1. Introduction
The interplay between the combinatorics of finite simple graphs
G and the algebra of the underlying edge ideals
has been studied by various researches during the last few decades. The algebraic invariants that have been particularly prone to combinatorial interpretation are regularity, projective dimension, depth, and Betti numbers. In this article, we study the depth of powers of edge ideals of bipartite graphs. Combinatorics of bipartite graphs have been particularly ripe with interesting algebraic counterparts in the edge ideals and their powers. Interested readers are referred to [
1,
2,
3], etc. In this paper, we continue the study pursued by the same authors in [
3]. We study the closely related topics of combinatorial characterization of regular elements and Cohen-Macaulayness of various powers of bipartite edge ideals.
In section two of this paper, we offer a new characterization of Cohen-Macaulay bipartite edge ideals. We characterize it using colon ideals of the form
, where
e is an edge/generator of
, somehow in the same way as it is done in [
3,
4], etc., in the study of regularity. An often quoted and important characterization of Cohen-Macaulay bipartite edge ideals is due to Herzog-Hibi in [
2]. In this article, we also give a new proof of this characterization ([
2]). One important feature of our proof is that it does not use Hall’s marriage theorem or any variant of it as it was done in [
2]. Throughout this article, we refer to
S as the polynomial ring
. Our main results in this section are as follows:
Theorem 1. Let G be a bipartite graph with partition and . Then the following are equivalent
- 1.
is Cohen-Macaulay
- 2.
and there exists a re-ordering of the vertex sets such that
- (a)
for all i
- (b)
If then .
- (c)
If then
- 3.
is unmixed and is connected in codimension one.
- 4.
and there exists exactly n edges such that and for , and are disjoint.
- 5.
and there exists exactly n edges such that is Cohen-Macaulay and for , and are disjoint.
For arbitrary bipartite edge ideals, it is often hard to compute the depth of powers of its edge ideals
for all
. Even if
G is Cohen-Macaulay, it is not so easy to compute the
for
. It is well known that
is asymptotically equal to the number of connected components of
G ([
5]). An important invariant related to the study of depth
is the dstab
which measures the minimal
t for which depth
equals the number of connected components of
G. To study such invariants the same authors in [
3] characterized regular elements on
for any unmixed bipartite graphs
G. In the third section of this paper we characterize elements of the form
that are regular on the powers
of a bipartite edge ideal
G. This is a generalization of the similar result proved in [
3]. Our characterization turns out to be the exactly same as the ☆-condition proved there. To signify its usefulness we call it the neighborhood properties (we refer to the definition in Definition 12) Our main result proved here is as follows:
Theorem 2. Let G be a bipartite graph and suppose that and satisfies the neighborhood properties. Then is an regular element on for all s.
2. Structure of Cohen-Macaulay and Unmixed Bipartite Graphs
A characterization theorem for Cohen-Macaulay bipartite graphs was given by Herzog-Hibi in [
2].
Theorem 3. (Herzog-Hibi, [
2])
Let G be a bipartite graph with partition and . Then the following are equivalent- 1.
is Cohen-Macaulay
- 2.
and there exists a re-ordering of the vertex sets such that
- (a)
for all i
- (b)
If then .
- (c)
If then
The following theorem is an improvement of the Herzog-Hibi characterization (Theorem 3). We are grateful to Prof. Huneke for the ideas presented in this proof. It is important to notice here that the following theorem does not make use of the Halls marriage theorem which is an important element of any proofs known to us of Theorem 3.
Definition 1. (Definition, p. 498, [6]) Let I be an ideal in a polynomial ring S such that . We say that the ring is connected in codimension one if for any two primes , there is a sequence of minimal primes such that for each , ht in . Theorem 4. Let G be a bipartite graph with partition and . Then the following are equivalent
- 1.
is Cohen-Macaulay
- 2.
and there exists a re-ordering of the vertex sets such that
- (a)
for all i
- (b)
If then .
- (c)
If then
- 3.
is unmixed and is connected in codimension one.
