals in general; an ideal IS is apolar to F if If.

Consider the special case d = 2 and F=Q={q=0} being a full rank quadric. Let VPS(Q,n)Hilbn(Pn1) denote the closure of the locus of degree n zero-dimensional schemes ΓPn1 satisfying IΓq. The result [Citation20, Corollary 2.2] and [Citation23, Proposition 6.5] both state that every ΓVPS(Q,n) is apolar to Q. It is not so, as the following example shows.

Example 1.1.

Let n = 4, let q=y1y4+y2y3, where yi is the dual to xi, and consider the ideal I=(x3x1x22,x2x3x1x4,x2x4,x3x4,x42,x32).

This ideal is saturated and has Hilbert function HS/I(i)=4 for every i1, hence I=IΓ for a finite, degree 4 subscheme ΓP3; in particular [Γ]VPS(Q,4). Consider now a k*-action on P3 corresponding to the grading by (0,1,0,1). The quadric Q is a semi-invariant for this action, so for every λk*, the subscheme Γλ:=λ·Γ corresponds to a point [Γλ] in VPS(Q,4) with ideal given by I(Γλ)=(x3x1λ2x22,λ(x2x3x1x4),λ2x2x4,λx3x4,λ2x42,x32).

The limit of those ideals at λ0, taken degree by degree, is the ideal Ilim:=(x24,x42,x2x4,x3x4,x12x4,x2x3x1x4,x32,x1x3),whose saturation is Ilimsat=(x3,x4,x24). Let Γ0=V(Ilimsat), then Γ0 is a limit of Γλ for λ0, so [Γ0]VPS(Q,n) while IΓ0=Ilimsat is not contained in q, so Γ0 is not apolar to Q.

The example implies in particular that several of the main results in [Citation20] are wrong as stated. The purpose of this article is to both discuss what can salvaged and how and, which is of independent interest, show unexpected connections between VPS(Q,n) and the moduli of Gorenstein algebras.

Weronika Buczyńska and Jarosław Buczyński introduced the multigraded Hilbert scheme into the subject of apolarity and observed that it is better than the usual Hilbert scheme also when one considers VPS [Citation3, Section 7.6].

The multigraded Hilbert scheme HilbH, defined in [Citation15], parametrizes homogeneous ideals with a given Hilbert function H. For a fixed form f, the condition If for an ideal I is a closed condition in the multigraded Hilbert scheme. Example 1.1 above shows that the apolarity condition IΓf is not closed in the usual Hilbert scheme. Hence, it is more natural to work in the multigraded Hilbert scheme.

We consider the Hilbert function H:=(1,n,n,) and, for the full rank quadric Q, define VPS(Q,H)={IHilbH|Iq}HilbH.

This is a closed subscheme of HilbH. Let VPSgood(Q,H)VPS(Q,H) be the locus of saturated ideals. This locus is open by [Citation17, Theorem 2.6]. Let VPSsbl(Q,H) be its closure, so VPSsbl(Q,H) is a union of irreducible components of VPS(Q,H). Let VPSuns(Q,H)=VPS(Q,H)VPSgood(Q,H) with its reduced scheme structure. In analogy with the locus of smoothable schemes in the Hilbert scheme, we say that VPSsbl(Q,H) is the locus of ideals that are Satura BLe, this is consistent with the notation of [Citation17] thanks to Corollary 3.9. The intersection VPSsbl(Q,H)VPSuns(Q,H) is the locus of unsaturated limit ideals: apolar unsaturated ideals that are limits of saturated apolar ideals.

Associating to each ideal IVPS(Q,H) the space I2 of quadrics in the ideal defines a forgetful map πG:VPS(Q,H)G((n2),q2),

into the Grassmannian of (n2)-dimensional subspaces in q2. We denote VPS(Q,H)G:=πG(VPS(Q,H))VPSgood(Q,H)G:=πG(VPSgood(Q,H))VPSsbl(Q,H)G:=πG(VPSsbl(Q,H))VPSuns(Q,H)G:=πG(VPSuns(Q,H)).

In this paper we will be concerned with these loci. The map πG restricted to VPSgood(Q,H) is an isomorphism (see Proposition 4.5), while the restriction to the other loci, in general, is not.

The scheme VPS(Q,H) is different from VPSsbl(Q,H) in general, and similarly for VPS(Q,H)G and VPSsbl(Q,H)G. For n3, the schemes VPS(Q,H) and VPSgood(Q,H) coincide and are isomorphic to VPS(Q,H)G, see Proposition 3.2. As soon as n4, both the complement of VPSgood(Q,H) in VPSsbl(Q,H), i.e. the intersection VPSsbl(Q,H)VPSuns(Q,H), and the complement of VPSsbl(Q,H) in VPS(Q,H) are nonempty. For n = 4 the ideals of the intersection VPSsbl(Q,H)VPSuns(Q,H) have an unexpected beautiful connection to the geometry of the inverse, quadric Q1={q1=0}. The latter defines the collineation q1:S1*S1 inverse to the symmetric collineation defined by q:S1S1*.

Example 1.2.

For n = 4, the loci are summarized in . In this case, VPSuns(Q,H) is given by ideals IΓq, where ΓL is finite of length 4 and L is a line in P3. The intersection VPSsbl(Q,H)VPSuns(Q,H) is 5-dimensional and consists of ideals IΓq where ΓL and L is a line in the quadric Q1 inverse to Q, see Section 3.

Fig. 1 Moduli spaces for n = 4.

Fig. 1 Moduli spaces for n = 4.

For n6, the locus VPSgood(Q,H)VPSsbl(Q,H) is singular (because there are obstructed Gorenstein schemes of degree n, see Theorem 1.4). For n3, the two loci coincide, are isomorphic to VPS(Q,H)G, and are smooth by Proposition 3.2 and [Citation18, Proposition 10]. We prove:

Theorem 1.3.

(Corollaries 4.3 and 4.10) For n = 4, 5 both VPSsbl(Q,H) and VPSsbl(Q,H)G are smooth.

There is a map from the multigraded Hilbert scheme to the Hilbert scheme. For n13, it sends the scheme VPSsbl(Q,H) to the variety of sums of powers VSP(Q,n). For n4, we do not know whether VSP(Q,n) is smooth; as noted above, the Hilbert scheme compactification seems to be not as interesting as the other two: Being apolar is not a closed property in the Hilbert scheme so VSP(Q,n) has no functorial interpretation, see Remark 4.4 for a bit more discussion.

1.1 Connections to the Gorenstein locus

A saturated apolar ideal of a finite scheme is locally Gorenstein [Citation2, Proof of Proposition 2.2]. Let Hilbngood(Pn1)Hilbn(Pn1) be the locus of ΓPn1 which satisfy h1(IΓ(1))=0. The locus VPSgood(Q,H)VPSsbl(Q,H) of saturated ideals is closely connected to the locus Hilbngood,Gor(Pn1) of locally Gorenstein Γ in Hilbngood(Pn1).

Theorem 1.4.

(Proposition 2.5) The natural map GLn×VPSgood(Q,H)Hilbngood(Pn1);(g,I)g·Iis smooth and its image is Hilbngood,Gor(Pn1). In particular, the singularity types encountered on VPSgood(Q,H) and on Hilbngood,Gor(Pn1) coincide.

The bridge given by the theorem allows us to apply results from both sides. First, we improve and make sharp the question about when VPSgood(Q,H) is reducible, raised in [Citation20, Theorem 1.3].

Corollary 1.5.

(Theorem 2.7) The locus VPSgood(Q,H) is irreducible exactly when n13.

Second, we use the result [Citation20, Corollary 5.16] to obtain properties of the Gorenstein locus itself.

Corollary 1.6.

The Gorenstein locus of every Hilbn(Pk) is reduced for every n6.

Proof.

Reducedness of a point [ZPk]Hilbn(Pk) depends only on the underlying Z, so Hilbn(Pk) is reduced if and only if Hilbngood(Pn1) is reduced. By Theorem 1.4 this holds if and only if VPSgood(Q,H) is reduced. The reducedness of the latter is proven in [Citation20, Corollaries 5.12 and 5.16]. □

Theorem 1.4 extends also to the whole VPSsbl(Q,H). The locus of saturated ideals in HilbH is open and its closure is called Sat¯H. This closure is a union of components of HilbH, for n13 it is a single component, called SlipH. This locus has recently gained much attention thanks to applications in border apolarity [Citation3, Citation17]. The theorem extends as follows.

Theorem 1.7.

(Proposition 3.8, Corollaries 3.9 and 4.3) The natural map GLn×VPSsbl(Q,H)Sat¯H;(g,I)g·Iis smooth. For n5, VPSsbl(Q,H) is smooth and the component SlipH is smooth along the image of this map.

