Properties of the Toric Rings of a Chordal Bipartite Family of Graphs
Abstract.
This work concerns the study of properties of a group of Koszul algebras coming from the toric ideals of a chordal bipartite infinite family of graphs (alternately, these rings may be interpreted as coming from determinants of certain ladder-like structures). We determine a linear system of parameters for each ring and explicitly determine the Hilbert series for the resulting Artinian reduction. As corollaries, we obtain the multiplicity and regularity of the original rings. This work extends results easily derived from lattice theory for a subfamily coming from a two-sided ladder to a family where, as we show, lattice theory no longer applies in any obvious way and includes constructive proofs which may be useful in future study of these rings and others.
Key words and phrases:
Koszul algebra, toric ring, graph, chordal bipartite, Castelnuovo-Mumford regularity, multiplicity2020 Mathematics Subject Classification:
05E40, 13C15, 13D02, 13F65, 13H15, 16S371. Introduction
In recent decades, there has been a growing interest in the investigation of algebraic invariants associated to combinatorial structures. Toric ideals of graphs (and the associated edge rings), a special case of the classical notion of a toric ideal, have been studied by various authors with regard to invariants such as depth, dimension, projective dimension, regularity, graded Betti numbers, Hilbert series, and multiplicity, usually for particular families of graphs (see for example [2, 3, 5, 7, 8, 9, 10, 12, 14, 16, 17, 18, 19, 22, 23, 26]). We note in Remarks 2.6 and 2.14 that the family we consider does not overlap at all or for large with those considered in [5], [8], [9], and [23]; it is more obviously distinct from other families that have been studied. We think it fitting to mention that the recent book by Herzog, Hibi, and Ohsugi ([15]) also investigates toric ideals of graphs as well as binomial ideals coming from other combinatorial structures.
In this work, we consider a family of graphs with iterated subfamilies and develop algebraic properties of the toric rings associated to the family which depend only on the number of vertices (equivalently, the number of edges) in the associated graphs. In the development of this project, we were particularly inspired by the work of Jennifer Biermann, Augustine O’Keefe, and Adam Van Tuyl in [3], where they establish a lower bound for the regularity of the toric ideal of any finite simple graph and an upper bound for the regularity of the toric ideal of a chordal bipartite graph. Our goal is to construct as “simple” a family of graphs as possible that still yields interesting toric ideals. It is our hope that our process and results will lead to further generalizations of properties of toric ideals for other (perhaps broader) families of graphs, or for graphs containing or arising from such graphs.
Herein, we introduce the infinite family of chordal bipartite graphs , where determines the number of edges and vertices and determines the structure of the graph, and establish some algebraic properties of the toric rings associated to the graphs . The use of bipartite graphs makes each normal and Cohen-Macaulay by [25] and [15]; we use the latter in Section 3. Our main results prove to be independent of and depend only on .
In Section 2, we construct the family of graphs from a family of ladder-like structures so that the toric ideals of the are generalized determinantal ideals of the . The ladder-like structures associated to a subfamily , introduced in Example 2.4, are in fact two-sided ladders (for large ), so that the family of rings is a generalization of the family of ladder determinantal rings coming from . While the rings arising from come from a distributive lattice and have easily derived properties (see for example [15]), we show that the rings associated to do not naturally arise from any lattice in general, and merit closer study.
In Section 3, we establish some algebraic properties of the , particularly Krull dimension, projective dimension, multiplicity, and regularity. To do so, we prove that the determinantal generators of the defining ideal are a Gröbner basis (it follows immediately from [15] that is Koszul) and work with the initial ideal . We also develop a system of parameters that allows us to work with Artinian reductions in part of our treatment, and their Hilbert series.
Our first result establishes the Krull dimension of the toric ring , where the ring is the polynomial ring over the edges of and is the toric ideal of .
Theorem 1.1 (Theorem 3.4).
