• Aucun résultat trouvé

Characterizations of freeness for Cohen-Macaulay spaces

N/A
N/A
Protected

Academic year: 2021

Partager "Characterizations of freeness for Cohen-Macaulay spaces"

Copied!
29
0
0

Texte intégral

(1)

HAL Id: hal-01248123

https://hal.archives-ouvertes.fr/hal-01248123v2

Preprint submitted on 28 Nov 2016

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Characterizations of freeness for Cohen-Macaulay spaces

Delphine Pol

To cite this version:

Delphine Pol. Characterizations of freeness for Cohen-Macaulay spaces. 2015. �hal-01248123v2�

(2)

DELPHINE POL

Abstract.

The purpose of this paper is to develop an analogue of freeness for reduced Cohen- Macaulay spaces embedded in a smooth manifold

S

which generalizes the notion of Saito free divisors. The main result of this paper is a characterization of freeness in terms of the projective dimension of the module of multi-logarithmic

k-forms, wherek

is the codimension of

X

in

S. We

then focus on quasi-homogeneous complete intersection curves, for which we compute explicitly a minimal free resolution of the module of multi-residues and the module of multi-logarithmic forms.

1. Introduction

Let D be the germ of a reduced hypersurface in a smooth analytic variety S. A meromorphic form with simple poles along D is called logarithmic if its differential has also simple poles along D. When the module of logarithmic 1-forms is free, the hypersurface D is called a free divisor. For example, normal crossing divisors, plane curves and discriminants of isolated hypersurface singularities are families of free divisors (see [Sai80]).

Several characterizations of freeness are known. Let us mention some of them. Thanks to the duality between the modules of logarithmic vector fields and logarithmic differential 1-forms proved in [Sai80], a hypersurface is free if and only if the module of logarithmic vector fields is free. Another characterization of freeness is the following: a hypersurface is free if and only if the Jacobian ideal is Cohen-Macaulay (see [Ale88]).

A theory of multi-logarithmic differential forms and multi-residues for a reduced complete inter- section is developed in [AT01], [AT08], [Ale12] and studied in [GS12]. A generalization of this theory to Cohen-Macaulay spaces is developed in [Ale14], which is inspired by the interpretation of regular meromorphic forms with residue symbols developed in [Ker84], which is also used in [Sch16].

Let X be a reduced Cohen-Macaulay space of codimension k contained in a smooth complex manifold S, and I

X

the defining ideal of X. There exists a reduced complete intersection C defined by a regular sequence f = (f

1

, . . . , f

k

) such that X ⊆ C. We set I

C

= P

k

i=1

f

i

O

S

and f = f

1

· · · f

k

. The module of multi-logarithmic differential q-forms with respect to the pair (X, C) is defined by:

(1) Ω

q

(log X/C, f) =

ω ∈ 1

f Ω

qS

; I

X

· ω ⊆ 1

f I

C

qS

and d(I

X

) ∧ ω ⊆ 1

f I

C

q+1S

.

Following the definition of freeness for complete intersection suggested in [GS12], we say that X is free if the restriction to X of the Jacobian ideal of C is Cohen-Macaulay. We prove in proposition 4.4 that this notion of freeness is independent from the choice of the complete intersection C.

The natural problem which then arises is to characterize the freeness of a Cohen-Macaulay space thanks to the module of multi-logarithmic forms. It leads to the main theorem of this paper (see theorem 4.6):

Theorem . A reduced Cohen-Macaulay space X contained in a reduced complete intersection C of the same dimension is free if and only if the projective dimension of Ω

k

(log X/C, h) is lower than or equal to k − 1, which is equivalent to the fact that the projective dimension of Ω

k

(log X/C ) is equal to k − 1.

Date : November 27, 2016.

1991 Mathematics Subject Classification. 14M10, 14B05, 13D02, 32A27.

Key words and phrases. logarithmic differential forms, logarithmic residues, complete intersections, curves.

1

(3)

For a free reduced Cohen-Macaulay space of codimension at least 2, the module Ω

k

(log X/C ) is no more free, however, it is of minimal projective dimension.

We then focus on a particular family of free singularities, namely the curves. The computation of multi-logarithmic forms is in general difficult. In this paper, we determine explicitly a free resolution of the module of multi-logarithmic forms for quasi-homogeneous complete intersection curves.

Let us describe more precisely the content of this paper.

We first consider in subsection 2.1 the complete intersection case, which is less technical and is used for the Cohen-Macaulay case.

Let C be a reduced complete intersection defined by a regular sequence (h

1

, . . . , h

k

). We set h = h

1

· · · h

k

, and Ω

q

(log C, h) = Ω

q

(log C/C, h).

In subsection 2.1.1 we first recall several definitions and properties of multi-logarithmic forms and multi-residues from [AT01], [AT08] and [Ale12]. We set R

q−kC,h

for the module of multi-residues of multi-logarithmic q-forms, and R

C,h

= R

0C,h

(see definition 2.7).

In subsections 2.1.1 and 2.1.2 we make explicit the dependence of the modules Ω

q

(log C, h) and R

q−kC,h

on the choice of the equations. We prove that the modules of numerators h

1

· · · h

k

·Ω

q

(log C, h) do not depend on the choice of the equations (see proposition 2.3). In addition, the modules R

q−kC,h

depends only on C (see proposition 2.9). We then fix a regular sequence (h

1

, . . . , h

k

) and we denote Ω

q

(log C) = Ω

q

(log C, h) and R

qC

= R

qC,h

.

Let C e be the normalization of C. A consequence of [Sch16, Proposition 2.1] is that O

Ce

⊆ R

C

. We give here another proof of this result which is similar to the proof of [Sai80, Lemma 2.8] (see proposition 2.13).

Following [GS12], we define the module of multi-logarithmic k-vector fields as the kernel of the map: V

k

Θ

S

→ J

C

, δ 7→ dh

1

∧ · · · ∧ dh

k

(δ), where J

C

is the Jacobian ideal of C.

In subsection 2.1.4, we show that the fact that C is reduced imposes a strong condition on the module of logarithmic 1-forms of the divisor D defined by h

1

· · · h

k

. We denote by D

i

the hypersurface defined by h

i

. Then we have (see proposition 2.23):

1

(log D) = Ω

1

(log D

1

) + · · · + Ω

1

(log D

k

).

In section 2.2, we consider first a reduced equidimensional space X of dimension n. Let C be a reduced complete intersection of dimension n containing X defined by a regular sequence (f

1

, . . . , f

k

). We first recall definitions and properties from [Ale14]. For all q ∈ N , the module f

1

· · · f

k

q

(log X/C, f) depends only on X and C, so that we set Ω

q

(log X/C) = Ω

q

(log X/C, f ).

We have the inclusion Ω

q

(log X/C ) ⊆ Ω

q

(log C). The multi-residue of a form ω ∈ Ω

q

(log X/C) is defined as the restriction to X of the multi-residue of ω seen as a multi-logarithmic form on C.