The component SlipH is not smooth in all points, not even for n = 4: the saturated ideal of a second-order neighborhood of a point of P3 yields a singular point of SlipH.

1.2 Salvaging [Citation20]

In this section we report on the state of the results in the paper [Citation20].

Our notation differs from that of [Citation20], we follow rather the notation of [Citation19]. The paper [Citation20] uses VPS(Q,n) for the Variety of Polar Simplices as a subscheme of the Hilbert scheme; what we call VSP(Q,n) the Variety of Sums of Powers. The paper [Citation20] uses VAPS(Q,n) for the image in the Hilbert scheme of the scheme VPSsbl(Q,H).

The crucial error is found in the proof of [Citation20, Corollary 2.2] that wrongly claims that VPSgood(Q,n)=VPS(Q,n),while we only have an open immersion VPSgood(Q,n)VPS(Q,n).

We therefore discuss the statements that depend on this error.

1.2.1 Introduction

The second part of [Citation20, Theorem 1.1] asserts that for n6, the variety VSP(Q,n) is singular, rational and (n2)-dimensional. This is true (after possibly reducing) with the same proof. The first part of Theorem 1.1 claims that for 2n5, the variety VSP(Q,n) is additionally smooth of Picard rank 1 and is Fano of index 2. For n = 2, 3, we have VPS(Q,H)=VPSsbl(Q,H) and πG is an isomorphism onto the image VPS(Q,H)G. In particular both are isomorphic to VSP(Q,n) and the argument of [Citation20] is correct. For n = 4, 5, we could ask the same question for VPS(Q,H) and VPSsbl(Q,H) and their images under πG. In both cases, by Theorem 1.3, VPSsbl(Q,H) is smooth and admits a birational morphism onto the smooth VPSsbl(Q,H)G, that contracts the boundary divisor VPSsbl(Q,H)VPSgood(Q,H) (cf. Remark 3.11), hence the Picard rank of VPSsbl(Q,H) is at least two. For VSP(Q,n) we do not know whether it is smooth, but if it were, see Remark 4.4, its Picard rank would also be at least two. When replacing VPSsbl(Q,H) by the Grassmannian subscheme VPSsbl(Q,H)G, however, we salvage also the first part of [Citation20, Theorem 1.1].

Salvaged Theorem 1.1. (Corollary 4.10) For 2n5, the Grassmannian subscheme VPSsbl(Q,H)G is a smooth rational (n2)-dimensional Fano variety of index 2 and Picard number 1.

The theorem [Citation20, Theorem 1.2] concerns VPSsbl(Q,H)G, the Grassmannian subscheme. We were able to obtain the result of this theorem for n = 4, with a correct degree, using a more nuanced machinery of excess intersections. The case n = 5 remains open. We refer to [Citation20] for the notation regarding the Gauss map; we will not use it in the present article.

Salvaged Theorem 1.2. (Proposition 4.15) The variety VPSsbl(Q,H)G contains the image TQ1 of the Gauss map. When n = 4 the restriction of the Plücker line bundle generates the Picard group of VPSsbl(Q,H)G and the degree is 362.

The theorem [Citation20, Theorem 1.3] concerns the linear span of VPSsbl(Q,H)G, and is wrong. The image VPSuns(Q,H)GVPSsbl(Q,H)G of unsaturated limit ideals does not lie in the span of TQ1 (loc.cit.). Whether VPSsbl(Q,H)G is a linear section of the Grassmannian therefore remains an open problem. It is true for n = 3, and we give a computational proof for n = 4.

Finally, [Citation20, Proposition 1.4] remains correct with the same proof.

1.2.2 Sections 2–5

The results of these sections are local and the proofs are not effected by mistakes concerning the compactifications up to restricting to the locus of linearly normal schemes. That is, up to replacing VAPS(Q,n) and VPS(Q,n) by VPSgood(Q,H) and its smoothable component, the statements are correct according to our knowledge, except the final part of Corollary 2.2.

1.2.3 Section 6

The theorem [Citation20, Theorem 6.3] concerns the degree of the Grassmannian subscheme VPSsbl(Q,H)G. The presence of unsaturated limit ideals in VPSsbl(Q,H) means that the claim of the theorem is wrong, instead the degree formula is a contribution in a computation of the degree of VPSsbl(Q,H)G using excess intersection. We show this in the case n = 4, see Proposition 4.15 and Remark 4.16.

1.3 Legend

In Section 2 we show that the locus VPSgood(Q,H) of saturated apolar ideals is irreducible if and only if the quadric Q has rank at most 13. In the following Section 3, we discuss the unsaturated ideals in the boundary VPSsbl(Q,H)VPSgood(Q,H) and give a precise characterization when n = 4, 5. In Section 4 we give our results on global properties both of the multigraded Hilbert scheme compactification VPSsbl(Q,H) and of the Grassmannian compactification VPSsbl(Q,H)G of VPSgood(Q,H). The Appendices A and B contain computer code [Citation22] in Macaulay2 [Citation11], used in our computational arguments and a result on VPSsbl(Q,H)G for n = 4, for which we only give a computational proof.

1.4 Notation

We let k be a field of characteristic zero. Our computations, see Appendices A and B, are performed over Q. Let S=k[x1,,xn] and T=k[y1,,yn] be polynomial rings. We view x1,,xn and y1,,yn as dual bases of dual spaces S1 and T1. Differentiation defines bilinear pairings Se×TdTde,de which induces isomorphisms SdTd* for any d0. In this sense we say that T is dual to S. We let S be the coordinate ring of Pn1, and T be the coordinate ring of the dual space Pˇn1. Thus we may set Pn1=P(T1) the projective space of 1-dimensional subspaces in T1.

For a homogeneous polynomial fTdSd* let fS denote its apolar ideal. For a subscheme ΓPn1, let IΓS denote its homogeneous ideal; note that this ideal is saturated. We say that an ideal IS is apolar to F={f=0} if If. We say that a zero-dimensional ΓPn1 is apolar to F if its homogeneous ideal is apolar, that is, if IΓf. We usually denote quadratic forms in T by q; it defines a quadric Q={q=0}Pˇn1 and a collineation: q:S1T1. When q is nondegenerate, then denote by q1S2 the quadratic form that defines the inverse collineation, q1:T1S1, and by Q1={q1=0}Pn1 the corresponding quadric.

We also identify qT2 with the associated map q:S2k. If qT2 is a full rank quadric, then qS is generated by degree two elements and q2=ker(q:S2k), so an ideal I is apolar to Q if and only if I1=0 and I2ker(q:S2k). A subscheme Γ as above is apolar to Q if and only if its homogeneous ideal IΓ is apolar.

2 Good points of VPS

The following result essentially appears in [Citation20], but due to its importance we provide a full proof.

Proposition 2.1.

[Citation20, Lemma 2.6] Let ΓPn1 have length n. The following are equivalent

  1. ΓPn1 is (locally) Gorenstein and linearly normal,

  2. there exists a full rank quadric Q such that Γ is apolar to Q.

Proof.

Let A=H0(OΓ). Suppose (1) holds, so the dualizing A-module ωA=HomA(A,k) is isomorphic to A. Let φωA be its generator, then the pairing A×Ak given by (a1,a2)φ(a1a2) is symmetric and perfect. Since ΓPn1 is linearly normal, the restriction of linear forms S1H0(OΓ(1))A is an isomorphism and we obtain a symmetric perfect pairing S1×S1k. Let qT2 be the corresponding quadric. It has full rank. Moreover, it arises from the following commutative diagram so q(IΓ)2, hence Γ is apolar to Q.

Assume (2) holds. Consider the diagram as above, where φ:Ak is the functional induced by q:S2k. Since q has full rank, the natural map S1A has to be injective, so an isomorphism by comparing dimensions. This shows that A×Ak from the diagram is a perfect pairing, hence ωA is isomorphic to A and, hence Γ is locally Gorenstein. □

Recall that HilbH is the multigraded Hilbert scheme parameterizing ideals with Hilbert function H=(1,n,n,n,). Let VPSgood(Q,H)HilbH denote the locus of ideals which are saturated and apolar to Q. Recall that we have a natural map HilbHHilbn(Pn1). Let Hilbngood(Pn1)Hilbn(Pn1) denote the open locus of Γ with H1(IΓ(1))=0; equivalently those are the subschemes which span Pn1, or, in yet other words, the subschemes with HS/IΓ=H.

Proposition 2.2.

The locus VPSgood(Q,H) is locally closed in HilbH. Under the map HilbHHilbn(Pn1) it maps isomorphically onto the locus of apolar subschemes in Hilbngood(Pn1).

Proof.

