March 6 in 2005
1991 Mathematics Subject Classification. 13D45, 13F20,13F55.
Combinatorial characterizations of generalized Cohen-Macaulay monomial ideals
Yukihide Takayama
Yukihide Takayama, Department of Mathematical Sciences, Ritsumeikan University, 1-1-1 Nojihigashi, Kusatsu, Shiga 525-8577, Japan E-mail address : takayama@se.ritsumei.ac.jp
-
Abstract.
We give a generalization of Hochster's formula for local cohomologies of square-free monomial ideals to monomial ideals, which are not necessarily square-free. Using this formula, we give combinatorial characterizations of generalized Cohen-Macaulay monomial ideals. We also give other applications of the generalized Hochster's formula.
Introduction
Let
be a field and let
be a polynomial ring with the standard grading. For a graded ideal
we set
. We denote by
the image of
in
for
and set
, the unique graded maximal ideal.
Also
denotes the local cohomology module of
with regard
. A residue class ring
is called a generalized Cohen-Macaulay ring (generalized CM ring), or FLC (Finite Length Cohomology) ring, if
has finite length for
. In this case, we will call the ideal
a generalized CM ideal.
As defining ideals of algebraic sets, we can find many examples of generalized CM ideals such as homogeneous coordinate rings of non-singular projective varieties.
For monomial ideals, which are not directly related to algebraic sets, the notions of generalized CM rings and Buchsbaum rings [9] coincide in the square-free case and the combinatorial characterization of generalied CM square-free monomial ideals (Stanley-Reisner ideals) has been given in terms Buchsbaum simplicial complexes [7, 8, 9] . However, as far as the author is concerned, the case of non-square-free monomial ideals has not been studied very much, and the aim of this paper is to give combinatorial characterizations of generalized CM monomial ideals, which are not always square-free.
We first give a generalization of Hochster's formula on local cohomologies for square-free monomial ideals [6] to monomial ideals that are not necessarily square-free (Theorem 1 ). From this formula, we can easily deduce several already known and probably new facts on vanishing degrees of local cohomologies. In particular, the vanishing degrees of generalized CM monomial ideals (Proposition 1 ). This result allows us to deduce combinatorial characterizations of generalized CM monomial ideals in terms of the exponents of variables in the monomial generators (Theorem 2 , Corollary 7 and Theorem 3 ). On the other hand, thanks to the generalized Hochster's formula we can compare local cohomologies for
and its radical
(Proposition 3 and 4 ), which, together with the combinatorial characterization of generalized CM property, suggests a method to construct generalized CM monomial ideals from Buchsbaum Stanley-Reisner ideals. Namely, by changing a square-free generator
of a Buchsbaum Stanley-Reisner ideal
to a monomial
, we make a generator of a generalized CM monomial ideal
with
, and the combinatorial characterization shows a right choice of the exponents
.
One way of the construction is changing all the occurrences of the variable
in the minimal set of generators to
with a fixed exponent
,
(Example 1 ). In some specific case, we can show, using our combinatorial characterization, that this is the only way of construction (Example 2 ).
For a finite set
we denote by
the cardinarity of
, and, for sets
and
,
means that
is a subset of
, which may be equal to
. The author thanks Jürgen Herzog for valuable discussions and detailed comments on the early version of the paper.
1 Local cohomologies of monomial ideals
1.1 Generalized Hochster's Formula
In this subsection, we give a natural extension of Hochster's formula on local cohomologies of Stanley-Reisiner ideals to monomial ideals. The proof goes along almost the same line as that for Stanley-Reisner ideals given, for example, in [3] chapter 5.3. But we will give a full detail for the readers' convenience.
Let
be a monomial ideal, which is not necessarily square-free. Then we have
where
is the
ech complex defined as follows:
and the differential
of this complex is induced by
where
is the natural homomorphism to localized rings and
, for example, denotes localization by
.
We can consider a
-grading to
,
and
induced by the multi grading of
. See for example [3] for more detailed information about this complex.
