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

FF-purity and the FF-pure threshold
as invariants of linkage

Vaibhav Pandey Department of Mathematics, Purdue University, 150 N University St., West Lafayette, IN 47907, USA [email protected] Dedicated to Professor Bernd Ulrich on the occasion of his seventieth birthday
Abstract.

The generic link of an unmixed radical ideal is radical (in fact, prime). We show that the squarefreeness of the initial ideal and FF-purity are, however, not preserved along generic links. On the flip side, for several important cases in liaison theory, including generic height three Gorenstein ideals and the maximal minors of a generic matrix, we show that the squarefreeness of the initial ideal, FF-purity, and the FF-pure threshold are each preserved along generic links by identifying a property of such ideals which propagates along generic links. We use this property to establish the FF-regularity of the generic links of such ideals. Finally, we study the FF-pure threshold of the generic residual intersections of a complete intersection ideal and answer a related question of Kim–Miller–Niu.

Key words and phrases:
generic link, FF-pure threshold, FF-pure, initial ideal, generic residual intersection, FF-regular
2010 Mathematics Subject Classification:
Primary 13C40, 13A35, 14M06; Secondary 14M10, 14M12.
The author was partially supported by the NSF–FRG grant DMS-1952366 and the AMS–Simons Travel Grant.

1. Introduction

Given a proper, unmixed ideal II in a polynomial ring RR over a field KK, its geometric link is an ideal JJ such that the ideal IJI\cap J is generated by a regular sequence, that is, the set theoretic union of the vanishing loci of II and JJ is a complete intersection. Clearly, the link of the ideal II depends on the choice of a regular sequence in II. When this regular sequence is chosen in the ‘most general’ manner, that is, as generic combinations of the generators of the ideal II, the link thus obtained, in a polynomial extension of RR, is called the (first) generic link L1(I)L_{1}(I) of II. Evidently, the generic link specializes to any link of II; it is the prototypical link to study since most ‘good properties’ are stable under deformations. Iterating this process, the nn-th generic link of II

Ln(I)\colonequalsL1(Ln1(I))L_{n}(I)\colonequals L_{1}(L_{n-1}(I))

is the generic link of its (n1)(n-1)-th generic link, and analogously, models any ideal which is linked to II in nn steps.

Since the beginning of the modern study of linkage, or liaison theory, by Peskine and Szpiro [PS73] and the foundational work by Huneke and Ulrich [HU85, HU87, HU88], the central research theme has been to understand when two given ideals can be linked in finitely many steps. This calls for studying how the algebraic and geometric properties of an ideal change when one passes to its link; in particular, one wishes to understand which properties remain invariant along links. For example, height and the Cohen-Macaulay property are invariants of linkage, and the canonical module of the link is ‘dual’ to that of R/IR/I [PS73]. While it is easily seen that the link of an unmixed, radical ideal may not be radical, by [HU85, Proposition 2.6], the generic link of an unmixed, radical ideal (even a generically complete intersection ideal) is indeed radical (in fact, prime). It is then natural to search for finer linkage invariants:

Question.

Let II be an unmixed ideal in a polynomial ring RR over a field KK and << a term order in RR. Let L1(I)L_{1}(I) denote the (first) generic link of II in a polynomial extension SS of RR.

  1. (1)

    If the initial ideal in<(I)\operatorname{in}_{<}(I) of II is squarefree, is the initial ideal of the generic link L1(I)L_{1}(I) also squarefree for some term order in SS?

  2. (2)

    Assume that KK has positive characteristic. If R/IR/I is an FF-pure ring, is the generic link S/L1(I)S/L_{1}(I) also FF-pure?

Note that if an ideal has a squarefree initial ideal with respect to some term order or if it defines an FF-pure ring, then it is radical to begin with.

We show that the non-maximal (even sub-maximal) minors of a generic matrix (Example 5.4), as well as the rational normal curve (more generally, the maximal minors of a Hankel matrix) (Example 5.5), answers the above question in the negative. In fact, these examples show that even if the ring R/IR/I is strongly FF-regular, its generic link may not even be FF-injective. This greatly strengthens the fact that generic linkage does not preserve rational singularities [Niu14, Corollary 3.4].

The bulk of this paper, however, deals with understanding under what hypothesis are the answers to the above questions in the affirmative. A major roadblock to our enquiry is that the generators of the generic link are known in very few cases. While a mapping cone construction exists (see [PS73]) to find a free resolution of the link if we know a resolution of the ideal II, it is quite cumbersome to carry out in practice. The mapping cone construction gives the explicit generators of the generic link of well-behaved ideals whose resolutions are short, like a perfect ideal of height 22, a Gorenstein ideal of height 33 ([KU92a, §4.10]), and a complete intersection ideal ([HU88, Example 3.4]).

To get around the issue of not knowing the generators of the generic link, for an unmixed, homogeneous ideal II in a polynomial ring RR, we identify a property 𝐏\mathbf{P} of the pair (R,I)(R,I) (see Definition 3.1) which propagates along generic links (see Lemma 3.3). Crucially, 𝐏\mathbf{P} only depends on the choice of a term order in RR and on the regular sequence defining the link, and not on the generators of the link.

The simple observation that the property 𝐏\mathbf{P} propagates along generic links is quite powerful: We show that if the pair (R,I)(R,I) has the property 𝐏\mathbf{P}, then the squarefreeness of the initial ideal, FF-purity, and the FF-pure threshold/log canonical threshold (a numerical measure of singularity in positive characteristic/birational geometry) are each preserved along the nn-th generic link of II for every n0n\geq 0 (Theorem 3.5). This theorem and each of the items (1)(1), (2)(2), and (3)(3) listed below provide evidence in favor of an affirmative answer to [KMN21, Question 4.5 (a)]. We now highlight the key results which we are able to prove using this point of view:

(1)

We show that the generic link of a generic height 33 Gorenstein ideal is strongly FF-regular in each prime characteristic (rational singularities in characteristic 0) (see Theorem 4.3). We remark that the generators of the generic link of height three Gorenstein ideals are known ([KU92a, §4.10], [KU92b, Theorem 4.7]); however, due to our approach of viewing the generic link via the property 𝐏\mathbf{P}, we do not need the generators of the generic link to establish its FF-regularity (or its FF-pure threshold, or the squarefreeness of its initial ideal). We also show the nn-th universal link of a generic height 33 Gorenstein ideal is FF-rational in each prime characteristic (rational singularities in characteristic 0) for every n>0n>0 (Corollary 4.5).

(2)

In [PT24, Theorem 6.3], the author and Tarasova showed that the generic link of the maximal minors of a generic matrix is strongly FF-regular (note that the generators of the generic link are not known). This result may now be viewed as an offshoot of the fact that the ideal of maximal minors of a generic matrix has the property 𝐏\mathbf{P} (Proposition 5.1). Remarkably, this perspective allows us to recover the FF-pure threshold/log canonical threshold of the maximal minors of a generic matrix calculated in [MSV14, Theorem 1.2], [Doc13, Theorem 5.6] as well as that of the generic link of the maximal minors of a generic matrix calculated in [KMN21, Theorem 1] by a simpler method, and, more importantly, to extend these calculations to the nn-th generic link for each n0n\geq 0 (see Remark 5.2).

(3)

The notion of residual intersections, in its modern form, essentially goes back to Artin and Nagata [AN72] and has broad applications in enumerative geometry, intersection theory, the study of Rees rings, and multiplicity theory ([Ful84, Ulr92, HU88]). Residual intersections are a vast generalization of linkage, where, under appropriate technical hypothesis, the regular sequence in consideration is replaced by an ideal of height ss with sht(I)s\geq\operatorname{ht}(I) (see §6.1 for details).

We calculate the FF-pure threshold/log canonical threshold of the generic ss-residual intersections of a complete intersection ideal having 𝐏\mathbf{P} (Corollary 6.5). Our key insight is that the property 𝐏\mathbf{P} is inherited by the generic ss-residual intersections of a complete intersection ideal for each sht(I)s\geq\operatorname{ht}(I) (Theorem 6.4). In particular, when s=ht(I)s=\operatorname{ht}(I), we answer [KMN21, Question 4.5(b)] of Kim–Miller–Niu by finding a class of complete intersection ideals—those possessing 𝐏\mathbf{P}—for which the FF-pure threshold is invariant under generic linkage. We also note that if the complete intersection ideal does not have 𝐏\mathbf{P}, then its FF-pure threshold may not be preserved under generic linkage (see Remark 6.6).

The paper is organized as follows: In §2, we recall the background of linkage and positive characteristic techniques needed for the paper. In §3, we discuss the property 𝐏\mathbf{P} and its consequences. In §4, we establish the FF-regularity of the generic link of generic height 33 Gorenstein ideals and discuss some immediate corollaries. In §5, we discuss some subtleties related to the property 𝐏\mathbf{P}. Finally, in §6, we recall the notion of residual intersections and study the FF-pure thresholds of the generic residual intersections of a complete intersection ideal.

2. Background

Linkage

While the notion of linkage holds more generally over Cohen–Macaulay rings (and most of the important results hold true over Gorenstein rings), for the purpose of this paper, and for simplicity of exposition, we restrict our attention to the situation where the ambient ring is polynomial (see the references in this subsection for the general statements).

Definition 2.1.

Let RR be a polynomial ring, and let II and JJ be proper RR-ideals. We say that II and JJ are linked (or R/IR/I and R/JR/J are linked) if there exists an RR-ideal 𝔞\mathfrak{a} generated by a regular sequence such that

J=𝔞:IandI=𝔞:J,J=\mathfrak{a}:I\qquad\text{and}\qquad I=\mathfrak{a}:J,

and use the notation I𝔞JI\sim_{\mathfrak{a}}J. Furthermore we say that the link is geometric if we have ht(I+J)ht(I)+1\operatorname{ht}(I+J)\geq\operatorname{ht}(I)+1 (where ht()\operatorname{ht}(-) denotes height).

It is clear that the ideal 𝔞\mathfrak{a} is contained in II and JJ. Note that the associated primes of II and JJ have the same height, that is, the ideals II and JJ are unmixed. Further, the heights of the ideals II, JJ, and 𝔞\mathfrak{a} are equal. Moreover, when the link is geometric, it follows that the ideal 𝔞\mathfrak{a} is the intersection of II and JJ.

The linked ideal is ‘dual’ to the given ideal in a sense made precise by the following foundational result of linkage theory:

Proposition 2.2.

[PS73] Let RR be a polynomial ring, II and JJ be RR-ideals, and 𝔞\mathfrak{a} is an ideal generated by a regular sequence such that I=𝔞:JI=\mathfrak{a}:J and J=𝔞:IJ=\mathfrak{a}:I. Suppose that R/IR/I is a Cohen–Macaulay ring. Then

  1.  (1)

    R/JR/J is a Cohen–Macaulay ring.

  2.  (2)

    If RR is a local ring, ωR/IJ/𝔞\omega_{R/I}\cong J/\mathfrak{a} and ωR/JI/𝔞\omega_{R/J}\cong I/\mathfrak{a}, where ωR/I\omega_{R/I} (respectively ωR/J\omega_{R/J}) denotes the canonical module of R/IR/I (respectively R/JR/J).

  3.  (3)

    If the ideals II and JJ are geometrically linked, then ht(I+J)=ht(I)+1\operatorname{ht}(I+J)=\operatorname{ht}(I)+1 and R/(I+J)R/(I+J) is a Gorenstein ring.

Definition 2.3.