Let HilbH,sat inside HilbH denote the subset of saturated ideals, this is an open subset [Citation17, Theorem 2.6] and so comes with a natural scheme structure. The condition Iq is closed in HilbH and so VPSgood(Q,H) is closed in HilbH,sat. By [Citation17, Theorem 3.9] the subset HilbH,sat maps isomorphically to Hilbngood(Pn1) under the natural map HilbHHilbn(Pn1). Restricting this isomorphism to VPSgood(Q,H), we get the final claim. □

We now describe the topology of VPSgood(Q,H).

Proposition 2.3.

The space VPSgood(Q,H) is connected.

Proof.

This follows from the argument in [Citation20, Section 6] which is correct if one restricts from VPS(Q,H) to VPSgood(Q,H). Essentially the same argument is given independently in [Citation13, Proposition 4.1]: any k-point Γ of VPSgood(Q,H) degenerates to a fixed one and this degeneration is possible inside VPSgood(Q,H) because it preserves orientation of Γ, as defined in [Citation13, 2.1]. □

Before we give more refined information about the topology, we need a technical idea of orthogonalization, which links VPSgood(Q,H) and Hilbngood(Pn1).

2.1 Orthogonalization

The following key technical theorem shows that any infinitesimal deformation of an apolar ideal can be “orthogonalized” to a deformation in q. In essence, it says the following: suppose that Iq is an apolar ideal and I is its deformation (in the multigraded sense) over a base with a nilpotent ε so I|ε=0=I. It may happen that I does not lie in q, for example when I comes from a infinitesimal change of coordinates that moves q. The theorem says that this is essentially the only possibility: there exists a coordinate change g=g(ε) so that g°I is contained in the ideal q. It follows from the fact that the GLn-action on full rank quadrics is transitive in the infinitesimal neighborhood of q.

Let us introduce the setup. For a k-algebra A let SA:=SkAA[x1,,xn]. For a surjection of k-algebras AA, we get a surjection SASA which in down-to-earth terms reduces the coefficients of the polynomials modulo J=ker(AA). Let T be a polynomial ring dual to S and let qT2(TA)2 be a full rank quadric, as in the setup at the beginning of Section 2. Let qSA be its apolar ideal. We denote by q also its image in SA.

Theorem 2.4

(orthogonalization). Let AA be a surjection of finite local k-algebras with kernel J satisfying J2=0. Let ISA, ISA be graded vector subspaces. Assume that

  1. I2q,

  2. (SA/I)2 is a free A-module,

  3. the image of I2 in (SA)2 is I2.

Then there exists an invertible matrix gGLn(A) such that g=Idnmod J and (g°I)2q.

Proof.

The duality of S and T induces a duality of free A-modules (SA)2 and (TA)2 and, as a consequence, the free A-modules (SA)2 and (TA)2. The space perpendicular to q(TA)2 is q(SA)2. Since SA/q is a free A-module of rank one, the surjections (SA)2(SA/I)2SA/qdualize to a composition A(SA/I)2(TA)2 which sends 1 to q.

The A-modules (SA/I)2, (SA/q)2 are free, hence HomA((SAI)2,(SAq)2)AHomA((SAI)2,(SAq)2).

Since I2q, the right-hand side contains the natural surjection (SA/I)2(SA/q)2. Let us lift it to a surjection φ:(SA/I)2(SA/q)2. The surjections

dualize to (1) A(SAI)2(TA)2.(1)

Let q˜(TA)2 denote the image of 1A. Since the sequences for I reduce to those for I modulo J, the quadric q˜ satisfies q˜qmod J. Let us identify quadrics with symmetric matrices. For an element gGLn(A), the action of g on (1) yields (2) A(SAg°I)2(TA)2(2) which maps 1A to the quadric g° q˜. We claim that there exists a g such that g=Idnmod J and g° q˜=q. We view the vector space T2 as an algebraic variety. Since q has full rank, the orbit map GLnT2:(hh°q) is smooth. The quadric \widetildeq can be interpreted as a map q˜:Spec (A)T2. Let Spec (A)GLn be the constant map to the identity matrix. Since q˜qmod J, we have a commutative diagram

From the infinitesimal lifting criterion [Citation24, Tag 02H6], there is a lift Spec (A)GLn:1h, with h°q= q˜.

Taking g=h1 yields the desired element such that g° q˜=q. Dualizing (2) we get that (g°I)2q. □

For a moment, we change focus from HilbH to Hilbngood(Pn1)Hilbn(Pn1). By Proposition 2.2 the reader can choose one of two perspectives: either think about Hilbngood(Pn1) or about the saturated locus of HilbH.

Proposition 2.5.

The action map GLn×VPSgood(Q,H)Hilbngood(Pn1) is smooth and its image is the Gorenstein locus in Hilbngood(Pn1).

Proof.

We view VPSgood(Q,H) as a locally closed subset of Hilbn(Pn1), using Proposition 2.2. Let UVPSgood(Q,H) be the restriction of the universal family. The points of VPSgood(Q,H) are linearly normal by Proposition 2.1. By Cohomology and Base Change, the OVPSgood(Q,H)-module (SCOVPSgood(Q,H)IU)2is locally free and its fiber over a point [Γ] is (S/IΓ)2. The claim now follows from Theorem 2.4 and the infinitesimal lifting criterion [Citation24, Tag 02H6]. □

Remark 2.6.

As explained in the introduction, the result of [Citation20, Corollary 5.16] yields interesting news about the point Spec (S/(y12+y22+y32+y42)), the so-called G-fat point, in Hilb6Gor(P5).

Theorem 2.7.

The scheme VPSgood(Q,H) is irreducible, when n13.

Proof.

First we show that Hilbngood(Pn1) is irreducible. For this, note that the deformation theory of a zero-dimensional subscheme is independent of its projective embedding, see for example [Citation1, p.4]. The Gorenstein locus is connected, see [Citation13, Proposition 4.1] or [Citation7], and the linear normality condition on the Gorenstein locus in Hilbngood(Pn1) is open in the Gorenstein locus Hilbn(Pn1). So when the latter is irreducible, so is the former. But the latter is irreducible when n13 by [Citation6, Theorem A].

Finally we prove that VPSgood(Q,H) is irreducible if and only if Hilbngood(Pn1) is irreducible. We observe that both spaces are connected (see Proposition 2.3) and related by smooth maps VPSgood(Q,H)GLn×VPSgood(Q,H)Hilbngood,Gor(Pn1),see Proposition 2.5, hence one of them contains a point which lies on two irreducible components if and only if the other does. □

3 Unsaturated limit ideals

Having discussed VPSgood(Q,H), in the following we analyze the boundary VPSsbl(Q,H)VPSgood(Q,H). We will see that the boundary is nonempty exactly when n4 and give a necessary condition for a point of VPS(Q,H) to lie in VPSsbl(Q,H). We will classify the points of the boundary for n = 4 and for n = 5. Two key inputs for the classifications are the results of [Citation17] and the orthogonalization, see Proposition 3.8. An important geometric consequence of these, which is new, is the connection with the inverse quadric Q1, see Corollary 3.5.

We begin with a general observation.

Proposition 3.1.

Each component of the boundary VPSsbl(Q,H)VPSgood(Q,H) is a divisor in VPSsbl(Q,H).

Proof.

Consider the natural projective map φ:HilbHHilbn(Pn1). By Proposition 2.2, the locus VPSgood(Q,H) maps isomorphically onto its image. Let X:=φ(VPSgood(Q,H))¯Hilbn(Pn1)be the closure of the image, so that X=φ(VPSsbl(Q,H)). Let DHilbn(Pn1) be the locus of Γ which are not linearly normal. Since h0(OΓ(1))=h0(OPn1(1)), the locus D is a divisor. In fact, on the universal scheme Γ˜Hilbn(Pn1)×Pn1, the map H0(OPn1(1))|Γ˜H0(OΓ˜(1)) between vector bundles of rank n drops rank on the divisor D. By Proposition 2.1, the intersection XD lies in the complement of φ(VPSgood(Q,H)) in X. Thus the boundary Xφ(VPSgood(Q,H)) is a union of divisors of X, so also VPSsbl(Q,H)VPSgood(Q,H) is a union of divisors. □

Now we pass to the more detailed description of the boundary.

Proposition 3.2.

Let n5 and let [I]VPSsbl(Q,H) be an unsaturated limit ideal. Then Γ=V(I) contains a length 4 subscheme contained in a line. In particular, for n3 there are no unsaturated limit ideals. Moreover, under the same assumptions, the Hilbert function of S/IΓ is (1,n2,n1,n,n,) or, for n = 5, (1,2,3,4,5,5,).

Proof.