Now we will consider the degree
subcomplex
of
for any
. Before that we will prepare the notation. For a monomial ideal
, we denote by
the minimal set of monomial generators. Let
be a monimial with
for all
, then we define
for
, and
. We set
and
for
.
Lemma 1.
Let
with
and set
. For all
we have
and the following are equivalent
-
-
and for all
there exists
such that
.
Notice that the condition
in
is redundant because this follows from the condition
. But it is written for the readers' convenience.
-
Proof.
The proof of
is verbatim the same as that of Lemma 5.3.6 (a) in [3] . Now we assume
, i.e.,
. This is equivalent to the condition that there exists a monomial
and
such that
-
for all
, and
-
,
where
denotes the multidegree. We know from
that we have
because a negative degree
in
must come from the denominator of the fraction
and
. Now we know that
is equivalent to the following condition:
for all
and for all
we have
where we set
with some integers
,
. Namely, for all
there exists
such that
. Furthermore, we know from the condition
that we have
for
since by
non-negative degrees in
must come from
. Consequently we obtain
.
Now we show the converse. Assume that we have
. Set
and
. Then since
there exists
and a monomial
in
such that
Now we show that
in
.
is equivalent to the condition that
for all
. As in the above discussion, this is equivalent to the condition
|
(2)
|
where we set
for some integers
,
. But by ( 1 ) we have
for
, so that
(i.e.,
) or
(i.e.,
). Hence we can replace “
” in ( 2 ) by “
” and then ( 2 ) is assured by the assumption. Thus we have
in
. Therefore
as requied. □
Let
. By Lemma 1 we see that
has a basis
Restricting the differentation of
to the
th graded piece, we obtain a complex
of finite dimensional
-vector spaces with differentation
given by
where the sum is taken over all
such that
with
and for all
there exists
such that
.
Also we define
if
and
. Then we describe the
th component of the local cohomology in terms of this subcomplex:
.
Now we fix our notation on simplicial complex. A simplicial complex
on a finite set
is a collection of subsets of
such that
whenever
for some
. Notice that, for the convenience in the later discussions, we do not assume the condition that
for
. We define
if
and
, which will be called the dimension of
. If we assume a linear order on
, say
, then we will call
oriented, and in this case we always denote an element
with the orderd sequence
. For a given oriented simplicial complex of dimension
, we denote by
the augumented oriented chain complex of
:
where
for all
. Here we define
for
. Now for an abelian group
, we define the
th reduced simplicial homology
of
to be the
th homology of the complex
for all
. Also we define the
th reduced simplicial cohomology
of
to be the
th cohomology of the dual chain complex
for all
. Notice that we have
and if
then
and
for all
.
Now we will establish an isomorphism between the complex
,
, and a dual chain complex. For any
, we define a simplicial complex
Notice that we may have
for some
.
Lemma 2.
For all
there exists an isomorphism of complexes
where
means shifting the degree of
by
.
-
Proof.
The assignment
induces an isomorphism
of
-vector spaces such that
, where
That this is a homomorphism of complexes can be checked in a straightforward way. □
Now we come to the main theorem in this section.
-
Proof.
By Lemma 2 and universal coefficient theorem for simplicial (co)homology, we have
| |
| |
It is clear from the definition that
if for all
we have
.
Moreover for all
with
for at least one index
we have
. To prove this fact we can assume without loss of generality that
and that
. Then we have
, and, for all
with
and
, we have
. In fact, since we have
for all
the existence of
with
implies
. Consequently we know that
is a cone by the vertex
so that, as is well known, we have
for all
as required. Thus we obtain
| |
Now if
, we must have
, i.e., for all
there exists
such that
, and this implies that
for all
, namely
is not a non-face of
, i.e.,
. Thus we finally obtain the required formula. □
The original Hochster's formula is a special case of Theorem 1 .
Corollary 1 (Hochster).
Let
be a simplicial complex and let
be the Stanley-Reisner ring corresponding to
. Then we have
where
.
-
Proof.
By Theorem 1 we have
where
and
| |
| |
| |
Then the rest of the proof is exactly as in Theorem 5.3.8 [3] . □
1.2 Vanishing degrees of local cohomolgies
In this subsection, we give some easy consequences of Theorem 1 . We define
if
and
if
.
