• 沒有找到結果。

As a consequence, this provides an alternative proof of the classical fact that Schubert varieties in flag varieties are normal and have rational singularities

N/A
N/A
Protected

Academic year: 2022

Share "As a consequence, this provides an alternative proof of the classical fact that Schubert varieties in flag varieties are normal and have rational singularities"

Copied!
15
0
0

加載中.... (立即查看全文)

全文

(1)

S 0019-2082

ON THE F -RATIONALITY AND COHOMOLOGICAL PROPERTIES OF MATRIX SCHUBERT VARIETIES

JEN-CHIEH HSIAO

Abstract. We characterize complete intersection matrix Schu- bert varieties, generalizing the classical result on one-sided lad- der determinantal varieties. We also give a new proof of the F -rationality of matrix Schubert varieties. Although it is known that such varieties are F -regular (hence F -rational) by the global F -regularity of Schubert varieties, our proof is of independent in- terest since it does not require the Bott–Samelson resolution of Schubert varieties. As a consequence, this provides an alternative proof of the classical fact that Schubert varieties in flag varieties are normal and have rational singularities.

1. Introduction

Matrix Schubert varieties (MSVs) were introduced by W. Fulton in his theory of degeneracy loci of maps of flagged vector bundles [Ful92]. Such va- rieties are reduced and irreducible. Classical (one-sided ladder) determinantal varieties are special examples of MSVs (they are so-called vexillary MSVs).

Just like one-sided ladder determinantal varieties [GL00], [GM00], MSVs can be identified (up to product of an affine space) as the opposite big cells of the corresponding Schubert varieties. This observation in [Ful92] implies that the MSVs are normal and Cohen–Macaulay, since Schubert varieties are (see [Ram85]).

The Cohen–Macaulay property of MSVs was re-established by A. Knutson and E. Miller [KM05] using the Gr¨obner basis theory, pipe dreams, and their theory of subword complexes. Interestingly, this gives a new proof of the Cohen–Macaulayness of Schubert varieties by the following principle.

Received September 1, 2011; received in final form September 8, 2013.

The author was partially supported by NSF under Grant DMS 0901123.

2010 Mathematics Subject Classification.13C40, 14M15, 14M10, 05E40, 13A35.

1

2014 University of Illinoisc

(2)

Theorem 1.1 ([KM05, Theorem 2.4.3]). Let C be a local condition that holds for a variety X whenever it holds for the product of X with any vector space. Then C holds for every Schubert variety in every flag variety if and only ifC holds for all MSVs.

1.1. F -rationality of MSVs. In the same spirit, the first part of this paper is devoted to a new proof of F -rationality of MSVs. F -rationality is a notion that arises from the theory of tight closure introduced by M. Hochster and C. Huneke [Hun96] in positive characteristic. The results of [Smi97]

and [Har98] establish a connection between F -rationality and the notion of rational singularity in characteristic 0: A normal variety in characteristic 0 has at most rational singularities if and only if it is of F -rational type. Therefore, by Theorem1.1the F -rationality of matrix Schubert varieties is equivalent to the classical fact that Schubert varieties are normal and have at most rational singularities (see, e.g., [Bri05] and [BK05] for the classical proofs of the later statement using the Bott–Samelson resolution).

Two other notions in tight closure theory will also be used later: F - regularity and F -injectivity. The relation between these properties is

regular = F -regular = F -rational = F -injective.

We remark that MSVs are in fact F -regular by Theorem1.1 and the global F -regularity of Schubert varieties [LRPT06] (again, this relies on the Bott–

Samelson resolution).

Our proof of F -rationality of MSVs utilizes the results of Schubert determi- nantal ideals in [KM05] as well as the techniques developed in [CH97], where A. Conca and J. Herzog prove that arbitrary (possibly two-sided) ladder de- terminant varieties are F -rational. However, it is still unknown whether such varieties are F -regular.

One of the key ingredients in our proof is the following theorem.

Theorem 1.2 ([CH97, Theorem 1.2]). Let R be a complete local Cohen–

Macaulay ring and c be a nonzero-divisor of R such that R[1c] is F -rational and R/cR is F -injective. Then R is F -rational.

After recalling several known facts in the theory of tight closure (Sec- tion3), we will see that the most essential step is to find c such that Rw[1c] is F -rational and that the initial ideal in<(c + Iw) of c + Iw is Cohen–

Macaulay (where < is any antidiagonal term order, Rw and Iw is the co- ordinate ring and the defining ideal of the MSV associated to the partial permutation w as defined in Section 2). This goal is achieved by choosing c = xi0,w(io) where i0 is the smallest number such that {(p, q) | p > i0, q >

w(i0)} ∩ E>0(w)= ∅. See Section2for unexplained notation.

