On the Castelnuovo-Mumford regularity of subspace arrangements
Abstract.
Let be the union of generic linear subspaces of codimension in . Improving an earlier bound due to Derksen and Sidman [DS02] we prove that the Castelnuovo-Mumford regularity of satisfies .
Key words and phrases:
subspace arrangements, Castelnuovo-Mumford regularity, linear resolutions, Hilbert functions, flat degenerations2020 Mathematics Subject Classification:
13D02,14M07,14B151. Introduction
The notion of Castelnuovo-Mumford regularity was introduced by Mumford in [MB66], to systematically control the vanishing of sheaf cohomology. Since then it has proved to be a fundamental notion throughout algebraic geometry and commutative algebra, both from the theoretical and computational perspectives. For instance, proofs for the existence of the Hilbert scheme rely on it [Kol99, MB66], while so do complexity bounds for Gröbner bases computations [BM93]. It even plays a role in the theory of subspace clustering [TV17] in machine learning.
In the present work we are concerned with the Castelnuovo-Mumford regularity of the reduced union of generic linear spaces of arbitrary dimensions in a projective space over an infinite field . When there are no pairwise intersections between the ’s, the regularity can be extracted from another important invariant; the Hilbert function. For points, i.e. when , the Hilbert function is easy to compute; e.g. see [GMR83]. However, for higher dimensions determining the Hilbert function is very challenging. A landmark achievement regarding this problem is the work of Hartshorne & Hirschowitz [HH82] who settled the case of lines. Another important step forward was taken by Derksen [Der07], who gave a formula for the Hilbert polynomial of and proved that it agrees with the Hilbert function at degrees . On the other hand, the Hilbert function is largely unknown at degrees .
The hardest part of the proof of Hartshorne & Hirschowitz [HH82] was to treat the case , which was done by degeneration techniques via smooth quadric surfaces. Recently a new proof was given by Aladpoosh & Catalisano [AC21], where degeneration techniques via linear spaces replaced those via quadrics. These have the potential of generalizing to higher dimensions and have inspired us in this article.
For generic lines the formula for the regularity has been derived by Rice [Ric22] from [HH82]. Apart from this, there has been a single general result regarding , due to Derksen and Sidman [DS02], who proved that for any reduced union of linear spaces, and provided examples of (special) configurations showing that their result is sharp; for special subspace arrangements there exist various results such as [BDMV18] and [BPS05].
In this paper we prove:
Theorem 1.
For generic linear spaces of codimension bigger than in , the regularity of their reduced union satisfies
At the heart of our proof lies a surprising result:
Theorem 2.
For generic linear spaces of codimension in , the regularity of their reduced union is equal to and the saturated ideal that defines has a linear graded minimal free resolution.
Another device that plays a role in the proof of Theorem 1, and of potential further significance to the problem of determining , is:
Theorem 3.
Let be any non-zero saturated homogeneous ideal of a polynomial ring of dimension over . For let be an ideal of generated by generic linear forms. Then
where .
In Theorem 3 one always has and in many cases the inequality is strict, e.g. if has a linear resolution then .
The bound of Theorem 1 is in general not sharp. As asserted though by Theorem 2, it is sharp when is the union of codimension- generic linear spaces in . For that particular configuration, Theorem 2 implies that the Hilbert function of is fully determined: it coincides with the Hilbert function of at degrees , while it agrees thereafter with the Hilbert polynomial of (described in [Der07]). Moreover, the ideal of has a linear resolution; we are not aware of other codimension- configurations where this phenomenon occurs. It further follows from our arguments that whenever the union of condimension- generic linear spaces in has regularity , then a union of generic linear spaces of codimension in has regularity .
2. Conventions
In the rest of the paper we set and we use for , where is the polynomial ring of dimension over the infinite field , endowed with the standard -grading. We denote by the unique maximal homogeneous ideal of . For a finitely generated graded -module we denote by the component of of degree , while the Hilbert function gives the dimension of the -vector space .
For a closed subscheme of , we let be the ideal sheaf of , and the unique graded saturated ideal of for which —here is the coherent sheaf induced by . When we say that a scheme contains a scheme , we will always mean an inclusion of ideals . Similarly, by the union of two schemes we will mean the scheme defined by the ideal .
With a non-negative integer and any integer, we denote the ordinary numerical binomial coefficient as , with the convention that it is if . With an element in a commutative ring that contains , we denote the binomial polynomial coefficient as
For any integer we have if and only if .
Finally, for a positive integer we set .