Similarly, we define and
if
and
if
.
Recall that
for
.
Corollary 2.
Let
be a monomial ideal. Then
for all
.
-
Proof.
By Theorem 1 , the terms in
with the highest total degree are at most
with
for
. Thus the total degree is at most
. □
From Corollary 2 , we can recover the following well known result.
Corollary 3.
Let
be a generalized CM Stanley-Reisner ideal.
Then
for all
.
-
Proof.
If
is square-free, then
for
. □
For a Stanley-Reisner generalized CM ideal
with
, it is well known that it is Buchsbaum and
for all
. The following theorem extends this result to monomial ideals in general.
Proposition 1.
Let
be a monomial ideal. Then following are equivalent:
-
-
for all
with
, in particular
-
for all
with
and
.
-
Proof.
The equivalence of
and
are immediate from Theorem 1 . We will prove the equivalence of
and
. Assume that
. Assume also that there exists
such that
,
and
. Now observe that by the definition of
, the condition is independent of the values
for
. This means that the total degree
can be any negative integer so that
is not of finite length, which contradicts the assumption. Thus we must have
for all such
. The converse implication is straightforward. □
Corollary 4.
Let
be a generalized CM monomial ideal with
. Then
for all
.
For a generalized CM ring
, there exists an integer
,
, such that
for
. If this condition holds, we will also call
, or
,
-Buchsbaum. An ideal
is generelized CM if and only if it is
-Buchsbaum for some
. If
is
-Buchsbaum but not
-Buchsbaum, then we will call
strict
-Buchsbaum.
Proposition 2.
Let
be a generalized CM monomial ideal. Then
is
-Buchsbaum.
-
Proof.
is
-Buchsbaum. Then the required result follows immediately from Corollary 2 and Corollary 4 . □
From Proposition 2 , we immediately know that a Stanley-Reisner ideal is 1-Buchsbaum if it is generalized CM, which is a weaker version of the well-known result that a generalized CM Stanley-Reisner ideal is Buchsbaum. The bound of
-Buchsbaumness given in Proposition 2 is best possible. In fact, we can construct strict
-Buchsbaum ideals as in the following example.
Example 1.
Let
be a Stanley-Reisner Buchsbaum ideal. Notice that such ideals can be constructed with the method presented in [
1]
and
(
) is a
-vector space for
.
Now consider a
-homomorphism
where
with
for
. We define
for a
-module, where a left-right
-module
is equal to
as a set, it is a right
-module in the ordinary sense and its left
-module structure is determined by
. Then we have
-
(1)
,
-
(2)
is an exact functor.
Thus, for
, we have
and since
is a direct sum of
,
is a direct sum of
.
Then we know that
but
with
.
Remark 1.
Bresinsky and Hoa gave a bound for
-Buchsbaumness for ideals generated by monomials and binomials
Theorem 4.5 [
2]
. For monomial ideals, our bound is stronger than that of Bresinsky and Hoa. Also, according to K. Yanagawa, Proposition 2 can also be deduced from his theory of square-free modules [
10]
.
Recall that Castelnuovo-Mumford regularity of the ring
is defined by
Let
. Then we have
for
. Then we have
Corollary 5.
Let
be a generalized CM monomial ideal with
and
. Then
for
. In particular, if
has
-linear resolution, we have
for
.
-
Proof.
First part is clear from Corollary 4 . If
has
-linear resolution, we have
. Thus the last part also follows immediately. □
2 Generalized Cohen-Macaulay monomial ideals
2.1 FLC property
In this subsection, we give a combinatorial characterization of FLC (finite length cohomology) property for monomial ideals, as an application of Theorem 1 . We prepare some notations. Let
be a monomial ideal. If
for some
and
, we easily know that
must be
. Then, by changing the name of the variables if necessary, we can write without loss of generality that
with
, where
. We denote by
the simplicial complex corresponding to a square-free monomial ideal
, which is a complex over the vertex set
. We regard
as a partially ordered set by defining
,
, to be
for
. We denote
. For
and a monomial
, we define
. Also for
with
and
, we define
as follows:
We abbreviate
as
for
.