Consider the Hilbert function H of S/(IΓ)2. Since I is unsaturated (in degree 2), h1(IΓ(2))0, so H(2)n1. From Macaulay’s Growth Theorem, see [Citation6, Section 2E and Lemma 2.9], it follows that

  1. when n3, we get H(m)2 for all m2, hence a contradiction,

  2. when n = 4, we get H(m)=m+1 for all m2, so (IΓ)2 defines a line L with ΓL,

  3. when n = 5, we get H(m)=m+1 or H(m)=m+2 for all m2, so V((IΓ)2) is a line or a subscheme with Hilbert polynomial m + 2; in the latter case it is either a line L with embedded point or a line L with a disjoint point. In either case ΓL has length at least n1=4. □

The following is our main tool to discern VPSsbl(Q,H) from VPS(Q,H).

Proposition 3.3.

[Citation17, Example 4.2] Let I be an unsaturated homogeneous ideal such that the Hilbert function of S/I is (1,n,n,n,,) while the Hilbert function of S/Isat is (1,n2,n1,n,n,). Assume furthermore that V(I2) is a line L and some, possibly embedded, points (this is automatic when n5, see Proposition 3.2).

If I is a limit of saturated ideals, then Isat·IL, where IL is the ideal of L, is contained in I.

For a linear space LT1, we denote by LS1 its perpendicular (this notion has nothing to do with q!). Recall also that the linear maps q:S1T1 and q1:T1S1 are nondegenerate and inverses of each other.

Lemma 3.4.

Let L,NT1 be linear subspaces with dimL+dimN=dimT1. Then the following are equivalent

  1. L·Nq,

  2. q(L·N)=0,

  3. q(L)=N,

  4. q1(N)=L,

  5. q1(N·L)=0,

  6. N·L(q1).

Proof.

The equivalences follows immediately from the definitions when noticing that q(L)(N)=q(L·N). □

Corollary 3.5

(a geometric condition for being in the VPSsbl(Q,H)). In the notation of Proposition 3.3 assume that IqS. Let LT1 be the line in V(I2) and let NT1 be the linear span of V(Isat). Then q1(L·N)=0.

Proof.

We have Isat·ILIq. The span N is equal to (Isat)1 and L is equal to (IL)1. By the assumption of Proposition 3.3, we have dimN+dimL=dimT1, so the claim follows from Lemma 3.4. □

Corollary 3.6.

Let [I]VPSsbl(Q,H),n5 be an unsaturated limit ideal. Then Γ=V(I) contains a subscheme of length 4 in a line L, and the line L is contained in the inverse quadric Q1. In particular, when n = 4, I=IΓq.

Proof.

Let I=IΓq, where, of course, IΓ=Isat. Then the linear span of N=V(IΓ)T1 contains L. But then L2N·L, and hence, by Corollary 3.5, q1(L2)=0, which means LQ1. When n = 4, IΓ has Hilbert function (1,2,3,4,4,) while IΓq has Hilbert function (1,4,4,4,4,), so I=IΓq. □

A finite apolar scheme of minimal length is Gorenstein. This property does not necessarily hold for unsaturated limit ideals.

Example 3.7

(Non-Gorenstein unsaturated limit ideal for n = 5). Let q=y1y4+y2y3+y52. Consider the family over A1=Spec C[t] given by the ideal (x4x5,x3x5,x1x5,x42,x3x4,x12t+x2x4,x1x4x52,x32,x2x3x52,x1x3,x14)

It is contained in q and the fiber over t=λ0 is saturated. The fiber over t = 0 is non-saturated and abstractly isomorphic to Spec C[ε1,ε2]/(ε14,ε1ε2,ε22), which in particular is not Gorenstein.

Recall that Sat¯HHilbH is the union of irreducible components of HilbH and it is defined as the closure of the locus of saturated [I]HilbH. The following theorem allows us to relate smoothness a point [I]VPSsbl(Q,H) and the same point [I]Sat¯H. This is useful both theoretically and in computations, see cases n = 4, 5 below.

Proposition 3.8.

For every GLn-stable locally closed subscheme Z of HilbH, the map GLn×(VPS(Q,H)Z)Z;(g,I)g·Iis smooth and has (n+12)-dimensional fibers.

Proof.

Smoothness of the map follows by Theorem 2.4 and infinitesimal lifting criterion [Citation24, Tag 02H6]. For the fiber, by GLn-equivariance it is enough to compute the dimension of the fiber over any point IVPS(Q,H)Z. The fiber consists of pairs (g,g1I) such that g1Iq. The containment is equivalent to I(g·q). Now, we have H(2)=n, so there is an n-dimensional space of quadrics q such that I(q). The fiber is isomorphic to the set of gGLn which map q into this space. The action of GLn on quadrics is transitive, so the dimension of the fiber is n+(n2)=(n+12) as claimed. □

Corollary 3.9

(being a limit of saturated is independent of being in VPS(Q,H)). We have VPSsbl(Q,H)=VPS(Q,H)Sat¯H as schemes. For n13, we have VPSsbl(Q,H)=VPS(Q,H)SlipH as schemes.

Proof.

By definition, we have VPSsbl(Q,H)VPS(Q,H)Sat¯H as schemes. To prove the other inclusion, pick a point [I]VPS(Q,H)Sat¯H and let A=ÔSat¯H,[I] be the complete local ring of [I]Sat¯H. Since A is complete, by applying Theorem 2.4 to finite order truncations of the map ϕ:Spec (A)Sat¯H, we obtain a pair of maps g:Spec (A)GLn and ψ:Spec (A)VPS(Q,H)Sat¯H with ϕ=g·ψ. Pick any generic point ηSpec (A). Since Sat¯H is the closure of saturated locus, the ideal ϕ(η) is saturated. We have ϕ(η)=g(η)·ψ(η), so also the ideal ψ(η) is saturated. This ideal is apolar to Q, thus ψ(η) lies in VPSgood(Q,H). This holds for every generic point η, so ψ factors through the closure of VPSgood(Q,H), that is, through VPSsbl(Q,H). This proves the other containment from the statement. For n13, all Gorenstein algebras are smoothable, hence the locus VPSgood(Q,H) is contained scheme-theoretically in SlipH and so its closure VPSsbl(Q,H) is contained in SlipH. This yields (still for n13) a chain of inclusions of schemes VPSsbl(Q,H)VPS(Q,H)SlipHVPS(Q,H)Sat¯H=VPSsbl(Q,H),which then have to be equalities. □

Corollary 3.10.

Let Γ be a scheme of length 4 contained in a line in Q1 and let I=IΓq, then I is an unsaturated limit ideal in VPSsbl(Q,H).

Proof.

I is an unsaturated limit ideal by [Citation17, Proposition 4.2] and it is apolar to Q, hence lies in VPSsbl(Q,H) by Corollary 3.9. □

3.1 Unsaturated limit ideals in the case n = 5

With reference to [Citation17, Sections 4.4.2 and 4.4.3], we will describe VPSsbl(Q,H)VPSuns(Q,H), when n = 5.

For [I]VPS(Q,H)VPSgood(Q,H), by Proposition 3.2, the length five scheme ΓP4 contains a length four subscheme Γ contained in a line. It follows that VPSuns(Q,H) is a union of two families F1,F2, described as follows.

A general point of F1 is an ideal I such that V(Isat)=Γ{p}, where p is a point and Γ has length four and is contained in a line. For general such Γ and p, the Hilbert function of S/Isat is (1,3,4,5,5,), so I=Isatq; the ideal I is determined by Isat as in the case of n = 4. The family F1 is irreducible and has dimension 6 + 4 + 4.

For the family F2, a general point is an ideal I such that Γ=V(Isat) is a length five scheme on a line. The ideal I is obtained from Isat as the intersection I=Isatcq where c is a cubic form such that HS/I(2)=5. We do not know whether the family F2 is irreducible or what is its dimension.

Let us now consider the saturable elements of both families. It follows from Corollaries 3.5 and 3.6 and from [Citation17, Section 4.4.2] that F1VPSsbl(Q,H) consists exactly of ideals I such that the line L=Γ is contained in Q1 and p is contained in the plane polar to L, here p is such that Supp Γ={p}Supp Γ counting multiplicities. The set of unsaturated limit ideals in F1 has dimension 3 + 4 + 2. In the case F2, again by Corollary 3.6 the line Γ must be contained in Q1. Moreover by [Citation17, Section 4.4.3], the cube c must be of the form l2·μ, where l is a linear form on the line and μ is a linear form on the polar plane. The set of unsaturated limit ideals in F2 has dimension 3 + 5 + 1.

Remark 3.11.

When n = 4, 5 and I is an unsaturated limit ideal, then the map πG:VPSsbl(Q,H)VPSgood(Q,H)VPSsbl(Q,H)G;[I][I2]is a forgetful map with positive dimensional fibers. The boundary VPSsbl(Q,H)VPSgood(Q,H) is a divisor by Proposition 3.1, and has dimension 5 and 9 when n = 4 and n = 5 respectively. The ideal I2 vanishes on the line that contains a subscheme of length four (or five) of V(Isat). In cases n = 4 and n = 5 with IF1, the ideal I2 depend only on a line, respectively a line and a point, so the family of ideals I2 in the image of πG has dimension one and five, respectively. In case n = 5 and IF2, the image I2VPSsbl(Q,H)G, depend on the line and a special cubic form, so the family of ideals I2 in the image of πG has dimension at most 6.