1.2. Complete intersection MSVs. Since MSVs are Cohen–Macaulay, it is then natural to ask when such varieties are smooth, complete intersection,

(3)

or Gorenstein. Classically, characterizations of Gorenstein ladder determi- nantal varieties are obtained in [Con95], [Con96] and [GM00]. In one-sided cases, the characterization can be generalized as the following. Recall that there exists a characterization of smooth (respectively, Gorenstein) Schubert varieties [LS90] (respectively, [WY06]). Since the singular (respectively, non- Gorenstein) locus of a Schubert variety is closed and invariant under the Borel subgroup action, the opposite big cell must be contained in the singular (respectively, non-Gorenstein) locus. Hence, a Schubert variety is smooth (re- spectively, Gorenstein) if and only if its corresponding MSV is so. Therefore, one can deduce a criterion of smooth (respectively, Gorenstein) MSVs by the corresponding result for Schubert varieties. See [WY06, Section 3.5] for more details.

The second goal of this paper is to characterize complete intersection MSVs.

We explain the characterization as the following. See Sections 2 and 5 for unexplained notation and more details.

Theorem (Theorem 5.2). The matrix Schubert variety Xw associated to a permutation w is a complete intersection if and only if, for any (p, q) in the diagram of w with rp,q(w) > 0, that is, (p, q)∈ D>0(w), w(p,q)is a permutation matrix in GLrp,q(w) such that Xw(p,q) is a complete intersection. Here, w(p,q)

is the connected (solid) square submatrix of size rp,q(w) whose southeast corner lies at (p− 1, q − 1). In fact, in this case

xp,q| (p, q) ∈ D=0(w) 

det X(p,q)| (p, q) ∈ D>0(w)

is a set of generators of Iw with cardinality|D(w)|, the codimension of Xwin Mn×n, where X(p,q)is the connected (solid) square submatrix of size rp,q(w) + 1 whose southeast corner lies at (p, q).

Theorem 5.2 generalizes a result in [GS95] for one-sided ladder determi- nantal varieties. The proof of Theorem 5.2uses Nakayama’s lemma and the properties of Schubert determinantal ideals established in [KM05].

After this work is finished, A. Woo and H. Ulfarsson give a criterion of lo- cally complete intersection Schubert varieties. Theorem5.2may be recovered by their criterion (see [UW13, Corollary 6.3] and the comment after that).´ 1.3. This paper is organized as follows. We will recall some preliminary facts about matrix Schubert varieties as well as tight closure theory in Sec- tions2and3, respectively. The proof of F -rationality of MSVs is in Section4.

Section5 is devoted to the characterization of complete intersection MSVs.

2. Matrix Schubert varieties

We recall some fundamental facts about matrix Schubert varieties (see [Ful92], [KM05] and [MS05] for more information).

(4)

Denote Ml×m the space of l× m matrices over a field K. An l × m matrix w∈ Ml×m is called a partial permutation if all entries of w are equal to 0 except for at most one entry equal to 1 in each row and column. If l = m and w∈ GLl, then w is called a permutation. An element w in the permutation group Sn will be identified as a permutation matrix (also denoted by w) in GLn via

wi,j=



1 if w(i) = j, 0 otherwise.

Let K[X] be the coordinate ring of Ml×mwhere X = (xi,j) is the generic l×m matrix of variables. For a matrix Z∈ Ml×m, denote Z[p,q] the upper-left p× q submatrix of Z. Similarly, X[p,q] denotes the upper-left p× q submatrix of X.

The rank of Z[p,q] will be denoted by rank(Z[p,q]) := rp,q(Z).

Given a partial permutation w∈ Ml×m, the matrix Schubert variety Xwis the subvariety

Xw=

Z∈ Ml×m| rp,q(Z)≤ rp,q(w) for all p, q

in Ml×m. The classical (one-sided ladder) determinantal varieties are special examples of MSVs.

It is known that MSVs are reduced and irreducible. Denote Rw= K[Xw] = K[X]/Iw

the coordinate ring of Xw. The defining ideal Iw of Xw (called Schubert determinantal ideal ) is generated by all minors of size rp,q(w) + 1 in X[p,q]. One can reduce the generating set of Iwas the following. Consider the diagram of w

D(w) =

(i, j)∈ [1, l] × [1, m] : w(i) > j and w−1(j) > i ,

that is, D(w) consists of elements that are neither due east nor due south of a nonzero entry of w. The essential set of w is defined to be

E(w) =

(p, q)∈ D(w) : (p + 1, q) /∈ D(w) and (p, q + 1) /∈ D(w) . One can check that (see [Ful92, Lemma 3.10])

Iw=

minors of size rp,q(w) + 1 in X[p,q]: (p, q)∈ D(w) (2.1)

=

minors of size rp,q(w) + 1 in X[p,q]: (p, q)∈ E(w) .

Also, the codimension of Xwin Ml×mis the cardinality|D(w)| of D(w) which is actually the Coxeter length of w when w is a permutation. We often need to consider certain subsets of D(w) orE(w). For that, we will put the conditions as subscripts to indicate the constraints. For examples, D=0(w) ={(p, q) ∈ D(w)| rp,q(w) = 0} and D>0(w) ={(p, q) ∈ D(w) | rp,q(w) > 0}.

Questions on Xw for a partial permutation w∈ Ml×m is often reduced to the cases where w is a permutation. More precisely, extend w to the

(5)

permutationw∈ Sn, n = l + m via

 w(i) =

⎧⎪

⎪⎩

j if wi,j= 1 for some j,

min{[m + 1, n] \ { w(1), . . . ,w(i − 1)}} if wi,j= 0 for all j, min{[1, n] \ { w(1), . . . ,w(i − 1)}} if i > l.