The dimension of is
As a corollary, since comes from a bipartite graph and is hence Cohen-Macaulay (Corollary 2.16), we obtain the projective dimension of .
Corollary 1.2 (Corollary 3.5).
The projective dimension of over is
We then develop a linear system of parameters for , using differences of elements on antidiagonals of the ladder-like structure .
Proposition 1.3 (Proposition 3.10).
Let . Then the image of
in is a system of parameters for .
Since is Cohen-Macaulay, the linear system of parameters above is a regular sequence (Corollary 3.12).
With the aim of obtaining the multiplicity and regularity of , we form an Artinian quotient of by the regular sequence above and call it . We note that does not denote the completion, and explain the choice of notation in Definition 3.7.
Using a convenient vector space basis for established in Lemma 3.13, we show the coefficients of the Hilbert series for .
Theorem 1.4 (Theorem 3.16).
If and , we have
In particular, when .
As a corollary, we obtain the regularity of , which is equal to the top nonzero degree of .
Corollary 1.5 (Corollary 3.18).
For ,
We include an alternate graph-theoretic proof of the result above at the end of this work. Beginning with an upper bound from [3] (or equivalently for our purposes, one from [14]) and then identifying the initial ideal with the edge ideal of a graph, we use results from [4] (allowing us to use instead of ) and then [13] for a lower bound which agrees with our upper bound.
From a recursion established in Lemma 3.15, we go on to prove a Fibonacci relationship between the lengths of the Artinian rings in Proposition 3.19, and obtain the multiplicity of as a corollary. In the following, we drop for convenience.
Corollary 1.6 (Corollary 3.21).
For , there is an equality of multiplicities
In particular,
For more background, detail, and motivation, we refer the reader to [1], but note that different notation and indexing conventions have been employed in this work. Throughout, is a field.
Acknowledgements. Macaulay2 [11] was used for computation and hypothesis formation. We would like to thank Syracuse University for its support and hospitality and Claudia Miller for her valuable input on the original project in [1] and this condensed version. We also acknowledge the partial support of an NSF grant.
2. The Family of Toric Rings
In the following, we define a family of toric rings coming from an iterative chordal bipartite family of graphs, . We show that although one subfamily of these rings comes from join-meet ideals of a (distributive) lattice and has some easily derived algebraic invariants, this is not true in general. The reader may find the definition of the toric ideal of a graph in Section 2.2, when it becomes relevant to the discussion. We recall for the reader that a chordal bipartite graph is a bipartite graph in which every cycle of length greater than or equal to six has a chord.
2.1. The Family of Graphs
Below, we define the family of chordal bipartite graphs iteratively from a family of ladder-like structures . We note that the quantities involved in the following definition follow patterns as follows:
0 | 2 | 2 |
1 | 2 | 3 |
2 | 3 | 3 |
3 | 3 | 4 |
⋮ | ⋮ | ⋮ |
Definition 2.1.
For each and each , we construct a ladder-like structure with rows and columns and with nonzero entries in the set . To do so, we use the notation for the first entries of , that is, all except the last entry. The construction is as follows, where throughout, indices of entries in are strictly increasing from left to right in each row and from top to bottom in each column. We note that does not depend on for , but does for .
-
•
For , the ladder-like structure is
-
•
For , to create (regardless of what is in ), we add another column with the entries and to the right of to obtain
-
•
For , to create , we add another row (column) with the entries below (to the right of) in the following way:
-
The entry is in the ultimate row and column, row and column .
-
The entry is in the new row (column) in a position directly below (to the right of) another nonzero entry in .
-
*
If the last entry of is , is directly beneath (to the right of) the first nonzero entry in the previous row (column).
-
*
If the last entry of is , is directly beneath (to the right of) the second nonzero entry in the previous row (column).
-
*
-
In this way, the entries in determine the choice at each stage for the placement of .
Remark 2.2.
We note a few things about this construction for (), which may be examined in the examples below:
-
•
We note that is directly beneath (to the right of) .