Proof. First we show
. We prove by induction on
n. If
, then
and hence clearly
is Cohen-Macaulay. Now assume that the result is true for
and let
G be a graph which satisfies the conditions
of
on
vertices (with partition
and
). Consider
Notice that
, where
is the graph obtained by deleting
and
from
G. Clearly
satisfies the conditions
of
and hence
is Cohen-Macaulay (on
vertices) by induction. So
is Cohen-Macaulay of dimension
n. Let
for some
. Let
for some
. As
by the condition
,
and hence
. So
, where
is the graph obtained from
G by deleting
. But by induction,
is Cohen-Macaulay of dimension
. Hence
is Cohen-Macaulay of dimension
n. Now in (
1), both
and
are Cohen-Macaulay of dimension
n, we have
is also Cohen-Macaulay of dimension
n ((Proposition 1.2.9, [
7]) and the fact that dimension of
is the maximum of the dimensions of
and
).
The implication
is a consequence of (Corollary 2.4, [
6]).
We finally show
. We first observe that
as
is unmixed and both
and
are minimal primes. Next, we prove that the existence of conditions
and
by induction. Let
and define
Note that for any subset . We now consider the ideals . If then which shows that . Since is unmixed, we have . Clearly for any , where for each , for some and for each , for some . Since is unmixed of height n and , we have .
Now choose with minimum vertex degree. Without loss of generality we may assume . Let be neighbors of and . Then as in the previous paragraph, consider . After relabeling, we may assume are only connected to . Let be the induced subgraph on . By our choice of , of minimal vertex degree t, notice that every other vertex has to have vertex degree at least t. In other words, since t is minimal, each vertex in has at least t neighbors and hence is a complete bipartite graph.
Since is connected in codimension one and , there exists a sequence of minimal primes such that in . If any minimal prime of does not contain some then it has to contain every (as ,as defined in the previous paragraph, is a complete bipartite graph). Let such that for all , contains all of (alternatively, ’s do not contain any of ). Now does not contain at least one of , hence it has to contain all . So in . Thus and hence is only connected to .
Now consider
. Since
is an intersection of minimal primes,
is an intersection of minimal primes of
containing
. Thus any minimal prime of
is a minimal prime of
, and so
is unmixed. We now show that
is connected at codimension one. Any minimal prime of
has to contain either
or
(as it is minimal it cannot contain both as
is only connected to
). Let
. As
, there exists a sequence of minimal primes
such that
. For any
,
The sequence defined as before has the property that and hence is connected in codimension one. Now notice that where is the graph obtained from G by deleting . By induction hypothesis, is Cohen-Macaulay. So there exists an ordering and satisfying of . As is only connected to , G also satisfies of .
To prove that condition holds, take and in such that are distinct. Assume that is not an edge. Then there is a minimal prime P that does not contain either or as the ideal generated by all x-variables except and all y-variables except is a prime ideal that contains and does not contain or . Now because is unmixed, height of this prime has to be n. Since and are not in P, we get that and are both in P. As P contains at least one of or for all m, one observes that height of P is strictly bigger than n, which is a contradiction. □
The following remark is extremely crucial for our work.
Remark 1. If G is a bipartite graph and is an edge then from (Theorem 6.7, [4]) we get . Theorem 5. Let G be a bipartite graph with partition and . Then the following are equivalent
- 1.
is Cohen-Macaulay
- 2.
and there exists exactly n edges such that and for , and are disjoint.
- 3.
and there exists exactly n edges such that is Cohen-Macaulay and for , and are disjoint.
Proof. First, we show . If is Cohen-Macaulay, we have ordering and of the vertices of G which satisfies the conditions of Theorem 4. Condition implies for all i, and conditions and implies for .
Now suppose there exist, after possible reordering,
which satisfied the conditions of
. First, we show that if
is the induced subgraph obtained by deleting
and
then the edge ideal
related to
satisfies the condition with
. Without loss of generality, we prove this for
. Clearly
for
. Suppose there exists an edge
such that
. Without loss of generality we may assume
. As
and
is an edge we can conclude that there exists a minimal generator of
which is an edge that is either of the form
or
(Theorem 6.7, [
4]). Again without loss of generality we may assume it is of the form
as the proof for the other follows simply by interchanging roles of
x and
y. So
and
are edges in
G (Theorem 6.7, [
4]). As
we conclude
is an edge in
G. As
we observe that
has to be an edge in
G. So
. Without loss of generality we may assume
. Now
so
has to be an edge in
G. Again
hence
is an edge in
G contradicting the assumption. So we may assume for all
i the edge ideal
of the graph
obtained by deleting
and
satisfies the conditions in
.