Let c

X

be the fundamental class of X (see notation 2.36). We prove the following characterization of multi-logarithmic differential forms with respect to the pair (X, C), which generalizes [Ale12, §3, theorem 1] and [Sai80, (1.1)] (see proposition 2.38): a meromorphic q-form ω ∈

f1

qS

is multi- logarithmic if and only if there exist g ∈ O

S

which induces a non zero divisor in O

C

, ξ ∈ Ω

q−kS

and η ∈ Ω e

qf

such that gω = c

X

f + η.

We suggest the following definition of multi-logarithmic k-vector fields: a holomorphic k-vector field δ is multi-logarithmic if δ(c

X

) ∈ I

X

. We then have the following perfect pairing which gener- alizes the duality of the hypersurface case (see proposition 3.3):

Der

k

(− log X/C) × Ω

k

(log X/C ) → 1 f I

C

.

We denote by J

X/C

the restriction to X of the Jacobian ideal of C. In subsection 3.2, using

an approach which is similar to the proof of [GS14, Proposition 3.4], and which uses the previous

perfect pairing, we prove that Hom

OC

(J

X/C

, O

C

) = R

X

, where R

X

is the module of multi-residues

of multi-logarithmic k-forms (see proposition 3.11).

(4)

Section 4 is devoted to the main results of this paper, namely, the characterizations of freeness for Cohen-Macaulay spaces. Since a hypersurface is free if the module of logarithmic vector fields is free, or equivalently, the module of logarithmic 1-forms is free, it is then natural to investigate conditions on the projective dimensions of Ω

k

(log X/C ) and Der

k

(− log X/C ) to characterize the freeness of a Cohen-Macaulay space.

We prove that a reduced Cohen-Macaulay space is free if and only if the projective dimension of Der

k

(− log X/C) is k − 1 (see proposition 4.5). Contrary to the hypersurface case, passing from Der

k

(− log X/C) to Ω

k

(log X/C) requires much more work. The proof of our main theorem 4.6 is developed in section 4, and uses Koszul complexes and change of rings spectral sequences.

We also give other characterizations of freeness involving the module of multi-residues in corol- lary 4.19.

In section 5 we study a particular family of free complete intersections: the complete intersec- tion curves. For quasi-homogeneous curves, we determine a minimal free O

S

-resolution of R

C

and Ω

m−1

(log C) (see theorems 5.9 and 5.11). We also give a characterization of plane homogeneous curves using multi-residues (see proposition 5.17). Several properties of complete intersection curves are studied in [Pol15], and are used in this section.

Acknowledgements. The author is grateful to Michel Granger for raising several questions on this subject and many helpful discussions. The author also thanks Mathias Schulze for inviting her in Kaiserslautern, and for his suggestion to use Ischebeck’s lemma instead of spectral sequences in section 3.2, and for pointing out the characterization of quasi-homogenous curves used in the proof of proposition 5.7.

2. Multi-logarithmic differential forms 2.1. Complete intersection case.

2.1.1. Definitions. We give here the definitions and several properties of multi-logarithmic differen- tial forms and their residues along a complete intersection, which are a natural generalization of the theory of logarithmic differential forms developed in [Sai80].

Let S be a smooth analytic complex variety of dimension m > 1. We denote by O

S

the sheaf of holomorphic functions on S. Let x ∈ S, and (h

1

, . . . , h

k

) be a regular O

S,x

-sequence such that the ideal I

C

= P

k

i=1

h

i

O

S,x

is radical. For the statements concerning sheaves, we will consider an open neighbourhood of p such that the equations (h

1

, . . . , h

k

) converge and define a reduced complete intersection.

Notation 2.1. For all i ∈ {1, . . . , k}, we denote by D

i

the hypersurface of S defined by h

i

, and D = D

1

∪· · ·∪ D

k

the hypersurface defined by h := h

1

· · · h

k

. For j ∈ {1, . . . , k}, we set h b

j

= Q

i6=j

h

i

and D c

j

the hypersurface of S defined by h b

j

.

We denote by C = D

1

∩· · ·∩D

k

the reduced complete intersection defined by the regular sequence (h

1

, . . . , h

k

). We set O

C

:= O

S

/I

C

. We denote n = dim(C) = m − k.

For q ∈ N , we denote by Ω

qS

, or Ω

q

for short, the sheaf of holomorphic differential forms on S.

For I ⊆ {1, . . . , m}, we set dx

I

= V

i∈I

dx

i

, and |I| the cardinality of I .

We first recall some definitions and results from [Ale12]. The following definition generalizes [Sai80, (1.1)] to the complete intersection case.

Definition 2.2 ([Ale12]). Let x ∈ S and ω ∈ 1

h Ω

qS,x

, with q ∈ N . We call ω a multi-logarithmic q-form along the complete intersection C for the equations (h

1

, . . . , h

k

) if

∀i ∈ {1, . . . , k}, dh

i

∧ ω ∈

k

X

j=1

1 h b

j

q+1S,x

= 1

h I

C

q+1S,x

.

We denote by Ω

qS

(log C, h) or Ω

q

(log C, h) the sheaf of germs of multi-logarithmic q-forms along

C with respect to the equations (h

1

, . . . , h

k

).

(5)

From the definition, one can see that h · Ω

q

(log C, h) is the kernel of the map of coherent sheaves ϕ : Ω

q

q+1

I

C

q+1

k

given by ϕ(ω) = (dh

1

∧ ω, . . . , dh

k

∧ ω). It implies the following result:

Proposition 2.3. The sheaves Ω

q

(log C, h) are coherent O

S

-modules and the modules h·Ω

q

(log C, h) do not depend on the choice of the defining equations (h

1

, . . . , h

k

).

Notation 2.4. To simplify the notations, we set Ω e

qh

:=

k

X

j=1

1 h b

j

qS

= 1 h I

C

qS

.

The following characterization of multi-logarithmic forms generalizes [Sai80, (1.1)]:

Theorem 2.5 ([Ale12, §3, theorem 1]). Let ω ∈ 1

h Ω

qS,x

, with q ∈ N and x ∈ S. Then ω is multi- logarithmic if and only if there exist a holomorphic function g ∈ O

S,x

which induces a non zero divisor in O

C,x

, a holomorphic differential form ξ ∈ Ω

q−kS,x

and a meromorphic q-form η ∈ Ω e

qh,x

such that:

(2) gω = dh

1

∧ · · · ∧ dh

k

h ∧ ξ + η.

Remark 2.6. For q < k, we have the equality Ω

qS

(log C, h) = Ω e

qh

.

2.1.2. Multi-residues. Let q ∈ N . The module of Kähler differentials of order q is:

qC

= Ω

qS

(h

1

, . . . , h

k

)Ω

qS

+ dh

1

∧ Ω

q−1S

+ · · · + dh

k

∧ Ω

q−1S C

= Ω

qS

dI

C

∧ Ω

q−1S C

.

Definition 2.7 ([Ale12, §4, Definition 1]). Let x ∈ S and ω ∈ Ω

qS,x

(log C, h), q > k. Let us assume that g, ξ, η satisfy the properties of theorem 2.5. Then the multi-residue of ω is:

res

C,h

(ω) := ξ

g

C

∈ M

C,x

OC,x

q−kC,x

=

π

M

Ce

O

Ce

q−k

Ce

x

where π : C e → C is the normalization of C and M

C

is the sheaf of meromorphic functions on C.

