This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

On the Castelnuovo-Mumford regularity of subspace arrangements

Aldo Conca Dipartimento di Matematica, Università di Genova,
Via Dodecaneso 35, 16146 Genova, Italy
[email protected]
 and  Manolis C. Tsakiris Academy of Mathematics and Systems Science,
Chinese Academy of Sciences, 100190, Beijing, China
[email protected]
Abstract.

Let XX be the union of nn generic linear subspaces of codimension >1>1 in d\mathbb{P}^{d}. Improving an earlier bound due to Derksen and Sidman [DS02] we prove that the Castelnuovo-Mumford regularity of XX satisfies reg(X)nn/(2d1)\operatorname*{reg}(X)\leq n-\lfloor{n/(2d-1)\rfloor}.

Key words and phrases:
subspace arrangements, Castelnuovo-Mumford regularity, linear resolutions, Hilbert functions, flat degenerations
2020 Mathematics Subject Classification:
13D02,14M07,14B15

1. 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 reg(X)\operatorname*{reg}(X) of the reduced union XX of nn generic linear spaces X1,,XnX_{1},\dots,X_{n} of arbitrary dimensions in a projective space kd\mathbb{P}_{k}^{d} over an infinite field kk. When there are no pairwise intersections between the XiX_{i}’s, the regularity can be extracted from another important invariant; the Hilbert function. For points, i.e. when dimXi=0\dim X_{i}=0, 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 XX and proved that it agrees with the Hilbert function at degrees n\geq n. On the other hand, the Hilbert function is largely unknown at degrees <n<n.

The hardest part of the proof of Hartshorne & Hirschowitz [HH82] was to treat the case d=3d=3, 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 reg(X)\operatorname*{reg}(X), due to Derksen and Sidman [DS02], who proved that reg(X)n\operatorname*{reg}(X)\leq n 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 nn generic linear spaces X1,,XnX_{1},\dots,X_{n} of codimension bigger than 11 in kd\mathbb{P}_{k}^{d}, the regularity of their reduced union XX satisfies

reg(X)nn/(2d1).\operatorname*{reg}(X)\leq n-\lfloor{n/(2d-1)\rfloor}.

At the heart of our proof lies a surprising result:

Theorem 2.

For n=2d1n=2d-1 generic linear spaces X1,,XnX_{1},\dots,X_{n} of codimension 22 in kd\mathbb{P}_{k}^{d}, the regularity of their reduced union XX is equal to n1n-1 and the saturated ideal IXI_{X} that defines XX 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 reg(X)\operatorname*{reg}(X), is:

Theorem 3.

Let JJ be any non-zero saturated homogeneous ideal of a polynomial ring SS of dimension rr over kk. For c=1,,rc=1,\dots,r let LcL_{c} be an ideal of SS generated by cc generic linear forms. Then

reg(JLc)={reg(J)+1, if cγreg(J), if c>γ,\operatorname*{reg}(J\cap L_{c})=\left\{\begin{matrix}&\operatorname*{reg}(J)+1,\,\,&\mbox{ if }c\leq\gamma\\ &\operatorname*{reg}(J),\,\,&\mbox{ if }c>\gamma\end{matrix}\right.,

where γ=max{c:reg(J+Lc)=reg(J)}\gamma=\max\{c:\operatorname*{reg}(J+L_{c})=\operatorname*{reg}(J)\}.

In Theorem 3 one always has γdepth(S/J)\gamma\geq\operatorname*{depth}(S/J) and in many cases the inequality is strict, e.g. if JJ has a linear resolution then γr1\gamma\geq r-1.

The bound of Theorem 1 is in general not sharp. As asserted though by Theorem 2, it is sharp when XX is the union of n=2d1n=2d-1 codimension-22 generic linear spaces in kd\mathbb{P}_{k}^{d}. For that particular configuration, Theorem 2 implies that the Hilbert function of XX is fully determined: it coincides with the Hilbert function of kd\mathbb{P}_{k}^{d} at degrees <2d2<2d-2, while it agrees thereafter with the Hilbert polynomial of XX(described in [Der07]). Moreover, the ideal of XX has a linear resolution; we are not aware of other codimension-22 configurations where this phenomenon occurs. It further follows from our arguments that whenever the union of n0n_{0} condimension-22 generic linear spaces in kd\mathbb{P}_{k}^{d} has regularity n0α\leq n_{0}-\alpha, then a union of nn generic linear spaces of codimension 2\geq 2 in kd\mathbb{P}_{k}^{d} has regularity nαn/n0\leq n-\alpha\lfloor{n/n_{0}\rfloor}.

In §2 we set up notational conventions and in §3 we review preliminaries. In §4 we prove Theorem 2 and in §5 we prove Theorem 3. Finally we prove Theorem 1 in §6.

2. Conventions

In the rest of the paper we set r=d+1r=d+1 and we use kr1=ProjS\mathbb{P}^{r-1}_{k}=\operatorname*{Proj}S for kd\mathbb{P}^{d}_{k}, where S=k[x1,,xr]S=k[x_{1},\dots,x_{r}] is the polynomial ring of dimension r3r\geq 3 over the infinite field kk, endowed with the standard \mathbb{Z}-grading. We denote by 𝔪=(x1,,xr)\mathfrak{m}=(x_{1},\dots,x_{r}) the unique maximal homogeneous ideal of SS. For a finitely generated graded SS-module MM we denote by [M]ν[M]_{\nu} the component of MM of degree ν\nu, while the Hilbert function HF(M,ν)\operatorname*{HF}(M,\nu) gives the dimension of the kk-vector space [M]ν[M]_{\nu}.

For YY a closed subscheme of kr1\mathbb{P}_{k}^{r-1}, we let Y\mathcal{I}_{Y} be the ideal sheaf of YY, and IYI_{Y} the unique graded saturated ideal of SS for which Y=I~Y\mathcal{I}_{Y}=\tilde{I}_{Y} —here I~Y\tilde{I}_{Y} is the coherent sheaf induced by IYI_{Y}. When we say that a scheme YY contains a scheme ZZ, we will always mean an inclusion of ideals IYIZI_{Y}\subset I_{Z}. Similarly, by the union YZY\cup Z of two schemes we will mean the scheme defined by the ideal IYIZI_{Y}\cap I_{Z}.

With bb a non-negative integer and nn any integer, we denote the ordinary numerical binomial coefficient as (nb)=n!/((nb)!b!){n\choose b}=n!/((n-b)!b!), with the convention that it is 0 if n<bn<b. With tt an element in a commutative ring that contains \mathbb{Q}, we denote the binomial polynomial coefficient as

[tb]=(tb+1)(t1)tb!.\left[\begin{matrix}t\\ b\end{matrix}\right]=\frac{(t-b+1)\cdots(t-1)t}{b!}.

For any integer aa we have [ab]=(ab)\left[\begin{smallmatrix}a\\ b\end{smallmatrix}\right]={a\choose b} if and only if a0a\geq 0.

Finally, for a positive integer \ell we set []={1,,}[\ell]=\{1,\dots,\ell\}.

3. Preliminaries

3.1. Regularity

The Castelnuovo-Mumford regularity of a closed scheme Ykr1Y\subset\mathbb{P}_{k}^{r-1} can be defined as the Castelunuovo-Mumford regularity reg(IY)\operatorname*{reg}(I_{Y}) of the saturated ideal IYSI_{Y}\subset S that defines YY. Here more generally we review the Castelnuovo-Mumford regularity of a finitely generated graded module MM over the polynomial ring SS [Eis94, BH98, Pee10], and quote basic properties that we will need.

In terms of local cohomology, the regularity of MM is defined as

reg(M)=max{i+j:[H𝔪i(M)]j0}.\operatorname*{reg}(M)=\max\{i+j:\,[H_{\mathfrak{m}}^{i}(M)]_{j}\neq 0\}.

In terms of the minimal graded free resolution of MM, reg(M)\operatorname*{reg}(M) is the maximum among the numbers biib_{i}-i, where bib_{i} is the largest degree of a minimal generator of the ii-th syzygy module of MM; that is

reg(M)=max{bi:[ToriS(M,k)]b0}.\operatorname*{reg}(M)=\max\{b-i:\,[{\operatorname*{Tor}}_{i}^{S}(M,k)]_{b}\neq 0\}.

The following fact is known as the regularity lemma:

Proposition 4.

Let 0MMM′′00\rightarrow M^{\prime}\rightarrow M\rightarrow M^{\prime\prime}\rightarrow 0 be a short exact sequence of finitely generated graded SS-modules and maps of degree zero. Then

reg(M)\displaystyle\operatorname*{reg}(M) max{reg(M),reg(M′′)}\displaystyle\leq\max\{\operatorname*{reg}(M^{\prime}),\operatorname*{reg}(M^{\prime\prime})\}
reg(M)\displaystyle\operatorname*{reg}(M^{\prime}) max{reg(M),reg(M′′)+1}\displaystyle\leq\max\{\operatorname*{reg}(M),\operatorname*{reg}(M^{\prime\prime})+1\}
reg(M′′)\displaystyle\operatorname*{reg}(M^{\prime\prime}) max{reg(M),reg(M)1}.\displaystyle\leq\max\{\operatorname*{reg}(M),\operatorname*{reg}(M^{\prime})-1\}.