Let RR be a polynomial ring and II an unmixed RR-ideal of height g>0g>0. Let 𝐟\colonequalsf1,,fn\mathbf{f}\colonequals f_{1},\ldots,f_{n} be a generating set of II. Let YY be a g×ng\times n matrix of indeterminates, and let 𝔞\mathfrak{a} be the ideal generated by the entries of the matrix Y[f1fn]TY[f_{1}\dots f_{n}]^{T} (where []T[\quad]^{T} denotes the transpose of the matrix). The ideal

L1(I)=L1(𝐟)\colonequals𝔞R[Y]:IR[Y]L_{1}(I)=L_{1}(\mathbf{f})\colonequals\mathfrak{a}R[Y]:IR[Y]

is a generic link of II.

We point out that, in the above definition, the entries of the matrix Y[f1fn]TY[f_{1}\dots f_{n}]^{T} form an R[Y]R[Y]-regular sequence due to [Hoc73]. A generic link of II is indeed a (geometric) link of II:

Proposition 2.4.

[PS73] Suppose that RR is a polynomial ring and that II is an unmixed ideal of RR of height gg. Let 𝔞\mathfrak{a} be an RR-ideal generated by a length gg regular sequence which is properly contained in the ideal II, and let J=𝔞:IJ=\mathfrak{a}:I, then I𝔞JI\sim_{\mathfrak{a}}J.

While the definition of generic link depends on a choice of a generating set of the ideal, it turns out that any two generic links are equivalent:

Definition 2.5.

Let (R,I)(R,I) and (S,J)(S,J) be pairs where RR and SS are Noetherian rings and IRI\subseteq R, JSJ\subseteq S are ideals. We say (R,I)(R,I) and (S,J)(S,J) are equivalent, and write (R,I)(S,J)(R,I)\equiv(S,J), if there exist finite sets of variables, XX over RR and ZZ over SS, and an isomorphism φ:R[X]S[Z]\varphi:R[X]\rightarrow S[Z] such that φ(IR[X])=JS[Z]\varphi(IR[X])=JS[Z].

Lemma 2.6.

[HU85, Proposition 2.4] Let II be an unmixed ideal in a polynomial ring RR. If J1R[X1]J_{1}\subseteq R[X_{1}] and J2R[X2]J_{2}\subseteq R[X_{2}] are generic links of II, then we have (R[X1],J1)(R[X2],J2)(R[X_{1}],J_{1})\equiv(R[X_{2}],J_{2}).

Due to the above lemma, we freely use the phrase “the generic link” of an ideal in this paper.

Definition 2.7.

Let RR be a polynomial ring and II be an unmixed RR-ideal of height g>0g>0. Let 𝐟\mathbf{f} be a generating set of II. The nn-th generic link of II is the ideal

Ln(I)=L1(Ln1(I))forn>0withL0(I)=IandL1(I)=L1(𝐟).L_{n}(I)=L_{1}(L_{n-1}(I))\quad\text{for}\;n>0\quad\text{with}\quad L_{0}(I)=I\quad\text{and}\quad L_{1}(I)=L_{1}(\mathbf{f}).

Assume that RR is a regular local ring. The first universal link L1(I)L^{1}(I) of II is obtained similarly as the generic link with respect to invertible indeterminates and the nn-th universal link is obtained iteratively:

Ln(I)=L1(Ln1(I))forn>0withL0(I)=IandL1(I)=L1(𝐟).L^{n}(I)=L^{1}(L^{n-1}(I))\quad\text{for}\;n>0\quad\text{with}\quad L^{0}(I)=I\quad\text{and}\quad L^{1}(I)=L^{1}(\mathbf{f}).

The linkage class of an ideal II in a polynomial ring RR is the set of all RR-ideals which can be obtained from II by a finite number of links.

The point of defining the nn-th generic link of II (respectively the nn-th universal link of II) is that the nn-th generic link (respectively the nn-th universal link) is a deformation (respectively, essentially a deformation) 111We recall that a ring AA is essentially a deformation of another ring BB if it is obtained by a finite sequence of deformations and localizations at prime ideals beginning from BB. of any RR-ideal linked to II in nn steps ([HU87, Proposition 2.14, Theorem 2.17]) and therefore controls the algebro-geometric properties of any link of II upto deformation. In fact, under mild assumptions, one can descend from a sequence of universal links to a sequence of links in the original ring and still preserve most of the good properties of universal linkage ([HU88, Lemma 2.1]).

Frobenius splittings and the FF-pure threshold

For a reduced Noetherian ring RR of prime characteristic p>0p>0, the (ee-fold) Frobenius endomorphism on RR is the map Fe:RRF^{e}:R\to R with Fe(r)=rpeF^{e}(r)=r^{p^{e}}. To avoid any confusion between the the domain and codomain, we instead use the notation Fe(R)F^{e}_{*}(R) for the codomain and Fe(r)F^{e}_{*}(r) for its elements. So Fe(R)F^{e}_{*}(R) is the same ring as RR with the RR-module structure obtained from the Frobenius map:

r.Fe(s)\colonequalsFe(rpes).r.F^{e}_{*}(s)\colonequals F^{e}_{*}(r^{p^{e}}s).

The ring RR is FF-finite if Fe(R)F^{e}_{*}(R) is a finite RR-module for some (equivalently, every) e>0e>0. A finitely generated algebra RR over a field KK is FF-finite if and only if F(K)F_{*}(K) is a finite field extension of KK—a fairly mild condition. The ring RR is FF-pure if the Frobenius endomorphism F:RF(R)F:R\to F_{*}(R) with F(r)=F(rp)F(r)=F_{*}(r^{p}) is pure, that is, for any RR-module MM, the map F1:RRMF(R)RMF\otimes 1:R\otimes_{R}M\to F_{*}(R)\otimes_{R}M is injective. Note that if RR is an FF-pure ring, the Frobenius map F:RFe(R)F:R\to F^{e}_{*}(R) splits as a map of RR-modules so that the RR-module F(R)F_{*}(R) admits a free RR-summand. An FF-finite ring RR is strongly FF-regular if it has a sufficiently large number of Frobenius splittings, made precise as follows: For every element cRc\in R^{\circ}, there exists an integer q=peq=p^{e} such that the RR-linear map RFe(R)R\longrightarrow F^{e}_{*}(R) sending 11 to Fe(c)F^{e}_{*}(c) splits as a map of RR-modules. Clearly, a strongly FF-regular ring is FF-pure.

The rings under consideration in this paper are \mathbb{N}-graded; since the various competing notions of FF-regularity (like weakly FF-regular, FF-regular, and strongly FF-regular) coincide in the graded case due to [LS99, Corollary 4.3], we make no distinction between them throughout the paper.

The following result will help us in showing that several ideals considered in this paper define FF-pure rings.

Lemma 2.8.

[PT24, Lemma 3.1, Corollary 3.3] Let R=K[x1,,xn]R=K[x_{1},\ldots,x_{n}] be a polynomial ring over the field KK of characteristic p>0p>0 and let II be an unmixed homogeneous RR-ideal. If 𝔞I{\mathfrak{a}}\subsetneq I is an ideal generated by a regular sequence of length equal to the height of II and J=𝔞:IJ={\mathfrak{a}}:I, then

𝔞[p]:𝔞(I[p]:I)(J[p]:J).{\mathfrak{a}}^{[p]}:{\mathfrak{a}}\subseteq(I^{[p]}:I)\cap(J^{[p]}:J).

In particular, if the ring S/𝔞S/{\mathfrak{a}} is FF-pure, then so are S/IS/I and S/JS/J.

Definition 2.9.

Let RR be a polynomial ring over a field of characteristic p>0p>0 and let 𝔪{\mathfrak{m}} denote its homogeneous maximal ideal. For a homogeneous proper ideal II and integer e>0e>0, set

νI(pe)=max{r|Ir𝔪[pe]},\nu_{I}(p^{e})=\max\{r\in\mathbb{N}\;|\;I^{r}\nsubseteq{\mathfrak{m}}^{[p^{e}]}\},

where 𝔪[pe]=(ape|a𝔪){\mathfrak{m}}^{[p^{e}]}=(a^{p^{e}}\;|\;a\in{\mathfrak{m}}). If II is generated by NN elements, it is readily seen that

0νI(pe)N(pe1).0\leq\nu_{I}(p^{e})\leq N(p^{e}-1).

Moreover, if fIr𝔪[pe]f\in I^{r}\setminus{\mathfrak{m}}^{[p^{e}]}, then fpIpr𝔪[pe+1]f^{p}\in I^{pr}\setminus{\mathfrak{m}}^{[p^{e+1}]}. Thus,

νI(pe+1)pνI(pe).\nu_{I}(p^{e+1})\geq p\nu_{I}(p^{e}).

It follows that the sequence of real numbers {νI(pe)/pe}e>0\left\{\nu_{I}(p^{e})/p^{e}\right\}_{e>0} is non-decreasing and bounded above; its limit is the FF-pure threshold of II, denoted fpt(I)\operatorname{fpt}(I).

The notion of FF-pure thresholds is due to Takagi and Watanabe [TW04]. We point out that the FF-pure threshold may be defined in a more general setup (see [MTW05, DSNnBP18, DSNnB18]), however the above definition is adequate for this paper. The FF-pure threshold is the positive characteristic analog of the log canonical threshold: a numerical measure of singularity in birational geometry. For simplicity of exposition, let II be a homogeneous ideal in a polynomial ring over the rational numbers. Using “II modulo pp” to denote the characteristic pp model, one has the inequality

fpt(I modulo p)lct(I)for allp0,\operatorname{fpt}(I\,\text{ modulo }\,p)\leq\operatorname{lct}(I)\quad\text{for all}\;p\gg 0,

where lct(I)\operatorname{lct}(I) denotes the log canonical threshold of II. Moreover,

limpfpt(I modulo p)=lct(I).\lim_{p\to\infty}\operatorname{fpt}(I\,\text{ modulo }\,p)=\operatorname{lct}(I).

These facts follow from the work of Hara and Yoshida [HY03]; see [MTW05, Theorems 3.3, 3.4]. The following result is well-known to experts; we include a proof for the convenience of the reader.

Lemma 2.10.

The FF-pure threshold of a homogeneous ideal II in a polynomial ring RR is bounded above by its height. Furthermore, if the homogeneous ideal is unmixed and radical, then

fpt(I)=ht(I)\operatorname{fpt}(I)=\operatorname{ht}(I)

implies that the ring R/IR/I is FF-pure.

Proof.

Let 𝔭{\mathfrak{p}} be a prime ideal in the polynomial ring R=K[x¯]R=K[\underline{x}], where the field KK has characteristic p>0p>0 and 𝔪{\mathfrak{m}} is the homogeneous maximal ideal of RR. Then R𝔭R_{{\mathfrak{p}}} is a regular local ring of dimension ht(𝔭)\operatorname{ht}({\mathfrak{p}}). For each e>0e>0, the pigeonhole principle gives

𝔭𝔭ht𝔭(pe1)+1𝔭𝔭[pe].{\mathfrak{p}}_{{\mathfrak{p}}}^{\operatorname{ht}{{\mathfrak{p}}}(p^{e}-1)+1}\subseteq{\mathfrak{p}}_{{\mathfrak{p}}}^{[p^{e}]}.

Contracting back to RR, by the flatness of the Frobenius map, we get

𝔭(htp(pe1)+1)𝔭[pe].{\mathfrak{p}}^{(\operatorname{ht}{p}(p^{e}-1)+1)}\subseteq{\mathfrak{p}}^{[p^{e}]}.

Therefore

𝔭r𝔭(r)𝔭[pe]𝔪[pe]forr>ht𝔭(pe1).{\mathfrak{p}}^{r}\subseteq{\mathfrak{p}}^{(r)}\subseteq{\mathfrak{p}}^{[p^{e}]}\subseteq{\mathfrak{m}}^{[p^{e}]}\quad\text{for}\quad r>\operatorname{ht}{{\mathfrak{p}}}(p^{e}-1).