4 The schemes VPSsbl(Q,H) and VPSsbl(Q,H)G for n5

As above, we let H=(1,n,n,). It follows from Proposition 3.2 that for n3, the schemes VPSsbl(Q,H) and VPSsbl(Q,H)G are both equal to VPSgood(Q,H) and hence smooth. We show they are both smooth also for n = 4, 5. For n6, already VPSgood(Q,H) is singular.

The idea of the proof is to prove that SlipH is smooth in the unsaturated limit ideals of VPSsbl(Q,H), then use Proposition 3.8 and Corollary 3.9 for Z=Sat¯H=SlipH to deduce the same for VPSsbl(Q,H) itself and finally prove the smoothness for their projections in the Grassmannian. We begin by showing that SlipH is smooth at the unsaturated limit ideals of VPS(Q,H).

Proposition 4.1.

Let n = 4, 5 and consider a homogeneous ideal [I]SlipH such that HS/Isat=(1,n2,n1,n,n,). Then [I] is a smooth point on SlipH.

Proof.

We will make an upper bound on the dimension of the tangent space to SlipH at [I]. Unfortunately, this space has no functorial interpretation, hence we begin with the tangent space T[I]HilbHHomS(I,S/I)0. The short exact sequence 0Isat/IS/IS/Isat yields 0HomS(I,Isat/I)0HomS(I,S/I)0HomS(I,S/Isat)0.

We have (Isat/I)3=0, while (Isat/I)2 is one-dimensional, so dimkHomS(I,Isat/I)0=dimkI2=(n2),and hence dimkT[I]HilbHdimkHomS(I,S/Isat)0+(n2).

The short exact 0IIsatIsat/I0 yields 0HomS(Isat,S/Isat)0HomS(I,S/Isat)0Ext1(Isat/I,S/Isat)0.

Using Proposition 3.2 and [Citation17, Example 4.1], we obtain that dimkExt1(Isat/I,S/Isat)0=1.

Since n23, by [Citation17, Corollary 3.16] we obtain that the point [Isat]HilbHS/Isat is smooth. The saturated locus of this Hilbert scheme maps isomorphically to its image in Hilbn(Pn1), see [Citation17, Proposition 3.9], hence has dimension dimkHomS(Isat,S/Isat)0=(n1)(n4)+2(n2)+4=(n1)n2n+4,where (n1)(n4) comes from a choice of n – 4 points, 2(n2)a choice of a line and 4 from a choice of four points on it. In total we obtain a bound (3) dimkT[I]HilbH(n1)n(2n4)+(n2)+1.(3)

We now pass to T[I]SlipH. The inclusion HomS(I,Isat/I)0HomS(I,S/I)0 corresponds to the tangent map for the embedding of φ1([Isat])HilbH, where φ:HilbHHilbn(Pn1) is the natural map.

By Proposition 3.3 we see that the intersection φ1([Isat])SlipH consists of I such that I1sat·(IL)1I2,where L is the line in V(I2sat). This is true even scheme-theoretically and shows that the intersection above is the Grassmannian G(1,I2sat/(I1sat·(IL)1)), which has dimension dimkI2sat1dimkI1sat·(IL)1=(n2)(2n5). Arguing as above, the estimate (3) yields dimkT[I]SlipH(n1)n(2n4)+(n2)+1(2n5)=(n1)n+12(n4)(n5).

The variety SlipH is birational to the smoothable component of Hilbn(Pn1), hence has dimension (n1)n. This proves that [I]SlipH is smooth for n = 4, 5. □

Proposition 4.2.

Let n = 5 and consider a homogeneous ideal [I]SlipH such that HS/Isat=(1,2,3,4,5,5,). Then [I] is a smooth point on SlipH.

Proof.

The strategy of the proof is the same as in Proposition 4.1. As in that proposition, we argue that the tangent space at [Isat]Hilb(1,2,3,4,,5) is 11-dimensional because the locus of saturated ideals with this Hilbert function is parameterized by 6-dimensional choice of line and 5-dimensional choice of five points. It follows that T[I]HilbH is an extension of the at most 12-dimensional space V:=HomS(I,S/Isat) by HomS(I,Isat/I). The latter space is the tangent space to the fiber π1([Isat]) of π:HilbHHilbn(Pn1). This fiber is smooth and rational of dimension 14, its parameterization is given in [Citation17, Section 4.4.3]. Also by [Citation17, Section 4.4.3] the intersection π1([I])SlipH is smooth and rational of dimension 8. It follows that inside the 26-dimensional space T[I]HilbH the intersection of T[I]SlipH with T[I]π1([Isat]) is 8-dimensional. The latter space is 14-dimensional, hence the dimension of T[I]SlipH is at most 2614+8=20, which coincides with the dimension of SlipH, so equality holds and |I] is a smooth point on SlipH. □

Corollary 4.3.

For n = 4, 5 the variety VPSsbl(Q,H) is smooth.

Proof.

Being smooth is preserved by smooth maps. It follows that the good part VPSgood(Q,H) is smooth by Proposition 2.5. The variety VPSsbl(Q,H) is smooth at unsaturated limit ideals by Proposition 3.8 with Corollary 3.9 and Propositions 4.1–4.2. □

Some natural simplified versions of Proposition 4.1 are false: for example HilbH can be singular at the points as in the statement and SlipH can be singular away from the locus in the Proposition; both pathologies occur for n = 4. It would be interesting to know whether Proposition 4.1 holds for higher n.

Remark 4.4.

We do not know whether VSP(Q,n)Hilbn(Pn1) is smooth even for n = 4, 5. The map φ:SlipHHilbn(Pn1) restricts to a map VPSsbl(Q,H)VSP(Q,n) which is an isomorphism on VPSgood(Q,H). For n = 4, this map is also bijective on points, because for an unsaturated limit ideal I we have I=Isatq. This shows that for n = 4, the variety VPSsbl(Q,H) is the normalization of VSP(Q,n).

Now we prove that also the Grassmannian compactification VPSsbl(Q,H)GG((n2),q2)G((n2),S2)is smooth for n = 4, 5. One cannot hope to have Proposition 3.8 for VPSsbl(Q,H)G and G((n2),S2) because the latter lacks structure. But there is an analogue of the Hilbert scheme in this setup. We define the syzygetic locus SyznG((n2),S2) as the closed subscheme given by the determinantal equations which on a point [V]G((n2),S2) boil down to HS/(V)(3)n, where (V)S is the ideal generated by VS2. The tangent space to [V]Syzn at a point where HS/(V)(3)=n is given by HomS((V)+S4,S(V)+S4)0.

The name compactification is justified by the following.

Proposition 4.5.

The map πG:Hilbngood(Pn1)Syzn, given by πG(I)=I+S4 is an open immersion. In particular, the map πG:VPSgood(Q,H)VPSsbl(Q,H)Gis an isomorphism onto VPSgood(Q,H)G.

Proof.

A point of Hilbngood(Pn1) corresponds to a saturated ideal I such that the quotient by I has Hilbert function H. It follows that the regularity of the quotient is one and the regularity of I is two. By [Citation8, Proposition 3.1] the complete local rings of [I]Hilbngood(Pn1) and πG([I])Syzn are isomorphic. It follows from [Citation24, Tag 039M] that πG is étale at [I]. From the regularity it also follows that I is generated by I2, so πG is injective on points. It follows from [Citation24, Tag 025G] that this map is an open immersion. The result on VPSgood(Q,H) follows by intersecting with the locus apolar to Q. □

Proposition 4.6.

Let n = 4, let [I]VPSsbl(Q,H) be an unsaturated limit ideal and let [I2]=πG([I])VPSsbl(Q,H)G. Then [I2] is a smooth point of the image of SlipH in G(6,S2).

Proof.

Every unsaturated limit ideal I defines a scheme Γ=V(I) of length four on a line on a quadric by Proposition 3.3 and Corollary 3.6. The ideal I itself is determined by Γ as I=IΓq. Using a torus action not changing the quadric, we may degenerate further so that Γ is supported only at a single point. There is then, up to coordinate change, only one unsaturated limit ideal. If q=y1y3+y2y4, then we may take it as I:=(x24,x3,x4)q.

Using the package VersalDeformations [Citation16], see Appendix A, we check that near the point [I2]Syzn the scheme Syzn is reduced and it has two irreducible components Z1, Z2 passing through [I2]; the dimensions are 12 and 10, respectively. Moreover, both these components are smooth at [I2]. The image of SlipH under πG:HilbHG(6,S2) is an integral 12-dimensional variety, hence it has to coincide with the larger component: πG(SlipH)=Z1. In particular, [I2] is a smooth point on it. □

Proposition 4.7.