Moreover, if MM^{\prime} has finite length, then

reg(M)=max{reg(M),reg(M′′)}.\operatorname*{reg}(M)=\max\{\operatorname*{reg}(M^{\prime}),\operatorname*{reg}(M^{\prime\prime})\}.

A linear form x[S]1x\in[S]_{1} is called almost regular on MM, if it is a non-zero divisor on M/H𝔪0(M)M/H_{\mathfrak{m}}^{0}(M). When M=S/JM=S/J for some homogeneous ideal JJ, this condition translates to xx being regular on S/JsatS/J^{\operatorname*{sat}}; here Jsat=J:𝔪J^{\operatorname*{sat}}=J:\mathfrak{m}^{\infty} is the saturation of JJ with respect to 𝔪\mathfrak{m}. We recall also that the saturation index sat(J)\operatorname*{sat}(J) of JJ is the smallest integer ss for which [Jsat]ν=[J]ν[J^{\operatorname*{sat}}]_{\nu}=[J]_{\nu} for every νs\nu\geq s, i.e. sat(J)=reg(Jsat/J)+1\operatorname*{sat}(J)=\operatorname*{reg}(J^{\operatorname*{sat}}/J)+1. 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 JJ be a homogeneous ideal of SS and xSx\in S a linear form almost regular on S/JS/J. Then

reg(J)=max{reg(J+(x)),sat(J)},\operatorname*{reg}(J)=\max\{\operatorname*{reg}(J+(x)),\,\operatorname*{sat}(J)\},

or equivalently

reg(S/J)=max{reg(S/J+(x)),reg(Jsat/J)}.\operatorname*{reg}(S/J)=\max\{\operatorname*{reg}(S/J+(x)),\,\operatorname*{reg}(J^{\operatorname*{sat}}/J)\}.

For a proof of the above folklore fact we refer to [CH03].

3.2. Subspace Arrangements

Given an arrangement of nn kk-vector subspaces Vi,i[n]V_{i},\,i\in[n] of [S]1[S]_{1}, and the ideals IiI_{i} that the ViV_{i}’s generate, we may consider the intersection i[n]Ii\cap_{i\in[n]}I_{i} or the product i[n]Ii\prod_{i\in[n]}I_{i} of the IiI_{i}’s. The intersection ideal is more geometric, because it is the radical ideal that defines the schematic union XX of the corresponding arrangement of linear spaces Xi=Proj(S/Ii)X_{i}=\operatorname*{Proj}(S/I_{i}) of kr1\mathbb{P}_{k}^{r-1}. But it is a much harder object to understand than the product ideal i[n]Ii\prod_{i\in[n]}I_{i}; already describing a set of generators of i[n]Ii\cap_{i\in[n]}I_{i} is a difficult task [Idà90]. On the other hand, many properties of i[n]Ii\prod_{i\in[n]}I_{i} 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 V1,,VnV_{1},\dots,V_{n} of [S]1[S]_{1} linearly general, if for every A[n]A\subseteq[n] we have

dimkiAVi=min{r,iAdimkVi}.\dim_{k}\sum_{i\in A}V_{i}=\min\left\{r,\,{\sum}_{i\in A}\dim_{k}V_{i}\right\}.

We call the subspace arrangement X1,,XnX_{1},\dots,X_{n} of kr1\mathbb{P}_{k}^{r-1} linearly general, if the ViV_{i}’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 Vi,i[n]V_{i},\,i\in[n], is linearly general. Then for every νn\nu\geq n

[i[n]Ii]ν=[i[n]Ii]ν.\textstyle[\cap_{i\in[n]}I_{i}]_{\nu}=\big{[}\prod_{i\in[n]}I_{i}\big{]}_{\nu}.

Using a formula for the primary decomposition that they developed, Conca & Herzog [CH03] proved that

reg(i[n]Ii)=n,\textstyle\operatorname*{reg}\big{(}\prod_{i\in[n]}I_{i}\big{)}=n,

while Derksen & Sidman proved:

Proposition 8 (Derksen & Sidman, [DS02]).
reg(i[n]Ii)n.\operatorname*{reg}(\cap_{i\in[n]}I_{i})\leq n.

Later on, Derksen gave a formula for the Hilbert polynomial and proved that it agrees with the Hilbert function from degrees nn and on:

Proposition 9 (Derksen, [Der07]).

Suppose that the subspace arrangement Vi,i[n]V_{i},\,i\in[n] is linearly general. For any subset A[n]A\subseteq[n] set dA=dimkiAVid_{A}=\dim_{k}\sum_{i\in A}V_{i}. Then for every νn\nu\geq n we have

HF(S/i[n]Ii,ν)=A[n](1)#A+1(ν+r1dAr1dA),\operatorname*{HF}\big{(}S/\cap_{i\in[n]}I_{i},\nu\big{)}=\sum_{\emptyset\neq A\subseteq[n]}(-1)^{\#A+1}{\nu+r-1-d_{A}\choose r-1-d_{A}},

where all terms with r1dA<0r-1-d_{A}<0 are zero by convention.

Remark 10.

The notion of a linearly general subspace arrangement in projective space given in Definition 6 is equivalent to the notion of a transversal subspace arrangement as defined in [Der07].

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 k3\mathbb{P}_{k}^{3} [HH82, AC21]— is that of a flat degeneration. This is a powerful tool due to the following fact:

Proposition 11.

Let S[λ]=Skk[λ]S[\lambda]=S\otimes_{k}k[\lambda] be the extended graded polynomial ring in r+1r+1 variables over kk, where the xix_{i}’s have degree 11 and λ\lambda has degree 0. Let JλJ_{\lambda} be a homogeneous ideal of S[λ]S[\lambda] and let JcJ_{c} be the ideal of SS obtained by replacing λ\lambda with ckc\in k. If S[λ]/JλS[\lambda]/J_{\lambda} is a flat k[λ]k[\lambda]-module, then the Hilbert function of JcJ_{c} does not depend on cc.

For a proof see the arguments given in Exercise 20.14 of [Eis94]. We will use Proposition 11 later on, in the proof of Lemma 17, in the following fashion: under the flatness hypothesis, JcJ_{c} does not contain a form of degree ν\nu if and only if J0J_{0} does not contain a form of degree ν\nu.

4. Proof of Theorem 2

4.1. First Hilbert Function Lemma

We begin by giving a formula for the Hilbert function at degree 1\ell-1 of a linearly general subspace arrangement of \ell codimension-22 planes in kr1\mathbb{P}_{k}^{r-1}.

Lemma 12.

Suppose that the subspace arrangement Vi,i[]V_{i},\,i\in[\ell], is linearly general and dimkVi=2\dim_{k}V_{i}=2 for every i[]i\in[\ell]. Then

HF(i[]Ii,1)=j0(1)j(j)(+r22j1).\operatorname*{HF}(\cap_{i\in[\ell]}I_{i},\ell-1)=\sum_{j\geq 0}(-1)^{j}{\ell\choose j}{\ell+r-2-2j\choose\ell-1}.
Proof.

Consider the short exact sequence

0Si[]IiSi[1]IiSISI+i[1]Ii0.0\rightarrow\frac{S}{\cap_{i\in[\ell]}I_{i}}\rightarrow\frac{S}{\cap_{i\in[\ell-1]}I_{i}}\oplus\frac{S}{I_{\ell}}\rightarrow\frac{S}{I_{\ell}+\cap_{i\in[\ell-1]}I_{i}}\rightarrow 0.

By Proposition 9

HF(Si[1]Ii,1)=\displaystyle\operatorname*{HF}\left(\frac{S}{\cap_{i\in[\ell-1]}I_{i}},\ell-1\right)=
A[1](1)|A|+1(1|A|)(1+r12|A|r12|A|)\displaystyle\sum_{\emptyset\neq A\subseteq[\ell-1]}(-1)^{|A|+1}{\ell-1\choose|A|}{\ell-1+r-1-2|A|\choose r-1-2|A|}
=j1(1)j+1(1j)(+r22jr12j).\displaystyle=\sum_{j\geq 1}(-1)^{j+1}{\ell-1\choose j}{\ell+r-2-2j\choose r-1-2j}.

By Proposition 7

[i[1]Ii]1=[i[1]Ii]1.\left[\bigcap_{i\in[\ell-1]}I_{i}\right]_{\ell-1}=\left[\prod_{i\in[\ell-1]}I_{i}\right]_{\ell-1}.

Thus, with S¯=S/I\bar{S}=S/I_{\ell} and I¯i=Ii+I/I\bar{I}_{i}=I_{i}+I_{\ell}/I_{\ell}, we have

[SI+i[1]Ii]1=[S¯i[1]I¯i]1.\left[\frac{S}{I_{\ell}+\bigcap_{i\in[\ell-1]}I_{i}}\right]_{\ell-1}=\left[\frac{\bar{S}}{{\prod}_{i\in[\ell-1]}\bar{I}_{i}}\right]_{\ell-1}.