Now we prove
Theorem 2.
Let
be a monomial ideal. If
then, for all
-face
and for all
such that
-
,
-
for all
, and
-
is maximal with the properties
and
for a fixed
,
we have the following: there exists
such that
-
(i)
, and
-
(ii)
for all
with
we have
.
-
Proof.
Assume that
. Then, by Proposition 1 , we have
for all
with
,
and
. This implies
.
Now for such
we set
. Notice that
since
. We also notice that, as far as the complex
is concerned, the values
for
are irrelevant. Thus we will change the values
and assume that
.
Notice that
, for the new
, is the same as before we change the values
for
.
We know that
is equivalent to the condition that there exists
such that
. Now we assume that
for all
, namely
.
The condition
is equivalent to the condition that
for some
, i.e., there exists
such that for all
we can find
satisfying
. Namely,
|
(3)
|
Under this condition we have, for any
with
and
,
for all
. Thus we can assume that
is maximal satisfying the condition that
for all
and
. Also, since
and
, we have
for all
. Hence we can replace '
' by '
' in the maximality condition for
. Now we have only to show that the condition ( 3 ) is equivalent to
and
in the statement.
Since we have
for all
, we can only find the index
as in 3
in
. Now for
, the existence of
satisfying
is always assured for every
. Moreover, if
for
, then
and this is
since
. Thus 3
is equivalent to the existence of
such that
|
(4)
|
Assume that
and set
as
for
and
.
Then by the maximality of
there exists
such that
, which contradicts the condition 3
. Thus we must have
for
as in 3
. If
is such that
, then
by assumption on
. Thus we can replace '
with
' in the condition ( 4 ) by '
with
'.
Consequently we know that 4
is equivalent to
and
. □
From Theorem 2 , we can recover a weaker version of the well-known result as follows.
Corollary 6.
If
is a generalized CM Stanley-Reisner ideal, i.e., Buchsbaum ideal, then
is pure, namely, every facet has the same dimension.
-
Proof.
Let
be a generalized CM Stanley-Reisner ideal. Then by Theorem 2 we know that for every
and for every
-face
there exists
with
such that
. From this we immediately know that
is pure. □
2.2 generalized CM monomial ideals of
If
,
is always (generalized) CM. For
, we can give combinatorial characterizations of generalized CM monomial ideals as follows. First we give the
case.
Corollary 7.
A monomial ideal
is generalized CM with
if and only if
-
, and
-
for all
and for all
such that
,
for all
, and
is maximal with the properties
and
for a fixed
, we have the following: there exists
such that
-
(i)
, and
-
(ii)
for all
with
we have
.
-
Proof.
As is well-known,
if and only if
. Then
is generalized CM if and only if
for
.
is always of finite length and
is of finite length if and only if
for all
with
and
by Proposition 1 . If
, we always have
. Now let
. Since
we must have
.
Also
if and only if
, which is equivalent to the condition in the statement by the proof of Theorem 2 . □
In
case, we need to give a combinatorial criterion for connectedness of simplicial complexes. Notice that a simplicial complex
over the vertex set
is not connected if and only if there exists non-empty disjoint subsets
such that
and for all
and all
there is no 1-face
.
Lemma 3.
Assume that
,
, and
. Then the set of vertices of
is
.
-
Proof.
The
th skeleton of
is
| |
| |
| |
where the first equation is because if
there is no
such that
and if
we always have the index
such that
for
, which is actually
. □
Now we show a combinatorial characterization of
generalized CM monomial ideals.
Theorem 3.
A monomial ideal
is generalized CM with
if and only if
-
, and
-
for all
and for all
such that
,
for all
, and
is maximal with the properties
and
for a fixed
, we have
-
there exists
such that
-
1.
, and
-
2.
for all
with
we have
-
there are no non-empty disjoint subsets
satisfying the following property:
-
1.
where
, and
-
2.
for all
and all
there exists
such that
-
for all 1-face
and for all
such that
,
for all
, and
is maximal with the properties
and
for a fixed
, we have the following: there exists
such that
-
1.