-
•
We note that the only entries in row (column ) of are , , and , so that the choices listed for placement of are the only cases. In particular, if and only if is directly beneath (to the right of) , and if and only if is directly beneath (to the right of) .
-
•
Finally, we note that the only entries in column (row ) of are , , and , and that the only entries in row (column ) of are and .
Example 2.3.
We have
In either of the cases above, we could go on to construct and in the following way: For , place to the right of and place to the right of either or , depending whether the last entry of is or , respectively. Then for , place below and place below either or , depending whether the last entry of is or , respectively.
Example 2.4.
In fact, when the entries of are all ones, we see that has a ladder shape (is a two-sided ladder for , shown below in the case when :
We denote the subfamily of graphs coming from by .
When the entries of are all zeros, has the following structure, shown below in the case when :
For a more varied example, we have below:
Definition 2.5.
If we associate a vertex to each row and each column and an edge to each nonzero entry of , we have a finite simple connected bipartite graph . The set of vertices corresponding to rows and the set of vertices corresponding to columns form a bipartition of the vertices of . We say a graph is in if for some and some .
Remark 2.6.
We note that by construction has no vertices of degree one, since each row and each column of has more than one nonzero entry. This ensures that for large our family is distinct from that studied in [5], since a Ferrers graph with bipartitation and with no vertices of degree one must have at least two vertices in of degree and at least two vertices in of degree , impossible for our graphs when , as the reader may verify. We also use the fact that has no vertices of degree one for an alternate proof of the regularity of at the end of this work.
Example 2.7.
When , is
We develop properties of the which allow us to show in Section 2.2 that certain minors of the are generators for the toric rings of the corresponding graphs .
Definition 2.8.
For this work, a distinguished minor of is a -minor involving only (nonzero) entries of the ladder-like structure , coming from a subarray of .
Proposition 2.9.
For each and each , the entry and the entry each appear in exactly two distinguished minors of . For , these minors are of the form
coming from the subarray
for some and
coming from the subarray
for some , and the only distinguished minor of with indices all less than is .
Proof.
The last statement is clear by Definition 2.1; we prove the remaining statements by induction on . For , we have the distinguished minors and coming from the subarrays
and
where and , so we have our base case. Now suppose the statement is true for with , and let and .
Case 1: If , then by Remark 2.2, is in the same column (row) as . By induction, we have the distinguished minor coming from the subarray
Then in fact we have a subarray of the form
so that we have the distinguished minors
with
by induction and with
Since the only entries in row (column ) of are and and since the only entries in column (row ) of are , , and by Remark 2.2, these are the only distinguished minors of containing either or , as desired.
Case 2 for is analogous and yields
and
∎
Definition 2.10.
Define the integers , for as in the statement of Proposition 2.9. We note in the remark below some properties of the .
Remark 2.11.
From the proof of Proposition 2.9, we note that , , and that for , we have the following:
For the sake of later proofs, we extend the notion of naturally to and say that , and note the following properties of the for :
-
•
We have and . Indeed, for , , and for , this is clear from the statement above.
-
•
We have . Indeed, for , , for , , and for , this follows from the statement above.
-
•
The form a non-decreasing sequence. Indeed, for , either or .
Remark 2.12.
We also note from the proof above that the following is a subarray of for all such that , which we use in the proof of the proposition below:
Proposition 2.13.
For , each graph is chordal bipartite with vertex bipartition of cardinalities
Proof.
We already know by Definition 2.5 that every graph is bipartite for , with the bipartition above coming from the rows and columns of . The cardinalities of the vertex sets follow from Definition 2.2. We prove the chordal bipartite property by induction on . It is clear for and that is chordal bipartite for , since these graphs have fewer than six vertices. Now suppose is chordal bipartite for with , and consider for . We know that the following array (or its transpose) is a subarray of by Remark 2.12, and we include for reference the corresponding subgraph of with vertices labeled by row and column.