Since the ideals I¯i,i[1]\bar{I}_{i},\,i\in[\ell-1] are generated by a linearly general subspace arrangement of [S¯]1[\bar{S}]_{1}, where S¯\bar{S} is a polynomial ring over kk in r2r-2 variables, we can compute the Hilbert function of the last term of the short exact sequence at degree 1\ell-1 again by Proposition 9:

HF(Si[1]I¯i,1)=\displaystyle\operatorname*{HF}\left(\frac{S}{\prod_{i\in[\ell-1]}\bar{I}_{i}},\ell-1\right)=
A[r2](1)|A|+1(1|A|)(1+r32|A|r32|A|)\displaystyle\sum_{\emptyset\neq A\subseteq[r-2]}(-1)^{|A|+1}{\ell-1\choose|A|}{\ell-1+r-3-2|A|\choose r-3-2|A|}
=j1(1)j+1(1j)(+r42jr32j).\displaystyle=\sum_{j\geq 1}(-1)^{j+1}{\ell-1\choose j}{\ell+r-4-2j\choose r-3-2j}.

Putting everything together, we arrive at

HF(i[]Ii,1)=HF(i[1]Ii,1)HF(i[1]I¯i,1)\displaystyle\operatorname*{HF}(\cap_{i\in[\ell]}I_{i},\ell-1)=\operatorname*{HF}(\cap_{i\in[\ell-1]}I_{i},\ell-1)-\operatorname*{HF}(\cap_{i\in[\ell-1]}\bar{I}_{i},\ell-1)
=j0(1)j(1j)(+r22jr12j)\displaystyle=\sum_{j\geq 0}(-1)^{j}{\ell-1\choose j}{\ell+r-2-2j\choose r-1-2j}
j0(1)j(1j)(+r42jr32j)\displaystyle-\sum_{j\geq 0}(-1)^{j}{\ell-1\choose j}{\ell+r-4-2j\choose r-3-2j}
=j0(1)j(1j)(+r22jr12j)\displaystyle=\sum_{j\geq 0}(-1)^{j}{\ell-1\choose j}{\ell+r-2-2j\choose r-1-2j}
+j1(1)j(1j1)(+r22jr12j)\displaystyle+\sum_{j\geq 1}(-1)^{j}{\ell-1\choose j-1}{\ell+r-2-2j\choose r-1-2j}
=j0(1)j[(1j)+(1j1)](+r22jr12j)\displaystyle=\sum_{j\geq 0}(-1)^{j}\bigg{[}{\ell-1\choose j}+{\ell-1\choose j-1}\bigg{]}{\ell+r-2-2j\choose r-1-2j}
=j0(1)j(j)(+r22jr12j).\displaystyle=\sum_{j\geq 0}(-1)^{j}{\ell\choose j}{\ell+r-2-2j\choose r-1-2j}.\qed

4.2. Second Hilbert Function Lemma

We recall a well-known binomial identity:

Lemma 13.

Let p(t)p(t) be a polynomial in tt and d>deg(p)d>\deg(p). Then

j=0d(1)j(dj)p(j)=0.\sum_{j=0}^{d}(-1)^{j}{d\choose j}p(j)=0.

We have the following interesting fact:

Lemma 14.

Through r1r-1 linearly general codimension-22 linear spaces of kr1\mathbb{P}_{k}^{r-1} passes a unique hypersurface of degree r2r-2, i.e.

HF(i[r1]Ii,r2)=1.\operatorname*{HF}(\cap_{i\in[r-1]}I_{i},r-2)=1.
Proof.

Lemma 12 gives us

HF(i[r1]Ii,r2)=\displaystyle\operatorname*{HF}(\cap_{i\in[r-1]}I_{i},r-2)=
=j0(1)j(r1j)(2r32jr2)\displaystyle=\sum_{j\geq 0}(-1)^{j}{r-1\choose j}{2r-3-2j\choose r-2}
=j=0(2r3)/2(1)j(r1j)[2r32jr2]\displaystyle=\sum_{j=0}^{\lfloor(2r-3)/2\rfloor}(-1)^{j}{r-1\choose j}\begin{bmatrix}2r-3-2j\\ r-2\end{bmatrix}
=j=0r2(1)j(r1j)[2r32jr2].\displaystyle=\sum_{j=0}^{r-2}(-1)^{j}{r-1\choose j}\begin{bmatrix}2r-3-2j\\ r-2\end{bmatrix}.

As a polynomial in tt, the degree of

p(t)=[2r32tr2]p(t)=\begin{bmatrix}2r-3-2t\\ r-2\end{bmatrix}

is r2r-2. Hence Lemma 13 gives

j=0r1(1)j(r1j)[2r32jr2]=0.\sum_{j=0}^{r-1}(-1)^{j}{r-1\choose j}\begin{bmatrix}2r-3-2j\\ r-2\end{bmatrix}=0.

Now we are done by observing that the last term in the summation above for j=r1j=r-1 is equal to 1-1. ∎

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 kr1\mathbb{P}^{r-1}_{k} consider two linear spaces YY and ZZ, of codimensions 22 and i+2i+2 respectively, which lie in a hyperplane HH, and intersect at a codimension i+3i+3 linear space WW. A (2,i+2)(2,i+2)-sundial is a scheme defined by an ideal sheaf of the form YZW2\mathcal{I}_{Y}\cap\mathcal{I}_{Z}\cap\mathcal{I}_{W}^{2}. We say that the hyperplane HH supports the sundial.

Lemma 16.

Let Y,ZY,Z be linearly general linear spaces in kr1\mathbb{P}^{r-1}_{k} of codimensions 22 and 2+i2+i, such that 4+ir4+i\leq r. Then there exists a flat family Ec,ckE_{c},\,c\in k of closed subschemes of kr1\mathbb{P}^{r-1}_{k}, such that Ec0E_{c\neq 0} is isomorphic to YZY\cup Z, while E0E_{0} is a (2,i+2)(2,i+2)-sundial.

Proof.

We may assume that IY=(x1,x2)I_{Y}=(x_{1},x_{2}) and IZ=(x3,,xi+4)I_{Z}=(x_{3},\dots,x_{i+4}). With λ\lambda as in Section 3.3, we define in the polynomial ring S[λ]S[\lambda] the ideal

Jλ=(x1,λ(x2x3)+x3)(x3,,xi+4).J_{\lambda}=\big{(}x_{1},\lambda(x_{2}-x_{3})+x_{3}\big{)}\cap(x_{3},\dots,x_{i+4}).

Since x1,λ(x2x3)+x3x_{1},\lambda(x_{2}-x_{3})+x_{3} is a regular sequence in S[λ]S[\lambda] and S[λ]/(x3,,xr)S[\lambda]/(x_{3},\dots,x_{r}), we have

Jλ=(x1,λ(x2x3)+x3)(x3,,xi+4).J_{\lambda}=\big{(}x_{1},\lambda(x_{2}-x_{3})+x_{3}\big{)}(x_{3},\dots,x_{i+4}).

Moreover, the canonical ring homomorphism φ:k[λ]S[λ]/Jλ\varphi:k[\lambda]\rightarrow S[\lambda]/J_{\lambda} is flat if and only if S[λ]/JλS[\lambda]/J_{\lambda} is torsion-free as a k[λ]k[\lambda]-module —now this follows immediately once one notes that (x1,λ(x2x3)+x3)\big{(}x_{1},\lambda(x_{2}-x_{3})+x_{3}\big{)} and (x3,,xi+4)(x_{3},\dots,x_{i+4}) are prime ideals of S[λ]S[\lambda], whose contraction in k[λ]k[\lambda] is zero.

The ideal JλJ_{\lambda} induces a flat family of closed subschemes Ec,ckE_{c},\,c\in k of kr1\mathbb{P}_{k}^{r-1}. The ideal of SS defining EcE_{c} is

Jc=(x1,c(x2x3)+x3)(x3,,xi+4).J_{c}=\big{(}x_{1},c(x_{2}-x_{3})+x_{3}\big{)}(x_{3},\dots,x_{i+4}).

For c0c\neq 0 we have that

Jc=(x1,c(x2x3)+x3)(x3,,xi+4),J_{c}=\big{(}x_{1},c(x_{2}-x_{3})+x_{3}\big{)}\cap(x_{3},\dots,x_{i+4}),

defines two codimension 22 and i+2i+2 linear spaces of kr1\mathbb{P}_{k}^{r-1}, which intersect at codimension i+4i+4. On the other hand,

J0\displaystyle J_{0} =(x1,x3)(x3,,xi+4)\displaystyle=\big{(}x_{1},x_{3}\big{)}(x_{3},\dots,x_{i+4})
=(x1,x3)(x3,,xi+4)(x1,x3,,xi+4)2\displaystyle=(x_{1},x_{3})\cap(x_{3},\dots,x_{i+4})\cap(x_{1},x_{3},\dots,x_{i+4})^{2}

defines a (2,i+2)(2,i+2)-sundial. ∎

4.4. A Variant of Castelnuovo’s Inequality

For a homogeneous ideal JJ of SS and any linear form xx, the short exact sequences

0S/J:(x)(1)xS/JS/J+(x)0,0\rightarrow S/J:(x)(-1)\stackrel{{\scriptstyle x}}{{\rightarrow}}S/J\rightarrow S/J+(x)\rightarrow 0,
0S(1)xSS/(x)0,0\rightarrow S(-1)\stackrel{{\scriptstyle x}}{{\rightarrow}}S\rightarrow S/(x)\rightarrow 0,

give

HF(J,ν)=HF(J+(x)/(x),ν)+HF(J:(x),ν1).\operatorname*{HF}(J,\nu)=\operatorname*{HF}\big{(}J+(x)/(x),\nu\big{)}+\operatorname*{HF}(J:(x),\nu-1).