Then D(w) = D(w), E(w) = E( w), and the defining ideals I w and Iw share the same set of generators. Therefore,

(2.2) Xw∼= Xw× Kn2−lm.

The following substantial results due to A. Knutson and E. Miller is in- dispensable in the proofs of our main theorems. Recall that a term order on K[X] is called antidiagonal if the initial term of every minor of X is its antidiagonal term. We will fix an antidiagonal term order < and simply write in(I), in(f ) as the initial ideal of an ideal I and the leading term of an ele- ment f , respectively. We will call an antidiagonal term of a minor of size r an antidiagonal of size r.

Theorem 2.1 ([KM05]). Fix any antidiagonal term order. Then (1) The generators of Iw in (2.1) constitute a Gr¨obner basis, that is,

in(Iw) =

antidiagonals of size rp,q(w) + 1 in X[p,q]: (p, q)∈ E(w)

; (2) in(Iw) is a Cohen–Macaulay square-free monomial ideal.

3. F -rationality and F -injectivity

Recall that in the theory of tight closure, a Noetherian ring is F -rational if all its parameter ideals are tightly closed. There is a weaker notion called F -injectivity. A Noetherian ring R is F -injective if for any maximal ideal m of R the map on the local cohomology module Hmi(R) induced by the Frobenius map is injective for all i. We collect some facts concerning F -rationality and F -injectivity. See [Hun96], or [BH93] for convenient resources.

Theorem 3.1. Let R be a Noetherian ring.

(1) R is F -rational if and only if Rm is F -rational for any maximal ideal m.

(2) If R is an F -rational ring that is a homorphic image of a Cohen–Macaulay ring, then RS is F -rational for any multiplicative closed set S of R.

Theorem 3.2. Let (R, m) be a Noetherian local ring.

(1) R is F -injective if and only if R is F -injective.

(2) Suppose in addition R is excellent, then R is F -rational if and only if R is F -rational.

Theorem3.3. Let R be a positive graded K-algebra, where K is a field of positive characteristic. Let m be the unique maximal graded ideal of R.

(1) R is F -injective if and only if Rm is F -injective.

(6)

(2) R is F -rational if and only if R[T ] is F -rational for any variable T over R.

(3) Suppose in addition that K is perfect. Then R is F -rational if and only if Rm is F -rational.

4. Matrix Schubert varties are F -rational

Fix an antidiagonal term order. Denote Jw= in(Iw) the initial ideal of Iw. In this section, the ground field K is perfect and of positive characteristic. As mentioned in theIntroduction, consider c := xi0,w(i0)where i0 is the smallest number such that{(p, q) | p > i0, q > w(i0)} ∩ E>0(w)= ∅. Note that such i0

exists exactly whenE>0(w)= ∅ (or equivalently Rw is not regular). We make this assumption (existence of c) for Lemmas 4.1, 4.2, 4.3 and set (p0, q0) = (i0, w(i0)). Note also that for this particular choice of i0,

(p, q)= (p0, q0)| p ≤ p0, q≤ q0

 

(p, q)| p ≤ p0

∩ D(w) (4.1)

⊆ D=0(w).

In particular, the only nonzero entry in w[p0,q0] is (p0, q0).

In the following, we use the notation [p1, . . . , pt| q1, . . . , qt] to denote the size t minor of the submatrix of X involving the rows of indices p1, . . . , ptand the columns of indices q1, . . . , qt.

Lemma4.1. Let Δ be any minor in X such that c| in Δ. Then Δ ∈ c+Jw

and hence so is Δ− in Δ.

Proof. Write Δ = [pt, . . . , p1, p0, p1, . . . , ps| qs, . . . , q1, q0, q1, . . . , qt], so in Δ =

 s



i=1

xp i,qi



· c ·

 t



j=1

xpj,qj

 ,

where ps>· · · > p1> p0> p1>· · · > ptand qs<· · · < q1 < q0< q1<· · · < qt. Use induction on t. When t = 0, then Δ = [p0, p1, . . . , ps| qs, . . . , q1, q0]. If s = 0, Δ = c = in Δ is obviously in c + Jw. Suppose s > 0, expanding Δ along the first row, we get

Δ = (−1)sc

p1, . . . , ps| qs, . . . , q1 +

s i=1

(−1)i+1xp0,qi

p1, . . . , ps| qs, . . . , qi, . . . , q1, q0

.

Since 0 = rp0,q1(w) =· · · = rp0,qs(w) by (4.1), xp0,q1, . . . , xp0,qs ⊆ Jw and hence Δ∈ c + Jw.

Suppose t > 0. For 1≤ i ≤ t, set Δi=

pt−1, . . . , p1, p0, p1, . . . , ps| qs, . . . , q1, q0, q1, . . . ,q i, . . . , qt

. Note that c| in Δi for 1≤ i ≤ t, since c is on the antidiagonal of Δi (the row deleted is above c and the column deleted is to the right of c). SoΔi| 1 ≤

(7)

i≤ t ⊆ c + Jw by the inductive hypothesis. Again, expanding Δ along the first row, we see that

Δ∈ Δi| 1 ≤ i ≤ t + xpt,qj | j = 1, . . . , s + xpt,q0.