The notion of multi-residue is well-defined with respect to the choices of g, ξ, η in (2) (see [Ale12,

§4 proposition 2]). We denote R

q−kC,h

:= res

C,h

(Ω

q

(log C, h)). If q = k, we set R

C,h

:= R

0C,h

⊂ M

C

. Proposition 2.8 ([Ale12, §4, lemma 1]). Let q ∈ N . We have the following exact sequence of O

S

-modules:

(3) 0 −→ Ω e

qh

−→ Ω

q

(log C, h) −

res

−−−

C,h

→ R

q−kC,h

−→ 0 In particular, the sheaf R

q−kC,h

is coherent.

From now on, we fix a point x ∈ S, and we consider only germs at X.

The modules of multi-residues do not depend on the choice of the defining equations:

Proposition 2.9. Let (h

1

, . . . , h

k

) and (f

1

, . . . , f

k

) be two regular sequences defining the same re- duced germ of complete intersection C of S. We set f = f

1

· · · f

k

. Let A = (a

ij

)

16i,j6k

∈ M

k

(O

S

) be such that (f

1

, . . . , f

k

)

t

= A(h

1

, . . . , h

k

)

t

. Then for all q ∈ N and for all α ∈ f · Ω

q

(log C, f ):

res

C,h

α h

= det(A)res

C,f

α

f

.

In particular, for all p > 0, the module R

pC,h

does not depend on the choice of the defining

equations.

(6)

Proof. Let α ∈ f

1

· · · f

k

· Ω

q

(log C, f ). Then there exists g ∈ O

S

which induces a non zero divisor in O

C

, ξ ∈ Ω

q−k

and η ∈ Ω e

qf

such that gα = df

1

∧ · · · ∧ df

k

∧ ξ + f η.

Since for all j ∈ {1, . . . , k}, df

j

= P

k

i=1

(h

i

da

ji

+ a

ji

dh

i

), there exists ν ∈ I

C

kS

such that:

(4) df

1

∧ · · · ∧ df

k

= ν + det(A)dh

1

∧ · · · ∧ dh

k

. Thus, gα = dh

1

∧ · · · ∧ dh

k

∧ det(A)ξ

+ ν ∧ ξ + f η.

Since f η ∈ I

C

qS

and ν ∈ I

C

kS

, we have ν ∧ ξ + f η

h

1

· · · h

k

∈ Ω e

qh

= 1

h I

C

qS

, so that res

C,h

α

h

= det(A)ξ

g = det(A)res

C,f

α

f

.

In the rest of the paper, we will use the notation Ω

(log C), R

C

and res

C

where the set of equations is implicitly (h

1

, . . . , h

k

).

The following proposition shows that we can assume that m is the embedding dimension of C.

Proposition 2.10. Let ` ∈ {1, . . . , k}. We assume that for all i ∈ {1, . . . , `} we have h

i

= x

i

and for i > `+ 1, h

i

∈ C {x

`+1

, . . . , x

m

} ∩m

2

. We set C

0

= D

`+1

∩ . . .∩ D

k

⊆ C

m−`

, so that C = {0} × C

0

. Then for all q > `:

q

(log C) = 1

x

1

· · · x

`

q−`

(log C

0

) ∧ dx

1

∧ · · · ∧ dx

`

+ Ω e

q(h

1,...,hk)

. Proof. Let ω =

1h

P

a

I

dx

I

. Then by definition ω ∈ Ω

q

(log C) if and only if for all i ∈ {1, . . . , k}, dh

i

∧ ω ∈ Ω e

q+1h

1,...,hk

. In particular, for all i ∈ {1, . . . , `}, dx

i

∧ ω ∈ Ω e

q+1(h

1,...,hk)

. This condition implies that for all I such that {1, . . . , `} 6⊆ I , a

I

∈ (h

1

, . . . , h

k

). We can write

ω = 1

x

1

· · · x

`

ω

0

∧ dx

1

∧ · · · ∧ dx

`

+ η

with η ∈ Ω e

q(h

1,...,hk)

and ω

0

= 1 h

`+1

· · · h

k

X

I0⊆{`+1,...,k}

|I0|=q−`

b

I0

dx

I0

, with b

I0

∈ C {x

`+1

, . . . , x

m

}.

Since h

`+1

, . . . , h

k

∈ C {x

`+1

, . . . , x

m

}, the conditions dh

i

∧ 1

x

1

· · · x

`

ω

0

∧dx

1

∧· · ·∧ dx

`

∈ Ω e

q+1(h

1,...,hk)

are equivalent to the conditions dh

i

∧ ω

0

∈ Ω e

q−`+1(h

`+1,...,hk)

. Therefore, ω ∈ Ω

q

(log C) if and only if

ω

0

∈ Ω

q−`

(log C

0

). Hence the result.

Corollary 2.11. With the same notations, for all q > k, the module R

q−kC

is equal to the O

C

-module R

q−kC0

.

Notation 2.12. We denote by J

C

the Jacobian ideal of C, which is the ideal of O

C

generated by the k × k-minors of the Jacobian matrix of (h

1

, . . . , h

k

). The Jacobian ideal does not depend on the choice of the equations (h

1

, . . . , h

k

).

The proof of the following result is similar to the proof of [Sai80, lemma 2.8]:

Proposition 2.13 (see [GS12]). Let π : C e → C be the normalization of C and x ∈ C. We have the inclusion O

Ce

⊆ R

C

. Proof. Let α ∈ O

Ce

. Then, thanks to [GLS10, Lemma 4.1], for every g ∈ J

C

, gα ∈ O

C

.

Thus, for every subset J ⊆ {1, . . . , m} with |J | = k, there exists a

J

∈ O

S

such that its class in O

C

satisfies ∆

J

α = a

J

∈ O

C

, where ∆

J

is the minor of the Jacobian matrix relative to the set J. Let I, J be two subsets of {1, . . . , m} with k elements. Then from the equality ∆

I

J

α − ∆

J

I

α = 0 ∈ O

C

we deduce:

I

a

J

− ∆

J

a

I

= h

1

b

IJ1

+ · · · + h

k

b

IJk

∈ O

S

.

(7)

Let us define ω =

P

|J|=kaJdxJ

h

h1

kS

. The previous equality gives:

I

ω = P

J

I

a

J

dx

J

h = a

I

dh

1

∧ · · · ∧ dh

k

h + η with η ∈ Ω e

k

.

Moreover, there exists a linear combination of the ∆

J

which does not induce a zero-divisor in O

C

(see [GLS10, Lemma 4.1]). Therefore, ω ∈ Ω(log C) and res

C

(ω) = α ∈ R

C

. Other proofs of propositions 2.9 and 2.13 are suggested respectively in [Ale12] and in [Sch16], which require the introduction of the modules of regular meromorphic forms and theorem 2.16 which is recalled below.