In particular, let 𝔭{\mathfrak{p}} be a minimal prime of the ideal II of the same height. We get

fpt(I)fpt(𝔭)limeht(I)(pe1)pe=ht(I).\operatorname{fpt}(I)\leq\operatorname{fpt}({\mathfrak{p}})\leq\lim_{e\to\infty}\frac{\operatorname{ht}(I)(p^{e}-1)}{p^{e}}=\operatorname{ht}(I).

The last assertion follows from [Tak04, Theorem 3.11] since the hypothesis

fpt(I)=ht(I)\operatorname{fpt}(I)=\operatorname{ht}(I)

implies that the pair (R,Iht(I))(R,I^{\operatorname{ht}(I)}) is FF-pure (see [Tak04, §3] for the notion of FF-purity of pairs). ∎

In the next section, we will calculate the FF-pure thresholds (and log canonical thresholds) of certain geometric links. We point out out that these calculations are independent of the choice of a generating set of the ideal essentially due to lemma 2.6 (alternatively, see [KMN21, Lemma 2.8]).

The next result will be useful in showing that several ideals considered in this paper have a squarefree initial ideal.

Theorem 2.11.

[KV23, Theorem 3.13] Let RR be a polynomial ring over a field. Let II be a radical ideal and << a term order in RR. Let gg be the maximum of the heights of the associated prime ideals of II.

If the initial ideal in<(I(g))\operatorname{in}_{<}(I^{(g)}) contains a squarefree monomial, then in<(I)\operatorname{in}_{<}(I) is a squarefree monomial ideal.

3. The property 𝐏\mathbf{P}

Definition 3.1.

Let RR be a polynomial ring over a field and II a homogeneous RR-ideal. Let 𝐏\mathbf{P} be the following property of the pair (R,I)(R,I): For a fixed term order << in RR, there exist homogeneous elements α¯\colonequalsα1,,αht(I)\underline{\alpha}\colonequals\alpha_{1},\ldots,\alpha_{\operatorname{ht}(I)} in the ideal II such that each initial term in<(αi)\operatorname{in}_{<}(\alpha_{i}) is squarefree and each pair of initial terms in<(αi),in<(αj)\operatorname{in}_{<}(\alpha_{i}),\operatorname{in}_{<}(\alpha_{j}) is mutually coprime for iji\neq j.

Remark 3.2.

Since each pair of initial terms in<(αi),in<(αj)\operatorname{in}_{<}(\alpha_{i}),\operatorname{in}_{<}(\alpha_{j}) is mutually coprime, the monomials in<(α1),,in<(αht(I))\operatorname{in}_{<}(\alpha_{1}),\ldots,\operatorname{in}_{<}(\alpha_{\operatorname{ht}(I)}) form an RR-regular sequence. Therefore the polynomials α¯=α1,,αht(I)\underline{\alpha}=\alpha_{1},\ldots,\alpha_{\operatorname{ht}(I)} also form an RR-regular sequence (see, for example, [BCRV22, Proposition 1.2.12]). Since II is homogeneous, as is (α¯)(\underline{\alpha}), we note that (α¯):I(\underline{\alpha}):I is also a homogeneous RR-ideal.

An important insight of this paper is that 𝐏\mathbf{P} propagates along generic links:

Lemma 3.3.

Let II be an unmixed homogeneous ideal in a polynomial ring R=K[x¯]R=K[\underline{x}] over the field KK. Assume that the pair (R,I)(R,I) has the property 𝐏\mathbf{P}.

  1.  (1)

    Let Ln(I)L_{n}(I) denote the nn-th generic link of II in a polynomial extension R[Y]R[Y]. For any integer n0n\geq 0, the pair (R[Y],Ln(I))(R[Y],L_{n}(I)) has 𝐏\mathbf{P}.

  2.  (2)

    If the ideal II is not generated by a regular sequence, then (α¯):I(\underline{\alpha}):I is a link of II in the polynomial ring RR. The pair (R,(α¯):I)(R,(\underline{\alpha}):I) has 𝐏\mathbf{P}.

Proof.

We first prove item (1)(1). Assume that the ideal II has height g>0g>0 and proceed by induction on nn. Since the nn-th generic link of II is the generic link of the (n1)(n-1)-th generic link, i.e.,

Ln(I)=L1(Ln1(I)),\ L_{n}(I)=L_{1}(L_{n-1}(I)),

it suffices to prove the assertion for n=1n=1. Note that the generic link is a homogeneous ideal (in some polynomial extension of RR) for each n>0n>0.

Let the elements α1,,αg\alpha_{1},\ldots,\alpha_{g} inside the ideal II be as given by 𝐏\mathbf{P}. Extend it to find generators α1,,αg,αg+1,,αr\alpha_{1},\ldots,\alpha_{g},\alpha_{g+1},\ldots,\alpha_{r} of II and fix this generating set. We proceed to find homogeneous elements β¯\colonequalsβ1,,βg\underline{\beta}\colonequals\beta_{1},\ldots,\beta_{g} inside the ideal L1(I)L_{1}(I) and a term order <1<_{1} in the polynomial ring R[Y]R[Y], where Y\colonequals(Yi,j)Y\colonequals(Y_{i,j}) is a g×rg\times r matrix of indeterminates, such that the pair (R[Y],L1(I))(R[Y],L_{1}(I)) has 𝐏\mathbf{P}. Recall that

L1(I)=Y[α1αgαg+1αr]TR[Y]:(α1,,αr)R[Y].L_{1}(I)=Y[\alpha_{1}\cdots\alpha_{g}\,\alpha_{g+1}\cdots\alpha_{r}]^{T}R[Y]:(\alpha_{1},\ldots,\alpha_{r})R[Y].

Define the following variable order in R[Y]R[Y]:

Y1,1>Y2,2>>Yg,g> the remaining Yi,j>xk,Y_{1,1}>Y_{2,2}>\cdots>Y_{g,g}>\text{ the remaining }Y_{i,j}>x_{k},

where the order << on the indeterminates xkx_{k} is the one given in the property 𝐏\mathbf{P} and the order on “the remaining Yi,jY_{i,j}” is arbitrary. Consider the lexicographical order <1<_{1} induced by this variable order in R[Y]R[Y].

Let β1,βg\beta_{1},\ldots\beta_{g} denote the entries of the matrix Y[α1αr]TY[\alpha_{1}\cdots\alpha_{r}]^{T}. Then,

in<1(β1)\displaystyle\operatorname{in}_{<_{1}}(\beta_{1}) =in<1(Y1,1α1+Y1,2α2++Y1,rαr)=in<1(Y1,1α1)=Y1,1in<(α1),\displaystyle=\operatorname{in}_{<_{1}}(Y_{1,1}\alpha_{1}+Y_{1,2}\alpha_{2}+\ldots+Y_{1,r}\alpha_{r})=\operatorname{in}_{<_{1}}(Y_{1,1}\alpha_{1})=Y_{1,1}\operatorname{in}_{<}(\alpha_{1}),
in<1(β2)\displaystyle\operatorname{in}_{<_{1}}(\beta_{2}) =in<1(Y2,1α1+Y2,2α2++Y2,rαr)=in<1(Y2,2α2)=Y2,2in<(α2),\displaystyle=\operatorname{in}_{<_{1}}(Y_{2,1}\alpha_{1}+Y_{2,2}\alpha_{2}+\ldots+Y_{2,r}\alpha_{r})=\operatorname{in}_{<_{1}}(Y_{2,2}\alpha_{2})=Y_{2,2}\operatorname{in}_{<}(\alpha_{2}),
\displaystyle\vdots
in<1(βg)\displaystyle\operatorname{in}_{<_{1}}(\beta_{g}) =in<1(Yg,1α1+Yg,2α2++Yg,rαr)=in<1(Yg,gαg)=Yg,gin<(αg).\displaystyle=\operatorname{in}_{<_{1}}(Y_{g,1}\alpha_{1}+Y_{g,2}\alpha_{2}+\ldots+Y_{g,r}\alpha_{r})=\operatorname{in}_{<_{1}}(Y_{g,g}\alpha_{g})=Y_{g,g}\operatorname{in}_{<}(\alpha_{g}).

Since the pair (R,I)(R,I) has 𝐏\mathbf{P}, we know that for the elements α¯=α1,,αg\underline{\alpha}=\alpha_{1},\ldots,\alpha_{g}, each initial term in<(αi)\operatorname{in}_{<}(\alpha_{i}) is squarefree and each pair of initial terms in<(αi),in<(αj)\operatorname{in}_{<}(\alpha_{i}),\operatorname{in}_{<}(\alpha_{j}) is mutually coprime for iji\neq j. It immediately follows from the above display that the same assertions also hold true for the elements β¯=β1,,βg\underline{\beta}=\beta_{1},\ldots,\beta_{g} of R[Y]R[Y].

We now prove item (2)(2). Since II is not generated by a regular sequence, the proper (homogeneous) RR-ideal α¯:I\underline{\alpha}:I is a link of the unmixed ideal II by Proposition 2.4. The pair (R,(α¯):I)(R,(\underline{\alpha}):I) has 𝐏\mathbf{P} because

α¯(α¯):I.\underline{\alpha}\subseteq(\underline{\alpha}):I.\qed
Remark 3.4.

Note that the above lemma also holds true if the phrase “each initial term in<(αi)\operatorname{in}_{<}(\alpha_{i}) is squarefree” is deleted from the property 𝐏\mathbf{P} or if the sequence of elements α¯\underline{\alpha} is shorter. However, the applications that we have in mind require 𝐏\mathbf{P} to be as stated.

The next result says that if the pair (R,I)(R,I) has 𝐏\mathbf{P}, then important properties of the ideal II like the initial ideal being squarefree, FF-purity, and the FF-pure threshold remain invariant under linkage. In fact, we get that the FF-pure threshold (or the log canonical threshold) of each generic link of II, and hence that of II itself, attains its maximal value. That is, the FF-pure threshold of the nn-th generic link is equal to the height—an invariant of linkage!

Theorem 3.5.

Let RR be a polynomial ring over a field KK. Fix an integer n0n\geq 0 and let II be an unmixed homogeneous RR-ideal. Assume that the pair (R,I)(R,I) has the property 𝐏\mathbf{P}.

  1.  (1)

    If the field KK has positive characteristic, then

    fpt(Ln(I))=ht(I).\operatorname{fpt}(L_{n}(I))=\operatorname{ht}(I).

    In particular, the FF-pure threshold of II is the height of II. Furthermore, the ideal Ln(I)L_{n}(I) defines an FF-pure ring (in a polynomial extension of RR).

  2.  (2)

    If the field KK has characteristic zero, then

    lct(Ln(I))=ht(I).\operatorname{lct}(L_{n}(I))=\operatorname{ht}(I).

    In particular, the log canonical threshold of II is the height of II. Furthermore, the ideal Ln(I)L_{n}(I) defines a log canonical singularity (in a polynomial extension of RR)

  3.  (3)

    The initial ideal of the nn-th generic link Ln(I)L_{n}(I) is squarefree. In particular, the initial ideal of II is squarefree.

Proof.

Assume throughout that the ideal II has height g>0g>0. Recall that since Ln(I)L_{n}(I) is linked to the ideal II, it also has height gg. Since the pair (R,I)(R,I) has the property 𝐏\mathbf{P}, by Lemma 3.3, the pair (R[Y],Ln(I))(R[Y],L_{n}(I)) also has 𝐏\mathbf{P}. Let the elements α¯=α1,,αg\underline{\alpha}=\alpha_{1},\ldots,\alpha_{g} in Ln(I)L_{n}(I) and the term order << in R[Y]R[Y] be as stated in 𝐏\mathbf{P}.