Once again [xpt,qj | j = 1, . . . , s + xpt,q0] ⊆ Jw, since 0 = rpt,q1(w) =· · · = rpt,q

s(w) = rpt,q0(w) by (4.1). Therefore, Δ∈ c + Jw as desired.  Lemma 4.2. in(c + Iw) =c + Jw.

Proof. The containment in(c + Iw)⊇ c + Jw is obvious. Conversely, let cf− g ∈ c + Iw for some f∈ K[X] and g ∈ Iw. If in(cf )= in(g), then in(cf− g) = in(cf) or in(−g). In either case, in(cf − g) ∈ c + Jw.

So we may assume that in(cf ) = in(g). Write g = m1Δ1+ m2Δ2+· · · + muΔuwhere the mi’s are monomial elements in K[X] and the Δi’s are minors in the generating set of Iw. If in(cf− g) is a term in cf, then in(cf − g) ∈ c.

Also, if in(cf− g) = in(miΔi) for some i, then in(cf− g) = miin(Δ˙ i)∈ Jw. Therefore, we may assume that in(cf− g) is a term of mi0Δi0 for some i0

and that in(cf− g) is neither in(mi0Δi0) nor a term of cf . This implies that in(mi0Δi0) is a term of cf and hence c| in(mi0Δi0) = mi0in(Δ˙ i0). If c| mi0, then in(cf− g) ∈ c since it is a term of mi0Δi0. Otherwise, c| in(Δi0). Then by Lemma 4.1, Δi0− in(Δi0)∈ c + Jw. Therefore, mi0Δi0− in(mi0Δi0)

c + Jw. Now, since in(cf− g) is a term of mi0Δi0− in(mi0Δi0) and since

c + Jw is a monomial ideal, we conclude that in(cf− g) ∈ c + Jw.  Lemma 4.3. Rw/cRw is F -injective.

Proof. By Theorem 2.1 in [CH97], it suffices to show that K[X]/ in(c+Iw) is Cohen–Macaulay and F -injective. By Lemma4.2, in(c + Iw) =c + Jw. Also, by Theorem2.1(2) Jwis a Cohen–Macaulay square-free monomial ideal.

Soc + Jw is also a square-free monomial ideal, and hence K[X]/(c + Jw) is F -injective by the discussion in the paragraph before corollary of [CH97]

involving Fedder’s criterion. The Cohen–Macaulayness of K[X]/(c + Jw) follows from that fact the c is a nonzero-divisor on K[X]/Jw.

To see this, first note that c = xi0,w(i0)= xp0,q0 ∈ J/ w. Suppose for some z∈ K[X] we have cz ∈ Jw. We will show that z∈ Jw. By Theorem 2.1, we may assume z is a monomial and cz = rD for some monomial r∈ K[X] and some antigonal D∈ Jw. If c| r, then z =rcD∈ Jw. Therefore, we may assume c r. Then c | D. We finish the proof by showing that Dc ∈ Jw.

Write

D =

 s



i=1

xpi,qi



· c ·

 t



j=1

xpj,qj

 ,

where p1> p2>· · · > ps> p0> p1>· · · > pt and q1 < q2<· · · < qs < q0<

q1<· · · < qt. Since c = xp0,q0∈ J/ w, either s > 0 or t > 0. Note also that D is

(8)

of size (s + t + 1), so rp1,qt(w)≤ s + t by Theorem2.1(1). On the other hand, as mentioned before the only nonzero entry in w[p,q] is (p0, q0), so

{nonzero entries in w[p1,qt]}



(p0, q0)

{nonzero entries in w[p1,qt]} {nonzero entries in w[p1,qs]}.

Therefore, 1 + rp1,qt(w) + rp1,q

s(w)≤ rp1,qt≤ s+t and hence either rp1,qt(w) <

t or rp1,q

s(w) < s. We conclude that either s

i=1xpi,qi∈ Jw ort

i=1xpi,qi

Jw, so Dc ∈ Jw. 

Theorem 4.4. Rw= K[X]/Iw is F -rational.

Proof. By (2.2) and Theorem 3.3(2), we may assume that w∈ Sn is a permutation. Use induction on n. If Rw is regular (this includes the cases n = 1, 2), then it is F -rational. Suppose n > 2 and Rwis not regular. Then the element c = xp0,q0 described as above exists. By Lemma 4.3, Rw/cRw is F - injective. Hence, Rw/c Rw is F -injective by Theorem3.2(1) and3.3(1). So by Theorems1.2,3.2(2) and3.3(3), it suffices to show that Rw[1c] is F -rational.

For (p, q) /∈ Γ = {(p, q) | p = p0 or q = q0}, consider the change of variables xp,q= xp,q− c−1xp,q0xp0,q.

Set X= (xp,q| (p, q) /∈ Γ). Then K[X]

1 c



= K X

xp,q| (p, q) ∈ Γ1 c



in the field of fraction of K[X].