Definition 2.14. The O

C

-module ω

m−kC

= Ext

kOS

(O

C

, Ω

mS

) is called the Grothendieck dualizing module of the germ (C, 0). Since C is a complete intersection, it is a free O

C

-module of rank 1.

Definition 2.15. The module ω

Cp

of regular meromorphic differential p-forms is:

ω

Cp

= Hom

OC

m−k−pC

, ω

m−kC

.

Theorem 2.16 ([AT08, Theorem 3.1]). For all p > 0, the module of multi-residues R

pC

is isomor- phic to the module ω

Cp

of regular meromorphic differential forms.

2.1.3. Multi-logarithmic vector fields. We introduce here a module of multi-logarithmic k-vector fields. We will prove in proposition 3.3 that there exists a perfect pairing between Ω

k

(log C) and Der

k

(− log C) with values in

1h

I

C

, which generalizes [Sai80, (1.6)].

Let Θ

S

denote the O

S

-module of holomorphic vector fields and Θ

kS

:= V

k

Θ

S

be the exterior power of order k of Θ

S

.

We can evaluate a k-form ω ∈ Ω

kS

on a k-vector field δ ∈ Θ

kS

, which gives a function denoted by ω(δ).

Definition 2.17 ([GS12]). A k-vector field δ ∈ Θ

kS

is called multi-logarithmic along C if it satisfies:

dh

1

∧ · · · ∧ dh

k

(δ) ∈ I

C

.

We denote by Der

k

(− log C) the O

S

-module of multi-logarithmic k-vector fields along C.

The following proposition is easy to prove with (4):

Proposition 2.18. The module Der

k

(− log C) does not depend on the choice of the equations.

The following proposition comes from definition 2.17, where J

C

is the Jacobian ideal defined in notation 2.12.

Proposition 2.19 ([GS12, (5.1)]). We have the following exact sequence of O

S

-modules:

(5) 0 → Der

k

(− log C) → Θ

kS

→ −−−−−−−→ J

dh1∧···∧dhk C

→ 0.

In particular, Der

k

(− log C) is a coherent O

S

-module.

Remark 2.20. The notion of logarithmic vector field studied in [Sai80] can be extended to spaces of higher codimension (see [HM93]). A holomorphic vector field η ∈ Θ

S

is logarithmic if η is tangent to the complete intersection at its smooth points, which is equivalent to η(I

C

) ⊆ I

C

. We denote by Der(− log C) the module of logarithmic vector fields. The following inclusion is a direct consequence of the definition:

(6) Der(− log C) ∧ Θ

k−1S

⊆ Der

k

(− log C).

Using the fact that (h

1

, . . . , h

k

) is a regular sequence, it is easy to prove that Der(− log D) ⊆ Der(− log C). In particular, it implies that Der(− log D) ∧ Θ

k−1S

⊆ Der

k

(− log C) (see [GS12, (5.3)]).

One can check that the inclusion (6) may be strict, by considering for example the complete

intersection curve defined by h

1

= x

3

− y

2

and h

2

= x

2

y − z

2

(see [Pol16, Exemple 3.2.10]).

(8)

2.1.4. A condition on the divisor associated to a regular sequence. We prove here that the module of logarithmic 1-forms on the divisor D = D

1

∪ · · · ∪ D

k

satisfies a decomposition property when D

1

∩ · · · ∩ D

k

is a reduced complete intersection (see proposition 2.23).

We first need the following lemma:

Lemma 2.21. If C is the germ of a reduced complete intersection defined by a regular sequence (h

1

, . . . , h

k

), then the complete intersection defined by (h

1

, . . . , h

k−1

) is also reduced.

In addition, for all i ∈ {1, . . . , k − 1}, the complete intersection defined by (h

1

, . . . , h

i−1

, Q

k

j=i

h

j

) is also reduced.

Proof. Let g ∈ p

(h

1

, . . . , h

k−1

). Then, since p

(h

1

, . . . , h

k−1

) ⊆ p

(h

1

, . . . , h

k

) = (h

1

, . . . , h

k

), there exists f

1

∈ (h

1

, . . . , h

k−1

) and α

1

∈ O

S

such that g = f

1

+ α

1

h

k

. We then have α

1

h

k

∈ p (h

1

, . . . , h

k−1

). Since (h

1

, . . . , h

k

) is a regular sequence, it implies that α

1

∈ p

(h

1

, . . . , h

k−1

).

Therefore, there exists f

2

∈ (h

1

, . . . , h

k−1

) and α

2

∈ O

S

such that g = f

2

+ α

2

h

2k

. By induction, we prove that for all i ∈ N

, there exist f

i

∈ (h

1

, . . . , h

k−1

) and α

i

∈ O

S

such that g = f

i

+ α

i

h

ik

. Hence, the class g of g in O

S

/(h

1

, . . . , h

k−1

) satisfies

g ∈ \

i>1

(h

ik

) ⊆ \

i>1

(m

i

) ⊂ O

S

/(h

1

, . . . , h

k−1

).

By Krull’s intersection theorem, it implies that g = 0, therefore, g ∈ (h

1

, . . . , h

k−1

). Hence the first statement.

For the second statement, it is sufficient to prove it for i = k − 1, the result will then follow by induction.

Let us consider g ∈ p

(h

1

, . . . , h

k−2

, h

k−1

· h

k

). Then, g ∈ (h

1

, . . . , h

k

), so that for all j ∈ {1, . . . , k} there exist α

j

∈ O

S

such that g = P

k

j=1

α

j

h

j

. Therefore, α

k−1

h

k−1

+ α

k

h

k

∈ p

(h

1

, . . . , h

k−2

, h

k−1

· h

k

).

Let q ∈ N

be such that (α

k−1

h

k−1

+ α

k

h

k

)

q

∈ (h

1

, . . . , h

k−2

, h

k−1

· h

k

). Since (h

1

, . . . , h

k

) is a regular sequence, we then have α

qk−1

∈ (h

1

, . . . , h

k−2

, h

k

) and α

qk

∈ (h

1

, . . . , h

k−1

). Since by the first

statement, these ideals are radical, it gives the result.

Proposition 2.22 (see [Ale12, Proposition 1]). For all q ∈ N we have the inclusion:

q

(log D) ⊆ Ω

q

(log C).

Therefore, res

C

(Ω

q

(log D)) ⊆ res

C

(Ω

q

(log C)).

Proposition 2.23. Let C be a reduced complete intersection. We keep the same notations as before.

The module of logarithmic 1-forms satisfies:

1

(log D) = Ω

1

(log D

1

) + · · · + Ω

1

(log D

k

).

Proof. We first prove the proposition for k = 2. The general case will then follow from this special case.

Let ω ∈ Ω

1

(log D). Then by proposition 2.22 and Remark 2.6, ω =

hη1

1

+

ηh2

2

with η

i

∈ Ω

1S

. Then:

dh ∧ ω = dh

1

∧ η

2

+ dh

2

∧ η

1

+ h

2

dh

1

∧ η

1

h

1

+ h

1

dh

2

∧ η

2

h

2

∈ Ω

2S

.