Note that item (2)(2) immediately follows from item (1)(1) since FF-pure rings in characteristic p>0p>0 are log canonical in characteristic zero ([HW02, Theorem 3.9]) and the log canonical threshold of an ideal is the limit of the FF-pure thresholds of its characteristic p>0p>0 models. In view of this, assume that KK has characteristic p>0p>0. We begin by proving item (1)(1).

Since (α)¯Ln(I)\underline{(\alpha)}\subseteq L_{n}(I) and, by Lemma 2.10, the FF-pure threshold of an ideal is bounded above by its height, we get

fpt((α)¯)fpt(Ln(I))g.\operatorname{fpt}(\underline{(\alpha)})\leq\operatorname{fpt}(L_{n}(I))\leq g.

We claim that fpt((α)¯)=g\operatorname{fpt}(\underline{(\alpha)})=g. This would give us

g=fpt((α)¯)fpt(Ln(I))g,g=\operatorname{fpt}(\underline{(\alpha)})\leq\operatorname{fpt}(L_{n}(I))\leq g,

so that we have equality throughout. We now show that fpt((α)¯)=g\operatorname{fpt}(\underline{(\alpha)})=g.

The initial term of

f\colonequalsα1αg(α)¯gf\colonequals\alpha_{1}\ldots\alpha_{g}\in\underline{(\alpha)}^{g}

is squarefree as the pair (R[Y],Ln(I))(R[Y],L_{n}(I)) has 𝐏\mathbf{P}. For any e>0e>0, 𝔪[pe]{\mathfrak{m}}^{[p^{e}]} is a monomial ideal. So, we have

(1) fpe1=(α1αg)pe1((α¯[pe]):(α)¯)𝔪[pe].f^{p^{e}-1}=(\alpha_{1}\ldots\alpha_{g})^{p^{e}-1}\in((\underline{\alpha}^{[p^{e}]}):\underline{(\alpha)})\smallsetminus{\mathfrak{m}}^{[p^{e}]}.

This gives

(α¯)g(pe1)𝔪[pe],while(α¯)g(pe1)+1𝔪[pe](\underline{\alpha})^{g(p^{e}-1)}\nsubseteq{\mathfrak{m}}^{[p^{e}]},\quad\text{while}\quad(\underline{\alpha})^{g(p^{e}-1)+1}\subseteq{\mathfrak{m}}^{[p^{e}]}

by the pigeonhole principle. We conclude that

ν(α¯)(pe)=max{r|(α¯)r𝔪[pe]}=g(pe1)\nu_{({\underline{\alpha}})}(p^{e})=\max\{r\;|\;(\underline{\alpha})^{r}\nsubseteq{\mathfrak{m}}^{[p^{e}]}\}=g(p^{e}-1)

for each e>0e>0. Therefore

fpt(α¯)=g,\operatorname{fpt}(\underline{\alpha})=g,

as required; we thus have

fpt(Ln(I))=ht(I).\operatorname{fpt}(L_{n}(I))=\operatorname{ht}(I).

The FF-purity of the ring R[Y]/Ln(I)R[Y]/L_{n}(I) follows from Lemma 2.10; alternatively, it also follows from Equation 1 in view of Lemma 2.8.

We now prove item (3)(3). Since the nn-th generic link R[Y]/Ln(I)R[Y]/L_{n}(I) is FF-pure, we get that Ln(I)L_{n}(I) is a radical ideal. Consider the polynomial

f\colonequalsα1αgLn(I)g.f\colonequals\alpha_{1}\ldots\alpha_{g}\in L_{n}(I)^{g}.

Then the initial term

in<(f)=i=1gin<(αi)in<((Ln(I)g)in<((Ln(I)(g))\operatorname{in}_{<}(f)=\prod_{i=1}^{g}\operatorname{in}_{<}(\alpha_{i})\in\operatorname{in}_{<}((L_{n}(I)^{g})\subseteq\operatorname{in}_{<}((L_{n}(I)^{(g)})

and is squarefree, as noted above. All the requirements of Theorem 2.11 are met; we conclude that Ln(I)L_{n}(I) has a squarefree initial ideal. ∎

Remark 3.6.

The reader may note that the conclusions of Theorem 3.5 hold for any pair (R,I)(R,I) which has 𝐏\mathbf{P}. This will be used in §6, where we show that the generic residual intersections of a complete intersection ideal inherit 𝐏\mathbf{P}.

We specialize to a link of II in the ambient polynomial ring for which the conclusions of Theorem 3.5 hold:

Corollary 3.7.

Let RR be a polynomial ring over a field KK and II an unmixed homogeneous RR-ideal which is not generated by a regular sequence. Assume that the pair (R,I)(R,I) has the property 𝐏\mathbf{P}. Then the ideal (α¯):I(\underline{\alpha}):I is a link of II. Moreover,

  1.  (1)

    If the field KK has positive characteristic, the ideal (α¯):I(\underline{\alpha}):I defines an FF-pure ring and

    fpt((α¯):I)=ht(I).\operatorname{fpt}((\underline{\alpha}):I)=\operatorname{ht}(I).
  2.  (2)

    If the field KK has characteristic zero, the ideal (α¯):I(\underline{\alpha}):I defines a log-terminal singularity and

    lct((α¯):I)=ht(I).\operatorname{lct}((\underline{\alpha}):I)=\operatorname{ht}(I).
  3.  (3)

    The initial ideal of (α¯):I(\underline{\alpha}):I is squarefree.

Proof.

Since II is not generated by a regular sequence, (α¯):I(\underline{\alpha}):I is a proper RR-ideal; it is a link of II by Remark 3.2. By the second assertion of Lemma 3.3, the pair (R,(α¯):I)(R,(\underline{\alpha}):I) has 𝐏\mathbf{P}. The arguments for each of the assertions are precisely the same as those of Theorem 3.5. ∎

In the next two sections, we list a number of examples, crucial in the study of linkage, of pairs (R,I)(R,I) which have the property 𝐏\mathbf{P}. We remind the reader that each of these examples satisfy the conclusions of Theorem 3.5 and Corollary 3.7.

In [KMN21, Question 4.5 (a)], Kim–Miller–Niu ask if the equality of the log canonical thresholds of a variety XX and that of its generic link implies the equality of the log canonical threshold of XX and those of the higher generic links of XX. In view of Theorem 3.5, the examples listed below provide evidence in favor of an affirmative answer: For each pair (R,I)(R,I) in the examples below, the log canonical threshold of the variety and that of its nn-th generic link are equal for each n>0n>0.

Remark 3.8.

In each of the examples discussed in the next two sections, the elements α¯\underline{\alpha} discussed in the property 𝐏\mathbf{P} form a part of a minimal generating set of the ideal of interest. Hence the nn-th (homogeneous) generic link that we get in each case is actually the nn-th minimal (homogeneous) generic link. Minimal linkage is generally preferable over arbitrary linkage: The minimal link of an almost complete intersection ideal defines a Gorenstein ring. For other favorable properties of minimal links, see [HU87, Remark 2.7].

4. FF-regularity of the generic link of generic height 33 Gorenstein ideals

Let II be a height 33 ideal in a polynomial ring RR such that R/IR/I is a Gorenstein ring. By the Buchsbaum–Eisenbud structure theorem [BE77], the generators of II are given by the size 2n2n pfaffians of a (2n+1)×(2n+1)(2n+1)\times(2n+1) alternating matrix. If the entries of this matrix XX are indeterminates, then I\colonequalsPf2n(X)I\colonequals\operatorname{Pf}_{2n}(X) is said to be a ‘generic’ height 33 Gorenstein ideal in the ring R=K[X]R=K[X]. Note that any other such ideal can be obtained by appropriately specializing the entries of XX.

Proposition 4.1.

The pair (K[X],Pf2n(X))(K[X],\operatorname{Pf}_{2n}(X)) has 𝐏\mathbf{P}.

Proof.

Let

α¯\colonequals[1,2,,2n],[1,2,,n,n+2,,2n+1],[2,3,,2n+1]\underline{\alpha}\colonequals[1,2,\ldots,2n],[1,2,\ldots,n,n+2,\ldots,2n+1],[2,3,\ldots,2n+1]

denote the pfaffians of XX with the given rows and columns and fix the lexicographical order in K[X]K[X] induced by the variable order

x1,2n+1>x1,2n>>x1,2>x2,2n+1>>x2,3>x3,2n+1>>x2n,2n+1x_{1,2n+1}>x_{1,2n}>\cdots>x_{1,2}>x_{2,2n+1}>\cdots>x_{2,3}>x_{3,2n+1}>\cdots>x_{2n,2n+1}

which moves from left to right along any row of the matrix XX and then vertically downwards. The initial terms

in<([1,2,,2n])\displaystyle\operatorname{in}_{<}([1,2,\ldots,2n]) =i=1nxi,2n+1i,\displaystyle=\prod_{i=1}^{n}x_{i,2n+1-i},
in<([1,2,,n,n+2,,2n+1])\displaystyle\operatorname{in}_{<}([1,2,\ldots,n,n+2,\ldots,2n+1]) =i=1nxi,2n+2i,\displaystyle=\prod_{i=1}^{n}x_{i,2n+2-i},
in<([2,3,,2n+1])\displaystyle\operatorname{in}_{<}([2,3,\ldots,2n+1]) =i=2n+1xi,2n+3i\displaystyle=\prod_{i=2}^{n+1}x_{i,2n+3-i}

are precisely as required for the property 𝐏\mathbf{P} to hold. Since the ideal Pf2n(X)\operatorname{Pf}_{2n}(X) has height 33, we are done. ∎

Corollary 4.2.

Let II be a generic height 33 Gorenstein ideal in a polynomial ring and n0n\geq 0. Then,

fpt(Ln(I))=3.\operatorname{fpt}(L_{n}(I))=3.

In particular, fpt(I)=3\operatorname{fpt}(I)=3. Furthermore, the initial ideal of Ln(I)L_{n}(I) is squarefree.

Proof.

This is immediate from Proposition 4.1 in view of Theorem 3.5. ∎

Proposition 4.1 has a remarkable consequence: The generic link of a generic height 33 Gorenstein ideal defines an FF-regular ring.

Theorem 4.3.

Let X\colonequals(xi,j)X\colonequals(x_{i,j}) be a (2n+1)×(2n+1)(2n+1)\times(2n+1) alternating matrix of indeterminates, KK a field, and R\colonequalsK[X]R\colonequals K[X]. Let Pf2n(X)\operatorname{Pf}_{2n}(X) denote the ideal generated by the size 2n2n pfaffians of XX.

  1. (1)

    If KK is an FF-finite field of characteristic p>0p>0, the generic link of R/Pf2n(X)R/\operatorname{Pf}_{2n}(X) is FF-regular.

  2. (2)

    If KK has characteristic 0, the generic link of R/Pf2n(X)R/\operatorname{Pf}_{2n}(X) has rational singularities.

Proof.

Assertion (2)(2) follows from (1)(1) since FF-regular rings are FF-rational and FF-rational rings have rational singularities by [Smi97, Theorem 4.3]. We therefore concentrate on the case where the characteristic of KK is positive.