Now by induction we may assume the result holds for . Pick such that has minimum degree. Let be the induced subgraph on vertices other than with edge ideal . As satisfies the condition it is Cohen-Macaulay by induction. Without loss of generality we may assume and ordering that gives ordering of previous theorem for is . As has degree one in it can have at most degree 2 in G. If is not an edge, due to minimality degree of is at most 1. If is an edge in G and is an edge in G for , as , we have is an edge in G and hence in contradicting the assumption. Now if and both are edges in G. Notice that also satisfies the hypothesis (). For, has to be connected to any neighbor of as is an edge and satisfies the hypothesis (). This leads to a contradiction and hence no for is connected to . This guarantees that conditions and of Theorem 4 is satisfied. The condition is satisfied as for all i, .
Next we show . To prove the if part, we pick, without loss of generality, with minimum degree and the corresponding edge . If degree of more than one then degree of any other vertex is more than one; as is Cohen-Macaulay this will be a contradiction to the fact that any Cohen-Macaulay bipartite graph should have a y-vertex of degree 1 (Theorem 3). So has degree one. Hence and is Cohen-Macaulay.
For the only if part let
be as the ordering prescribed by the Herzog-Hibi (Theorem 3) characterization. All we need to show is that
is not Cohen-Macaulay for
. This follows as
for
(which is a minimal monomial generator of
J) as well as for
. To see this first we show that
for all
k. Here at every step we use the description of colon ideal provided by (Theorem 6.7, [
4]). If
is a minimal monomial generator of
which is not in
J then
and
are in
J. Both of them cannot belong to
as from
that will imply
belongs to
and as a result will belong to
J, contradicting the assumption. Without loss of generality assume
does not belong to
. Then
and
is in
. If
does not belong to
then
and
belong to
. If
is in
as
is in
and
we have
is in
. In either case we have
and
belong to
. Hence
belongs to
J contradicting our assumption.
Next we show that . If is a minimal monomial generator of which is not in J then and is in J. As is in J it is either in or is a neighbor of in G. If is in as we have is in . By symmetry is in . Hence is in J contrary to the assumption. Hence J is not Cohen-Macaulay. □
The next theorem gives insight into the associated graded ring of a Cohen-Macaulay bipartite edge ideal. The proof of this theorem uses the description of the colon of the
power of an edge ideal with
edges introduced in [
4].
Theorem 6. Let be Cohen-Macaulay bipartite edge ideal with an ordering of vertices satisfying Theorem 3(2) and for . Then for all i and for all k, . Hence s are non zero divisors in the associated graded ring of .
Proof. Let
be a minimal monomial generator of
. By induction
. So
where
s are minimal monomial generators of
and
g any monomial. So
. As
f is a minimal monomial generator, without loss of generality we may assume
g is of degree 2 and
is a minimal monomial generator of
. Let
. If
g is an edge we are done. Otherwise by ([
4], Theorem 6.7),
and
are even connected with respect to
. If
is an edge and for some
,
and
is an edge. Then by Theorem 4(2(c))
is an edge and hence proceeding inductively we show
g is an edge and the result follows. □
We illustrate this theorem for .
Example 1. Let and . One can check using Macaulay 2, that and .
In a private communication, Prof. Villarreal mentioned that results similar to Theorems 5 and 6 can be found in [
8,
9].
3. Regular Elements in Powers of Bipartite Edge Ideals
This section presents methods to recognize regular elements on the power of bipartite edge ideals based on the combinatorics of the graph. We first present some examples to motivate the definition and the results.