3. Preliminaries
3.1. Regularity
The Castelnuovo-Mumford regularity of a closed scheme can be defined as the Castelunuovo-Mumford regularity of the saturated ideal that defines . Here more generally we review the Castelnuovo-Mumford regularity of a finitely generated graded module over the polynomial ring [Eis94, BH98, Pee10], and quote basic properties that we will need.
In terms of local cohomology, the regularity of is defined as
In terms of the minimal graded free resolution of , is the maximum among the numbers , where is the largest degree of a minimal generator of the -th syzygy module of ; that is
The following fact is known as the regularity lemma:
Proposition 4.
Let be a short exact sequence of finitely generated graded -modules and maps of degree zero. Then
Moreover, if has finite length, then
A linear form is called almost regular on , if it is a non-zero divisor on . When for some homogeneous ideal , this condition translates to being regular on ; here is the saturation of with respect to . We recall also that the saturation index of is the smallest integer for which for every , i.e. . With that, the behavior of the regularity upon taking quotients by almost regular elements is extremely useful for inductive arguments on the dimension:
Proposition 5.
Let be a homogeneous ideal of and a linear form almost regular on . Then
or equivalently
For a proof of the above folklore fact we refer to [CH03].
3.2. Subspace Arrangements
Given an arrangement of -vector subspaces of , and the ideals that the ’s generate, we may consider the intersection or the product of the ’s. The intersection ideal is more geometric, because it is the radical ideal that defines the schematic union of the corresponding arrangement of linear spaces of . But it is a much harder object to understand than the product ideal ; already describing a set of generators of is a difficult task [Idà90]. On the other hand, many properties of are by now well understood, including a complete description of its minimal graded free resolution and of a minimal primary decomposition [CT22].
Definition 6.
We call the subspace arrangement of linearly general, if for every we have
We call the subspace arrangement of linearly general, if the ’s are linearly general.
The following result allows valuable access from the product ideal to the intersection ideal.
Proposition 7 (Conca & Herzog, [CH03]).
Suppose that the subspace arrangement , is linearly general. Then for every
Using a formula for the primary decomposition that they developed, Conca & Herzog [CH03] proved that
while Derksen & Sidman proved:
Proposition 8 (Derksen & Sidman, [DS02]).
Later on, Derksen gave a formula for the Hilbert polynomial and proved that it agrees with the Hilbert function from degrees and on:
Proposition 9 (Derksen, [Der07]).
Suppose that the subspace arrangement is linearly general. For any subset set . Then for every we have
where all terms with are zero by convention.
3.3. Flat Degenerations and Hilbert functions
A typical technique that is employed in the study of families of schemes —such as unions of generic lines in [HH82, AC21]— is that of a flat degeneration. This is a powerful tool due to the following fact:
Proposition 11.
Let be the extended graded polynomial ring in variables over , where the ’s have degree and has degree . Let be a homogeneous ideal of and let be the ideal of obtained by replacing with . If is a flat -module, then the Hilbert function of does not depend on .
4. Proof of Theorem 2
4.1. First Hilbert Function Lemma
We begin by giving a formula for the Hilbert function at degree of a linearly general subspace arrangement of codimension- planes in .
Lemma 12.
Suppose that the subspace arrangement , is linearly general and for every . Then
Proof.
By Proposition 7
Thus, with and , we have
Since the ideals are generated by a linearly general subspace arrangement of , where is a polynomial ring over in variables, we can compute the Hilbert function of the last term of the short exact sequence at degree again by Proposition 9:
Putting everything together, we arrive at
4.2. Second Hilbert Function Lemma
We recall a well-known binomial identity:
Lemma 13.
Let be a polynomial in and . Then
We have the following interesting fact:
Lemma 14.
Through linearly general codimension- linear spaces of passes a unique hypersurface of degree , i.e.
4.3. Sundials
The notion of a sundial played a crucial role in the proof of [HH82]. It was also used in [Hir81] and has more recently been generalized in various ways; e.g. see [AC21]. Here we discuss yet another generalization that we will employ in the proof of Lemma 17.
Definition 15.
In consider two linear spaces and , of codimensions and respectively, which lie in a hyperplane , and intersect at a codimension linear space . A -sundial is a scheme defined by an ideal sheaf of the form . We say that the hyperplane supports the sundial.
Lemma 16.
Let be linearly general linear spaces in of codimensions and , such that . Then there exists a flat family of closed subschemes of , such that is isomorphic to , while is a -sundial.
Proof.