Let w be the permutation obtained by deleting the p0th row and the q0th column of w and let Iw be the corresponding Schubert determinantal ideal in K[X]. Denote I = Iw· K[X][1c] the extended ideal of Iwand set

I= Iw+

xp,q| (p, q) ∈ Γ, p < p0or q < q0

· K X

xp,q| (p, q) ∈ Γ1 c

 . We claim that

(4.2) I = I.

It follows from (4.2) that K[X]

Iw

1 c



=K[X] Iw

xp,q| (p, q) ∈ Γ, p ≥ p0 and q≥ q0

1 c

 .

By inductive hypothesis, K[X]/Iw is F -rational. So Theorem 3.1(2) and Theorem3.3(2) imply that Rw[1c] is F -rational. Therefore, it suffices to prove (4.2).

We prove (4.2) by showing that the generators of I belongs to I and conversely. First observe that

(a) For (p, q)∈ Γ satisfying p < p0 or q < q0, by (4.1) xp,q∈ Iw.

(9)

(b) Fix a (p, q) satisfying p < p0 or q < q0. Let Δ be an r-minor (r≥ 1) of X[p,q] that does not involve the p0th row and the q0th column. Denote Δthe corresponding r-minor in X[p,q](replace xp,qby xp,q). Then direct computation shows

Δ− Δ

xp,q| (p, q) ∈ Γ, p < p0 or q < q0 .

By (a), xp,q| (p, q) ∈ Γ, p < p0 or q < q0 ⊆ I, so we see that Δ − Δ I∩ I.

(c) Let Δ be any r-minor in X that involves Γ but does not involve c. Then Δ = Δ where Δ is obtained from Δ by replacing xp,q((p, q) /∈ Γ) by xp,q. (d) Let Δ be any r-minor in X that does not involve Γ and let Δ the cor- responding r-minor in X (replace xp,q by xp,q). Denote Δ and Δ the (r + 1)-minors obtained by adding the p0th row and q0th column to Δ and Δ , respectively. Then

cΔ = Δ and = Δ.

Now, we are ready to prove (4.2), I = I.

We first show that I⊆ I. Fix (p, q)∈ D=0(w)∪ E>0(w). We need to show that the following set

xp,q| (p, q) ∈ D=0(w)

 

(p,q)∈E>0(w)

rp,q(w) + 1

-minors in X[p,q]

is contained in I. Consider the following cases.

(i) (p, q)∈ D=0(w). We must have p < p0 or q < q0. (i.1) If (p, q)∈ Γ, then xp,q∈ I by construction.

(i.2) If (p, q) /∈ Γ, then (p, q) ∈ D=0(w) and hence xp,q∈ Iw⊆ I. There- fore, xp,q= xp,q+ c−1xp0,qxp,q0∈ I, since either xp0,q or xp,q0 is in I.

(ii) (p, q)∈ E>0(w), say rp,q(w) = r. Let Δ be any (r + 1)-minor in X[p,q]. In this case, p > p0 by (4.1).

(ii.1) q < q0. In this case, (p, q)∈ E=r(w). If Δ involves the p0th row, expanding along this row we see that Δ∈ xp0,q| q < q0 ⊆ I. Oth- erwise, let Δ be the corresponding (r + 1)-minor in X[p,q] and we have Δ− Δ∈ xp0,q| q < q0 ⊆ I by (b). But (p, q)∈ E=r(w) im- plies Δ∈ Iw⊆ I. So Δ∈ I.

(ii.2) q > q0. In this case, (p, q)∈ E=r−1(w).

(ii.2.1) If Δ involves c, then Δ = c(Δ\ [p0| q0]) by (d), where Δ\ [p0| q0] is the r-minor obtained from Δ by deleting the row and the column involving c. But (p, q)∈ E=r−1(w) implies (Δ\ [p0| q0])∈ Iw⊆ I. So Δ∈ I.

(ii.2.2) If Δ involves Γ but does not involve c, then by (c) Δ = Δ where Δis obtained from Δ by replacing xp,q((p, q) /∈ Γ) by

(10)

xp,q. Expanding Δalong the row or column involving Γ, we see that Δ∈ r-minors in X[p,q]. But (p, q) ∈ E=r−1(w) impliesr-minors in X[p,q] ⊆ Iw⊆ I. So Δ = Δ∈ I. (ii.2.3) If Δ does not involve Γ, then cΔ = Δ by (d). Expand-

ing the (r + 2)-minor Δ along the row and the column involving Γ, we see that Δ∈ r-minors in X[p,q]. Again,

r-minors in X[p,q] ⊆ I since (p, q)∈ E=r−1(w). So Δ = c−1∈ I.

Conversely, we show that I⊆ I. Fix (p, q) ∈ D=0(w)∪ E>0(w). Again, we show that the set

xp,q| (p, q) ∈ D=0

w

 

(p,q)∈E>0(w)

rp,q w

+ 1

-minors in X[p,q]

is contained in I.

(i) (p, q)∈ D=0(w).