Example 2. Consider the ring and the bipartite edge ideal corresponding to
Macaulay2 computations show that is a regular element on for . Notice that is Cohen-Macaulay. This can also be recovered from (Theorem 3.8, [3]). One would be tempted to generalize that is always a regular element for bipartite graphs. But it is not always the case as it is shown in this example.
Example 3. Consider the ring and the bipartite edge ideal corresponding to
Macaulay2 computations show that is not a regular element or .
Studying more such examples, we came up with the following definition involving the combinatorial nature of the graphs.
Definition 2. Let G be a bipartite graph. Then satisfies the neighborhood condition if Remark 2. Condition (
2)
of Definition 2 is equivalent to the following conditionSuppose (
2)
of Definition 2 is true. Then , where . This means where . In other words, , where . The other direction is analogous. We show in [
3] that
is a regular element on
for all
when
G is an unmixed bipartite graph. Of course, when
G is unmixed bipartite,
and
satisfies the neighborhood conditions. In this section, we show that the difference of vertices which satisfies the neighborhood condition are the right candidates for being a regular element on
for any bipartite graph
G.
Theorem 8 is the main theorem we study in this section. We break up the proof of this theorem into three main parts, where Theorem 7, Lemma 1 provide all the tools required to prove Theorem 8.
Theorem 7. Let G be a bipartite graph and suppose that and satisfies the neighborhood properties. If m is a monomial such that , then for .
Proof. We prove by induction on k. Suppose . Then . As , then either or for some and . If , then the claim is obviously true.
Suppose
with
. Let
for
. We assume
. Since
, we have
. Thus
Suppose a neighbor of divides a, then clearly . Now suppose a neighbor of divides a. Since and satisfies the neighborhood properties, any neighbor of is also a neighbor of and hence .
Suppose that no neighbor of
or
divide
a. Now in the decomposition in (
3), if
does not divide
, then
divides
and hence
. Now if
does not divide
, then
divides
. Thus
. If
divides
, then
divides
and hence
. Now suppose
divides, say
. Again, since
and
satisfy the neighborhood properties, any neighbor of
is a neighbor of
and hence
.
Now suppose that
divides
. Since
divides
, we assume, without loss of generality,
and
. Thus we have
Now a neighbor of
, say
, divides
a, then
Similarly if a neighbor of
, say
, divides
a, then
Now suppose no neighbor of
or
divides
a. Consider (
4). If
does not divide
, then
Deleting
on both sides, we get
. Thus we assume
divides
and hence assume, without loss of generality
. Now we have
Now if
does not divides
, then
Deleting on both sides we get .
Thus we assume
divide
and hence assume, without loss of generality,
. We now have
We continue in the same fashion and arrive at the
j-th decomposition
Also
and
for
. Now if a neighbor of
, say
, divides
a, then
If a neighbor of
, say
, divides
a, then
Now suppose no neighbor of
or
divides
a. Now consider (
7). If
does not divide
, then
Deleting
on both sides we have
. Thus we assume
divides
and hence assume, without loss of generality,
. We now have
Again, if
does not divide
, then
Deleting on both sides, we get . Thus we assume divides and hence assume, without loss of generality, .
Continuing in the same fashion we may reach the final decomposition
Recall that every stage we make sure that none of the neighbors of the
y’s appearing in
divide
a. Thus a neighbor of
or
divides
a. Now we can use the decomposition in (
8) and (
9) to show that
depending on whether
or
is odd or even. This concludes the proof of claim of this theorem in
case.
Now assume by induction, that if for , then . Suppose . We also assume that . For, if , then and hence by induction hypothesis, we have .
We claim that it is enough to show that or . Suppose we show that . We now have . Thus and hence . Since is a monomial, we use induction hypothesis to conclude that . Thus we now have . As before, we have . Again, since is a monomial, we use induction hypothesis to conclude that . We now have . We continue the process to get . We still have . Since is a monomial, by induction hypothesis, we get . We now have . This is the case. We now use the induction hypothesis to get . On the other hand, if we show that , then we can analogously show that .
Now we go to the induction step. We have . Since and for any , we have where and not all may be distinct. Suppose a neighbor of divides a, then .
Now suppose no neighbor of
divide
a. Since
we have
We observe that may be written divisible by many minimal monomial generators of . We can take such that has smallest number of x variables in common with .