Castelnuovo’s inequality —usually quoted in sheaf-theoretic form— is an easy consequence of the equality above:

HF(J,ν)HF((J+(x)/(x))sat,ν)+HF(J:(x),ν1).\operatorname*{HF}(J,\nu)\leq\operatorname*{HF}\big{(}\big{(}J+(x)/(x)\big{)}^{\operatorname*{sat}},\nu\big{)}+\operatorname*{HF}(J:(x),\nu-1).

Call YY the closed subscheme of kr1\mathbb{P}_{k}^{r-1} defined by JJ and HH the hyperplane defined by the linear form xx. Then the closed subscheme of kr1\mathbb{P}_{k}^{r-1} defined by the ideal J:(x)J:(x) is called the residual scheme of YY along HH, denoted as ResH(Y)\operatorname*{Res}_{H}(Y). The closed subscheme of kr2\mathbb{P}_{k}^{r-2} defined by the ideal J+(x)/(x)J+(x)/(x) is called the trace of YY along HH, and is denoted by TrH(Y)\operatorname*{Tr}_{H}(Y) —note (J+(x)/(x))sat(J+(x)/(x))^{\operatorname*{sat}} also defines TrH(Y)\operatorname*{Tr}_{H}(Y).

In the proof of Lemma 17 we will be concerned with showing HF(J,ν)=0\operatorname*{HF}(J,\nu)=0 for certain ideals JJ and degrees ν\nu. Often, the scheme YY defined by the ideal JJ will be the reduced union in kr1\mathbb{P}_{k}^{r-1} of linear spaces YiY_{i}, together with a (2,j+)(2,j+\ell)-sundial UU, for some j,0j,\ell\geq 0. That is, J=(iIYi)IUJ=(\cap_{i}I_{Y_{i}})\cap I_{U}. The linear form xx will always be chosen to define a hyperplane that supports the sundial; e.g., in the notation of the proof of Lemma 16 IU=J0I_{U}=J_{0} and we would take x=x3x=x_{3}. Instead of dealing with the trace scheme TrH(Y)\operatorname*{Tr}_{H}(Y) defined by the possibly non-saturated ideal of S/(x)S/(x)

(iIYi)IU+(x)(x),\frac{(\cap_{i}I_{Y_{i}})\cap I_{U}+(x)}{(x)},

it will be more convenient to work with its subscheme TrH(Y)\operatorname*{Tr}^{\dagger}_{H}(Y), defined by the larger ideal

(iIYi+(x)(x))(IU+(x)(x)).\left(\bigcap_{i}\frac{I_{Y_{i}}+(x)}{(x)}\right)\cap\left(\frac{I_{U}+(x)}{(x)}\right).

Since the sundial will always be of type (2,j+)(2,j+\ell) and (x)(x) defines a hyperplane that supports the sundial, we see that TrH(U)\operatorname*{Tr}_{H}(U) — defined by the ideal (IU+(x))/(x)(I_{U}+(x))/(x)— is the union in kr2\mathbb{P}_{k}^{r-2} of a hyperplane together with a codimension-(j+1)(j+\ell-1) linear space ZZ. Thus TrH(U)\operatorname*{Tr}_{H}(U) always contributes a hyperplane as an irreducible component of TrH(Y)\operatorname*{Tr}^{\dagger}_{H}(Y), and possibly one more, whenever j+1=1j+\ell-1=1. Let 1α21\leq\alpha\leq 2 be the number of the hyperplanes that appear as irreducible components of TrH(U)\operatorname*{Tr}_{H}(U). Denote by TrH(Y)\operatorname*{Tr}^{\dagger\dagger}_{H}(Y) the scheme obtained by removing these components from TrH(Y)\operatorname*{Tr}^{\dagger}_{H}(Y). It becomes clear that to show no hypersurface of kr1\mathbb{P}_{k}^{r-1} of degree ν\nu contains YY, it suffices to show that i) there is no hypersurface of kr1\mathbb{P}_{k}^{r-1} of degree ν1\nu-1 that contains ResH(Y)\operatorname*{Res}_{H}(Y), and ii) there is no hypersurface of kr2\mathbb{P}_{k}^{r-2} of degree (να)(\nu-\alpha) that contains TrH(Y)\operatorname*{Tr}^{\dagger\dagger}_{H}(Y). 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 n=2r3n=2r-3, we have

HF(i[n]Ii,n2)=0.\operatorname*{HF}(\cap_{i\in[n]}I_{i},n-2)=0.

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 nn codimension-22 generic linear spaces of kr1\mathbb{P}_{k}^{r-1}. We will be denoting the statement of Lemma 17 by

2r3, 2r5,r\mathscr{H}_{2r-3,\,2r-5,\,r}

—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 rr. As a base for the induction, we take r=3r=3 —the statement is then trivial, since no line in 2\mathbb{P}^{2} contains three generic points. In the sequel we assume that i) r4r\geq 4, and ii) 2r3,2r5,r\mathscr{H}_{2r^{\prime}-3,2r^{\prime}-5,r^{\prime}} is true for any 3r<r3\leq r^{\prime}<r.

Within this induction hypothesis, we prove an auxiliary statement:

Lemma 18.

Let 0jr40\leq j\leq r-4. Consider the union in kr1\mathbb{P}_{k}^{r-1} of 2r5j2r-5-j codimension-22 generic linear spaces, together with a codimension-(j+3)(j+3) generic linear space. Then there is no hypersurface of degree 2r6j2r-6-j containing this union.

We denote the statement of Lemma 18 by

2r5j,j+3, 2r6j,r.\mathscr{H}^{\prime}_{2r-5-j,\,j+3,\,2r-6-j,\,r}.

We prove it by ascending induction on rr and descending induction on jj. For the base of our induction on rr we take r=4r=4 —the only possibility is j=0j=0 and the statement follows from the well-known fact that there is a unique quadric surface in 3\mathbb{P}^{3} through three generic lines (thus no quadric contains three generic lines and a generic point). For any rr, the base for the induction on jj is for j=r4j=r-4. The statement then becomes r1,r1,r2,r\mathscr{H}^{\prime}_{r-1,r-1,r-2,r}. This follows immediately from Lemma 14, because through r1r-1 generic codimension-22 linear spaces in kr1\mathbb{P}_{k}^{r-1} passes a unique hypersurface of degree r2r-2, and the codimension-(r1)(r-1) space is a generic point, which can be taken to be outside that hypersurface. In the sequel we will assume that i) 2r5j,j+3, 2r6j,r\mathscr{H}^{\prime}_{2r^{\prime}-5-j,\,j+3,\,2r^{\prime}-6-j,\,r^{\prime}} is true for any r<rr^{\prime}<r and any jj such that 0jr40\leq j\leq r^{\prime}-4, and ii) 2r5j,j+3, 2r6j,r\mathscr{H}^{\prime}_{2r-5-j^{\prime},\,j^{\prime}+3,\,2r-6-j^{\prime},\,r} is true for any jj^{\prime} for which j<jr4j<j^{\prime}\leq r-4.

With 0j<r40\leq j<r-4, let YY be the scheme of Lemma 18. Since 2+(j+3)r2+(j+3)\leq r, we apply Lemma 16 to degenerate YY into a scheme that consists of 2r6j2r-6-j codimension-22 generic linear spaces and a (2,j+3)(2,j+3)-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 kr1\mathbb{P}_{k}^{r-1} of degree 2r7j2r-7-j that contains the residual scheme ResH(Y)\operatorname*{Res}_{H}(Y) of YY with respect to a hyperplane HH that supports the sundial, and ii) there is no hypersurface of kr2\mathbb{P}_{k}^{r-2} of degree 2r7j2r-7-j that contains TrH(Y)\operatorname*{Tr}^{\dagger\dagger}_{H}(Y).

The residual scheme ResH(Y)\operatorname*{Res}_{H}(Y) is the union in kr1\mathbb{P}_{k}^{r-1} of 2r6j2r-6-j codimension-22 generic linear spaces, together with a codimension-(j+4)(j+4) generic linear space. The fact that such a scheme lies in no hypersurface of degree 2r7j2r-7-j is the statement

2r5(j+1),(j+1)+3, 2r6(j+1),r.\mathscr{H}^{\prime}_{2r-5-(j+1),\,(j+1)+3,\,2r-6-(j+1),\,r}.

As j+1r4,j+1\leq r-4, this is true by our induction hypothesis on jj for \mathscr{H}^{\prime}.

Next, TrH(Y)\operatorname*{Tr}^{\dagger\dagger}_{H}(Y) is the union in kr2\mathbb{P}_{k}^{r-2} of 2r6j2r-6-j codimension-22 generic linear spaces, together with a codimension-(j+2)(j+2) generic linear space. We claim that there is no degree-(2r7j)(2r-7-j) hypersurface of kr2\mathbb{P}_{k}^{r-2} that contains TrH(Y)\operatorname*{Tr}^{\dagger\dagger}_{H}(Y). If j=0j=0, this is the statement

2(r1)3,2(r1)5,r1,\mathscr{H}_{2(r-1)-3,2(r-1)-5,r-1},

which is true by our induction hypothesis on rr for \mathscr{H}. If j>0j>0, it is the statement

2(r1)5(j1),(j1)+3, 2(r1)6(j1),r1,\mathscr{H}^{\prime}_{2(r-1)-5-(j-1),\,(j-1)+3,\,2(r-1)-6-(j-1),\,r-1},

which is true by our induction hypothesis on rr for \mathscr{H}^{\prime} (note that j1<(r1)4j-1<(r-1)-4). This concludes the proof of Lemma 18.