Therefore, there exists θ ∈ Ω

2S

such that:

h

22

dh

1

∧ η

1

+ h

21

dh

2

∧ η

2

= h

1

h

2

θ.

Since (h

1

, h

2

) is a regular sequence, h

2

divides dh

2

∧ η

2

and h

1

divides dh

1

∧ η

1

, which means that η

1

h

1

∈ Ω

1

(log D

1

) and η

2

h

2

∈ Ω

1

(log D

2

).

The general case is obtained by induction from the case k = 2. Indeed, by lemma 2.21, for all i ∈ {1, . . . k − 1}, (h

1

, . . . , h

i−1

, Q

k

j=i

h

j

) is a regular sequence defining a reduced complete intersection,

so that Ω

1

(log D) = Ω

1

(log D

1

) + · · · + Ω

1

(log D

i−1

) + Ω

1

(log(D

i

∪ · · · ∪ D

k

)).

(9)

Remark 2.24. In particular, it gives a necessary condition on Ω

1

(log D) which has to be satisfied when C is a reduced complete intersection.

A divisor is called splayed if in some coordinate system (y

1

, . . . , y

m

) it has an equation of the form g = g

1

· g

2

with g

1

∈ C {y

1

, . . . y

`

} and g

2

∈ C {y

`+1

, . . . , y

m

}, with ` ∈ {1, . . . , m − 1} (see [AF13] and [Sch16]). The following corollary is then a consequence of [AF13, Theorem 2.12] and proposition 2.23:

Corollary 2.25. Let C be a reduced complete intersection of codimension 2 defined by (h

1

, h

2

). If the divisor D defined by h

1

· h

2

is free, then D is a splayed divisor.

2.2. Generalization to reduced equidimensional spaces. We generalize to reduced equidimen- sional spaces the notions of the previous part. We treated first the complete intersection case because the reduced equidimensional case rely partially on it.

The reduced equidimensional case uses a description of regular meromorphic differential forms by residue symbols which has been introduced in [Ker84]. This description was already used by M.

Schulze in a talk given at Oberwolfach in 2012, and then in [Sch16], in order to generalize logarithmic residues to reduced equidimensional spaces. A definition of multi-logarithmic differential forms is suggested in [Ale14], where the interpretation of regular meromorphic forms with residue symbols is also used.

2.2.1. Multi-logarithmic differential forms. Let X ⊆ S be the germ of a reduced equidimensional analytic subset of S of dimension n defined by a radical ideal I

X

. We set k = m − n.

One can prove that there exists a regular sequence (f

1

, . . . , f

k

) ⊆ I

X

such that the ideal generated by f

1

, . . . , f

k

is radical (see [AT08, Remark 4.3] or [Pol16, Proposition 4.2.1] for a detailed proof of this result). We fix such a sequence (f

1

, . . . , f

k

). We denote by C the complete intersection defined by the ideal I

C

generated by (f

1

, . . . , f

k

). In particular, C = X ∪ Y , where Y is of pure dimension n and does not contain any component of X. We set f = f

1

· · · f

k

, and I

Y

the radical ideal defining Y .

We recall that Ω e

qf

=

f1

I

C

qS

.

Definition 2.26 ([Ale14, Definition 10.1]). Let q ∈ N . We define the module of multi-logarithmic q-forms with respect to the pair (X, C) as :

q

(log X/C) =

ω ∈ 1

f

1

· · · f

k

qS

; I

X

ω ⊆ Ω f

qf

and (dI

X

) ∧ ω ⊆ Ω e

q+1f

. Remark 2.27. For all q ∈ N , we have :

Ω e

qf

⊆ Ω

q

(log X/C) ⊆ Ω

q

(log C).

In addition, if C is a reduced complete intersection, we have Ω

q

(log C/C) = Ω

q

(log C).

Proposition 2.28. Let q ∈ N . The module Ω

q

(log X/C ) is coherent and the module of numerators f · Ω

q

(log X/C) depends only on X and C.

Proof. Let q ∈ N and (h

1

, . . . , h

r

) be a generating family of I

X

. The module f · Ω

q

(log X/C ) is the kernel of the map β : Ω

qS

→ Ω

qS

/I

C

qS

r

q+1S

/I

C

q+1S

r

defined for ω ∈ Ω

qS

by β(ω) = (h

1

ω, . . . , h

r

ω, dh

1

∧ ω, . . . , dh

r

∧ ω). Hence the result.

Remark 2.29. The modules f · Ω

q

(log X/C) depend on the choice of the complete intersection C, since I

C

· Ω

qS

⊆ f · Ω

q

(log X/C).

Definition 2.30 ([Ale14, Proposition 10.1]). The multi-residue map res

X/C

: Ω

q

(log X/C) → M

X

OX

q−kX

is defined as the restriction of the map res

C

.

For q ∈ N , we set R

qX/C

:= res

X/C

(Ω

q+k

(log X/C )).

Proposition 2.31 ([Ale14, Theorem 10.2]). We have the following exact sequence of O

S

-modules:

(7) 0 → Ω e

qf

→ Ω

q

(log X/C) −−−−→ R

resX/C q−kX/C

→ 0.

(10)

We have the following theorem:

Theorem 2.32 ([Ale14, Theorem 10.2]). Let q ∈ N . The module R

qX/C

is isomorphic to the module of regular meromorphic forms ω

qX

. In particular, R

qX/C

does not depend on the choice of the complete intersection C.

The proof of this theorem uses [Kun86, (E.20)] and [Kun86, (E.21)], and the following character- ization of regular meromorphic forms with residue symbols:

Proposition 2.33 ([Ker84, (1.2)]). Let q ∈ N . The module of regular meromorphic forms ω

qX

satisfies:

ω

Xq

=

α f

1

, . . . , f

k

; α ∈ Ω

q+kS

, (f

1

, . . . , f

k

) ⊆ I

X

a regular sequence, I

X

α ⊆ X

f

i

k+qS

, dI

X

∧ α ⊆ X

f

i

q+k+1S

o

where

α f

1

, . . . , f

k

denotes the residue symbol (see [Ker83] for more details).

We end this section with a characterization of multi-logarithmic q-forms with respect to the pair (X, C), which generalizes theorem 2.5. We first need to introduce the fundamental form of X (see for example [Ker84, (1.3)]). We recall that C = X ∪ Y is an irredundant decomposition of C.

Notation 2.34. Let β

f

∈ O

Ce

be such that β

f X

= 1 and β

f Y

= 0.

A direct consequence of [GLS10, Lemma 4.1] is the following lemma:

Lemma 2.35. The form β

f

df

1

∧ · · · ∧ df

k

∈ Ω

kS

OS

M

C

satisfies β

f

df

1

∧ · · · ∧ df

k

∈ Ω

kS

OS

O

C

Notation 2.36. We fix c

X

∈ Ω

kS

such that c

X

= β

f

df

1

∧ · · · ∧ df

k

∈ Ω

kS

OS

O

C

.

The following result is a consequence of the definition of c

X

and [Ker84, (1.3)]:

Proposition 2.37 (see [Ker84, (1.3)]). We have

cfX

∈ Ω

k

(log X/C ) and res

X/C

cX

f

= 1 ∈ M

X

. The following proposition gives a characterization of multi-logarithmic forms which generalizes theorem 2.5.