We know the only difference between and is one vertex corresponding to row (column ) and two edges and , where corresponds to the column (row) containing and corresponds to column (row ). By Remark 2.2, , since the only entries in row (column ) are and . Then any even cycle containing must also contain and . Similarly, by the same remark, the only other edges with endpoint are and , the entries added to make , so we know that any even cycle containing and must contain either or . We see that any even cycle containing and is either a 4-cycle or has as a chord, and any even cycle containing and is either a 4-cycle or has as a chord. Thus every graph is chordal bipartite for , with the bipartition above. ∎
Remark 2.14.
We note that the previous proposition ensures that our graphs are distinct from those studied in [8], [9], and [23], which are not chordal bipartite except for the first family in [8], in which every four-cycle shares exactly one edge with every other four-cycle (also distinct from our family except for the trivial case with only one four-cycle, corresponding to ).
2.2. Toric Rings for
In this section, we develop the toric ring for each of the chordal bipartite graphs in the family . We first show that the toric ideal of the graph is the same as the ideal generated by the distinguished minors of . We then demonstrate that for some and , these ideals do not arise from the join-meet ideals of lattices in a natural way, so that results in lattice theory do not apply to the general family in an obvious way.
We first define the toric ideal of a graph. For any graph with vertex set and edge set , there is a natural map taking an edge to the product of its endpoints. The polynomial subring in generated by the images of the edges under the map is denoted and is called the edge ring of . The kernel of is denoted and is called the toric ideal of ; the ring is isomorphic to . In this work, we consider the toric ring and the toric ideal for our particular graphs. It is known in general that is generated by binomial expressions coming from even closed walks in [27, Prop 3.1] and that the toric ideal of a chordal bipartite graph is generated by quadratic binomials coming from the -cycles of (see [20, Th 1.2]).
Let be the polynomial ring in the edges of . The edge ring for is denoted by and is isomorphic to the toric ring
where is the toric ideal of [15, 5.3]. For the general construction of a toric ideal, we refer the reader to [24, Ch 4]. Our goal is to show that the toric ideal of is equal to
Proposition 2.15.
Let . For , we have
where
Proof.
Corollary 2.16.
The rings are normal Cohen-Macaulay rings.
Proof.
Because we know the distinguished minors of , we are now able to characterize the generators for the toric ideal of .
Remark 2.17.
By Proposition 2.9, the generators for may be summarized as follows. For integers such that , set
where the nonnegative integers are as in Remark 2.11, that is, , , and for , we have
We note that the number of generators depends on and that the depend on , but we may ignore dependence on when working with general . We sometimes call the standard generators of , and show in Section 2.3 that for certain and , they are not equal to the usual generators for the join-meet ideal of any lattice .
Example 2.18.
We consider the toric ideal of a graph in . For and , by Remark 2.11 we have , , and for , so that
where is generated by the distinguished minors of :
2.3. Distinction From Join-Meet Ideals of Lattices
We saw in Example 2.4 and Proposition 2.15 that if , then is a ladder determinantal ideal for . It is known that a ladder determinantal ideal is equal to the join-meet ideal of a (distributive) lattice (indeed, with a natural partial ordering which decreases along rows and columns of we obtain such a lattice). Some algebraic information such as regularity and projective dimension may be easily derived for some join-meet ideals of distributive lattices (see, for example, Chapter 6 of [15]). We spend some time in this section establishing that not all rings arise from a lattice in a natural way (see Remark 2.20), and so there does not seem to be any obvious way to obtain our results in Section 3 from the literature on join-meet ideals of distributive lattices or on ladder determinantal ideals. The results in Section 3 may be viewed as an extension of what may already be derived for the family from the existing literature.
The following five lemmas serve to provide machinery to show that there is at least one ring in the family , namely , whose toric ideal does not come from a lattice on the set in any obvious way. That is, we show that the standard generators of , the from Remark 2.17, are not equal to the standard generators (see Definition 2.19) for any lattice on .