We may assume that and . With as in Section 3.3, we define in the polynomial ring the ideal
Since is a regular sequence in and , we have
Moreover, the canonical ring homomorphism is flat if and only if is torsion-free as a -module —now this follows immediately once one notes that and are prime ideals of , whose contraction in is zero.
The ideal induces a flat family of closed subschemes of . The ideal of defining is
For we have that
defines two codimension and linear spaces of , which intersect at codimension . On the other hand,
defines a -sundial. ∎
4.4. A Variant of Castelnuovo’s Inequality
For a homogeneous ideal of and any linear form , the short exact sequences
give
Castelnuovo’s inequality —usually quoted in sheaf-theoretic form— is an easy consequence of the equality above:
Call the closed subscheme of defined by and the hyperplane defined by the linear form . Then the closed subscheme of defined by the ideal is called the residual scheme of along , denoted as . The closed subscheme of defined by the ideal is called the trace of along , and is denoted by —note also defines .
In the proof of Lemma 17 we will be concerned with showing for certain ideals and degrees . Often, the scheme defined by the ideal will be the reduced union in of linear spaces , together with a -sundial , for some . That is, . The linear form will always be chosen to define a hyperplane that supports the sundial; e.g., in the notation of the proof of Lemma 16 and we would take . Instead of dealing with the trace scheme defined by the possibly non-saturated ideal of
it will be more convenient to work with its subscheme , defined by the larger ideal
Since the sundial will always be of type and defines a hyperplane that supports the sundial, we see that — defined by the ideal — is the union in of a hyperplane together with a codimension- linear space . Thus always contributes a hyperplane as an irreducible component of , and possibly one more, whenever . Let be the number of the hyperplanes that appear as irreducible components of . Denote by the scheme obtained by removing these components from . It becomes clear that to show no hypersurface of of degree contains , it suffices to show that i) there is no hypersurface of of degree that contains , and ii) there is no hypersurface of of degree that contains . The arguments in Section 4.5 make repeated use of this idea.
4.5. Third Hilbert Function Lemma
The hardest part of proving Theorem 2 is establishing the following key fact:
Lemma 17.
With , we have
Recalling Proposition 11 and the discussion in Section 3.3, it suffices to show that this holds for a special member of a flat family of schemes, whose general member is the union of codimension- generic linear spaces of . We will be denoting the statement of Lemma 17 by
—the first index refers to the number of linear spaces in the arrangement, the second index refers to the degree of interest of the ideal of the arrangement, and the last index is the number of ambient variables.
To prove Lemma 17, we proceed by induction on . As a base for the induction, we take —the statement is then trivial, since no line in contains three generic points. In the sequel we assume that i) , and ii) is true for any .
Within this induction hypothesis, we prove an auxiliary statement:
Lemma 18.
Let . Consider the union in of codimension- generic linear spaces, together with a codimension- generic linear space. Then there is no hypersurface of degree containing this union.
We denote the statement of Lemma 18 by
We prove it by ascending induction on and descending induction on . For the base of our induction on we take —the only possibility is and the statement follows from the well-known fact that there is a unique quadric surface in through three generic lines (thus no quadric contains three generic lines and a generic point). For any , the base for the induction on is for . The statement then becomes . This follows immediately from Lemma 14, because through generic codimension- linear spaces in passes a unique hypersurface of degree , and the codimension- space is a generic point, which can be taken to be outside that hypersurface. In the sequel we will assume that i) is true for any and any such that , and ii) is true for any for which .
With , let be the scheme of Lemma 18. Since , we apply Lemma 16 to degenerate into a scheme that consists of codimension- generic linear spaces and a -sundial. Now we apply the variant of Castelnuovo’s inequality described in Section 4.4 —it is enough to show that i) there is no hypersurface of of degree that contains the residual scheme of with respect to a hyperplane that supports the sundial, and ii) there is no hypersurface of of degree that contains .
The residual scheme is the union in of codimension- generic linear spaces, together with a codimension- generic linear space. The fact that such a scheme lies in no hypersurface of degree is the statement
As this is true by our induction hypothesis on for .
Next, is the union in of codimension- generic linear spaces, together with a codimension- generic linear space. We claim that there is no degree- hypersurface of that contains . If , this is the statement
which is true by our induction hypothesis on for . If , it is the statement
which is true by our induction hypothesis on for (note that ). This concludes the proof of Lemma 18.