, and
-
2.
for all
with
we have
.
-
Proof.
if and only if
. Now assume that
. Then
is generalized CM if and only if
and
, which is equivalent to
|
(5)
|
and
|
(6)
|
for all
with
and
by Proposition 1 . The condition 5
is equivalent to
by Corollary 7 and the condition 6
is equivanent to
|
(7)
|
and
|
(8)
|
since
for
. The condition 7
exactly means the connectedness of the simplicial complex
. Let
be the set of vertices of
. By what we noticed just before Lemma 3 , this is equivalent to the condition that there exist disjoint no non-empty subsets
such that
and for all
and all
we have
. By Lemma 3 we have
where
. Since we have
,
implies
. Thus
. The condition
is equivalent to the condition that
for some
. This can also be refined to the condition that
for some
. In fact, first of all we have
, for all
, since
and
. Thus we can replace '
' in the above condition by '
'. Also, since
,
and
for all
. Thus
is equivalent to
as required. This is the condition
Now we will show that if
is such that
,
,
for all
and
, then
is also connected. We prove the contrapositon: if
is disjoint then
is disjoint too. Assume that there exist disjoint non-empty subsets
such that
and for all
and all
we have
for some
. First of all, for
we have
so that
and thus
. Now for all
and
,
for some
.
Also since
we must have
and
. Then we know that we must have
. Thus, by setting
and
, we obtain the non-empty disjoint subsets
showing the disjointness of
. Consequently, we can assume
to be maximal as in the statement
. Finally, by the proof of Theorem 2 , we know that the condition 8
is equivalent to
. □
Remark 2.
Unfortunately we do not know a good combinatorial characterization for
for
, which is needed to obtain similar results to Theorem 3 for
.
2.3 Construction from Buchsbaum Stanley-Reisner ideals
In this subsection, we compare local cohomologies of monomial ideals
and
. It is well known that
for all
with
, which is an immediate consequence from the original version of Hochster's formula for Stanley-Reisner ideals. On the other hand, we may have
for such
. But for multi-degrees
with
, we have an isomorphism.
Proposition 3 (Herzog-Takayama-Terai [4] ).
Let
be a monomial ideal. Then we have the following isomorphisms of
-vector spaces
for all
with
.
Proposition 4 (Herzog-Takayama-Terai [4] ).
Let
be a monomial ideal. Then
implies
. In particular, if
is generalized CM, then
is also generalized CM (Buchsbaum).
-
Proof.
We will give here a new proof, which is different from that in [4] . Assume that
. Then by Proposition 1
for all
with
. Thus if
and
, we must have
. Now by Proposition 3 we have
for all
with
, and this is non-zero if and only if
. Thus we have
. □
Corollary 8.
Let
be a generalized CM monimial ideal and assume that
is not Cohen-Macaulay (but generalized CM by Proposition 4 ).
Then
is not Cohen-Macaulay.
-
Proof.
Assume that
is Cohen-Macaulay. Then, by Proposition 3 and the comment before Proposition 3 , we have
for all
and for all
, namely
for all
and
is Cohen-Macaulay. □
These results suggests a method for constructing (non-CM) generalized CM monomial ideals from Buchsbaum Stanley-Reisner ideals: given a Buchsbaum Stanley-Reisner ideal
, make monomials
,
, for each generator
. In general, one can make more than one monomial generators from a single square-free generator. If we choose suitable exponents
, the ideal generated by the monomials is (non-CM) generalized CM. Theorem 2 , Corollary 7 and Theorem 3 give the criteria for suitable exponents.
Example 2.
Let
,
. It is easy to check that
is Buchsbaum of dimension
and depth
. Let
for some
.
Then
is generalized CM ideal if and only if
for all
and
for all
, namely
is an image of Frobenius map in the sense of Example 1 .
Remark 3.
Notice that if
then both
and
are Cohen-Macaulay for all
.
-
Proof of Example 2 .