We now finish the proof of Lemma 17. So let Y=i[2r3]YiY=\bigcup_{i\in[2r-3]}Y_{i} be the union in kr1\mathbb{P}_{k}^{r-1}of 2r32r-3 codimension-22 generic linear spaces YiY_{i}. Since r4r\geq 4, by Lemma 16 we degenerate Y2r4Y2r3Y_{2r-4}\cup Y_{2r-3} within a flat family to a (2,2)(2,2)-sundial. The resulting scheme YY^{\prime} is the union in kr1\mathbb{P}_{k}^{r-1} of 2r52r-5 codimension-22 generic linear spaces, together with a generic (2,2)(2,2)-sundial. By Proposition 11, it is enough to show that there is no hypersurface of degree 2r52r-5 that contains YY^{\prime}. We do this by applying the variant of Castelnuovo’s inequality described in Section 4.4. So let HH be a hyperplane that supports the (2,2)(2,2)-sundial. The residual scheme ResH(Y)\operatorname*{Res}_{H}(Y^{\prime}) is the union in kr1\mathbb{P}_{k}^{r-1}of 2r52r-5 codimension-22 generic linear spaces together with a codimension-33 generic linear space. Saying that there is no hypersurface of degree 2r62r-6 that contains ResH(Y)\operatorname*{Res}_{H}(Y^{\prime}), is the same as saying that

2r5, 3, 2r6,r\mathscr{H}^{\prime}_{2r-5,\,3,\,2r-6,\,r}

holds true —now this follows from Lemma 18 with j=0j=0. Next, TrH(Y)\operatorname*{Tr}^{\dagger\dagger}_{H}(Y^{\prime}) is the union in kr2\mathbb{P}_{k}^{r-2} of 2r52r-5 codimension-22 generic linear spaces. We must show that there is no hypersurface of degree 2r72r-7 that contains TrH(Y)\operatorname*{Tr}^{\dagger\dagger}_{H}(Y^{\prime}) (note that α=2\alpha=2 in the notation of Section 4.4). But this is statement

2(r1)3, 2(r1)5,r1,\mathscr{H}_{2(r-1)-3,\,2(r-1)-5,\,r-1},

which is true by our induction hypothesis on rr for \mathscr{H}.

4.6. Saturation

The last ingredient is:

Lemma 19.

Let Ji,i[]J_{i},\,i\in[\ell], be ideals generated by linear forms; suppose dimS/Ji2\dim S/J_{i}\geq 2 for at least one ii. Then for a generic linear form hh

[i[]Ji+(h)]sat=i[](Ji+(h))\displaystyle\big{[}\cap_{i\in[\ell]}J_{i}+(h)\big{]}^{\operatorname*{sat}}=\cap_{i\in[\ell]}\big{(}J_{i}+(h)\big{)}
Proof.

Consider the short exact sequence

0Si[]Jiφi[]S/JiC0,0\rightarrow\frac{S}{\cap_{i\in[\ell]}J_{i}}\stackrel{{\scriptstyle\varphi}}{{\rightarrow}}\prod_{i\in[\ell]}S/J_{i}\rightarrow C\rightarrow 0,

where φ\varphi takes the class of aSa\in S in S/i[]JiS/\cap_{i\in[\ell]}J_{i} to the classes of aa mod JiJ_{i} for i[]i\in[\ell], and CC is the cokernel. Tensoring with S/(h)S/(h) and using the fact that hh is S/JiS/J_{i}-regular for every ii, we get

0Tor1S(C,S(h))Si[]Ji+(h)φi[]SJi+(h)ChC0.0\rightarrow{\operatorname*{Tor}}_{1}^{S}\left(C,\frac{S}{(h)}\right)\rightarrow\frac{S}{\cap_{i\in[\ell]}J_{i}+(h)}\stackrel{{\scriptstyle\varphi}}{{\rightarrow}}\prod_{i\in[\ell]}\frac{S}{J_{i}+(h)}\rightarrow\frac{C}{hC}\rightarrow 0.

We conclude that

M:=i[](Ji+(h))i[]Ji+(h)=Tor1S(C,S(h)).M:=\frac{\cap_{i\in[\ell]}\big{(}J_{i}+(h)\big{)}}{\cap_{i\in[\ell]}J_{i}+(h)}={\operatorname*{Tor}}_{1}^{S}\left(C,\frac{S}{(h)}\right).

Now, we can compute Tor1S(C,S/(h))\operatorname*{Tor}_{1}^{S}(C,S/(h)) as the kernel of the multiplication map C(1)hCC(-1)\stackrel{{\scriptstyle h}}{{\rightarrow}}C. Since hh is generic, it is almost regular on C(1)C(-1) and so MM has finite length. Thus with A=i[](Ji+(h))A=\cap_{i\in[\ell]}\big{(}J_{i}+(h)\big{)} and B=i[]Ji+(h)B=\cap_{i\in[\ell]}J_{i}+(h) we have BABsatB\subset A\subseteq B^{\operatorname*{sat}}. This gives Bsat=AsatB^{\operatorname*{sat}}=A^{\operatorname*{sat}}. But AA is radical because the ideals Ji+(h)J_{i}+(h) are prime and at least one of them is properly contained in the maximal homogeneous ideal of SS. ∎

4.7. Finishing the Proof of Theorem 2

We proceed by induction on rr; the case r=3r=3 is a simple exercise, so we assume r4r\geq 4.

Let h[S]1h\in[S]_{1} be a generic linear form. As such, hh is regular on S/i[n]IiS/\cap_{i\in[n]}I_{i}, and so Proposition 5 gives

reg(Si[n]Ii)=reg(Si[n]Ii+(h)).\operatorname*{reg}\left(\frac{S}{\cap_{i\in[n]}I_{i}}\right)=\operatorname*{reg}\left(\frac{S}{\cap_{i\in[n]}I_{i}+(h)}\right).

It thus suffices to show that

reg(Si[n]Ii+(h))=n2.\operatorname*{reg}\left(\frac{S}{\cap_{i\in[n]}I_{i}+(h)}\right)=n-2.

Towards that end, we consider the short exact sequence

0i[n](Ii+(h))i[n]Ii+(h)Si[n]Ii+(h)Si[n](Ii+(h))0.0\rightarrow\frac{\cap_{i\in[n]}\big{(}I_{i}+(h)\big{)}}{\cap_{i\in[n]}I_{i}+(h)}\rightarrow\frac{S}{\cap_{i\in[n]}I_{i}+(h)}\rightarrow\frac{S}{\cap_{i\in[n]}\big{(}I_{i}+(h)\big{)}}\rightarrow 0.

By Lemma 19, the first module in the exact sequence has finite length. Thus Proposition 4 gives

reg(Si[n]Ii+(h))=\operatorname*{reg}\left(\frac{S}{\cap_{i\in[n]}I_{i}+(h)}\right)=
max{reg(i[n](Ii+(h))i[n]Ii+(h)),reg(Si[n](Ii+(h)))}.\max\left\{\operatorname*{reg}\left(\frac{\cap_{i\in[n]}\big{(}I_{i}+(h)\big{)}}{\cap_{i\in[n]}I_{i}+(h)}\right),\operatorname*{reg}\left(\frac{S}{\cap_{i\in[n]}\big{(}I_{i}+(h)\big{)}}\right)\right\}.

Set S¯=S/(h)\bar{S}=S/(h) and I¯i=Ii+(h)/(h)\bar{I}_{i}=I_{i}+(h)/(h). Then

Si[n2](Ii+(h))=S¯i[n2]I¯i,\frac{S}{\cap_{i\in[n-2]}(I_{i}+(h))}=\frac{\bar{S}}{\cap_{i\in[n-2]}\bar{I}_{i}},

and our induction hypothesis on rr gives

reg(S¯i[n2]I¯i)=n4.\operatorname*{reg}\left(\frac{\bar{S}}{\cap_{i\in[n-2]}\bar{I}_{i}}\right)=n-4.

Hence reg(i[n2]I¯i)=n3\operatorname*{reg}(\cap_{i\in[n-2]}\bar{I}_{i})=n-3. Since taking quotient with a generic linear form can not increase the regularity (Proposition 5), we have

reg(i[n2]I¯i+I¯n1)n3.\operatorname*{reg}(\cap_{i\in[n-2]}\bar{I}_{i}+\bar{I}_{n-1})\leq n-3.

This, together with the short exact sequence

0S¯i[n1]I¯iS¯i[n2]I¯iS¯I¯n1S¯i[n2]I¯i+I¯n10,0\rightarrow\frac{\bar{S}}{\cap_{i\in[n-1]}\bar{I}_{i}}\rightarrow\frac{\bar{S}}{\cap_{i\in[n-2]}\bar{I}_{i}}\oplus\frac{\bar{S}}{\bar{I}_{n-1}}\rightarrow\frac{\bar{S}}{\cap_{i\in[n-2]}\bar{I}_{i}+\bar{I}_{n-1}}\rightarrow 0,

and the regularity lemma, give

reg(i[n1]I¯i)n2.\operatorname*{reg}(\cap_{i\in[n-1]}\bar{I}_{i})\leq n-2.