Let n = 5 and let [I]VPSsbl(Q,H)VPSuns(Q,H) be an unsaturated limit ideal. Then its image [I2]=πG([I])VPSsbl(Q,H)G is a smooth point of the image of SlipH in G(10,S2).

Proof.

The argument is analogous to Proposition 4.6, although reducing to finitely many possible limits requires much more work (we need finitely many of them to compute the tangent space dimensions). By Corollary 3.6 and Proposition 3.2, an unsaturated limit ideal I corresponds to a scheme Γ=V(Isat) of length five with a subscheme Γ0Γ of length four on a line LQ1 and a possibly embedded fifth point. Up to coordinate change, we may assume that q=y1y3+y2y4+y52 and the line L is x1=x4=x5=0. We have q1=4x1x3+4x2x4+x52. Consider two cases for Γ=V(Isat):

  1. the scheme Γ is contained in L,

  2. the scheme Γ is not contained in L.

We begin with the case (2). By Corollary 3.5 the fifth point lies on the plane Π=(x1=x4=0). Consider a coordinate change x5x5+λ4x4+λ1x1,x3x314λ12x114λ1λ4x412λ1x5,x2x214λ42x414λ1λ4x112λ4x5.

It preserves Q1 and L and maps a point [0:0:0:0:1]Π to [λ1:0:0:λ4:1]. Using such an action, we may assume that the fifth point is not [0:0:0:0:1]. Consider the torus action by t°[y1:y2:y3:y4:y5]=[ty1:ty2:ty3:ty4:t2y5].

When t0, the fifth point converges to a point of L. The set of singular points of the image is closed, so we may reduce to the case of the fifth point on a line; either the point is embedded or we reduce to case (1). We consider the case (1) below, so assume that the point is embedded.

For every λC, the coordinate change x2x2+λx3, x1x1λx4 maps Q1, L and Π to themselves. Using such an action for general λ, we may assume that Γ is not supported at [0:1:0:0:0]. Consider the torus action by t°[y1:y2:y3:y4:y5]=[y1:ty2:y3:t1y4:y5].

Again it preserves Q1, L and Π. The limit with t0 of any point [0:y2:y3:0:0]L{[0:1:0:0:0]} is [0:0:1:0:0].

Hence, we reduce to the case of Γ supported at [0:0:1:0:0] having an embedded point, so that Γ0=(x1=x24=x4=x5=0) and Γ spans Π. Here I is determined by I=I(Γ)q. We check its smoothness in Section A. (This is the case from Example 3.7.)

The case (1) is similar, but slightly more algebraic, as in this case the coordinate changes need to take into account the structure of I rather than Isat. By [Citation17, Proposition 4.4(3)], the space I3 is spanned by y23,y22y3,y2y32,y32 and y5·l2, where lC[y2,y3]1. Making the coordinate change as above, we may assume that Γ=V(Isat) is not supported at [0:1:0:0:0] and that l is not proportional to y2. Then a torus degeneration as above reduces us to the case y5y32 and Γ supported at [0:0:1:0:0], thus I is determined as I=I(Γ)(y5y32,q). Again, we check its smoothness in Section A. □

Remark 4.8.

Actually, according to the calculation in Section A, near [I2] the variety πG(HilbH) is equal to Syzn for n = 4, 5.

Proposition 4.9.

Let πG:HilbHSyzn be the natural map. The map GLn×VPSsbl(Q,H)GπG(SlipH);(g,I2)g·I2is smooth, so in particular VPSsbl(Q,H)G is smooth for n = 4, 5.

Proof.

We have πG(VPSsbl(Q,H))=VPSsbl(Q,H)G and the map πG sends I to I2, so the smoothness of the map follows by Theorem 2.4. For n5, the locus VPSgood(Q,H) is smooth so πG(VPSgood(Q,H)) is smooth by Proposition 4.5. Moreover, πG(SlipH) is smooth at the unsaturated limit ideals by Propositions 4.6–4.7, hence VPSsbl(Q,H)G is smooth by the smoothness of the map. □

When n = 1 and n = 2 the variety VPSsbl(Q,H)G is P1, respectively a 3-fold general linear section of G(2,5), cf. [Citation18]. In particular, in both cases, the Grassmannian subscheme VPSsbl(Q,H)G is a Fano variety of index 2. The same holds for n = 4, 5.

Corollary 4.10.

VPSsbl(Q,H)G is a Fano variety of index 2, when n = 4, 5.

Proof.

We follow the argument of [Citation20, Theorem 6.1]. Consider the set Hh={[I2]VPSsbl(Q,H)G | V(I2)h=}where hP(T1) is a hyperplane. In [Citation20, Lemma 4.9] the set Hh is defined for schemes ΓVPS(Q,n), but the definition extends naturally. Similarly, the proof of the lemma extends to show that Hh is the restriction of a Plücker divisor on G((n2),S2). Notice that the divisors Hh contains the image of all unsaturated limit ideals when n = 4, 5, since V(I2) in these cases contains a line that necessarily intersects the hyperplane h. By [Citation20, Proposition 5.11], the complement of Hh is isomorphic to an affine space, while Hh is very ample by Proposition 4.9. The image BLVPSsbl(Q,H)G of the set of all unsaturated limit ideals is not a divisor, so the argument of the proof of [Citation20, Theorem 6.1] applies: Since the complement of Hh is isomorphic to an affine space, the Picard group of VPSsbl(Q,H)G is Z as soon as the restriction Hh of the special Plücker divisor is irreducible. So assume Hh=H1+H2.

Any apolar scheme Γ lies in the complement of some h, so the subset BL is the base locus of the special divisors Hh. But BL is not a divisor, so H1 and H2 must both move with base locus contained in BL.

Now Hh·l=1 for every line lVPSsbl(Q,H)GBL, so only one of the two components can have positive intersection with l. In particular, one of the components, say H2 has intersection H2·l=0 and contains every line l that it intersects. By, [Citation20, Lemma 6.2], any two smooth apolar schemes Γ,ΓVPSsbl(Q,H)GBL are connected by a sequence of lines, so H2 would contain all of VPSsbl(Q,H)G. Therefore Hh is irreducible.

The computation of the Fano-index in [Citation20, Theorem 6.1] when n = 4, 5, applies similarly to VPSsbl(Q,H)G, see also Remark 4.14. □

4.1 The scheme VPSsbl(Q,H)G for n = 4

In the case n = 4, we give a more precise description of the Grassmannian compactification, starting with the image BLVPSsbl(Q,H)G of the family of unsaturated limit ideals in VPSsbl(Q,H) by the map π:VPSsbl(Q,H)Syzn. By Corollaries 3.6 and 3.10 each unsaturated limit ideal is contained in the ideal of a line in Q1, and the degree two part of such an ideal depends only on the line: it is the space of those quadrics in the ideal of the line that lie in q. Therefore BL=VPSsbl(Q,H)GVPSuns(Q,H)G=C1C2is the disjoint union of two rational curves C1 and C2, one for each pencil of lines in Q1.

Lemma 4.11.

Let n = 4. The two curves C1 and C2 are rational normal curves of degree 6 in VPSsbl(Q,H)GG(6,q2).

Proof.

For each line LQ1, we let IL be the intersection of its ideal with the space of quadrics in the apolar ideal q. Thus IL is 6-dimensional in q, and its orthogonal is a 3-dimensional subspace IL(q2)*. We proceed by arguing in G(3,(q2)*). Projectively the subspace P(IL) is the span of the conic v2(L)v2(Q1) under the veronese embedding v2:P(T1)P(T2). The rational P2-scroll X formed by these spans as L moves in one pencil of lines in Q1 is a rational normal scroll: For each triple of lines in one pencil in Q1, the corresponding conics on X is not contained in any hyperplane, so the degree of X is at least 6, the minimal degree of a 3-fold in P((q2)*). On the other hand v2(Q1) is linearly normal, hence also X, so X has degree 6. The Grassmannian embedding of a rational normal scroll of planes X in the Grassmannian G(3,(q2)*) of planes is a rational normal curve and has the same degree as X. The corresponding embedding in G(6,q2) of the pencil of spaces P(IL) dual to P(IL) is projectively equivalent and therefore also has degree 6. □

Lemma 4.12.

For each rational normal curve Ci,i=1,2 in VPSsbl(Q,H)G, the tangent bundle of VPSsbl(Q,H)G restricted to Ci is O(2)6 and the normal bundle is O(2)5.

Proof.