It is clear that must divide , otherwise for some and hence . Recall the no neighbor of divides a. Thus we can assume that no neighbor of , divides a as that will make . So without loss of generality we may assume for where for every j, .
Now we observe that if any neighbor of for divide a then, clearly, . For, without loss of generality, say is an edge where divides a. As is an edge, so is (by neighborhood properties). Thus we have . Hence this will force . So we assume no neighbor of for divide a.
As there are s many x variables in and , some of the x variables of divides a. We also have that no neighbor of any divides a and divides . Let where divides a and . We may write where . But this is an expression of with having less number of x variables in common with than a which is a contradiction. Thus, one of the neighbors of for some divides a and hence . □
Lemma 1. Let G be a bipartite graph and suppose that and satisfies the neighborhood properties. Now assume are monomials of the same degree such that . Further suppose,Then for . Proof. First, assume that
. We prove by induction on
k. If
, then clearly the claim is true by Theorem 7. By induction, assume the claim is true for
satisfying (15)–(17) and
. Now suppose we have
satisfying (15)–(17). We show that
. This will show that
satisfying (15)–(17). Thus by induction hypothesis we have
for
proving the claim.
From (
15), we have
and
where
, a monomial. From (16), we have
. Subsequently, we show that
Since , we have or for some where .
Suppose . Since , divides a or one of the ’s. If divides a, then .
Now suppose
divides, say
for some
. Since
, we have
. Using this equality in (
18), we have
Since
does not divide
a, then
divides some of the
and hence we have
or
. Without loss of generality, assume
divides
. Thus
Let
. Now as
, we have
. Also, notice that
Since is a monomial, we have , by Theorem 7. Now and hence we are done. □
We now prove one of the main results of this section. In this theorem, we attempt to rearrange the sum into a configuration shown in the previous lemma.
Theorem 8. Let G be a bipartite graph and suppose that and satisfies the neighborhood properties. Then is an regular element on for all s.
Proof. Consider where ’s are monomials of the same degree. We prove by induction on k.
Suppose and . Thus . Now we use Theorem 7, to show that proving the base case of induction.
Suppose
for
implies
. Now consider
where all
’s are distinct. We show
for
.
Observe that if , then we have and . Now we use induction hypothesis to show that for .
Now we first consider the following configuration, i.e., after possible re-ordering of
’s we have
We refer to this case as the
k-cancellation case. Using (
20), we get
and
. Using this and (20), we get
Thus . Using this description in (22) we get , a contradiction.
Now consider (
19). Without loss of generality, after possible reordering, assume that
. If
, then we get
and
. Now using induction hypothesis, we get
.
Suppose, if
we introduce the re-ordering
Notice that . Thus it is enough to show that . Under this re-ordering and . If , then we get and . Now using induction hypothesis, we get and hence .
Now if
, we introduce a new ordering
As before we consider if or and introduce new ordering, if necessary.
We now continue this process and arrive at the
j-th re-ordering defined as follows
with the following configuration
First, suppose . As before, we consider two cases or . If , then we arrive at the k-cancellation case discussed above, which leads to a contradiction. So we have which is discussed separately in Lemma 1, showing that .
Now we assume and for . If for , then we have and and we use induction hypothesis to conclude that and hence for .
Thus assume
for some
. Now we use the ordering
with the configuration
for
.
Now if for , then and and we use induction hypothesis to conclude that and hence for .
Now if
for some
, then we use the ordering as before
with the configuration
for
.
We continue in the same fashion to reach
-th re-ordering to get
with the following configuration
Suppose , then . Now using Lemma 1 we have and hence for .
If , then there exists a term such that and and hence we are done by induction. □
Corollary 1. Let G be a bipartite graph. Suppose . Then and satisfies the neighborhood properties, if and only if is regular on for all s.
Proof. Suppose and satisfies the neighborhood properties, then is regular on for all s by Theorem 8.
Now if
and
does not satisfy the neighborhood properties, then there exists
such that
and
where
. Thus for all
s and
,
Since
, we get
. Thus
is not a regular element on
. □