(i.1) If p < p0 or q < q0, then (p, q)∈ D=0(w) and hence xp,q∈ I. By (a), either xp0,q or xp,q0 is in I. Therefore,

xp,q= xp,q− c−1xp0,qxp,q0∈ I.

(i.2) If p < p0 and q > q0, then (p, q)∈ E=1(w). Hence, the 2-minor cxp,q= cxp,q− xp0,qxp,q0∈ Iw⊆ I. So xp,q∈ I.

(ii) (p, q)∈ E>0(w). In this case, p > p0 by (4.1). Suppose rp,q(w) = r and let Δ be any (r + 1)-minor in X[p,q].

(ii.1) If q < q0, then (p, q)∈ E=r(w). By (d), cΔ = Δ. Expanding Δ along the q 0th column, we see Δ∈ (r + 1)-minors in X[p,q].

But (p, q)∈ E=r(w) implies(r + 1)-minors in X[p,q] ⊆ I. So Δ= c−1Δ∈ I.

(ii.2) If q > q0, then (p, q)∈ E=r+1(w). Again, cΔ= Δ by (d). This time Δ ∈ Iw⊆ I since (p, q) ∈ E=r+1(w). Therefore, Δ= c−1Δ ∈ I. 

Example4.5. Consider w = 35142 in S5. Use the same notation as in the proof of Theorem 4.4. We have c = x13, Iw=x11, x12, x21, x22, [12| 34], [34 | 12] and Iw=x21, x22, x24, [34| 12]. Check that the following elements lie in x11, x12 and hence in I ∩ I:

x21− x21, x22− x22, [34| 12] − [34 | 12], [12| 34] − c−1x24. Therefore, we see that I = I and Rw[c−1] = Rw[x11, x12, c−1].

In the following diagram, the 1’s indicate the permutation and the dots indicate the elements in D>0(w).

(11)

1

1

1

1

1

5. Complete intersection matrix Schubert varieties

We want to characterize the complete intersection MSVs. By (2.2), we may assume w∈ Sn. Denote w(p,q) the rp,q(w)× rp,q(w) submatrix of w involving the rows of indices p− rp,q(w), . . . , p− 1 and the columns of in- dices q− rp,q(w), . . . , q− 1. Define the submatrix X(p,q) of X similarily.

Furthermore, denote X(p,q) the (rp,q(w) + 1)× (rp,q(w) + 1) submatrix of X involving the rows of indices p− rp,q(w), . . . , p and the columns of in- dices q− rp,q(w), . . . , q. If rp,q(w) = 0, X(p,q)= det X(p,q)= xp,q. However, in order to make our proof more transparent, we will only use X(p,q) for (p, q)∈ D>0(w).

Recall that the codimension of Xwin Mn×nis|D(w)|. So Xwis a complete intersection if and only if Iw can be generated by|D(w)| many elements. We need the following lemma.

Lemma 5.1. Let w be such that Xw is a complete intersection. Then for any (p, q)∈ D>0(w) and any 1≤ i ≤ rp,q(w),

(p− i, q) /∈ D(w) and (p, q − i) /∈ D(w).

In particular,

D>0(w)⊆ E(w), that is, D>0(w) =E>0(w).

Proof. We only have to prove the first statement. The second statement follows by definition. Suppose the first statement does not hold, then by symmetry we may assume that there exist (p0, q0) in D>0(w) and some 1≤ i0≤ rp0,q0(w) so that (p0− i0, q0)∈ D(w) and (p0− j, q0) /∈ D(w) for all 1 ≤ j < i0.

Denote r := rp0,q0(w). Consider the (r + 1)-minor

Δ = [p0− r − 1, . . . , p0− i0, . . . , p0| q0− r, . . . , q0]

in the (r + 2)× (r + 1) submatrix XΔ of X involving the rows of indices p0− r − 1, . . . , p0 and the columns of indices q0− r, . . . , q0.

Consider the set G =

xp,q| (p, q) ∈ D=0(w) 

det X(p,q)| (p, q) ∈ D>0(w) {Δ}.

Observe that the unions are disjoint and |G| = |D(w)| + 1. Denote m the maximal graded ideal of K[X]. We claim that

the image of G in Iw/mIw form a (5.1)

K

= K[X]/m

-linearly independent set.

(12)

By Nakayama’s lemma, Iwis generated by at least |D(w)| + 1 elements. This contradicts the assumption that Xw is a complete intersection.

It remains to show the claim (5.1). Suppose

 

(p,q)∈D=0(w)

cp,qxp,q

 +

 

(p,q)∈D>0(w)

cp,qdet X(p,q)



+ cΔΔ∈ mIw, where cp,q and cΔ are in K. Notice mIw is a homogeneous ideal whose gen- erators are of degree at least 2. This implies cp,q= 0 for all (p, q)∈ D=0(w) since otherwise we will have an element in mIw that has a nonzero degree 1 part. Therefore,

F :=

 

(p,q)∈D>0(w)

cp,qdet X(p,q)



+ cΔΔ∈ mIw.

Fix any antidiagonal term order on K[X]. Then in(F ) is the antidiagonal term of one of the minors in

det X(p,q)| (p, q) ∈ D>0(w) {Δ}.