We now finish the proof of Lemma 17. So let be the union in of codimension- generic linear spaces . Since , by Lemma 16 we degenerate within a flat family to a -sundial. The resulting scheme is the union in of codimension- generic linear spaces, together with a generic -sundial. By Proposition 11, it is enough to show that there is no hypersurface of degree that contains . We do this by applying the variant of Castelnuovo’s inequality described in Section 4.4. So let be a hyperplane that supports the -sundial. The residual scheme is the union in of codimension- generic linear spaces together with a codimension- generic linear space. Saying that there is no hypersurface of degree that contains , is the same as saying that
holds true —now this follows from Lemma 18 with . Next, is the union in of codimension- generic linear spaces. We must show that there is no hypersurface of degree that contains (note that in the notation of Section 4.4). But this is statement
which is true by our induction hypothesis on for .
4.6. Saturation
The last ingredient is:
Lemma 19.
Let , be ideals generated by linear forms; suppose for at least one . Then for a generic linear form
Proof.
Consider the short exact sequence
where takes the class of in to the classes of mod for , and is the cokernel. Tensoring with and using the fact that is -regular for every , we get
We conclude that
Now, we can compute as the kernel of the multiplication map . Since is generic, it is almost regular on and so has finite length. Thus with and we have . This gives . But is radical because the ideals are prime and at least one of them is properly contained in the maximal homogeneous ideal of . ∎
4.7. Finishing the Proof of Theorem 2
We proceed by induction on ; the case is a simple exercise, so we assume .
Let be a generic linear form. As such, is regular on , and so Proposition 5 gives
It thus suffices to show that
Towards that end, we consider the short exact sequence
By Lemma 19, the first module in the exact sequence has finite length. Thus Proposition 4 gives
Set and . Then
and our induction hypothesis on gives
Hence . Since taking quotient with a generic linear form can not increase the regularity (Proposition 5), we have
This, together with the short exact sequence
and the regularity lemma, give
One more application of this argument with and gives
We have thus shown that
and so it remains to prove
Since , we have that is saturated from degree and on. Hence in view of Proposition 19, and agree from degree and on, and so it is enough to prove that
For the Hilbert function on the left we note
and now
To get a handle on , we work with the short exact sequence
whose degree component gives
By Lemma 17,
and so
To finish the proof we have to show that
Both these Hilbert function values are accessible via Lemma 12. Indeed, they are equal:
Lemma 20.
We have that
Proof.
The statement is equivalent to
and recalling that , equivalent to
The degree of the polynomial
is , so Lemma 13 gives
In the summation above there are terms. For , we claim that the terms corresponding to and are equal. To see this, with a positive integer, first recall the polynomial identity
Now, the th term is
which is precisely the th term. Consequently,
For those values of , the binomial polynomial coincides with the binomial coefficient, hence
Finally, the binomial coefficient
is already zero for any , so that
∎
5. Proof of Theorem 3
Recalling that is an ideal of generated by generic linear forms , we begin by recording some basic observations:
Lemma 21.
With any homogeneous ideal, we have:
-
(i)
.
-
(ii)
has finite length.
-
(iii)
If then .
Proof.
(i) We have a short exact sequence:
| (1) |
Since is generated by generic linear forms, Proposition 5 gives . Furthermore . Now the claim follows from the regularity lemma (Proposition 4).
(ii) The sequence is almost regular on . Hence the Koszul homology has finite length for all . In particular has finite length.
(iii) This follows from the above exact sequence together with the regularity lemma. ∎
The next step, even though intuitively non-surprising, is the hardest part behind the proof of Theorem 3:
Proposition 22.
For every non-zero saturated homogeneous ideal and every we have:
Proof.
We discuss first the case . Suppose, by contradiction that . Let us write , where . Set
Note that we have:
Since by Lemma 21 has finite length, also has finite length, and so by Proposition 4
We have an exact complex
and we further have a canonical projection
Note that the restriction of to composed with gives the canonical projection
Now we have
where the first inequality follows from Lemma 21(i) applied to the ideal , and the second inequality is by hypothesis.
Let and be such that
Since is saturated, is as well saturated; thus .
Consider the maps induced in cohomology by the above short exact sequence:
where we have used the fact that
via the map induced by the projection . Now
since
Moreover,
since
Therefore we have a surjective map
which, as observed above, is induced by the canonical projection . But is the composition of the two canonical projections
where is an ideal generated by generic linear forms, contained in and containing . The above composition induces maps
whose composition is the above surjective map . But as we have already seen, the middle term is zero, which is a contradiction.
We can now refine Lemma 21:
Proposition 23.
Suppose that is saturated. Then if and only if .
Proof.