Let Y\colonequals(Yi,j)Y\colonequals(Y_{i,j}) be a generic 3×(2n+1)3\times(2n+1) matrix of indeterminates and S\colonequalsR[Y]S\colonequals R[Y]. The proof will proceed by induction on nn. Note that if n=1n=1, the ideal Pf2n(X)=Pf2(X)\operatorname{Pf}_{2n}(X)=\operatorname{Pf}_{2}(X) is the homogeneous maximal ideal of a polynomial ring in 33 variables. So the FF-regularity of the generic link is a special case of [PT24, Theorem 5.1]. For the remainder of the proof, assume n2n\geq 2. Let Δ1,,Δ2n+1\Delta_{1},\ldots,\Delta_{2n+1} denote the pfaffians of XX with

Δ2n=[1,3,,2n+1]andΔ2n+1=[2,3,,2n+1].\Delta_{2n}=[1,3,\ldots,2n+1]\quad\text{and}\quad\Delta_{2n+1}=[2,3,\ldots,2n+1].

We have,

J=Y[Δ1,,Δ2n+1]TS:(Δ1,,Δ2n+1)S,J=Y[\Delta_{1},\ldots,\Delta_{2n+1}]^{T}S:(\Delta_{1},\ldots,\Delta_{2n+1})S,

and S/JS/J is the generic link of R/Pf2n(X)R/\operatorname{Pf}_{2n}(X).

We first show that the ring (S/J)x1,2(S/J)_{x_{1,2}} is FF-regular. To prove the inductive step n>1n>1, we show that the localization (S/J)x1,2(S/J)_{x_{1,2}} is a faithfully flat extension of the generic link S/JS^{\prime}/J^{\prime} of the size 2n22n-2 pfaffians of a generic alternating (2n1)×(2n1)(2n-1)\times(2n-1) matrix XX^{\prime}, with S=K[X]S^{\prime}=K[X^{\prime}].

Let αi\alpha_{i} denote the ii-th entry of the matrix Y[Δ1,,Δ2n+1]TY[\Delta_{1},\ldots,\Delta_{2n+1}]^{T} for 1i2n+11\leq i\leq 2n+1; set αi\colonequalsαi/x1,2\alpha_{i}^{\prime}\colonequals\alpha_{i}/x_{1,2} for i=1,2,3i=1,2,3 and Δj\colonequalsΔj/x1,2\Delta_{j}^{\prime}\colonequals\Delta_{j}/x_{1,2} for j=1,,2n+1j=1,\ldots,2n+1. Consider the linear change of coordinates,

xi2,j2\colonequalsxi,j+x1,jx2,ix1,ix2,jx1,2i,j=3,,2n+1x^{\prime}_{i-2,j-2}\colonequals x_{i,j}+\frac{x_{1,j}x_{2,i}-x_{1,i}x_{2,j}}{x_{1,2}}\qquad i,j=3,\ldots,2n+1

and set X\colonequals(xi,j)X^{\prime}\colonequals(x^{\prime}_{i,j}), a generic alternating (2n1)×(2n1)(2n-1)\times(2n-1) matrix. By [Bar95, Lemma 1.3], there exists an invertible matrix EE, with entries in Rx1,2R_{x_{1,2}}, such that

ETXE=(0110000x1,2x1,2n1x1,20x2,2n1x1,2n1x2,2n10).E^{T}XE=\left(\begin{array}[]{@{}c|c@{}}\begin{matrix}0&1\\ -1&0\end{matrix}&\mbox{\Large 0}\\ \hline\cr\mbox{\Large 0}&\begin{matrix}0&x^{\prime}_{1,2}&\cdots&x^{\prime}_{1,2n-1}\\ -x^{\prime}_{1,2}&0&\cdots&x^{\prime}_{2,2n-1}\\ \vdots&\vdots&\ddots&\vdots\\ -x^{\prime}_{1,2n-1}&-x^{\prime}_{2,2n-1}&\cdots&0\end{matrix}\end{array}\right).

A routine, albeit tedious, calculation gives

αi=Yi,1Δ1++Yi,2n1Δ2n1i=1,2,3,\alpha_{i}^{\prime}=Y_{i,1}^{\prime}\Delta_{1}^{\prime}+\ldots+Y_{i,2n-1}^{\prime}\Delta_{2n-1}^{\prime}\qquad i=1,2,3,

where Δ1,,Δ2n1\Delta_{1}^{\prime},\ldots,\Delta_{2n-1}^{\prime} are the size 2n22n-2 pfaffians of XX^{\prime} and the elements Yi,jY^{\prime}_{i,j} are algebraically independent over the field KK; note that

Pf2n(X)=Pf2n(ETXE)=Pf2n2(X).\operatorname{Pf}_{2n}(X)=\operatorname{Pf}_{2n}(E^{T}XE)=\operatorname{Pf}_{2n-2}(X^{\prime}).

Let Y\colonequals(Yi,j)Y^{\prime}\colonequals(Y^{\prime}_{i,j}) be a 3×(2n1)3\times(2n-1) (generic) matrix; set R\colonequalsK[X]R^{\prime}\colonequals K[X^{\prime}], S\colonequalsR[Y]S^{\prime}\colonequals R[Y^{\prime}], and

J\colonequalsY[Δ1,,Δ2n1]S:(Δ1,,Δ2n1)S.J^{\prime}\colonequals Y^{\prime}[\Delta_{1}^{\prime},\ldots,\Delta_{2n-1}^{\prime}]S^{\prime}:(\Delta_{1}^{\prime},\ldots,\Delta_{2n-1}^{\prime})S^{\prime}.

Then we get an isomorphism

(S/J)x1,2S/J[x1,3,,x1,2n+1,x2,3,,x2,2n+1,Yi,j|i=1,2,3;j=2n,2n+1],(S/J)_{x_{1,2}}\cong S^{\prime}/J^{\prime}[x_{1,3},\ldots,x_{1,2n+1},x_{2,3},\ldots,x_{2,2n+1},Y_{i,j}|i=1,2,3;\;j=2n,2n+1],

as desired (where ΔiΔi\Delta_{i}\mapsto\Delta_{i}^{\prime} for i=1,,2n1i=1,\ldots,2n-1). Therefore, the ring (S/J)x1.2(S/J)_{x_{1.2}} is FF-regular by induction.

We now construct an S/JS/J-linear splitting of the map

S/JF(S/J)where1F(x1,2)S/J\to F_{*}(S/J)\qquad\text{where}\qquad 1\mapsto F_{*}(x_{1,2})

in order to establish the FF-regularity of the generic link S/JS/J. After reindexing, let

Δ1=[1,,2n],Δ2=[1,,n,n+2,,2n+1],andΔ3=[2,,2n+1]\Delta_{1}=[1,\ldots,2n],\;\Delta_{2}=[1,\ldots,n,n+2,\ldots,2n+1],\;\text{and}\;\Delta_{3}=[2,\ldots,2n+1]

denote the pfaffians of XX with the given rows and columns and let β1,β2,β3\beta_{1},\beta_{2},\beta_{3} denote the first three entries of the matrix Y[Δ1,,Δ2n+1]TY[\Delta_{1},\ldots,\Delta_{2n+1}]^{T}. By Lemma 3.3 and Proposition 4.1, we get that the initial term of the polynomial

f\colonequalsβ1β2β3f\colonequals\beta_{1}\beta_{2}\beta_{3}

is squarefree with respect to the term order <1<_{1} in Lemma 3.3 (constructed by extending the term order << in Proposition 4.1). Furthermore, we have,

fp1J[p]:Jf^{p-1}\in J^{[p]}:J

by Lemma 2.8. Note that x1,2x_{1,2} does not divide the initial term in<1(f)\operatorname{in}_{<_{1}}(f). In view of this, consider the polynomial

g\colonequalsYi,j,xk,lin<1(f)(k,l)(1,2)Yi,jxk,lfg\colonequals\prod_{\begin{subarray}{c}Y_{i,j},x_{k,l}\notin\operatorname{in}_{<_{1}}(f)\\ (k,l)\neq(1,2)\end{subarray}}Y_{i,j}x_{k,l}f

and note that gp1J[p]:Jg^{p-1}\in J^{[p]}:J. The Frobenius trace map

Tr(F(x1,2p2gp1)):F(S/J)S/JsendsF(x1,2)1\operatorname{Tr}(F_{*}(x_{1,2}^{p-2}g^{p-1})):F_{*}(S/J)\to S/J\quad\text{sends}\quad F_{*}(x_{1,2})\mapsto 1

to give the required splitting. The FF-regularity of the generic link S/JS/J now follows from [HH94a, Theorem 5.9(a)]. ∎

Remark 4.4.

The generators of the generic link of height 33 Gorenstein ideals are known: [KU92a, §4.10] gives the generators of the generic link as nested pfaffians of varying sizes of an ‘almost alternating’ matrix (see [KU92b, Theorem 4.7] for an easier proof). We emphasize that due to our approach of viewing the generic link via the property 𝐏\mathbf{P}, we did not need its generators to establish the FF-regularity of the generic link (or its FF-pure threshold, or the squarefreeness of its initial ideal).

Corollary 4.5.

Let II be a generic height 33 Gorenstein ideal of R=K[x¯]R=K[\underline{x}] over an FF-finite field KK of characteristic p>0p>0 and Lk(I)L^{k}(I) the kk-th universal link of II in a transcendental extension SS of RR. Then S/Lk(I)S/L^{k}(I) is an FF-rational ring for each k0k\geq 0.

Proof.

Due to [HU87, Corollary 2.15], the even universal links S/L2k(I)S/L^{2k}(I) (respectively the odd universal links S/L2k+1(I)S/L^{2k+1}(I)) are essentially a deformation of R/IR/I (respectively that of S/L1(I)S/L^{1}(I)). That is, the universal links are obtained by a finite sequence of deformations and localizations at prime ideals beginning from R/IR/I or S/L1(I)S/L^{1}(I), in the respective cases. Note that generic pfaffian rings are FF-regular and the generic links of generic height 33 Gorenstein ideals are FF-regular by Theorem 4.3. Since the universal link is a localization of the corresponding generic link at a prime ideal, the assertion follows as FF-regular rings are FF-rational and FF-rationality deforms (and is a local property) due to [HH94a, Theorem 4.2 (h)]. ∎

If α¯\underline{\alpha} is a regular sequence in a polynomial ring RR such that the pair (R,α¯)(R,\underline{\alpha}) has 𝐏\mathbf{P}, then, by Theorem 3.5, any ideal in the linkage class of (α¯)(\underline{\alpha}) has a deformation (the nn-th generic link of (α¯)(\underline{\alpha})) which is FF-pure. By Proposition 4.1, a class of examples of such licci ideals (i.e., ideals linked to a complete intersection) are generic height 33 Gorenstein ideals (see [Wat73]). Another class of licci ideals is that of height 22 perfect ideals, subsumed in Proposition 5.1, which we prove next.

5. Minors of a generic matrix: A curious dichotomy

The aim of this section is to discuss some subtleties related to the property 𝐏\mathbf{P}. We begin by studying the minors of a generic matrix: Let X\colonequals(xij)X\colonequals(x_{ij}) be an m×nm\times n matrix of indeterminates for positive integers mnm\leq n and R\colonequalsK[X]R\colonequals K[X]. Let It(X)I_{t}(X) denote the RR-ideal generated by the size tt-minors of the matrix XX.

Proposition 5.1.

The pair (K[X]),Im(X))(K[X]),I_{m}(X)) has 𝐏\mathbf{P}.

Proof.

Let

α¯\colonequals[1,m],[2,m+1],,[nm+1,n]\underline{\alpha}\colonequals[1,m],[2,m+1],\ldots,[n-m+1,n]

denote the size mm minors of XX with the given adjacent columns and fix the lexicographical order in K[X]K[X] induced by the variable order

x1,1>x1,2>>x1,n>x2,1>>x2,n>>xm,n.x_{1,1}>x_{1,2}>\dots>x_{1,n}>x_{2,1}>\dots>x_{2,n}>\dots>x_{m,n}.