Before we begin, we introduce some definitions and notation that we use extensively throughout.
Definition 2.19.
The join-meet ideal of a lattice is defined from the join (least upper bound) and meet (greatest lower bound) of each pair of incomparable elements . In this work, a standard generator of the join-meet ideal of a lattice is a nonzero element of one of the following four forms:
for . We sometimes refer to such an element as a standard generator of (the join-meet ideal is defined by analogous generators in the literature, though sometimes instead of and ). We note that for a standard generator, the pair is an incomparable pair, and the pair is a comparable pair.
Though we are in a commutative ring, we provide all possible orderings for factors within the terms of a standard generator to emphasize that either factor of the monomial
may be the join or the meet of and .
Remark 2.20.
We give an explanation for why it makes sense to focus only on the standard generators of a join-meet ideal. We recall that the standard generators for from Remark 2.17 come from distinct arrays within the ladder-like structure and recognize that either monomial of determines its array. Then an element of the form in with must be equal to for some , since a nontrivial sum of with coefficients in either has more than two terms or is equal to for some , and other coefficients would be extraneous. Then any generating set for where each element has the form in with must consist of all the (up to sign). We conclude that it is natural to check whether the are standard generators of a lattice , instead of non-standard generators.
Definition 2.21.
Given a standard generator of a lattice , where , let be defined as follows:
-
•
If , the elements in the first (positive) monomial of are not comparable in (so the elements in the second (negative) monomial of are comparable in ).
-
•
If , the elements in the negative monomial of are not comparable in (so the elements in the positive monomial of are comparable in ).
For a given list of standard generators of a lattice , we use
where , to encode the comparability of the variables in these generators.
We note that exactly one of or happens for each ; we are merely encoding which monomial in corresponds to , and which to .
Notation 2.22.
We use the notation if and in a lattice , and if and in .
In the first lemma, we begin by showing what restrictions we must have on a lattice whose join-meet ideal contains the 2-minors of the following array as standard generators:
Lemma 2.23.
Suppose
are standard generators of a lattice . Let be defined for these three elements as in Definition 2.21. Then up to relabeling of variables,
Proof.
We first note that some of the cases we consider are equivalent. If we relabel variables according to the permutation , we see that
This limits the cases we need to consider to
That is, we only need to show that the case is impossible.
Let . Then without loss of generality, up to reversing the order in the lattice (which does not affect the join-meet ideal), we have . If , we have , so and hence . We conclude that the case is impossible. ∎
In the second lemma, we show what restrictions we must have on a lattice whose join-meet ideal contains the 2-minors of the ladder
as standard generators, and which meets certain comparability conditions.
Lemma 2.24.
Suppose
are standard generators of a lattice , and that ,,, and are comparable pairs in . Let be defined for these five elements as in 2.21. Then up to relabeling of variables, .
Proof.
We first note that with natural relabeling, both and satisfy the hypotheses of Lemma 2.23, so if we let be defined as in Definition 2.21, this limits the cases we need to consider to 5-tuples whose first three elements and whose last three elements satisfy the conclusion of Lemma 2.23. We note that some of the cases we consider are equivalent. If we relabel variables according to the permutation , we see that , and if we relabel the variables according to the permutation , we have . The permutation yields . Then by Lemma 2.23 we have the eighteen cases
Case 1: . Since , without loss of generality (reversing the order on the entire lattice if needed) we have . Then and , with the ordering chosen, yield
If , then , but then is not a standard generator of , and this is a contradiction. If , then so that both and from are comparable pairs, but this is a contradiction. We conclude that the case is impossible.
Because of the comparability of the pair , the case forces comparability of or and hence yields a contradiction, as the reader may verify. In the case , the comparability of forces either the comparability of (a contradiction), or , up to reversing the order in the lattice, since is a standard generator. Since is a comparable pair, it immediately follows that either is comparable (a contradiction) or , which is a contradiction since is a standard generator, as the reader may verify. These cases are compatible with the given relabelings and thus conclude our proof. ∎
In the third lemma, we show what restrictions we must have on a lattice whose join-meet ideal contains the 2-minors of the ladder
as standard generators and which meets certain comparability conditions.