One more application of this argument with i[n1]I¯i\cap_{i\in[n-1]}\bar{I}_{i} and I¯n\bar{I}_{n} gives

reg(i[n]I¯i)n1.\operatorname*{reg}(\cap_{i\in[n]}\bar{I}_{i})\leq n-1.

We have thus shown that

reg(Si[n](Ii+(h)))n2,\operatorname*{reg}\left(\frac{S}{\cap_{i\in[n]}(I_{i}+(h))}\right)\leq n-2,

and so it remains to prove

reg(i[n](Ii+(h))i[n]Ii+(h))n2.\operatorname*{reg}\left(\frac{\cap_{i\in[n]}\big{(}I_{i}+(h)\big{)}}{\cap_{i\in[n]}I_{i}+(h)}\right)\leq n-2.

Since reg(i[n]Ii+(h))n\operatorname*{reg}(\cap_{i\in[n]}I_{i}+(h))\leq n, we have that i[n]Ii+(h)\cap_{i\in[n]}I_{i}+(h) is saturated from degree nn and on. Hence in view of Proposition 19, i[n](Ii+(h))\cap_{i\in[n]}\big{(}I_{i}+(h)\big{)} and i[n]Ii+(h)\cap_{i\in[n]}I_{i}+(h) agree from degree nn and on, and so it is enough to prove that

HF(i[n](Ii+(h)),n1)=HF(i[n]Ii+(h),n1).\operatorname*{HF}(\cap_{i\in[n]}(I_{i}+(h)),n-1)=\operatorname*{HF}(\cap_{i\in[n]}I_{i}+(h),n-1).

For the Hilbert function on the left we note

HF(Si[n](Ii+(h)),n1)=HF(S¯i[n]I¯i,n1),\operatorname*{HF}\left(\frac{S}{\cap_{i\in[n]}(I_{i}+(h))},n-1\right)=\operatorname*{HF}\left(\frac{\bar{S}}{\cap_{i\in[n]}\bar{I}_{i}},n-1\right),

and now

HF(i[n](Ii+(h)),n1)=HF(S,n2)+HF(i[n]I¯i,n1).\operatorname*{HF}(\cap_{i\in[n]}(I_{i}+(h)),n-1)=\operatorname*{HF}(S,n-2)+\operatorname*{HF}(\cap_{i\in[n]}\bar{I}_{i},n-1).

To get a handle on HF(i[n]Ii+(h),n1)\operatorname*{HF}(\cap_{i\in[n]}I_{i}+(h),n-1), we work with the short exact sequence

0Si[n]Ii(1)hSi[n]IiSi[n]Ii+(h)0,0\rightarrow\frac{S}{\cap_{i\in[n]}I_{i}}(-1)\stackrel{{\scriptstyle h}}{{\rightarrow}}\frac{S}{\cap_{i\in[n]}I_{i}}\rightarrow\frac{S}{\cap_{i\in[n]}I_{i}+(h)}\rightarrow 0,

whose degree n1n-1 component gives

HF(i[n]Ii+(h),n1)=\displaystyle\operatorname*{HF}(\cap_{i\in[n]}I_{i}+(h),n-1)= HF(i[n]Ii,n1)HF(i[n]Ii,n2)\displaystyle\operatorname*{HF}(\cap_{i\in[n]}I_{i},n-1)-\operatorname*{HF}(\cap_{i\in[n]}I_{i},n-2)
+HF(S,n2).\displaystyle+\operatorname*{HF}(S,n-2).

By Lemma 17,

HF(i[n]Ii,n2)=0,\operatorname*{HF}(\cap_{i\in[n]}I_{i},n-2)=0,

and so

HF(i[n]Ii+(h),n1)=HF(S,n2)+HF(i[n]Ii,n1).\operatorname*{HF}(\cap_{i\in[n]}I_{i}+(h),n-1)=\operatorname*{HF}(S,n-2)+\operatorname*{HF}(\cap_{i\in[n]}I_{i},n-1).

To finish the proof we have to show that

HF(i[n]I¯i,n1)=HF(i[n]Ii,n1).\operatorname*{HF}(\cap_{i\in[n]}\bar{I}_{i},n-1)=\operatorname*{HF}(\cap_{i\in[n]}I_{i},n-1).

Both these Hilbert function values are accessible via Lemma 12. Indeed, they are equal:

Lemma 20.

We have that

j0(1)j(nj)(n+r22jn1)=j0(1)j(nj)(n+r32jn1).\sum_{j\geq 0}(-1)^{j}{n\choose j}{n+r-2-2j\choose n-1}=\sum_{j\geq 0}(-1)^{j}{n\choose j}{n+r-3-2j\choose n-1}.
Proof.

The statement is equivalent to

j0(1)j(nj)(n+r32jn2)=0,\sum_{j\geq 0}(-1)^{j}{n\choose j}{n+r-3-2j\choose n-2}=0,

and recalling that n=2r3n=2r-3, equivalent to

j0(1)j(2r3j)(3r62j2r5)=0.\sum_{j\geq 0}(-1)^{j}{2r-3\choose j}{3r-6-2j\choose 2r-5}=0.

The degree of the polynomial

[3r62t2r5]\begin{bmatrix}3r-6-2t\\ 2r-5\end{bmatrix}

is 2r52r-5, so Lemma 13 gives

j=02r3(1)j(2r3j)[3r62j2r5]=0.\sum_{j=0}^{2r-3}(-1)^{j}{2r-3\choose j}\begin{bmatrix}3r-6-2j\\ 2r-5\end{bmatrix}=0.

In the summation above there are 2r22r-2 terms. For j=0,,r2j=0,\dots,r-2, we claim that the terms corresponding to jj and 2r3j2r-3-j are equal. To see this, with bb a positive integer, first recall the polynomial identity

[tb]=(1)b[bt1b].\begin{bmatrix}t\\ b\end{bmatrix}=(-1)^{b}\begin{bmatrix}b-t-1\\ b\end{bmatrix}.

Now, the (2r3j)(2r-3-j)th term is

(1)2r3j(2r32r3j)[3r62(2r3j)2r5]\displaystyle(-1)^{2r-3-j}{2r-3\choose 2r-3-j}\begin{bmatrix}3r-6-2(2r-3-j)\\ 2r-5\end{bmatrix}
=(1)3j(2r3j)[r+2j2r5]\displaystyle=(-1)^{-3-j}{2r-3\choose j}\begin{bmatrix}-r+2j\\ 2r-5\end{bmatrix}
=(1)3j(2r3j)(1)2r5[2r5+r2j12r5]\displaystyle=(-1)^{-3-j}{2r-3\choose j}(-1)^{2r-5}\begin{bmatrix}2r-5+r-2j-1\\ 2r-5\end{bmatrix}
=(1)j(2r3j)[3r62j2r5],\displaystyle=(-1)^{j}{2r-3\choose j}\begin{bmatrix}3r-6-2j\\ 2r-5\end{bmatrix},

which is precisely the jjth term. Consequently,

j=0r2(1)j(2r3j)[3r62j2r5]=0.\sum_{j=0}^{r-2}(-1)^{j}{2r-3\choose j}\begin{bmatrix}3r-6-2j\\ 2r-5\end{bmatrix}=0.

For those values of jj, the binomial polynomial coincides with the binomial coefficient, hence

j=0r2(1)j(2r3j)(3r62j2r5)=0.\sum_{j=0}^{r-2}(-1)^{j}{2r-3\choose j}{3r-6-2j\choose 2r-5}=0.

Finally, the binomial coefficient

(3r62j2r5){3r-6-2j\choose 2r-5}

is already zero for any j>r2j>r-2, so that

j0(1)j(2r3j)(3r62j2r5)=0.\sum_{j\geq 0}(-1)^{j}{2r-3\choose j}{3r-6-2j\choose 2r-5}=0.

5. Proof of Theorem 3

Recalling that LcL_{c} is an ideal of SS generated by cc generic linear forms h¯=h1,,hc\underline{h}=h_{1},\dots,h_{c}, we begin by recording some basic observations:

Lemma 21.

With JJ any homogeneous ideal, we have:

  1. (i)

    reg(JLc)reg(J)+1\operatorname*{reg}(J\cap L_{c})\leq\operatorname*{reg}(J)+1.

  2. (ii)

    (JLc)/JLc(J\cap L_{c})/JL_{c} has finite length.

  3. (iii)

    If reg(J+Lc)<reg(J)\operatorname*{reg}(J+L_{c})<\operatorname*{reg}(J) then reg(JLc)=reg(J)\operatorname*{reg}(J\cap L_{c})=\operatorname*{reg}(J).

Proof.

(i) We have a short exact sequence:

(1) 0SJLcSJSLcSJ+Lc0.\displaystyle 0\to\frac{S}{J\cap L_{c}}\to\frac{S}{J}\oplus\frac{S}{L_{c}}\to\frac{S}{J+L_{c}}\to 0.

Since LcL_{c} is generated by generic linear forms, Proposition 5 gives reg(S/(J+Lc))reg(S/J)\operatorname*{reg}(S/(J+L_{c}))\leq\operatorname*{reg}(S/J). Furthermore reg(S/Lc)=0\operatorname*{reg}(S/L_{c})=0. Now the claim follows from the regularity lemma (Proposition 4).