For 1inm+11\leq i\leq n-m+1, by determinant expansion along the first row of XX, we get

in<([i,m+i1])=x1,ix2,i+1xm,m+i1,\operatorname{in}_{<}\big{(}[i,m+i-1]\big{)}=x_{1,i}x_{2,i+1}\dots x_{m,m+i-1},

which is the product of the variables in the main diagonal of the minor [i,m+i1][i,m+i-1]. Since the determinantal ideal Im(X)I_{m}(X) has height nm+1n-m+1, we are done. ∎

Remark 5.2.

In view of Theorem 3.5, Proposition 5.1 recovers the FF-pure threshold (respectively log canonical threshold) of determinantal rings of maximal minors calculated in [MSV14, Theorem 1.2] (respectively [Doc13, Theorem 5.6], [Joh03]) as well as the log canonical threshold of the generic links of maximal minors calculated in [KMN21, Theorem 1] by a simpler method and also extends these calculations to the nn-th generic link for each n>0n>0.

We now illustrate that each of the conclusions of Theorem 3.5 typically fails for the generic links of the non-maximal minors of a generic matrix as well as that of the (maximal) minors of a Hankel matrix. The following lemma will be useful to this end:

Lemma 5.3.

Let R=K[x1,,xn]R=K[x_{1},\ldots,x_{n}] and II be a homogeneous RR-ideal of height gg for which each generator has degree dd. Then the (top) aa-invariant of the first universal link of II is d(g1)nd(g-1)-n.

Proof.

Let Y:=(Yi,j)Y:=(Y_{i,j}) be a g×ng\times n matrix and S:=R(x1,,xn)(Yi,j)S:=R_{(x_{1},\ldots,x_{n})}(Y_{i,j}). Let

L1(I)\colonequalsY[f1,,fn]TS:(f1,,fn)S=(α¯):ISL^{1}(I)\colonequals Y[f_{1},\ldots,f_{n}]^{T}S:(f_{1},\ldots,f_{n})S=(\underline{\alpha}):IS

denote the (first) universal link of II. Note that, since the ideal II is generated in the same degree, the factor ring S/L1(I)S/L^{1}(I) is the localization of a graded ring at its homogeneous maximal ideal. Hence its top local cohomology module

H(x1,,xn)ng(S/L1(I))H^{n-g}_{(x_{1},\ldots,x_{n})}(S/L^{1}(I))

is a \mathbb{Z}-graded S/L1(I)S/L^{1}(I)-module in a natural manner. By [KU92b, Lemma 2.3], the graded canonical module of the universal link is

ωS/L1(I)IS/(α¯)[n+i=1gdeg(αi)].\omega_{S/L^{1}(I)}\cong IS/(\underline{\alpha})[-n+\sum_{i=1}^{g}\deg(\alpha_{i})].

Since the indeterminates YijY_{ij} are units in SS and so have degree zero, it follows that the graded canonical module ωS/L1(I)\omega_{S/L^{1}(I)} is generated in degree

d(n+gd)=nd(g1).d-(-n+gd)=n-d(g-1).

Therefore the aa-invariant

a(S/L1(I))=d(g1)na(S/L^{1}(I))=d(g-1)-n

is as claimed. ∎

Example 5.4.

Let XX be a 3×73\times 7 matrix of indeterminates and R=K[X]R=K[X]. The pair (R,I2(X))(R,I_{2}(X)) clearly does not have the property 𝐏\mathbf{P}: The ideal I2(X)I_{2}(X) has height 1212 and does not contain any sequence of elements α1,,α12\alpha_{1},\ldots,\alpha_{12} as required for 𝐏\mathbf{P} to hold since that would require atleast 2424 indeterminates.

By Lemma 5.3, the aa-invariant of the (first) universal link is

a(R(Y)/L1(I2(X)))=2(121)21=1>0.a(R(Y)/L^{1}(I_{2}(X)))=2(12-1)-21=1>0.

Since an FF-injective ring has a non-positive aa-invariant, it follows that the universal link L1(I2(X))L^{1}(I_{2}(X)) does not define an FF-injective ring. As the universal link is a localization of the generic link at a prime ideal and since FF-injectivity is a local property [DM19, Proposition 3.3], it also follows that the generic link L1(I2(X))L_{1}(I_{2}(X)) is not FF-injective. Therefore it is not FF-pure (though, perhaps interestingly, one may check that, the aa-invariant of the generic link is negative). This observation stands in sharp contrast to the recent result [PT24, Theorem 5.6] which says that the generic link of the maximal minors of a generic matrix is FF-regular.

Since a squarefree initial ideal implies FF-injectivity (see, for example, [KV23, Corollary 4.11(2)]), it follows that the generic link L1(I2(X))L_{1}(I_{2}(X)) does not have a squarefree initial ideal. However, the generic determinantal ideal of minors of any size has a squarefree initial ideal since the minors form a Gröbner basis with respect to a diagonal term order by [Stu90]. We conclude that (generic) linkage does not preserve the squarefreeness of the initial ideal.

In addition, since generic determinantal rings of minors of any size are FF-regular ([HH94b, §7]) and hence have rational singularities ([Smi97, Theorem 4.3]), we recover the fact that (generic) linkage does not preserve rational singularities (see [Niu14, Corollary 3.4]).

Although the generic link L1(I2(X))L_{1}(I_{2}(X)) is not FF-pure, its FF-pure threshold still makes sense. By [KMN21, Theorem 1] (see also [Doc13, Theorem 5.6]),

fpt(L1(I2(X)))=10.5<12=ht(I2(X)).\operatorname{fpt}(L_{1}(I_{2}(X)))=10.5<12=\operatorname{ht}(I_{2}(X)).

We now show that the property 𝐏\mathbf{P} does not hold for Hankel determinantal rings, which are linear specializations of the maximal minors of a generic matrix (see [CMSV18, §1] for an introduction to Hankel determinantal rings):

Example 5.5.

Let HH be a t×nt\times n Hankel matrix with t2t\geq 2 and n2t+2n\geq 2t+2. Let It(H)I_{t}(H) denote the ideal generated by the size tt minors of HH in the polynomial ring R=K[H]R=K[H]. By Lemma 5.3, the aa-invariant of the universal link

a(R(Y)/L1(It(H)))=t(nt)(n+t1)=(n1)(t1)t2t2t1>0.a(R(Y)/L^{1}(I_{t}(H)))=t(n-t)-(n+t-1)=(n-1)(t-1)-t^{2}\geq t^{2}-t-1>0.

By the same arguments as in Example 5.4, we conclude that the generic link L1(It(H))L_{1}(I_{t}(H)) does not define an FF-injective ring and does not have a squarefree initial ideal. However, the Hankel determinantal ring R/It(H)R/I_{t}(H) is FF-rational if the field KK has characteristic pp with ptp\geq t, and therefore has rational singularities (see [CMSV18, Theorem 2.1]). Furthermore, the ideal It(H)I_{t}(H) has a squarefree initial ideal as its generators form a Gröbner basis with respect to the usual lexicographical ordering on the variables by [Con98, Proposition 3.4].

Note that the case t=2t=2 gives us that the generic link of the rational normal curve is not FF-injective and does not have a squarefree initial ideal for n6n\geq 6.

The natural nullcone of the symplectic group

We illustrate that the choice of the term order appearing in 𝐏\mathbf{P} may be quite subtle.

Let X\colonequals(xi,j)X\colonequals(x_{i,j}) be a 2t×n2t\times n matrix of indeterminates for positive integers tt and nn. Let Ω\Omega denote the size 2t2t standard symplectic block matrix

Ω\colonequals(0𝕀𝕀0),\Omega\colonequals\begin{pmatrix}0&\mathbb{I}\\ -\mathbb{I}&0\end{pmatrix},

where 𝕀\mathbb{I} is the size tt identity matrix. Let

𝔓(X)\colonequals(XTΩX)\mathfrak{P}(X)\colonequals(X^{T}\Omega X)

be the ideal generated by the entries of the alternating matrix XTΩXX^{T}\Omega X. The ideal 𝔓(X)\mathfrak{P}(X) is Hilbert’s nullcone ideal for the natural action of the Symplectic group (see [HJPS23, §6]) given by

M:XMXforMSp2t(K).M:X\to MX\quad\text{for}\quad M\in\operatorname{Sp}_{2t}(K).

The Symplectic nullcone is a Cohen–Macaulay domain with

dimK[X]/𝔓(X)\displaystyle\dim K[X]/\mathfrak{P}(X) =\displaystyle= {2nt(n2)if nt+1,nt+(t+12)if nt\displaystyle\begin{cases}2nt-\displaystyle{\binom{n}{2}}&\text{if }\ n\leq t+1,\\ nt+\displaystyle{\binom{t+1}{2}}&\text{if }\ n\geq t\\ \end{cases}

by [HJPS23, Theorem 6.8]. We describe the generators of 𝔓(X)\mathfrak{P}(X). For ease of notation in the next Proposition, we relabel the entries of the lower half of the 2t×n2t\times n matrix X=(xi,j)X=(x_{i,j}) as follows: Let wi,j=xi+t,nj+1w_{i,j}=x_{i+t,n-j+1} for 1it1\leq i\leq t and 1jn1\leq j\leq n. Then

X=(x1,1x1,2x1,nxt,1xt,2xt,nw1,nw1,n1w1,1wt,nwt,n1wt,1).X=\begin{pmatrix}x_{1,1}&x_{1,2}&\cdots&x_{1,n}\\ \vdots&\vdots&&\vdots\\ x_{t,1}&x_{t,2}&\cdots&x_{t,n}\\ w_{1,n}&w_{1,n-1}&\cdots&w_{1,1}\\ \vdots&\vdots&&\vdots\\ w_{t,n}&w_{t,n-1}&\cdots&w_{t,1}\end{pmatrix}.

The entries of the nullcone ideal (XTΩX)(X^{T}\Omega X) are sums of the size two minors

di,j\colonequalsdet(x1,ix1,jw1,ni+1w1,nj+1)++det(xt,ixt,jwt,ni+1wt,nj+1)d_{i,j}\colonequals\det\begin{pmatrix}x_{1,i}&x_{1,j}\\ w_{1,n-i+1}&w_{1,n-j+1}\end{pmatrix}+\cdots+\det\begin{pmatrix}x_{t,i}&x_{t,j}\\ w_{t,n-i+1}&w_{t,n-j+1}\end{pmatrix}

coming from columns ii and jj of the matrix XX.

Proposition 5.6.

The pair (K[X],𝔓(X))(K[X],\mathfrak{P}(X)) has 𝐏\mathbf{P}.

Proof.

Let

α¯\colonequals{di,j| 1i<jn and jit}\underline{\alpha}\colonequals\{d_{i,j}\;|\;1\leq i<j\leq n\;\text{ and }\;j-i\leq t\}

be the subset of the generators of the nullcone ideal 𝔓(X)\mathfrak{P}(X) with the given columns. The reader may verify that the above elements (or any such subset of the generators of 𝔓(X))\mathfrak{P}(X)) do not satisfy 𝐏\mathbf{P} under the ‘usual’ lex or revlex order on the variables. A more delicate term order is needed.