Proposition 2.38. Let ω ∈

f1

qS

. Then ω ∈ Ω

q

(log X/C ) if and only if there exist g ∈ O

S

which induces a non zero divisor in O

C

, ξ ∈ Ω

q−kS

and η ∈ Ω e

qf

such that:

(8) gω = c

X

f ∧ ξ + η In addition, we then have res

X/C

(ω) =

ξ

g X

.

Proof. Let ω ∈ Ω

q

(log X/C ). Then ω ∈ Ω

q

(log C). Let g, ξ, η satisfying theorem 2.5 such that:

gω = df

1

∧ · · · ∧ df

k

f ∧ ξ + η.

Then gres

X/C

(ω) = ξ = res

X/C

cX

f

∧ ξ

. By proposition 2.31, there exists η

0

∈ Ω e

qf

such that gω = c

X

f ∧ ξ + η

0

.

Conversely, let ω ∈

1f

qS

be such that gω =

cfX

∧ ξ + η with g, ξ, η as in the statement of the

proposition. Let h ∈ I

X

. Since

cfX

∈ Ω

k

(log X/C), we deduce that hgω ∈ Ω e

qf

and dh ∧ gω ∈ Ω e

q+1f

.

Therefore, since g induces a non zero divisor in O

C

, we have ω ∈ Ω

q

(log X/C ).

(11)

We deduce from [Sch16, (2.14)] the following equality:

(9) R

X

= {ρ

X

; ρ ∈ R

C

such that ρ

Y

= 0} .

In particular, it implies that O

Xe

⊆ R

X

.

As a consequence, we have the following proposition:

Proposition 2.39. Let ω ∈ Ω

k

(log C). Then:

ω ∈ Ω

k

(log X/C) ⇐⇒ res

C

(ω)|

Y

= 0.

Proof. Let ω ∈ Ω

k

(log C), and ρ = res

C

(ω). Then:

ρ

Y

= 0 ⇐⇒ ∃ω

0

∈ Ω

k

(log X/C), ρ

X

= res

X/C

0

)

⇐⇒ ∃w

0

∈ Ω

k

(log X/C ), ∃η ∈ Ω e

kf

, ω = ω

0

+ η ∈ Ω

k

(log X/C).

Hence the result.

2.2.2. Multi-logarithmic vector fields. We suggest the following definition of multi-logarithmic k- vector fields, which generalizes definition 2.17:

Definition 2.40. Let δ ∈ Θ

kS

. We say that δ is a multi-logarithmic k-vector field with respect to the pair (X, C) if c

X

(δ) ∈ I

X

.

We denote by Der

k

(− log X/C) the module of multi-logarithmic k-vector fields with respect to (X, C).

Remark 2.41. Since c

X

∈ I

Y

kS

, for all δ ∈ Θ

kS

, c

X

(δ) ∈ I

Y

. Therefore, Der

k

(− log X/C ) = n

δ ∈ Θ

kS

; c

X

(δ) ∈ I

C

o . As a consequence, we have the following proposition:

Proposition 2.42. We have the following inclusion:

Der

k

(− log C) ⊆ Der

k

(− log X/C ).

Remark 2.43. If C is a reduced complete intersection, we have Der

k

(− log C/C ) = Der

k

(− log C).

3. Duality results

3.1. Perfect pairing between multi-logarithmic vector fields and multi-logarithmic forms.

For a reduced hypersurface D, a O

S

-duality is satisfied between Ω

1

(log D) and Der(− log D) (see [Sai80, Lemma (1.6)]).

Let us prove that there is a perfect pairing between Ω

k

(log X/C) and Der

k

(− log X/C), and in particular, for a complete intersection C, between Ω

k

(log C) and Der

k

(− log C). We fix a point x ∈ S and consider germs at this point, without mentioning it.

We first need the following lemma, which is a direct consequence of proposition 2.38 and the definition of Der

k

(− log X/C).

To simplify the notations, we set Σ =

1f

I

C

.

Lemma 3.1. Let δ ∈ Der

k

(− log X/C) and ω ∈ Ω

k

(log X/C ). Then ω(δ) ∈ Σ.

Thanks to lemma 3.1, we see that we have a natural pairing Der

k

(− log X/C ) × Ω

k

(log X/C) → Σ.

Let us notice the following lemma:

Lemma 3.2. We have the following perfect pairings:

Ω e

kf

× Θ

kS

→ Σ, (10)

1 f Ω

kS

×

k

X

i=1

f

i

Θ

kS

→ Σ.

(11)

(12)

Proof. Let us notice that Ω e

kf

= Ω

kS

⊗ Σ and P

f

i

Θ

kS

= f Θ

kS

⊗ Σ. Let M be either Ω

kS

or f Θ

kS

. It is easy to prove that Hom

OS

(M ⊗ Σ, Σ) = Hom

OS

(M, O

S

) and Hom

OS

(M, Σ) = Hom

OS

(M, O

S

)⊗Σ.

Hence the result.

Proposition 3.3. We have the following perfect pairing:

(12) Ω

k

(log X/C ) × Der

k

(− log X/C) → Σ.

Proof. We have the following inclusions:

(13) Ω e

kf

⊆ Ω

k

(log X/C ) ⊆ 1

f Ω

kS

(14) Θ

kS

⊇ Der

k

(− log X/C) ⊇ Der

k

(− log X/C ).

We deduce from lemmas 3.1, 3.2 and the inclusions (13) and (14) the following:

Der

k

(− log X/C ) ⊆ Hom

OS

k

(log X/C), Σ

⊆ Θ

kS

,

k

(log X/C) ⊆ Hom

OS

Der

k

(− log X/C ), Σ

⊆ 1 f Ω

kS

. Let us prove that the left-hand-side inclusions are equalities.

Since

cfX

∈ Ω

k

(log X/C), for all δ ∈ Hom

OS

k

(log X/C ), Σ

we have c

X

(δ) ∈ I

C

so that δ ∈ Der

k

(− log X/C). Therefore, Hom

OS

k

(log X/C ), Σ

= Der

k

(− log X/C ).

Let ω ∈ Hom

OS

Der

k

(− log X/C), Σ

. Let us prove that ω ∈ Ω

k

(log X/C). Let us set ω = P

|I|=k 1

f

ω

I

dx

I

. Let h ∈ I

X

. Then for all I, h∂

xI

∈ Der

k

(− log X/C ), so that h∂

xI

(ω) = h

1f

ω

I

∈ Σ.

Therefore, hω ∈ Ω e

kf

.

We set for J ⊆ {1, . . . , m} with |J | = k + 1, δ

J

= P

k+1

`=1

(−1)

`−1∂x∂h

j`

∂x

j1

∧ · · · ∧ ∂x d

j`

∧ · · · ∧ ∂x

jk+1

. We thus have ω(δ

J

) = (dh ∧ ω)(∂

xJ

). Let us prove that δ

J

∈ Der

k

(− log X/C). We have:

f

df

1

∧ · · · ∧ df

k

) (δ

J

) = β

f k+1

X

`=1

(−1)

`−1

∂h

∂x

j`

j

1...bj`...jk+1

= β

f

∂h

∂xj1

. . .