So in(F ) is in the generating set of in(Iw) described in Theorem 2.1(1). On the other hand, F ∈ mIw implies that there exists an antidiagonal δ in the generating set of in(Iw) described in Theorem 2.1(1) such that δ is a factor of in(F ) but δ= in(F ). This means that δ ∈ in(Iw) is an antidiagonal in one of the submatrices of the matrices in

X(p,q)| (p, q) ∈ D>0(w)

{XΔ},

and that δ is of size≤ rp,q(w) (if δ is in X(p,q)) or of size≤ rp0,q0(w) (if δ is in XΔ). This is impossible in view of Theorem2.1(1). Therefore, we conclude that cp,q(w) = cΔ= 0 for all (p, q)∈ D(w) and the claim (5.1) is proved. 

Now we are ready for the following theorem.

Theorem 5.2. Let w∈ Sn. The following conditions are equivalent:

(1) Xw is a complete intersection,

(2) for any (p, q)∈ D>0(w), w(p,q) is a permutation in Srp,q(w) such that Xw(p,q) is a complete intersection.

In this case,

xp,q| (p, q) ∈ D=0(w) 

det X(p,q)| (p, q) ∈ D>0(w) is a set of generators for Iw.

Proof. The conditions in (2) shows that for any (p, q)∈ D>0(w) the only nonzero entries of w[p,q]appear in w(p,q), so all size (rp,q(w) + 1) minors except det X(p,q) belong to xp,q| (p, q) ∈ D=0(w). Therefore, the last statement follows immediately from (2.1) and the equivalence of (1) and (2).

(13)

We prove (2) implies (1). Let (p1, q1), . . . , (pt, qt) be all the elements in D>0(w) satisfying

(p, q)| p ≥ pi, q≥ qi

∩ D>0(w) =∅.

Denote ri= rpi,qi(w) and wi= w(pi,qi)∈ Sri. By the assumptions in (2), the diagram D(wi) of wi is contained in D(w). Also, the ideal Iwi of Xwi is generated by |D(wi)| many elements, since Xwi is a complete intersection.

Note also that the conditions in (2) and the choice of (pi, qi) imply that D(w) can be decomposed as



(p, q) (p, q) ∈ D=0(w)

t

i=1

D(wi)

 t



i=1

D(wi)

 

(pi, qi)| 1 ≤ i ≤ t ,

where the unions are disjoint. Furthermore, by construction one can check using (2.1) that

Iw=

!

xp,q (p, q) ∈ D=0(w)

t

i=1

D(wi)

"

+

t i=1

Iwi+

t i=1

det X(pi,qi).

So Iw is generated by







(p, q)∈ D=0(w) (p, q) /∈

t i=1

D(wi) +

t i=1

D(wi)+ t =D(w)

many elements. Therefore, Xw is a complete intersection.

To prove (1) implies (2), let (p, q)∈ D>0(w) and use induction on rp,q(w).

When rp,q(w) = 1, we must have wp−1,q−1= 1 since otherwise either (p− 1, q)∈ D=1(w) or (p, q− 1) ∈ D=1(w) which contradicts Lemma 5.1. Also, Xw(p,q) is the affine line, so we are done for rp,q(w) = 1.

Suppose rp,q(w) > 1 and denote r = rp,q(w). By Lemma5.1, (p− i, q) /∈ D(w) and (p, q − i) /∈ D(w), for i = 1, . . . , r.

This implies that rank(w(p,q)) = r and w(p,q)∈ Sr. Moreover, consider the diagram D(w(p,q)). This is exactly the part of the diagram D(w) involving the rows of indices p− r, . . . , p − 1 and the columns of indices q − r, . . . , q − 1.

For any (p, q)∈ D>0(w(p,q)), rp,q(w(p,q)) = rp,q(w) < r. So by inductive hypothesis, w(p,q)∈ Sr satisfies the conditions in (2). Therefore, by the im- plication of (2) ⇒ (1) we just proved, Xw(p,q) is a complete intersection as

desired. 

Example 5.3. Consider w∈ S6. Again, In the following diagrams the 1’s indicate the permutation and the dots indicate the elements in D>0(w).

(1) If w = 361452, then Xw is not a complete intersection since D>0(w) = {(2, 4), (2, 5), (4, 2), (5, 2)} but (2, 5) /∈ E>0(w).

(14)

1

1

1

1

1

1

(2) If w = 352614, then D>0(w) ={(2, 4), (4, 4)} = E>0(w). But Xw is still not a complete intersection since w(4,4)∈ S/ 2. So we see that the condition D>0(w) =E>0(w) is not sufficient for Xw to be a complete intersection.

1

1

1

1

1

1

(3) If w = 462153, then Xw is a complete intersection, and Iw is generated by{xi,j| 1 ≤ i ≤ 2, 1 ≤ j ≤ 3} ∪ {x31, det X(2,5), det X(5,3)}.

1

1

1 1

1

1

Acknowledgments. The author would like to thank A. Knutson, E. Miller, K. Smith, U. Walther and A. Woo for their comments and suggestions about this work. Special thanks go to the referee for carefully reading this paper and many useful suggestions on the presentation of the results.