Fix coordinates so that Q=y1y4+y2y3 and C is the rational normal curve containing the point p=[(x3,x4)q]. Explicitly, we have (4) (x3,x4)q=(x1x3,x2x3x1x4,x32,x2x4,x3x4,x42).(4)

Consider the SL2-action on k[x1,,x4] where a matrix [aij]SL2 acts by (a000a1000a000a10a010a1100a010a11)

The quadric Q is invariant under this action and so SL2 acts on VPS(Q,H)G and on C. Actually, the action on C is transitive; the point p is sent to (x1a01+x3a11,x2a01+x4a11)q; thus the induced map SL2Aut(C)PGL2 is the usual one. The action above induces SL2-linearizations on tangent bundle and normal bundle from the statement. The action of the standard torus of SL2 decomposes the degree two part of (4) into weight spaces x1x3,x2x3x1x4,x2x4,andx32,x3x4,x42of weights 0 and –2, respectively. The tangent space TpVPS(Q,H)G is equal to HomS(((x3,x4)q)+S4, q(x3,x4)q+S4)0, which is 9-dimensional. Let W=x12,x1x2,x22q, it is a weight space of weight two. Any linear map ρ2 from the weight zero space in (x3,x4)q to W has weight two. To see that ρ2 lifts to a tangent vector ρ it is enough to check that it induces a linear map ρ3 in degree three that is zero on the linear syzygies of (x3,x4)q, since the linear syzygies generate all syzygies of (x3,x4)q. Since ρ2 has weight two, it induces, in degree three, a map ρ3 of the weight one space x1,x2·x1x3,x2x3x1x4,x2x4 to a weight three space of cubics.

But any linear syzygy of (x3,x4)q has weight ±1 so extending ρ2 to the zero map on the weight (2) space x32,x3x4,x42, means that ρ3 is zero on the syzygies. We conclude that ρ2 lifts to a tangent vector ρ.

The space of maps ρ2 is 9-dimensional, hence equal to TpVPS(Q,H)G, the tangent space to VPS(Q,H)G at p. It follows that the weights of the action of the standard torus of SL2 on TpVPS(Q,H)G are all equal to 2. The fiber of the tangent bundle from the statement is a subrepresentation of TpVPS(Q,H)G for this torus, so it also has all weights equal to 2. The fiber of the normal bundle is its quotient representation, so the same holds here. By the classification of vector bundles on P1, we find that the tangent bundle is O(2)6, while the normal bundle is O(2)5. □

Remark 4.13.

The degree and normal bundle of the curves C1 and C2 is also found in the Macaulay2 computation [Citation22].

Remark 4.14.

The normal bundle of the two curves C1C2=VPSsbl(Q,H)GVPSuns(Q,H)G may be used to compute the index of the former variety as a Fano variety. The Picard group is generated by H, so the canonical divisor is K=aH for some integer a. Since the normal bundle in VPSsbl(Q,H)G of each of the two rational curves Ci in BL has degree 10 while the degree of each curve is 6, the adjunction formula implies 2=degKCi=10deg(aHCi), so a = 2 and VPSsbl(Q,H)G has Fano index 2.

Proposition 4.15.

Let n = 4. The subscheme VPSsbl(Q,H)G inside G(6,q2) is an arithmetically Cohen-Macaulay variety and a smooth linear section VPSsbl(Q,H)G=P38G(6,q2),of degree 362. The image of the set of unsaturated limit ideals VPSsbl(Q,H)GVPSuns(Q,H)G form a union C1C2 of two disjoint rational normal curves each of degree 6.

A general codimension 6 linear space containing C1C2 intersects the locus VPSgood(Q,H)G in 310 points.

Proof.

It remains to prove the linear section and Cohen-Macaulay properties and the degree computations. This involves the Macaulay2 computation [Citation22], see Theorem B.1 in Appendix B. It proves that the linear section X:=P38G(6,q2), where P38 is the linear span of VPSsbl(Q,H)G, is a 6-dimensional arithmetically Cohen-Macaulay variety of degree 362.

Now, we argue that the degree of X coincides with the degree of VPSsbl(Q,H)G, so that VPSsbl(Q,H)G is the only component of maximal dimension in X. But, since X is arithmetically Cohen-Macaulay, every component is maximal, so X and VPSsbl(Q,H)G coincide if their degrees do.

The degree argument uses excess intersection. Let L be a general linear space of codimension 6 that contain the two curves C1C2. The excess intersection of the two curves in the intersection LVPSsbl(Q,H)G is computed by the formula [Citation9, Prop 9.1.1.(2)]: deg(L·VPSsbl(Q,H)G)Ci=deg(c1(NL))c1(NCi)=6(H·Ci)10=6210,where NL is the normal bundle of L restricted to Ci and H is the class of a hyperplane. So the excess intersection along C1C2 has degree 2(6210)=52. By the below remark, the linear space L intersects VPSgood(Q,H)G=VPSsbl(Q,H)G(C1C2) in 310 points. So the intersection of VPSsbl(Q,H)G with a general linear space of codimension 6 is 310+52=362, which coincides with the computation of the degree of X. □

Remark 4.16.

In [Citation20] special Plücker hyperplanes are considered in a computation of VPSsbl(Q,H)G. They already appeared in the proof of Corollary 4.10. For a hyperplane h=V(l)P(T1), the set Hh={[I2]VPSsbl(Q,H)G | V(I2)h=}is a Plücker hyperplane section of VPSsbl(Q,H)G. Since V(I2) contains a line for every unsaturated limit ideal I, every divisor Hh contains the image of all unsaturated limit ideals. According to [Citation20, Theorem 6.3], when n = 4, the intersection of six general hyperplanes Hh contains 310 ideals of polar simplices, i.e. points on VPSgood(Q,H)G. That theorem claims that this is also the degree of VPSgood(Q,H)G, based on the false assumption that this variety is closed. The locus VPSsbl(Q,H)GVPSgood(Q,H)G=C1C2, which is contained in every hyperplane Hh. The intersection of six general hyperplane sections Hh therefore contains C1C2 in addition to the 310 points on VPSgood(Q,H)G. So to compute the degree of VPSsbl(Q,H)G using the formula in [Citation20, Theorem 6.3], becomes the excess intersection problem of finding the contribution of C1C2 to the degree of the intersection.

Acknowledgments

We thank Tomasz Mańdziuk and Emanuele Ventura for pointing out an issue in an earlier version of the paper, and anonymous referees for helpful comments on presentation.

Additional information

Funding

JJ is supported by National Science Centre grant 2020/39/D/ST1/00132. This work is partially supported by the Thematic Research Programme “Tensors: geometry, complexity and quantum entanglement”, University of Warsaw, Excellence Initiative - Research University and the Simons Foundation Award No. 663281 granted to the Institute of Mathematics of the Polish Academy of Sciences for the years 2021–2023.