Lemma 2.25.
Suppose
are standard generators of a lattice , and that , , , , , , , and are comparable pairs in . Let be defined for these seven elements as in Definition 2.21. Then up to relabeling of variables, .
Proof.
We first note that with natural relabeling, both and satisfy the hypotheses of Lemma 2.24, so if we let be defined as in Definition 2.21, this limits the cases we need to consider to 7-tuples whose first five elements and whose last five elements satisfy the conclusion of Lemma 2.24. The only possible cases are and . If we relabel variables according to the permutation , we see that , so that these two cases are equivalent. Then up to relabeling of variables, . ∎
In the fourth lemma, we show what restrictions we must have on a lattice whose join-meet ideal contains the 2-minors of the ladder
as standard generators, and which meets certain comparability conditions.
Lemma 2.26.
Suppose
are standard generators of a lattice , and that , , , , , , , , , , , and are comparable pairs in . Let be defined for these nine elements as in Definition 2.21. Then .
Proof.
We first note that with natural relabeling, both and
satisfy the hypotheses of Lemma 2.25, so if we let be defined as in Definition 2.21, this limits the cases we need to consider to 9-tuples whose first seven entries and whose last seven entries satisfy the conclusion of Lemma 2.25. We see by Lemma 2.25 that .
∎
We now have the machinery necessary to show that for , does not come from a lattice. In our proof, we use the previous four lemmas and the fact that is generated by the distinguished minors of :
Proposition 2.27.
Let and . Then the set of standard generators for is not equal to the complete set of standard generators (up to sign) of any (classical) lattice.
Proof.
By Remark 2.17 and choice of and , the generators of are
Suppose a lattice exists whose complete set of standard generators (up to sign) equals {}. We note that if the monomial does not appear in any of the , then is a comparable pair, since otherwise would be in the set of standard generators of . Thus the pairs , , , , , , , , , , , , , and are comparable pairs in . Let be defined as in Definition 2.21. Then with natural relabeling of the first nine relations, this lattice satisfies the hypotheses of Lemma 2.26, so the only cases we need to consider are .
Since , without loss of generality, we have . Then with the ordering chosen, the fact that yields
The reader may verify that and both yield contradictions based on inspecting the standard generator in light of the comparability of the pairs and , respectively, using the same technique employed in Case 1 of the proof of Lemma 2.24.
We conclude that there is no lattice whose complete set of standard generators (up to sign) equals the set of standard generators of . ∎
3. Properties of the Family of Toric Rings
In Section 2, we defined a family of toric rings, the rings coming from the family , and we demonstrated some context for these rings in the area of graph theory. Now we investigate some of the algebraic properties of each . We develop proofs to establish dimension, regularity, and multiplicity.
3.1. Dimension and System of Parameters
We use the the degree reverse lexicographic monomial order with throughout this section, and denote it by . We show that the standard generators given in Remark 2.17 are a Gröbner basis for with respect to .
Lemma 3.1.
If are as in Remark 2.17, then is a Gröbner basis for with respect to .
Proof.
This is a straightforward computation using Buchberger’s Criterion and properties of the from Remark 2.11. By Remark 2.17, for the ideal is generated by
where , , and for , we have
If we adopt -polynomial notation for the -polynomial of and , then the cases to consider are
To give a flavor of the computation involved, we show the case for , and leave the remaining cases to the reader. We show in each subcase that is equal to a sum of basis elements with coefficients in , so that the reduced form of is zero in each subcase. We have
Case 1: If and , then and , so we have
Case 1.1: If in addition and , then and , so we have
Case 1.2: If in addition or and , then and , so we have
Case 2: If or if and , then and , so we have
This concludes the case for . The remaining cases are similar.
∎
Corollary 3.2.