To define such a term order << in K[X]K[X], we first define an order on the variables. Sort the entries of the matrix XX into blocks B0,B1,,Bn1B_{0},B_{1},\dots,B_{n-1} according to the following formula:

xi,j,wi,j are in block B where ={2j+i2if 1j<ni+12,2n2jiif ni+12j<ni+1,0otherwise.\displaystyle\text{$x_{i,j},w_{i,j}$ are in block $B_{\ell}$ where }\ell=\begin{cases}2j+i-2&\text{if \; }1\leq j<\dfrac{n-i+1}{2},\\ 2n-2j-i&\text{if \; }\dfrac{n-i+1}{2}\leq j<n-i+1,\\ 0&\text{otherwise.}\end{cases}

Now, for aBa\in B_{\ell} and bBkb\in B_{k}, set a<ba<b if <k\ell<k. Then, within each set BB_{\ell}, fix an arbitrary order among the variables. This gives us a total variable order in K[X]K[X]. Our term order << is the reverse lexicographical order induced by this variable order. The initial terms of the elements of α¯\underline{\alpha} are as required for the property 𝐏\mathbf{P} to hold by [PTW23, Lemmas 3.4, 3.5] (see [PTW23, Example 3.3] for an explicit example). ∎

Remark 5.7.

One can show that the generic link of the symplectic nullcone is FF-regular in positive prime characteristic. The proof follows along the lines of Theorem 4.3 by making use of [PTW23, Lemma 3.2]. We leave the details to the interested reader.

6. Generic residual intersections of a complete intersection inherit 𝐏\mathbf{P}

In this section, we show that the property 𝐏\mathbf{P} is inherited by the generic residual intersections of an ideal generated by a regular sequence in a polynomial ring. This observation is quite surprising in light of the discussion so far since residual intersections do not preserve height unless they are links.

6.1. Brief recall of residual intersections

The notion of residual intersections essentially goes back to Artin and Nagata [AN72]. Let XX and YY be two irreducible closed subschemes of a Noetherian scheme ZZ with codimZXcodimZY=s\operatorname{codim}_{Z}X\leq\operatorname{codim}_{Z}Y=s and YXY\nsubseteq X, then YY is called a residual intersection of XX if the number of equations needed to define XYX\cup Y as a subscheme of ZZ is the smallest possible, namely ss. However, in order to include the case where XX and YY are reducible with XX possibly containing some component of YY, the following algebraic definition is more suited:

Definition 6.1.

Let RR be a polynomial ring and II an RR-ideal. Given an ideal 𝔞I\mathfrak{a}\subsetneq I generated by ss elements, set J=𝔞:IJ=\mathfrak{a}:I. If ht(J)sht(I)\operatorname{ht}(J)\geq s\geq\operatorname{ht}(I), then JJ is an ss-residual intersection of II (or R/JR/J is an ss-residual intersection of R/IR/I).

If furthermore I𝔭=𝔞𝔭I_{{\mathfrak{p}}}={\mathfrak{a}}_{{\mathfrak{p}}} for all 𝔭V(I){\mathfrak{p}}\in V(I) with ht𝔭s\operatorname{ht}{{\mathfrak{p}}}\leq s, then JJ is called a geometric ss-residual intersection of II.

Residual intersections are natural generalizations of linkage: If an ideal II is unmixed of height gg, then the gg-residual intersections of II are precisely the links of II and the geometric gg-residual intersections are precisely the geometric links of II by Proposition 2.4. Furthermore, as in the case of links, we may define the ‘most general’ residual intersections if the following technical requirement is met:

Definition 6.2.

We say that an ideal II in a Noetherian ring satisfies the condition GsG_{s} if μ(I𝔭)ht(𝔭)\mu(I_{{\mathfrak{p}}})\leq\operatorname{ht}({\mathfrak{p}}) for all prime ideals 𝔭{\mathfrak{p}} containing II such that ht(𝔭)s1\operatorname{ht}({\mathfrak{p}})\leq s-1 (where μ()\mu(-) denotes the minimal number of generators). We say that II satisfies GG_{\infty} if it satisfies GsG_{s} for each ss. It is clear that this is equivalent to the condition that μ(I𝔭)ht(𝔭)\mu(I_{{\mathfrak{p}}})\leq\operatorname{ht}({\mathfrak{p}}) for each prime ideal 𝔭{\mathfrak{p}} containing II.

Definition 6.3.

Let RR be a polynomial ring and II be an ideal of height g>0g>0 satisfying Gs+1G_{s+1}, where sgs\geq g. Choose any generating set f¯\colonequalsf1,,fn\underline{f}\colonequals f_{1},\ldots,f_{n} of II and let YY be an s×ns\times n matrix of indeterminates. Let 𝔞\mathfrak{a} be the ideal generated by the entries of the matrix Y[f1fn]TY[f_{1}\dots f_{n}]^{T}. Then we set

RI(s;I)=RI(s;f¯)\colonequals𝔞R[Y]:IR[Y]\operatorname{RI}(s;I)=\operatorname{RI}(s;\underline{f})\colonequals\mathfrak{a}R[Y]:IR[Y]

and call this ideal a generic ss-residual intersection of II.

Generic residual intersections are in fact geometric residual intersections [HU88, Theorem 3.3]. As is the case with generic linkage, the generic residual intersections are essentially independent of the generating set for the ideal [HU90, Lemma 2.2]. The generic ss-residual intersection specializes to any ss-residual intersection of the given ideal and tracks the homological properties of the ideal via deformation.

6.2. FF-pure threshold of the generic residual intersections of a complete intersection

We layout the setup for this subsection: Let α¯\colonequalsα1,,αg\underline{\alpha}\colonequals\alpha_{1},\ldots,\alpha_{g} be a homogeneous regular sequence, with each αi\alpha_{i} having the same degree, in the polynomial ring K[x¯]K[\underline{x}] such that the pair (K[x¯],α¯)(K[\underline{x}],\underline{\alpha}) has 𝐏\mathbf{P}. Fix an integer sgs\geq g and let Y\colonequals(Yi,j)Y\colonequals(Y_{i,j}) be a matrix of indeterminates of size s×gs\times g. Let

RI(s;α¯)\colonequalsY[α1,,αg]TR[Y]:(α¯)R[Y]\operatorname{RI}(s;\underline{\alpha})\colonequals Y[\alpha_{1},\cdots,\alpha_{g}]^{T}R[Y]:(\underline{\alpha})R[Y]

denote the generic ss-residual intersection of the ideal (α¯)(\underline{\alpha}); set M\colonequalsY[α1,,αg]TM\colonequals Y[\alpha_{1},\cdots,\alpha_{g}]^{T}. Since an ideal generated by a regular sequence satisfies GG_{\infty}, the residual intersection RI(s;α¯)\operatorname{RI}(s;\underline{\alpha}) exists for each sgs\geq g.

We now show that the generic residual intersections of complete intersections inherit 𝐏\mathbf{P}. We need the generators of the residual intersection to do this, which were calculated in [HU88, Example 3.4] (see also [BKM90]).

Theorem 6.4.

If the pair (R,α¯)(R,\underline{\alpha}) has the property 𝐏\mathbf{P}, then (R[Y],RI(s;α¯))(R[Y],\operatorname{RI}(s;\underline{\alpha})) also has 𝐏\mathbf{P} for each s|α¯|s\geq|\underline{\alpha}|.

Proof.

Since α¯=α1,,αg\underline{\alpha}=\alpha_{1},\ldots,\alpha_{g} is a regular sequence, by the structure theorem for the residual intersections of a complete intersection ring ([HU88, Example 3.4]), we have

RI(s,α¯)=(entries of the matrix M)R[Y]+Ig(Y)R[Y]RI(s,\underline{\alpha})=(\text{entries of the matrix $M$})R[Y]+I_{g}(Y)R[Y]

is a homogeneous R[Y]R[Y]-ideal of height ss. We shall construct a term order <1<_{1} in R[Y]R[Y] such that the pair (R[Y],β¯)(R[Y],\underline{\beta}) satisfies 𝐏\mathbf{P} with

β¯\colonequals{M1,1,,Mg1,1}{[i,i+1,,gi+1]| 1isg+1}.\underline{\beta}\colonequals\{M_{1,1},\ldots,M_{g-1,1}\}\bigcup\{[i,i+1,\ldots,g-i+1]\;|\;1\leq i\leq s-g+1\}.

That is, β¯\underline{\beta} consists of the first g1g-1 entries of the matrix MM and all the gg-minors of the matrix Y\colonequals(Yi,j)Y\colonequals(Y_{i,j}) with adjacent rows. Clearly, the set β¯\underline{\beta} has

(g1)+(sg+1)=s=ht(RI(s;α¯))(g-1)+(s-g+1)=s=\operatorname{ht}(\operatorname{RI}(s;\underline{\alpha}))

elements, as required; hereafter, we call these elements β1,,βg1,βg,,βs\beta_{1},\ldots,\beta_{g-1},\beta_{g},\ldots,\beta_{s} with the indexing according to the above display. We now construct the required term order.

Define the following variable order in R[Y]R[Y]:

Ys,g>Ys,g1>>Ys,1>\displaystyle Y_{s,g}>Y_{s,g-1}>\cdots>Y_{s,1}>
\displaystyle\vdots
Yg1,g>Yg1,g1>>Yg1,1>\displaystyle Y_{g-1,g}>Y_{g-1,g-1}>\cdots>Y_{g-1,1}>
Yg2,g1>Yg2,g>Yg2,g2>>Yg2,1>\displaystyle Y_{g-2,g-1}>Y_{g-2,g}>Y_{g-2,g-2}>\cdots>Y_{g-2,1}>
Yg3,g2>Yg3,g>Yg3,g1>>Yg3,1>\displaystyle Y_{g-3,g-2}>Y_{g-3,g}>Y_{g-3,g-1}>\cdots>Y_{g-3,1}>
\displaystyle\vdots
Y1,2>Y1,g>Y1,g1>>Y1,1>xi.j.\displaystyle Y_{1,2}>Y_{1,g}>Y_{1,g-1}>\cdots>Y_{1,1}>x_{i.j}.

That is, we begin the variable order with the last row of the matrix YY, move from right to left, and then vertically up along the rows of YY for (the bottom) sgs-g rows, thereafter we follow the same order for the remaining rows except that the (i,i+1)(i,i+1) entry appears first in the ordering of the ii-th row. In addition, the ordering on the indeterminates xi,jx_{i,j} is the one given in the property 𝐏\mathbf{P} for the pair (R,α¯)(R,\underline{\alpha}). Our term order is the lexicographical order <1<_{1} induced by this variable order in R[Y]R[Y]. Note that the initial terms of the first g1g-1 entries of the matrix MM are

in<1(β1)\displaystyle\operatorname{in}_{<_{1}}(\beta_{1}) =in<1(Y1,1α1+Y1,2α2++Y1,gαg)=Y1,2in<(α2),\displaystyle=\operatorname{in}_{<_{1}}(Y_{1,1}\alpha_{1}+Y_{1,2}\alpha_{2}+\cdots+Y_{1,g}\alpha_{g})=Y_{1,2}\operatorname{in}_{<}(\alpha_{2}),
in<1(β2)\displaystyle\operatorname{in}_{<_{1}}(\beta_{2}) =in<1(Y2,1α1+Y2,2α2++Y2,gαg)=Y2,3in<(α3),\displaystyle=\operatorname{in}_{<_{1}}(Y_{2,1}\alpha_{1}+Y_{2,2}\alpha_{2}+\cdots+Y_{2,g}\alpha_{g})=Y_{2,3}\operatorname{in}_{<}(\alpha_{3}),
\displaystyle\vdots
in<1(βg1)\displaystyle\operatorname{in}_{<_{1}}(\beta_{g-1}) =in<1(Yg1,1α1+Yg1,2α2++Yg1,gαg)=Yg1,gin<(αg).\displaystyle=\operatorname{in}_{<_{1}}(Y_{g-1,1}\alpha_{1}+Y_{g-1,2}\alpha_{2}+\cdots+Y_{g-1,g}\alpha_{g})=Y_{g-1,g}\operatorname{in}_{<}(\alpha_{g}).