References

[BH93] W. Bruns and J. Herzog, Cohen–Macaulay rings, Cambridge Studies in Advanced Mathematics, vol. 39, Cambridge Univ. Press, Cambridge, 1993.MR 1251956

[BK05] M. Brion and S. Kumar, Frobenius splitting methods in geometry and repre- sentation theory, Progress in Mathematics, vol. 231, Birkh¨auser, Boston, MA, 2005.MR 2107324

[Bri05] M. Brion, Lectures on the geometry of flag varieties, Topics in cohomological studies of algebraic varieties, Trends Math., Birkh¨auser, Basel, 2005, pp. 33–

85.MR 2143072

[CH97] A. Conca and J. Herzog, Ladder determinantal rings have rational singularities, Adv. Math. 132 (1997), no. 1, 120–147.MR 1488240

[Con95] A. Conca, Ladder determinantal rings, J. Pure Appl. Algebra 98 (1995), no. 2, 119–134.MR 1319965

[Con96] A. Conca, Gorenstein ladder determinantal rings, J. Lond. Math. Soc. (2) 54 (1996), no. 3, 453–474.MR 1413891

[Ful92] W. Fulton, Flags, Schubert polynomials, degeneracy loci, and determinantal formulas, Duke Math. J. 65 (1992), no. 3, 381–420.MR 1154177

(15)

[GL00] N. Gonciulea and V. Lakshmibai, Singular loci of ladder determinantal varieties and Schubert varieties, J. Algebra 229 (2000), no. 2, 463–497.MR 1769284 [GM00] N. Gonciulea and C. Miller, Mixed ladder determinantal varieties, J. Algebra

231 (2000), no. 1, 104–137.MR 1779595

[GS95] D. Glassbrenner and K. E. Smith, Singularities of certain ladder determinantal varieties, J. Pure Appl. Algebra 101 (1995), no. 1, 59–75.MR 1346428 [Har98] N. Hara, A characterization of rational singularities in terms of injectivity of

Frobenius maps, Amer. J. Math. 120 (1998), no. 5, 981–996.MR 1646049 [Hun96] C. Huneke, Tight closure and its applications, CBMS Regional Conference

Series in Mathematics, vol. 88, published for the Conference Board of the Mathematical Sciences, Washington, DC, 1996. With an appendix by Melvin Hochster.MR 1377268

[KM05] A. Knutson and E. Miller, Gr¨obner geometry of Schubert polynomials, Ann. of Math. (2) 161 (2005), no. 3, 1245–1318.MR 2180402

[LRPT06] N. Lauritzen, U. Raben-Pedersen and J. F. Thomsen, Global F -regularity of Schubert varieties with applications to D-modules, J. Amer. Math. Soc. 19 (2006), no. 2, 345–355 (electronic).MR 2188129

[LS90] V. Lakshmibai and B. Sandhya, Criterion for smoothness of Schubert vari- eties in Sl(n)/B, Proc. Indian Acad. Sci. Math. Sci. 100 (1990), no. 1, 45–

52.MR 1051089

[MS05] E. Miller and B. Sturmfels, Combinatorial commutative algebra, Graduate Texts in Mathematics, vol. 227, Springer, New York, 2005.MR 2110098

[Ram85] A. Ramanathan, Schubert varieties are arithmetically Cohen–Macaulay, Invent.

Math. 80 (1985), no. 2, 283–294.MR 0788411

[Smi97] K. E. Smith, F -rational rings have rational singularities, Amer. J. Math. 119 (1997), no. 1, 159–180.MR 1428062

[ ´UW13] H. ´Ulfarsson and A. Woo, Which Scubert varieties are local complete intersec- tions? Proc. Lond. Math. Soc. (3) 107 (2013), 1004–1052.MR 3126390 [WY06] A. Woo and A. Yong, When is a Schubert variety Gorenstein? Adv. Math. 207

(2006), no. 1, 205–220.MR 2264071

Jen-Chieh Hsiao , Department of Mathematics, National Cheng Kung Univer- sity, No.1 University Road, Tainan 70101, Taiwan

E-mail address:jhsiao@mail.ncku.edu.tw

參考文獻

相關文件

By the Lebesgue’s theorem we know that if f is (Riemann) integrable on A, then the discontinuities of f

Compute Df (x) and

Lemma 86 0/1 permanent, bipartite perfect matching, and cycle cover are parsimoniously equivalent.. We will show that the counting versions of all three problems are in

The first typ e is the unsafe drivers, who disregard traffic regulations and make life difficult for other drivers and people walking in the street.. Unsafe drivers do

This paper proposes that the popularization and secularization of Buddhism are problems that have emerged, and are also related to the heritage of folk traditions in

The classification theory of varieties usually reduced to the study of varieties of the following three types: varieties of general type, varieties with Kodaira dimension zero

In the two dimensional case, singularities admitting projective crepant resolu- tions are exactly canonical singularities, and have been studied extensively for a long

In §2, we prove a more general version of Theorem 2.1.3 below, which applies to irregular varieties not necessarily of maximal Albanese dimension.. Using this result we are able