References

  • Artin, M. (1976). Deformations of Singularities. Notes by C.S. Seshadri and Allen Tannenbaum. Bombay, India: Tata Institute of Fundamental Research.
  • Buczyńska, W., Buczyński, J. (2014). Secant varieties to high degree veronese reembeddings, catalecticant matrices and smoothable gorenstein schemes. Duke Math. J. 23(1): 63–69.
  • Buczyńska, W., Buczyński, J. (2021). Apolarity, border rank, and multigraded Hilbert scheme. Duke Math. J. 170(16): 3659–3702. 10.1215/00127094-2021-0048
  • Bolognesi, M., Massarenti, A. (2018). Varieties of sums of powers and moduli spaces of (1, 7)-polarized abelian surfaces. J. Geom. Phys. 125: 23–32. 10.1016/j.geomphys.2017.12.004
  • Carlini, E., Catalisano, M. V., Oneto, A. (2017). Waring loci and the Strassen conjecture. Adv. Math. 314: 630–662. 10.1016/j.aim.2017.05.008
  • Casnati, G., Jelisiejew, J., Notari, R. (2015). Irreducibility of the Gorenstein loci of Hilbert schemes via ray families. Algebra Number Theory 9(7): 1525–1570. 10.2140/ant.2015.9.1525
  • Casnati, G., Notari, R. (2009). On the Gorenstein locus of some punctual Hilbert schemes. J. Pure Appl. Algebra 213(11): 2055–2074. 10.1016/j.jpaa.2009.03.002
  • Erman, D. (2012). Murphy’s law for Hilbert function strata in the Hilbert scheme of points. Math. Res. Lett. 19(6): 1277–1281. 10.4310/MRL.2012.v19.n6.a8
  • Fulton, W. (1998). Intersection Theory, volume 2 of Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics], 2nd ed. Berlin: Springer-Verlag.
  • Gallet, M., Ranestad, K., Villamizar, N. (2018). Varieties of apolar subschemes of toric surfaces. Ark. Mat. 56(1): 73–99. 10.4310/ARKIV.2018.v56.n1.a6
  • Grayson, D., Stillman, M. Macaulay2, a software system for research in algebraic geometry, ver. 1.21. Available at: https://macaulay2.com.
  • Hauser, H. (1983). An algorithm of construction of the semiuniversal deformation of an isolated singularity. In: Singularities, Part 1 (Arcata, Calif., 1981), volume 40 of Proceedings of Symposia in Pure Mathematics. Providence, RI: American Mathematical Society, pp. 567–573.
  • Hoyois, M., Jelisiejew, J., Nardin, D., Yakerson, M. (2022). Hermitian k-theory via oriented gorenstein algebras. J. Reine Angew. Math. 793: 105–142. 10.1515/crelle-2022-0063
  • Huang, H., Michałek, M., Ventura, E. (2020). Vanishing Hessian, wild forms and their border VSP. Math. Ann. 378(3–4): 1505–1532. 10.1007/s00208-020-02080-8
  • Haiman, M., Sturmfels, B. (2004). Multigraded Hilbert schemes. J. Algebraic Geom. 13(4): 725–769. 10.1090/S1056-3911-04-00373-X
  • Ilten, N. O. (2012). Versal deformations and local Hilbert schemes. J. Softw. Algebra Geom. 4: 12–16. 10.2140/jsag.2012.4.12
  • Jelisiejew, J., Mańdziuk, T. (2022). Limits of saturated ideals. arXiv:2210.13579.
  • Mukai, S. (1992). Fano 3-folds. London Math. Soc., Lect. Notes Ser. 179: 255–263.
  • Ranestad, K., Schreyer, F.-O. (2000). Varieties of sums of powers. J. Reine Angew. Math. 525: 147–181. 10.1515/crll.2000.064
  • Ranestad, K., Schreyer, F.-O. (2013). The variety of polar simplices. Doc. Math. 18: 469–505. 10.4171/dm/406
  • Ranestad, K., Voisin, C. (2017). Variety of power sums and divisors in the moduli space of cubic fourfolds. Doc. Math. 22: 455–504. 10.4171/dm/571
  • Schreyer, F.-O. (2023). vsp4, a Macaulay2 package for the computation of the variety of polar simplices of a quadric in P 3. Available at: https://www.math.uni-sb.de/ag/schreyer/index.php/computeralgebra.
  • Skjelnes, R. M., Ståhl, G. S. (2022). Explicit projective embeddings of standard opens of the Hilbert scheme of points. J. Algebra 590: 254–276. 10.1016/j.jalgebra.2021.10.007
  • The Stacks project authors. (2023). The stacks project. Available at: https://stacks.math.columbia.edu.

Appendix A:

Computer code

In this section we exhibit the computer code used to prove smoothness of specific points in VPSsbl(Q,H)G for n = 4, 5. The computations are performed over Q. See package [Citation16] for details.

loadPackage("VersalDeformations", Reload= >true);

computeComponents = I -> (

– check that Syz and Hilb agree

assert(hilbertFunction(3, I) == dim ring I);

J = ideal select(flatten entries mingens I, x->sum degree x< =2);

– the ideal I_2 + S_{\geq 4}

Jtr = ideal mingens(J + (ideal gens ring J)^4);

tg = normalMatrix(0, Jtr);

reducedtg = CT^1(0, Jtr);

ob = CT^2(0, Jtr);

(F,R,G,C) = versalDeformation(mingens Jtr,reducedtg,ob);

IG = ideal mingens ideal sum G;

pd = primaryDecomposition IG;

assert(#pd == 2);

– actually, those are linear spaces

assert(dim ideal singularLocus pd_0 == -1);

assert(dim ideal singularLocus pd_1 == -1);

use ring I; – reset the default ring

return (rank source tg - codim pd_0,

rank source tg - codim pd_1);

);

The code for n = 4 is as follows.

QQ[x_0. x_3];

Q = x_0*x_2 + x_1*x_3;

Qperp = inverseSystem(Q);

I = intersect(Qperp, ideal(x_1^4, x_2, x_3)); –case n = 4

computeComponents(I) == (12,10) – true

The code for n = 5 is very similar.

QQ[z_0. z_4];

Qz = z_0*z_2 + z_1*z_3 + z_4^2;

Qzperp = inverseSystem(Qz);

I = ideal(z_3*z_4, z_2*z_4, z_0*z_4, z_3^2, z_2*z_3, z_1*z_3,

z_0*z_3-z_4^2, z_2^2, z_1*z_2 - z_4^2, z_0*z_2, z_0^4);

computeComponents(I) == (20,20) – true

I = intersect(ideal(z_0, z_3, z_4, z_1^5),

inverseSystem(z_4*z_2^2), inverseSystem(Qz));

computeComponents(I) == (20,20) – true

The code for n = 4 is as follows.

QQ[x_1. x_4];

Q = x_1*x_3 + x_2*x_4;

Qperp = inverseSystem(Q);

I = intersect(Qperp, ideal(x_2^4, x_3, x_4)); –case n = 4

computeComponents(I) == (12,10) – true

The code for n = 5 is very similar.

QQ[x_1. x_5];

Qz = x_1*x_3 + x_2*x_4 + x_5^2;

Qzperp = inverseSystem(Qz);

I = ideal(x_4*x_5, x_2*x_5, x_1*x_5, x_4^2, x_1*x_4,x_3*x_4,

2*x_2*x_4-x_5^2, x_1^2,x_1*x_2, 2*x_1*x_3 - x_5^2, x_2^4);

computeComponents(I) == (20,20) – true

I = intersect(ideal(x_1, x_4, x_5, x_2^5),

inverseSystem(x_5*x_3^2), inverseSystem(Qz));

computeComponents(I) == (20,20) – true

Appendix B:

Remarks on the package vsp4.m2

In our package we prove computationally the following theorem. The computations are performed over Q.

Theorem B.1.

Consider VPSsbl(Q,H) with H=(1,4,4,). The Grassmannian model VPSsbl(Q,H)GG(6,q2)P(96)1is smooth. Then VPSsbl(Q,H)G is the intersection of G(6,q2) with the linear span of VPSsbl(Q,H)G inside P(96)1. It is an arithmetically Cohen-Macaulay variety of degree 362 with h-vector (1,32,148,148,32,1).

The proof is computational. The first step consists in computing the unfolding [Citation12] of the ideal (x02,x0x1,x12,x0x2x1x3,x1x2,x0x3,x24) and the computation of the flatness relations via a Gröbner basis computation. As it turns out the base space of the family consists of two components of dimension 8 and 6, respectively. The 8-dimensional family consists of lines in P3 together with 4 points. The general element of the 6-dimensional family consists of 4 distinct points not on a line, they vary as sets of points apolar to q. The intersection of these families consists of 4 points on a line on the inverse quadric Q1. So the intersection has two components corresponding to the two family of lines on Q1. They are of dimension 5.

In the next step we compute the image of these families in the Grassmannian G(6,q2). It is computational accessible to compute the linear equations in P(96)1. To see that image of the 8-dimensional family coincides with the 3-uple embedding of G(2,4)P5 is straightforward.

Analyzing the image of the 6-dimensional family is computationally difficult. The linear span is a P38 and the 1050 Plücker quadrics restrict to 380 independent quadrics on this P38. Moreover, the quadrics define a saturate ideal J of dimension 7 and degree 362. So J corresponds to a 6-dimensional scheme X of degree 362. The 6-th difference function of the Hilbert function of X takes values (1,32,148,148,32,1).

This lead us to conjecture that X is an arithmetically Gorenstein subscheme of P38P(96)1. Note that (31+22)=148+380. Thus if X is arithmetically Gorenstein, its homogeneous ideal in P38 is generated by these 380 quadrics. However we could not check directly that the homogeneous ideal of the image of the 6-dimensional family is generated by quadrics.

Instead we establish that X is arithmetically Cohen-Macaulay by testing that for a carefully chosen sparse sequence l1,,l6 of linear forms the ideals J+(l1,,li) are still saturated for i=1,6. Since dim(J+(l1,,l6))=1 (which we check computationally) the ideal J+(l1,,l6) defines the homogeneous ideal of a zero dimensional subscheme of a P32 of degree 362. Thus there exists a final linear form l7 such that l1,,l7 are a regular sequence for the homogeneous coordinate ring of X.

This proves that X is arithmetically Cohen-Macaulay and therefore unmixed.

In Proposition 4.15, we apply excess intersection theory to show that the two varieties X and VPSsbl(Q,H)G have the same degree, hence must coincide.

By the Macaulay2 computation, the image of the intersection of the two families consists of two rational normal curves C1, C2 of degree 6, as above. Furthermore it checks that X is smooth along the normal curves CiP1 and that their normal bundles are isomorphic to 15OP1(2), in accordance with Lemmas 4.11 and 4.12. Finally, by inspecting the 6 × 9 matrix which defines the map of the 6-dimensional family into the Grassmannian we see that VPSsbl(Q,H)G is smooth outside C1C2.