Assume in the following that
is generalized CM. Then
must satisfy the conditions in Theorem 2 , in particular the condition for
.
Now we set
and
for
and
and
for
. Notice that
and
denote the maximal and the minimal exponents of the variable
,
. Let
be the simplicial complex corresponding to
, which is the disjoint union of two
-simplices over the vertex set
and
. Then an
-face
is either
for
or
for
.
We know that the condition for
such that
to be
for all
is as follows:
-
Case (1)
for
Since
| |
| |
we must have
-
(1)
for all
, and
-
(2)
for every
we have at least one of the followings: (a)
, (b)
.
-
Case (2)
for
Since
| |
| |
we must have
-
(1)
for all
, and
-
(2)
for every
we have at least one of the followings: (a)
, (b)
.
According to Theorem 2 , for a maximal
, where
and
are as above, there must exist an index
such that
and
for every
with
we have
. Moreover, by the proof of the Theorem 2 we know that
can be replaced by
We now consider the condition for
,
and
satisfying
and
.
-
Case (3)
with
By
we must have
for all
. This holds if and only if
and
.
-
Case (4)
with
Similarly, we have and
for any
with
.
Now in case (1), an
satisfying
and
must be
,
or
. Assume that
. Then the condition of case (1) must imply the condition of case (4). Since
| |
| |
we have
so that
is constant for all
with
. Now if
for some, and equivalently all,
, then we have
so that
. Then we have
, a contradiction. Thus we know
for all
, and the index
as in the condition of case (4) can be at least
with
but this contradicts the condition
. Consequently, an
satisfying
and
cannot be from
and we must have
with
.
Now the condition of Case (1) must imply the condition of Case (3). Comparing Case (1) 1 with the condition of Case (3), we know that we must have
|
(9)
|
for every
with
. Now we consider similarly with the Case (2) and obtain the condition
|
(10)
|
for every
with
. Finally, the condition 9
for
together with the condition 10
for
entails
are constant for all
and
are constant for all
, i.e.,
is obtained from
by Frobenius transformation in the sense of Example 1 . □
Example 3.
If we allow to make more than two monomial generators from a single square-free generator, we can construct more generalized CM monomial ideals from the same Stanley-Reisner ideals as in Example 2 , For example, from
we make
Also from
we make
and
are both generalized CM, but for example
is not generalized CM.
References
-
A. Björner and T. Hibi, Betti numbers of Buchsbaum complexes, Math. Scand. 67 (1990) no. 2, 193–196.
-
H. Bresinsky and L. T. Hoa, The
-Buchsbaum property for some polynomial ideals, J. Math. Kyoto Univ. 43-4 (2003), 699–717.
-
W. Bruns and J. Herzog, Cohen-Macaulay rings, Cambridge studies in advanced mathematics 39, Cambridge, 1993.
-
J. Herzog, Y. Takayama and N. Terai, On the radical of a monomial ideals, preprint, (2004).
-
J. Herzog, D. Popescu and Marius Vladoiu, On the Ext-modules of ideals of Borel-type, Contemp. Math. 331 (2003), 171-186.
-
M. Hochster, Cohen-Macaulay rings, combinatorics, and simplicial complexes. In B. R. McDonald and R. A. Morris (eds.), Ring theory II. Lect. Notes in Pure and Appl. Math. 26, M. Dekker, (1977) 171–223.
-
P. Schenzel, On the number of faces of simplicial complexes and the purity of Frobenius, Math. Z., 178, (1981), 125–142.
-
R. P. Stanley, Combinatorics and commutative algebra, Birkhäuser, second edition, 1996.
-
J. Stückrad and W. Vogel, Buchsbaum rings and applications, Springer, 1986.
-
K. Yanagawa, Alexander duality for Stanely-Reisner rings and squarefree
-graded modules, J. Algebra 225, No. 2, (2000), 630-645.
Yukihide Takayama, Department of Mathematical Sciences, Ritsumeikan University, 1-1-1 Nojihigashi, Kusatsu, Shiga 525-8577, Japan E-mail address : takayama@se.ritsumei.ac.jp