By Lemma 21(iii) we know that implies . We argue for the reverse direction. For the sake of a contradiction, suppose
We may assume ; otherwise we simply replace with , noting that implies by Proposition 22. Set . The generic hyperplane section of can be realized by taking , i.e. . Since is almost regular on , the kernel of the multiplication map has finite length; thus Proposition 20.20 in [Eis94] gives
In our setting, by hypothesis we have , so Now the short exact sequence in the proof of Lemma 21(i), together with the assumption that is saturated, implies that is a submodule of . Since is a submodule of , we have
which is a contradiction. ∎
6. Proof of Theorem 1
For we let be generated by generic linear forms. We first prove:
Lemma 24.
Suppose that . Then .
Proof.
For each let be an ideal generated by two generic linear forms. By Theorem 2 and Lemma 21(i), we have . If for every , we are done. So we may assume . We may write and . Set , and . Then Proposition 22 gives
Continuing in a similar fashion by inductively applying Proposition 22 to the ideal , proves the statement. ∎
Theorem 1 now follows immediately from Lemma 24 and the following sub-additivity property of ideals of subspace arrangements:
Lemma 25.
Let and be ideals generated by generic linear forms, where and are positive integers. Then
References
- [AC21] T. Aladpoosh and M.V. Catalisano, On the Hartshorne–Hirschowitz theorem, Journal of Pure and Applied Algebra 225 (2021), no. 12, 106761.
- [BDMV18] B. Benedetti, M. Di Marca, and M. Varbaro, Regularity of line configurations, Journal of Pure and Applied Algebra 222 (2018), no. 9, 2596–2608.
- [BH98] W. Bruns and H. J. Herzog, Cohen-Macaulay Rings, no. 39, Cambridge University Press, 1998.
- [BM93] D. Bayer and D. Mumford, What can be computed in algebraic geometry?, Computational Algebraic Geometry and Commutative Algebra, Cambridge University Press, 1993, pp. 1–48.
- [BPS05] A. Björner, I. Peeva, and J. Sidman, Subspace arrangements defined by products of linear forms, Journal of the London Mathematical Society 71 (2005), no. 2, 273–288.
- [Cav07] G. Caviglia, Bounds on the Castelnuovo-Mumford regularity of tensor products, Proceedings of the American Mathematical Society 135 (2007), no. 7, 1949–1957.
- [CH03] A. Conca and J. Herzog, Castelnuovo-Mumford regularity of products of ideals, Collectanea Mathematica 54 (2003), no. 2, 137–152.
- [CT22] A. Conca and M.C. Tsakiris, Resolution of ideals associated to subspace arrangements, Algebra and Number Theory 16 (2022), no. 5, 1121–1140.
- [Der07] H. Derksen, Hilbert series of subspace arrangements, Journal of Pure and Applied Algebra 209 (2007), no. 1, 91–98.
- [DS02] H. Derksen and J. Sidman, A sharp bound for the Castelnuovo–Mumford regularity of subspace arrangements, Advances in Mathematics 172 (2002), no. 2, 151–157.
- [Eis94] D. Eisenbud, Commutative Algebra with a View Toward Algebraic Geometry, Springer, 1994.
- [GMR83] A. V. Geramita, P. Maroscia, and L.G. Roberts, The Hilbert function of a reduced k-algebra, Journal of the London Mathematical Society 2 (1983), no. 3, 443–452.
- [HH82] R. Hartshorne and A. Hirschowitz, Droites en position générale dans l’espace projectif, Algebraic Geometry, Springer, 1982, pp. 169–188.
- [Hir81] A. Hirschowitz, Sur la postulation générique des courbes rationnelles, Acta Mathematica 146 (1981), 209–230.
- [Idà90] M. Idà, On the homogeneous ideal of the generic union of lines , Journal für die Reine und Angewandte Mathematik 403 (1990), 67–153.
- [Kol99] J. Kollár, Rational Curves on Algebraic Varieties, vol. 32, Springer Science & Business Media, 1999.
- [MB66] D. Mumford and G. M. Bergman, Lectures on curves on an algebraic surface, no. 59, Princeton University Press, 1966.
- [Pee10] I. Peeva, Graded Syzygies, vol. 14, Springer, 2010.
- [Ric22] J. A. Rice, Generic lines in projective space and the Koszul property, Nagoya Mathematical Journal (2022), 1–30.
- [TV17] M. C. Tsakiris and R. Vidal, Filtrated algebraic subspace clustering, SIAM Journal on Imaging Sciences 10 (2017), no. 1, 372–415.