(ii) The sequence h¯\underline{h} is almost regular on S/JS/J. Hence the Koszul homology Hi(h¯,S/J)H_{i}(\underline{h},S/J) has finite length for all i>0i>0. In particular H1(h¯,S/J)=Tor1S(S/J,S/Lc)=(JLc)/JLcH_{1}(\underline{h},S/J)=\operatorname*{Tor}^{S}_{1}(S/J,S/L_{c})=(J\cap L_{c})/JL_{c} 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 JSJ\subset S and every c[r1]c\in[r-1] we have:

reg(JLc)reg(JLc+1)\operatorname*{reg}(J\cap L_{c})\geq\operatorname*{reg}(J\cap L_{c+1})
Proof.

We discuss first the case c<r1c<r-1. Suppose, by contradiction that reg(JLc)<reg(JLc+1)\operatorname*{reg}(J\cap L_{c})<\operatorname*{reg}(J\cap L_{c+1}). Let us write Lc+1=Lc+L1L_{c+1}=L_{c}+L_{1}^{\prime}, where L1=(hc+1)L_{1}^{\prime}=(h_{c+1}). Set

K=(JLc)+(JL1).K=(J\cap L_{c})+(J\cap L_{1}^{\prime}).

Note that we have:

JLc+1=JL1+JLcKJLc+1.JL_{c+1}=JL_{1}^{\prime}+JL_{c}\subseteq K\subseteq J\cap L_{c+1}.

Since by Lemma 21 (JLc+1)/JLc+1(J\cap L_{c+1})/JL_{c+1} has finite length, JLc+1/KJ\cap L_{c+1}/K also has finite length, and so by Proposition 4

reg(SJLc+1)reg(SK).\operatorname*{reg}\left(\frac{S}{J\cap L_{c+1}}\right)\leq\operatorname*{reg}\left(\frac{S}{K}\right).

We have an exact complex

0SJLcL1SJL1SJLcfSK0\displaystyle 0\to\frac{S}{J\cap L_{c}\cap L_{1}^{\prime}}\to\frac{S}{J\cap L_{1}^{\prime}}\oplus\frac{S}{J\cap L_{c}}\stackrel{{\scriptstyle f}}{{\to}}\frac{S}{K}\to 0

and we further have a canonical projection

g:SKSJLc+1.g:\frac{S}{K}\to\frac{S}{J\cap L_{c+1}}.

Note that the restriction of ff to S/(JL1)S/(J\cap L_{1}^{\prime}) composed with gg gives the canonical projection

π1,c+1:SJL1SJLc+1.\pi_{1,c+1}:\frac{S}{J\cap L_{1}^{\prime}}\to\frac{S}{J\cap L_{c+1}}.

Now we have

reg(SJLcL1)reg(SJLc)+1reg(SJLc+1),\displaystyle\operatorname*{reg}\left(\frac{S}{J\cap L_{c}\cap L_{1}^{\prime}}\right)\leq\operatorname*{reg}\left(\frac{S}{J\cap L_{c}}\right)+1\leq\operatorname*{reg}\left(\frac{S}{J\cap L_{c+1}}\right),

where the first inequality follows from Lemma 21(i) applied to the ideal JLcJ\cap L_{c}, and the second inequality is by hypothesis.

Let i[r]i\in[r] and aa\in\mathbb{N} be such that

[H𝔪i(SJLc+1)]a0 and i+a=reg(SJLc+1).\left[H^{i}_{\mathfrak{m}}\left(\frac{S}{J\cap L_{c+1}}\right)\right]_{a}\neq 0\mbox{ and }i+a=\operatorname*{reg}\left(\frac{S}{J\cap L_{c+1}}\right).

Since JJ is saturated, JLc+1J\cap L_{c+1} is as well saturated; thus i>0i>0.

Consider the maps induced in cohomology by the above short exact sequence:

[H𝔪i(SJL1)]a[H𝔪i(SJLc)]a[H𝔪i(SJLc+1)]a[H𝔪i+1(SJLcL1)]a,\begin{matrix}\left[H^{i}_{\mathfrak{m}}\left(\frac{S}{J\cap L_{1}^{\prime}}\right)\right]_{a}\\ \oplus\\ \left[H^{i}_{\mathfrak{m}}\left(\frac{S}{J\cap L_{c}}\right)\right]_{a}\end{matrix}\to\left[H_{\mathfrak{m}}^{i}\left(\frac{S}{J\cap L_{c+1}}\right)\right]_{a}\to\left[H^{i+1}_{\mathfrak{m}}\left(\frac{S}{J\cap L_{c}\cap L_{1}^{\prime}}\right)\right]_{a},

where we have used the fact that

H𝔪i(SK)Hmi(SJLc+1)H_{\mathfrak{m}}^{i}\left(\frac{S}{K}\right)\simeq H_{m}^{i}\left(\frac{S}{J\cap L_{c+1}}\right)

via the map induced by the projection ff. Now

[H𝔪i(SJLc)]a=0,\left[H^{i}_{\mathfrak{m}}\left(\frac{S}{J\cap L_{c}}\right)\right]_{a}=0,

since

a+i=reg(SJLc+1)>reg(SJLc).a+i=\operatorname*{reg}\left(\frac{S}{J\cap L_{c+1}}\right)>\operatorname*{reg}\left(\frac{S}{J\cap L_{c}}\right).

Moreover,

[H𝔪i+1(SJLcL1)]a=0,\left[H^{i+1}_{\mathfrak{m}}\left(\frac{S}{J\cap L_{c}\cap L_{1}^{\prime}}\right)\right]_{a}=0,

since

i+1+a=1+reg(SJLc+1)>reg(SJLcL1).i+1+a=1+\operatorname*{reg}\left(\frac{S}{J\cap L_{c+1}}\right)>\operatorname*{reg}\left(\frac{S}{J\cap L_{c}\cap L_{1}^{\prime}}\right).

Therefore we have a surjective map

ε:[H𝔪i(SJL1)]a[H𝔪i(SJLc+1)]a0,\displaystyle\varepsilon:\left[H^{i}_{\mathfrak{m}}\left(\frac{S}{J\cap L_{1}^{\prime}}\right)\right]_{a}\to\left[H_{\mathfrak{m}}^{i}\left(\frac{S}{J\cap L_{c+1}}\right)\right]_{a}\neq 0,

which, as observed above, is induced by the canonical projection π1,c+1\pi_{1,c+1}. But π1,c+1\pi_{1,c+1} is the composition of the two canonical projections

SJL1π1,cSJLcπc,c+1SJLc+1,\frac{S}{J\cap L_{1}^{\prime}}\stackrel{{\scriptstyle\pi_{1,c}}}{{\to}}\frac{S}{J\cap L_{c}^{\prime}}\stackrel{{\scriptstyle\pi_{c,c+1}}}{{\to}}\frac{S}{J\cap L_{c+1}},

where LcL_{c}^{\prime} is an ideal generated by cc generic linear forms, contained in Lc+1L_{c+1} and containing L1L_{1}^{\prime}. The above composition induces maps

[H𝔪i(SJL1)]a[H𝔪i(SJLc)]a[H𝔪i(SJLc+1)]a,\displaystyle\left[H^{i}_{\mathfrak{m}}\left(\frac{S}{J\cap L_{1}^{\prime}}\right)\right]_{a}\to\left[H_{\mathfrak{m}}^{i}\left(\frac{S}{J\cap L_{c}^{\prime}}\right)\right]_{a}\to\left[H_{\mathfrak{m}}^{i}\left(\frac{S}{J\cap L_{c+1}}\right)\right]_{a},

whose composition is the above surjective map ϵ\epsilon. But as we have already seen, the middle term is zero, which is a contradiction.

It remains to discuss the case c=r1c=r-1. In this special case, since Lc+1=Lr=𝔪L_{c+1}=L_{r}=\mathfrak{m} and JLc+1=JJ\cap L_{c+1}=J, we have to prove that reg(JLr1)reg(J)\operatorname*{reg}(J\cap L_{r-1})\geq\operatorname*{reg}(J). We have J+Lr1=(p)+Lr1J+L_{r-1}=(p)+L_{r-1} where deg(p)=min{a:Ja0}\deg(p)=\min\{a:J_{a}\neq 0\}. Note that reg(J+Lr1)=deg(p)\operatorname*{reg}(J+L_{r-1})=\deg(p). If deg(p)<reg(J)\deg(p)<\operatorname*{reg}(J) we conclude by Lemma 21iii) that reg(JLr1)=reg(J)\operatorname*{reg}(J\cap L_{r-1})=\operatorname*{reg}(J). Otherwise deg(p)=reg(J)\deg(p)=\operatorname*{reg}(J) (one has that JJ has a linear resolution) and the short exact sequence in the proof of Lemma 21(i) gives:

0H𝔪0(SJ+Lr1)=SJ+Lr1H𝔪1(SJLr1)0\to H^{0}_{\mathfrak{m}}\left(\frac{S}{J+L_{r-1}}\right)=\frac{S}{J+L_{r-1}}\to H^{1}_{\mathfrak{m}}\left(\frac{S}{J\cap L_{r-1}}\right)

It follows that

reg(SJLr1)1+reg(SJ+Lr1)=reg(J),\operatorname*{reg}\left(\frac{S}{J\cap L_{r-1}}\right)\geq 1+\operatorname*{reg}\left(\frac{S}{J+L_{r-1}}\right)=\operatorname*{reg}(J),

that is reg(JLr1)reg(J)+1\operatorname*{reg}(J\cap L_{r-1})\geq\operatorname*{reg}(J)+1. ∎

We can now refine Lemma 21:

Proposition 23.