∂x∂h

∂f1 jk+1

∂xj1

. . .

∂x∂f1

..

jk+1

. .. .

∂fk

∂xj1

. . .

∂x∂fk

jk+1

.

Since the codimension of X is k and (h, f

1

, . . . , f

k

) ⊆ I

X

, the restriction of the previous de- terminant to X is zero. Given that β

f Y

= 0, we have (β

f

df

1

∧ · · · ∧ df

k

)(δ

J

) = 0 ∈ M

C

. Thus, δ

J

∈ Der

k

(− log X/C ). Since ω(δ

J

) = (dh ∧ ω)(∂

xJ

), we deduce that dh ∧ ω ∈ Ω e

k+1f

and Hom

OS

Der

k

(− log X/C ), Σ

= Ω

k

(log X/C ).

Remark 3.4. By lemma 3.2 and proposition 3.3, if C is a reduced complete intersection, we have the following perfect pairings between modules which all depend only on C and not on the choice of equations:

I

C

kS

× Θ

kS

→ I

C

kS

× I

C

Θ

kS

→ I

C

f · Ω

k

(log C) × Der

k

(− log C) → I

C

3.2. Multi-residues and Jacobian ideal. Let C be a reduced complete intersection. A conse- quence of [Sch16, Lemma 5.4] and theorem 2.16 is that the dual of the Jacobian ideal J

C

of C is the module of multi-residues R

C

.

In this subsection, we give another proof this duality, which does not depend on the isomorphism

of theorem 2.16. In addition, we extend it to the case of reduced equidimensional spaces. Our proof

uses the perfect pairings of proposition 3.3 and is analogous to the proof of [GS14, Proposition 3.4].

(13)

Notation 3.5. Let X be a reduced equidimensional space of codimension k in S and C be a reduced complete intersection of codimension k containing X. We denote by J

X/C

⊆ O

X

the restriction of the Jacobian ideal J

C

of C to the space X.

Remark 3.6. Since β

f X

= 1, the ideal J

X/C

is given by J

X/C

= n

δ(c

X

) ∈ O

X

; δ ∈ Θ

kS

o

. The following result is a consequence of the definitions:

Proposition 3.7. We have the following exact sequence of O

S

-modules:

0 → Der

k

(− log X/C ) → Θ

kS

→ J

X/C

→ 0.

Remark 3.8. Several notions of Jacobian ideals are associated to a reduced equidimensional space:

the ideal J

X

⊆ O

X

generated by the k × k minors of the Jacobian matrix of (h

1

, . . . , h

r

), where P

r

i=1

h

i

O

S

= I

X

, the ω-Jacobian J

Xω

(see for example[Sch16]), and the ideal J

X/C

considered above.

These three notions may not coincide, as one can check on the example of the non Gorenstein curve of C

3

defined by h

1

= y

3

− x

2

z, h

2

= x

3

y − z

2

and h

3

= x

5

− y

2

z, which is the irreducible curve parametrized by (t

5

, t

7

, t

11

) (see [Pol16, Exemple 4.2.35]).

We first need the following lemma:

Lemma 3.9. Let X be a reduced space of pure dimension n = m − k. We assume k > 2. Then Ext

1O

S

J

X/C

, O

S

= 0 and Ext

1O

S

J

X/C

, Σ

= Hom

OC

J

X/C

, O

C

.

Proof. We apply the functor Hom

OS

(J

X/C

, −) to the exact sequence 0 → Σ −−→ O

×f S

→ O

C

→ 0.

It gives:

0 → Hom

OS

(J

X/C

, O

C

) → Ext

1OS

J

X/C

, Σ

→ Ext

1OS

J

X/C

, O

S

→ . . .

The depth of O

S

is m and since J

X/C

is a fractional ideal of O

X

, the dimension of J

X/C

is m − k = dim O

X

. Thus, by Ischebeck’s lemma (see [Mat80, 15.E]), we have Ext

1OS

J

X/C

, O

S

= 0.

Hence the result.

Notation 3.10. Let I ⊂ M

C

be an ideal. We set I

= Hom

OC

(I, O

C

).

Proposition 3.11. We have J

X/C

' R

X

. In particular, if C is a reduced complete intersection, J

C

= R

C

.

Proof. We assume k > 2. For k = 1, we refer to [GS14, Proposition 3.4].

We consider the double complex Hom

OS

Der

k

(− log X/C) , → Θ

kS

, f : Σ → O

S

, which gives al- most the same diagram as the dual of (3.8) in [GS14]. By lemma 3.9, Ext

1O

S

(J

X/C

, O

S

) = 0, so that we obtain the following commutative diagram:

0 0

0 Hom

OS

Θ

kS

, Σ

Hom

OS

Der

k

(− log X/C), Σ

Ext

1OS

J

X/C

, Σ 0

0 Hom

OS

Θ

kS

, O

S

Hom

OS

Der

k

(− log X/C ), O

S

0

0 Hom

OS

J

X/C

, O

C

Hom

OS

Θ

kS

, O

C

Hom

OS

Der

k

(− log X/C ), O

C

Ext

1OS

J

X/C

, O

C

0

Ext

1O

S

J

X/C

, Σ

0 Ext

1O

S

Der

k

(− log X/C ), Σ

Ext

2O

S

J

X/C

, Σ 0

0

λ

σ µ

·f

π

(14)

Let ϕ : Der

k

(− log X/C) → Σ. Thanks to the isomorphism β of the proof of proposition 3.3, ϕ coresponds to ω =

f1

P

I

ϕ(f ∂x

I

)dx

I

∈ Ω

k

(log X/C ), and ϕ(δ) = ω(δ).

By a diagram chasing process, we obtain the map:

k

(log X/C ) → Hom

OS

J

X/C

, O

C

ω 7→

a 7→ res

C

(ω)a .

The map a 7→ res

C

(ω)a ∈ O

C

is well defined since by proposition 2.39, res

C

(ω)|

Y

= 0.

Similarly, the same diagram chasing process starting from the lower left Hom

OS

J

X/C

, O

C

to the upper right Hom

OS

Der

k

(− log X/C), Σ

, show that the map θ : R

X

→ J

X/C

ρ 7→ θ

ρ

:

J

X/C

→ O

C

a 7→ ρa

is an isomorphism.

4. Freeness for Cohen-Macaulay spaces

We prove here our main result, namely, theorem 4.6, which is a characterization of freeness for Cohen-Macaulay spaces by the projective dimension of the module Ω

k

(log X/C ), which generalizes the hypersurface case.

4.1. Definition and statements. In [Sai80], K. Saito studies a particular family of hypersurfaces for which the module of logarithmic vector fields is free. This kind of hypersurfaces are called free divisors. Among the different characterizations of free divisors, let us mention the following one:

Theorem 4.1 ([Ale88]). The germ of a reduced singular divisor is free if and only if its singular locus is Cohen Macaulay of codimension 1 in D. Equivalently, a reduced divisor is free if and only if the Jacobian ideal is Cohen-Macaulay.