Further, by determinant expansion along the last row of YY, the initial terms of the remaining sg+1s-g+1 elements of β¯\underline{\beta},

in<1(βi)=in<1([i,i+1,,g+i1])=Yi,1Yi+1,2Yg+i1,g,1isg+1,\operatorname{in}_{<_{1}}(\beta_{i})=\operatorname{in}_{<_{1}}([i,i+1,\cdots,g+i-1])=Y_{i,1}Y_{i+1,2}\ldots Y_{g+i-1,g},\quad 1\leq i\leq s-g+1,

are just the products of the variables occurring in the main diagonal of the matrix [i,i+1,,g+i1][i,i+1,\cdots,g+i-1]. Since (R,α¯)(R,\underline{\alpha}) has 𝐏\mathbf{P} by hypothesis, we get that for the elements β¯=β1,,βs\underline{\beta}=\beta_{1},\ldots,\beta_{s} of R[Y]R[Y], each initial term is squarefree and each pair of initial terms is mutually coprime, as required for (R[Y],RI(s;α¯))(R[Y],\operatorname{RI}(s;\underline{\alpha})) to inherit 𝐏\mathbf{P}. ∎

While the FF-pure threshold of the generic link of an ideal is atleast as large as that of the ideal itself by [Niu14, Proposition 3.7], it is clearly bounded above by the height of the ideal due to Lemma 2.10. On the other hand, the FF-pure threshold of the generic residual intersections may behave quite differently:

Corollary 6.5.

If the pair (R,α¯)(R,\underline{\alpha}) has 𝐏\mathbf{P}, then, for each s|α¯|s\geq|\underline{\alpha}|, we have

lct(RI(s;α¯))=fpt(RI(s;α¯))=s|α¯|=fpt(α¯)=lct(α¯).\operatorname{lct}({\operatorname{RI}(s;\underline{\alpha}))}=\operatorname{fpt}({\operatorname{RI}(s;\underline{\alpha}))}=s\geq|\underline{\alpha}|=\operatorname{fpt}{(\underline{\alpha})}=\operatorname{lct}{(\underline{\alpha})}.
Proof.

The assertion immediately follows from Theorem 6.4 in view of Theorem 3.5 and Remark 3.6. The outer equality is [HY03]. ∎

Remark 6.6.

Corollary 6.5 with s=|α¯|s=|\underline{\alpha}| answers [KMN21, Question 4.5(b)] by giving a class of complete intersection ideals—those possessing the property 𝐏\mathbf{P}—for which the log canonical threshold is preserved under generic linkage.

We point out that the log canonical threshold, may not be preserved under generic linkage (and therefore under generic residual intersections) for complete intersection ideals II in a polynomial ring RR if (R,I)(R,I) does not have 𝐏\mathbf{P}: Let R=K[x1,x2,x3]R=K[x_{1},x_{2},x_{3}] and I=(x12x2,x33)I=(x_{1}^{2}x_{2},x_{3}^{3}) be a homogeneous RR-ideal. By [KMN21, Example 2.4(a)] and the linear programming methods in [ST09], we get

lct(L1(I))=11/6>5/6=lct(I).\operatorname{lct}(L_{1}(I))=11/6>5/6=\operatorname{lct}(I).

We end with a reformulation of [KMN21, Question 4.5(b)]:

Question 6.7.

Let II be an unmixed, height g>0g>0 ideal in a polynomial ring over a field of positive characteristic. Let s>gs>g be an integer such that II satisfies Gs+1G_{s+1} (so that the generic residual intersection RI(s;I)\operatorname{RI}(s;I) exists). When do we have the equality

fpt(RI(s;I))=fpt(I)?\operatorname{fpt}(\operatorname{RI}(s;I))=\operatorname{fpt}(I)?

Acknowledgments

I thank Manav Batavia, Linquan Ma, Yevgeniya Tarasova, Bernd Ulrich, and Matteo Varbaro for several insightful discussions. I am grateful to Samiksha Bidhuri for her unwavering support and encouragement.

References

  • [AN72] Michael Artin and Masayoshi Nagata. Residual intersections in Cohen-Macaulay rings. J. Math. Kyoto Univ., 12:307–323, 1972.
  • [Bar95] Margherita Barile. Arithmetical ranks of ideals associated to symmetric and alternating matrices. J. Algebra, 176(1):59–82, 1995.
  • [BCRV22] Winfried Bruns, Aldo Conca, Claudiu Raicu, and Matteo Varbaro. Determinants, Gröbner bases and cohomology. Springer Monographs in Mathematics. Springer, Cham, [2022] ©2022.
  • [BE77] David A. Buchsbaum and David Eisenbud. Algebra structures for finite free resolutions, and some structure theorems for ideals of codimension 33. Amer. J. Math., 99(3):447–485, 1977.
  • [BKM90] Winfried Bruns, Andrew R. Kustin, and Matthew Miller. The resolution of the generic residual intersection of a complete intersection. J. Algebra, 128(1):214–239, 1990.
  • [CMSV18] Aldo Conca, Maral Mostafazadehfard, Anurag K. Singh, and Matteo Varbaro. Hankel determinantal rings have rational singularities. Adv. Math., 335:111–129, 2018.
  • [Con98] Aldo Conca. Straightening law and powers of determinantal ideals of Hankel matrices. Adv. Math., 138(2):263–292, 1998.
  • [DM19] Rankeya Datta and Takumi Murayama. Permanence properties of FF-injectivity. to appear in Math. Res. Lett., https://arxiv.org/abs/1906.11399, 2019.
  • [Doc13] Roi Docampo. Arcs on determinantal varieties. Trans. Amer. Math. Soc., 365(5):2241–2269, 2013.
  • [DSNnB18] Alessandro De Stefani and Luis Núñez Betancourt. FF-thresholds of graded rings. Nagoya Math. J., 229:141–168, 2018.
  • [DSNnBP18] Alessandro De Stefani, Luis Núñez Betancourt, and Felipe Pérez. On the existence of FF-thresholds and related limits. Trans. Amer. Math. Soc., 370(9):6629–6650, 2018.
  • [Ful84] William Fulton. Intersection theory, volume 2 of Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)]. Springer-Verlag, Berlin, 1984.
  • [HH94a] Melvin Hochster and Craig Huneke. FF-regularity, test elements, and smooth base change. Trans. Amer. Math. Soc., 346(1):1–62, 1994.
  • [HH94b] Melvin Hochster and Craig Huneke. Tight closure of parameter ideals and splitting in module-finite extensions. J. Algebraic Geom., 3(4):599–670, 1994.
  • [HJPS23] Melvin Hochster, Jack Jeffries, Vaibhav Pandey, and Anurag K. Singh. When are the natural embeddings of classical invariant rings pure? Forum Math. Sigma, 11:Paper No. e67, 2023.
  • [Hoc73] Melvin Hochster. Properties of Noetherian rings stable under general grade reduction. Arch. Math. (Basel), 24:393–396, 1973.
  • [HU85] Craig Huneke and Bernd Ulrich. Divisor class groups and deformations. Amer. J. Math., 107(6):1265–1303, 1985.
  • [HU87] Craig Huneke and Bernd Ulrich. The structure of linkage. Ann. of Math. (2), 126(2):277–334, 1987.
  • [HU88] Craig Huneke and Bernd Ulrich. Residual intersections. J. Reine Angew. Math., 390:1–20, 1988.
  • [HU90] Craig Huneke and Bernd Ulrich. Generic residual intersections. In Commutative algebra (Salvador, 1988), volume 1430 of Lecture Notes in Math., pages 47–60. Springer, Berlin, 1990.
  • [HW02] Nobuo Hara and Kei-Ichi Watanabe. F-regular and F-pure rings vs. log terminal and log canonical singularities. J. Algebraic Geom., 11(2):363–392, 2002.
  • [HY03] Nobuo Hara and Ken-Ichi Yoshida. A generalization of tight closure and multiplier ideals. Trans. Amer. Math. Soc., 355(8):3143–3174, 2003.
  • [Joh03] Amanda Ann Johnson. Multiplier ideals of determinantal ideals. ProQuest LLC, Ann Arbor, MI, 2003. Thesis (Ph.D.)–University of Michigan.
  • [KMN21] Youngsu Kim, Lance Edward Miller, and Wenbo Niu. Log canonical thresholds of generic links of generic determinantal varieties. Proc. Amer. Math. Soc., 149(7):2777–2787, 2021.
  • [KU92a] Andrew R. Kustin and Bernd Ulrich. A family of complexes associated to an almost alternating map, with applications to residual intersections. Mem. Amer. Math. Soc., 95(461):iv+94, 1992.
  • [KU92b] Andrew R. Kustin and Bernd Ulrich. If the socle fits. J. Algebra, 147(1):63–80, 1992.
  • [KV23] Mitra Koley and Matteo Varbaro. Gröbner deformations and FF-singularities. Math. Nachr., 296(7):2903–2917, 2023.
  • [LS99] Gennady Lyubeznik and Karen E. Smith. Strong and weak FF-regularity are equivalent for graded rings. Amer. J. Math., 121(6):1279–1290, 1999.
  • [MSV14] Lance Edward Miller, Anurag K. Singh, and Matteo Varbaro. The FF-pure threshold of a determinantal ideal. Bull. Braz. Math. Soc. (N.S.), 45(4):767–775, 2014.
  • [MTW05] Mircea Mustaţǎ, Shunsuke Takagi, and Kei-ichi Watanabe. F-thresholds and Bernstein-Sato polynomials. In European Congress of Mathematics, pages 341–364. Eur. Math. Soc., Zürich, 2005.
  • [Niu14] Wenbo Niu. Singularities of generic linkage of algebraic varieties. Amer. J. Math., 136(6):1665–1691, 2014.
  • [PS73] Christian Peskine and Lucien Szpiro. Dimension projective finie et cohomologie locale. Applications à la démonstration de conjectures de M. Auslander, H. Bass et A. Grothendieck. Inst. Hautes Études Sci. Publ. Math., (42):47–119, 1973.
  • [PT24] Vaibhav Pandey and Yevgeniya Tarasova. Linkage and FF-regularity of determinantal rings. to appear in Int. Math. Res. Not. IMRN, https://doi.org/10.1093/imrn/rnae040, 2024.
  • [PTW23] Vaibhav Pandey, Yevgeniya Tarasova, and Uli Walther. On the natural nullcones of the symplectic and general linear groups. https://arxiv.org/pdf/2310.01816.pdf, 2023.
  • [Smi97] Karen E. Smith. FF-rational rings have rational singularities. Amer. J. Math., 119(1):159–180, 1997.
  • [ST09] Takafumi Shibuta and Shunsuke Takagi. Log canonical thresholds of binomial ideals. Manuscripta Math., 130(1):45–61, 2009.
  • [Stu90] Bernd Sturmfels. Gröbner bases and Stanley decompositions of determinantal rings. Math. Z., 205(1):137–144, 1990.
  • [Tak04] Shunsuke Takagi. F-singularities of pairs and inversion of adjunction of arbitrary codimension. Invent. Math., 157(1):123–146, 2004.
  • [TW04] Shunsuke Takagi and Kei-ichi Watanabe. On F-pure thresholds. J. Algebra, 282(1):278–297, 2004.
  • [Ulr92] Bernd Ulrich. Remarks on residual intersections. In Free resolutions in commutative algebra and algebraic geometry (Sundance, UT, 1990), volume 2 of Res. Notes Math., pages 133–138. Jones and Bartlett, Boston, MA, 1992.
  • [Wat73] Junzo Watanabe. A note on Gorenstein rings of embedding codimension three. Nagoya Math. J., 50:227–232, 1973.