Suppose that JJ is saturated. Then reg(JLc)=reg(J)+1\operatorname*{reg}(J\cap L_{c})=\operatorname*{reg}(J)+1 if and only if reg(J+Lc)=reg(J)\operatorname*{reg}(J+L_{c})=\operatorname*{reg}(J).

Proof.

According to Lemma 21(i) and Proposition 22, the regularity of JLcJ\cap L_{c} can either be reg(J)+1\operatorname*{reg}(J)+1 or reg(J)\operatorname*{reg}(J).

By Lemma 21(iii) we know that reg(JLc)=reg(J)+1\operatorname*{reg}(J\cap L_{c})=\operatorname*{reg}(J)+1 implies reg(J+Lc)=reg(J)\operatorname*{reg}(J+L_{c})=\operatorname*{reg}(J). We argue for the reverse direction. For the sake of a contradiction, suppose

reg(J+Lc)=reg(J)=reg(JLc).\operatorname*{reg}(J+L_{c})=\operatorname*{reg}(J)=\operatorname*{reg}(J\cap L_{c}).

We may assume reg(J+Lc+1)<reg(J)\operatorname*{reg}(J+L_{c+1})<\operatorname*{reg}(J); otherwise we simply replace LcL_{c} with Lc+1L_{c+1}, noting that reg(JLc)=reg(J)\operatorname*{reg}(J\cap L_{c})=\operatorname*{reg}(J) implies reg(JLc+1)=reg(J)\operatorname*{reg}(J\cap L_{c+1})=\operatorname*{reg}(J) by Proposition 22. Set M=S/(J+Lc)M=S/(J+L_{c}). The generic hyperplane section M/xMM/xM of MM can be realized by taking x=hc+1x=h_{c+1}, i.e. M/xM=S/J+Lc+1M/xM=S/J+L_{c+1}. Since xx is almost regular on MM, the kernel 0:Mx0:_{M}x of the multiplication map MxMM\stackrel{{\scriptstyle x}}{{\rightarrow}}M has finite length; thus Proposition 20.20 in [Eis94] gives

reg(M)=max{reg(M/xM),reg(0:Mx)}.\operatorname*{reg}(M)=\max\{\operatorname*{reg}(M/xM),\operatorname*{reg}(0:_{M}x)\}.

In our setting, by hypothesis we have reg(M/xM)<reg(M)\operatorname*{reg}(M/xM)<\operatorname*{reg}(M), so reg(M)=reg(0:Mx).\operatorname*{reg}(M)=\operatorname*{reg}(0:_{M}x). Now the short exact sequence in the proof of Lemma 21(i), together with the assumption that JJ is saturated, implies that H𝔪0(M)H_{\mathfrak{m}}^{0}(M) is a submodule of H𝔪1(S/JLc)H_{\mathfrak{m}}^{1}(S/J\cap L_{c}). Since 0:Mx0:_{M}x is a submodule of H𝔪0(M)H_{\mathfrak{m}}^{0}(M), we have

reg(J)1\displaystyle\operatorname*{reg}(J)-1 =reg(S/JLc)1+reg(H𝔪0(M))\displaystyle=\operatorname*{reg}(S/J\cap L_{c})\geq 1+\operatorname*{reg}(H_{\mathfrak{m}}^{0}(M))
1+reg(0:Mx)=1+reg(M)=reg(J),\displaystyle\geq 1+\operatorname*{reg}(0:_{M}x)=1+\operatorname*{reg}(M)=\operatorname*{reg}(J),

which is a contradiction. ∎

Theorem 3 now immediately follows from Proposition 23 and Lemma 21(iii), together with the remark that if cdepth(S/J)c\leq\operatorname*{depth}(S/J), then LcL_{c} is generated by an S/JS/J-regular sequence and so reg(S/J)=reg(S/J+Lc)\operatorname*{reg}(S/J)=\operatorname*{reg}(S/J+L_{c}).

6. Proof of Theorem 1

For i[n]i\in[n] we let IiI_{i} be generated by ci2c_{i}\geq 2 generic linear forms. We first prove:

Lemma 24.

Suppose that n2r3n\geq 2r-3. Then reg(i[n]Ii)n1\operatorname*{reg}(\cap_{i\in[n]}I_{i})\leq n-1.

Proof.

For each i[n]i\in[n] let IiIiI_{i}^{\prime}\subset I_{i} be an ideal generated by two generic linear forms. By Theorem 2 and Lemma 21(i), we have reg(i[n]Ii)n1\operatorname*{reg}(\cap_{i\in[n]}I_{i}^{\prime})\leq n-1. If ci=2c_{i}=2 for every i[n]i\in[n], we are done. So we may assume cn>2c_{n}>2. We may write In=(h1,h2,,hcn)I_{n}=(h_{1},h_{2},\dots,h_{c_{n}}) and In=(h1,h2)I_{n}^{\prime}=(h_{1},h_{2}). Set J=i[n1]IiJ=\cap_{i\in[n-1]}I_{i}^{\prime}, L2=InL_{2}=I_{n}^{\prime} and Lcn=InL_{c_{n}}=I_{n}. Then Proposition 22 gives

n1reg(i[n]Ii)=reg(JL2)reg(JLcn)=reg((i[n1]Ii)In).n-1\geq\operatorname*{reg}(\cap_{i\in[n]}I_{i}^{\prime})=\operatorname*{reg}(J\cap L_{2})\geq\operatorname*{reg}(J\cap L_{c_{n}})=\operatorname*{reg}((\cap_{i\in[n-1]}I_{i}^{\prime})\cap I_{n}).

Continuing in a similar fashion by inductively applying Proposition 22 to the ideal (i[n1]Ii)In(\cap_{i\in[n-1]}I_{i}^{\prime})\cap I_{n}, 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 Ia𝔪,a𝒜=[]I_{a}\subsetneq\mathfrak{m},\,a\in\mathcal{A}=[\ell] and Jb𝔪,b=[m]J_{b}\subsetneq\mathfrak{m},\,b\in\mathcal{B}=[m] be ideals generated by generic linear forms, where \ell and mm are positive integers. Then

reg((a𝒜Ia)(bJb))reg(a𝒜Ia)+reg(bJb).\operatorname*{reg}\left((\cap_{a\in\mathcal{A}}I_{a})\cap(\cap_{b\in\mathcal{B}}J_{b})\right)\leq\operatorname*{reg}(\cap_{a\in\mathcal{A}}I_{a})+\operatorname*{reg}(\cap_{b\in\mathcal{B}}J_{b}).
Proof.

For convenience, we set I𝒜=a𝒜IaI_{\mathcal{A}}=\cap_{a\in\mathcal{A}}I_{a} and J=bJbJ_{\mathcal{B}}=\cap_{b\in\mathcal{B}}J_{b}. We have inclusions of ideals

(a𝒜Ia)(bIb)I𝒜JI𝒜J.\textstyle(\prod_{a\in\mathcal{A}}I_{a})(\prod_{b\in\mathcal{B}}I_{b})\subset I_{\mathcal{A}}J_{\mathcal{B}}\subset I_{\mathcal{A}}\cap J_{\mathcal{B}}.

By Proposition 3.4 in [CH03], we have

[(a𝒜Ia)(bIb)]sat=I𝒜J,\textstyle\left[(\prod_{a\in\mathcal{A}}I_{a})(\prod_{b\in\mathcal{B}}I_{b})\right]^{\operatorname*{sat}}=I_{\mathcal{A}}\cap J_{\mathcal{B}},

and hence also

(I𝒜J)sat=I𝒜J.\textstyle(I_{\mathcal{A}}J_{\mathcal{B}})^{\operatorname*{sat}}=I_{\mathcal{A}}\cap J_{\mathcal{B}}.

Thus the SS-module

Tor1S(SI𝒜,SJ)=I𝒜JI𝒜J,{\operatorname*{Tor}}_{1}^{S}\left(\frac{S}{I_{\mathcal{A}}},\frac{S}{J_{\mathcal{B}}}\right)=\frac{I_{\mathcal{A}}\cap J_{\mathcal{B}}}{I_{\mathcal{A}}J_{\mathcal{B}}},

is zero-dimensional, and so [Cav07] give

reg(SI𝒜+J)=reg(SI𝒜SSJ)reg(SI𝒜)+reg(SJ).\operatorname*{reg}\left(\frac{S}{I_{\mathcal{A}}+J_{\mathcal{B}}}\right)=\operatorname*{reg}\left(\frac{S}{I_{\mathcal{A}}}\otimes_{S}\frac{S}{J_{\mathcal{B}}}\right)\leq\operatorname*{reg}\left(\frac{S}{I_{\mathcal{A}}}\right)+\operatorname*{reg}\left(\frac{S}{J_{\mathcal{B}}}\right).

With this, the short exact sequence

0SI𝒜JSI𝒜SJSI𝒜+J00\rightarrow\frac{S}{I_{\mathcal{A}}\cap J_{\mathcal{B}}}\rightarrow\frac{S}{I_{\mathcal{A}}}\oplus\frac{S}{J_{\mathcal{B}}}\rightarrow\frac{S}{I_{\mathcal{A}}+J_{\mathcal{B}}}\rightarrow 0

and the regularity lemma (Proposition 4) give the assertion. ∎

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 3\mathbb{P}^{3}, 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.