Our purpose here is to extend the notion and characterizations of freeness to Cohen-Macaulay spaces. We give the following definition for freeness, which is inspired by theorem 4.1:

Definition 4.2. A reduced Cohen-Macaulay space X contained in a reduced complete intersection C of the same dimension is called free if J

X/C

is Cohen-Macaulay.

Remark 4.3. Definition 4.2 generalizes the notion of freeness for complete intersections defined in [GS12, Definition 5.1].

Proposition 4.4. If there exists a reduced complete intersection C of dimension n containing X such that J

X/C

is Cohen-Macaulay, then for all reduced complete intersection C

0

of dimension n containing X, J

X/C0

is Cohen-Macaulay. In other words, the notion of freeness does not depend on the choice of C.

Proof. Let C and C

0

be two reduced complete intersections of dimension n containing X. Let C

00

be a reduced complete intersection containing C and C

0

. Let A be a transition matrix from (f

1

, . . . , f

k

) to (f

100

, . . . , f

k00

), where (f

100

, . . . , f

k00

) are defining equations for C

00

. There exists ν ∈ I

C

kS

such that:

df

100

∧ · · · ∧ df

k00

= det(A)df

1

∧ · · · ∧ df

k

+ ν.

Therefore, we have: J

X/C00

= det(A)J

X/C

⊆ O

X

. In addition, since X ⊆ C ∩ C

00

and X is not included in the singular locus of C or C

00

, det(A) is a non zero divisor of O

X

. Therefore, J

X/C

and J

X/C00

are isomorphic. Similarly, J

X/C0

and J

X/C00

are isomorphic, which gives us the result.

We now state our main results:

Proposition 4.5. Let X be a reduced Cohen-Macaulay space of codimension k in S, and C be a

reduced complete intersection of codimension k containing X. The following statements are equiva-

lent:

(15)

(1) X is free,

(2) O

X

/J

X/C

is Cohen-Macaulay of dimension m − k − 1 or O

X

/J

X/C

= (0), (3) projdim(Der

k

(− log X/C )) 6 k − 1,

(4) projdim(Der

k

(− log X/C )) = k − 1.

The first three equivalences are given in [GS12] for reduced complete intersections.

We add to this list other characterizations, which leads to the following theorem:

Theorem 4.6. We can extend the list of equivalences of proposition 4.5 with:

(5) projdim Ω

k

(log X/C )

6 k − 1, (6) projdim Ω

k

(log X/C )

= k − 1.

In particular, for k = 1, we recognize the several characterizations of freeness for divisors we mentioned before. In the hypersurface case, the duality between Der(− log D) and Ω

1

(log D) gives immediately the fact that if one of the two modules is free, the other is also free, whereas for for Cohen-Macaulay spaces of codimension greater than 2, the statement on the projective dimension of Ω

k

(log X/C) needs much more work, that’s why we consider it separately from the others.

4.2. Proof of proposition 4.5. We prove the equivalences of proposition 4.5. They are based on the depth lemma as stated in [dJP00, Lemma 6.5.18] and the Auslander-Buchsbaum formula.

The equivalence 1. ⇐⇒ 2. is proved in [Sch16, Proposition 5.6] for Gorenstein spaces. Our proof is completely similar.

If J

X/C

= O

X

, the statement is clear. Let us assume that J

X/C

6= O

X

. Let us consider the following exact sequence of O

X

-modules:

(15) 0 → J

X/C

→ O

X

→ O

X

/J

X/C

→ 0.

By assumption, O

X

is Cohen-Macaulay of dimension n. Moreover, since we assume C to be reduced, the singular locus of C is of dimension at most n− 1, and therefore the depth of O

X

/J

X/C

is at most n − 1. We deduce from the depth lemma that depth(J

X/C

) = n ⇐⇒ depth(O

X

/J

X/C

) = n − 1.

We now prove 1. ⇒ 4. We recall the following exact sequence:

(16) 0 → Der

k

(− log X/C ) → Θ

kS

→ J

X/C

→ 0.

Then, thanks to the depth lemma, since the depth of J

X/C

is m − k and the depth of Θ

kS

is m, we have depth(Der

k

(− log X/C)) = m − k + 1. By the Auslander-Buchsbaum formula, we have projdim(Der

k

(− log X/C )) = k − 1.

The implication 4 ⇒ 3. is trivial.

Let us prove 3. ⇒ 1. This implication is mentioned in [GS12] for complete intersections. By the Auslander-Buchsbaum formula, depth(Der

k

(− log X/C)) > m − k + 1. In addition, we have depth(Θ

kS

) = m, and depth(J

X/C

) 6 m − k. As a consequence of the exact sequence (16) and of the depth lemma we have depth(Der

k

(− log X/C )) = m − k + 1 and depth(J

X/C

) = m − k, so that J

X/C

is maximal Cohen-Macaulay.

4.3. Preliminary to the proof of theorem 4.6. The methods used to prove proposition 4.5 applied to the short exact sequence (17) are not sufficient to prove directly theorem 4.6.

(17) 0 → Ω e

kf

→ Ω

k

(log X/C ) → R

X

→ 0.

The proof of theorem 4.6 is based on the explicit computation of some modules and morphisms of the long exact sequence obtained by applying the functor Hom

OS

(−, O

S

) to the short exact sequence (17):

(18)

0 → Hom

OS

(R

X

, O

S

) → Hom

OS

(Ω

k

(log X/C ), O

S

) → Hom

OS

( Ω e

kf

, O

S

) → Ext

1OS

(R

X

, O

S

) → . . . The structure of the proof is the following. Thanks to the Koszul complex, we compute the mod- ules Ext

qO

S

Ω e

kf

, O

S

. We then determine the modules Ext

qO

S

(R

X

, O

S

) for q 6 k using the change

Références

Documents relatifs

The aim of the present paper is ( i ) to present general results about freeness problems, ( ii ) to study the decidability of freeness problems over various particular

A BSTRACT - We showthat the map sending a sheaf to its dual gives an isomorphism of the moduli space of semistable sheaves with fixed multiplicity and Euler char- acteristic

From the definition of the Cohen-Macaulay homological dimensions, the connection to Cohen-Macaulayness is less than obvious, but as described in the introduction, we shall prove

Mainly we deal with the codimension two case in which we show that the Dubreil bound t G d 1 1 1 is a necessary and sufficient condition to have such an ideal and we give a sharp

In order to obtain these characterizations we first show that a closed bounded absolutely convex subset B of an arbitrary Frechet space F is uniformly polar and hence pilar in F if

Huckaba, Reduction number bounds on analytic deviation two ideals and Cohen-Macaulayness of associated graded rings, Comm.. Huneke, Reduction numbers, Rees algebras,

- wild, which can be defined in two ways: either as those algebras having n-parameter families of indecomposable modules of a fixed rank for arbitrary big n, or

Verdier, Reflexive modules over rational double points, Math.. Auslander, Functors and morphisms determined by objects,