The ring is Koszul for all and all .
Proof.
Since has a quadratic Gröbner basis, the ring is Koszul for all and all due to [15, Th 2.28]. ∎
Corollary 3.3.
The initial ideal for with respect to the degree reverse lexicographic monomial order is
We note that does not depend on , which will be useful for the following sections.
We use the initial ideal from Corollary 3.3 and direct computation to show the Krull dimension of . As a corollary, we obtain the projective dimension of . We note that the Krull dimension, like the initial ideal, does not depend on . We refer the reader to Remark 2.17 for a reminder of how to think of the toric ring
in the context of this work.
Theorem 3.4.
The Krull dimension of is
Proof.
Let be the degree reverse lexicographic monomial order with
By Corollary 3.3, the initial ideal of with respect to is
Since and are known to have the same Krull dimension (see for example [6, Props 9.3.4 and 9.3.12]), it suffices to prove that
To see that the dimension is at least , we construct a chain of prime ideals in containing . Since every monomial generator of contains a variable of odd index, we begin with , a prime ideal containing . Then we have the chain of prime ideals , so that
To see that the dimension is at most , we find a sequence of elements in such that the quotient by the ideal they generate has dimension zero. Let
in , and take the quotient of by the image of to obtain the following. In the last step, we rewrite the quotient of and by by setting and equal to zero and replacing with for :
Since
the above ring has dimension zero. Thus,
We conclude that . ∎
Corollary 3.5.
The projective dimension of over is
Proof.
We know the Krull dimension of the polynomial ring is . The result follows from the fact that is Cohen-Macaulay (Corollary 2.16) and from the graded version of the Auslander-Buchsbaum formula. ∎
Remark 3.6.
The proof of the previous theorem shows that the image of
in is a system of parameters for . We prove in the next theorem that the image of in (which we call ) is also a system of parameters for . Before doing so, we introduce some notation and a definition which allows us to better grapple with the quotient ring .
Definition 3.7.
Consider the isomorphism
We view taking the quotient by as setting and equal to zero and replacing with for to obtain
By the same process (detailed below), we obtain the ideal . We further define the quotient
We find this notation natural since it is often used for the removal of variables, and the quotient by may be viewed as identifying and removing variables. Since this work has no completions in it, there should be no conflict of notation.
Definition 3.8.
To define in particular, we recall the standard generators of and introduce further notation to describe the generators of . By Remark 2.17, the standard generators of are
for , where the nonnegative integers are as in Remark 2.11.
Let be the largest index such that . By Remark 2.11, we see that the are defined recursively and form a non-decreasing sequence. Then
and since we view taking the quotient by as setting and equal to zero and replacing with for , we define by replacing with (defined below) for to obtain
for , where
We note that for each , since by Remark 2.11. By properties of the original from Remark 2.11, we know that , , and for ,
Example 3.9.
We construct the ring . For the graph , we have the toric ring
coming from the ladder-like structure
from Example 2.3. We know
so that is isomorphic to
Now we show that is also a system of parameters for , and not just for the quotient by the initial ideal.
Proposition 3.10.
Let and let
so that the image of in is the system of parameters from Remark 3.6. Then the image of in is a system of parameters for .
Proof.
Let be defined as above. Then by Theorem 3.4 and Definition 3.7 we need only show that . We have for that
for
and for
where
for from Definition 3.8. We know . We claim that
This is clear for . For , we prove this by induction. Since and are in , we have . Since
and , we get , so that . Now suppose for . We have
But by Definition 3.8 and by induction, so that , and hence . We conclude that
so we have equality. Since has dimension zero, so does . Thus, the image of is a system of parameters for . ∎
Remark 3.11.
We note that as a consequence of the proof of the preceding theorem, the ring is Artinian, which will be relevant in Section 3.2.
Corollary 3.12.
The image of
in is a regular sequence for .
3.2. Length, Multiplicity, and Regularity
In this section, we determine the multiplicity and Castelnuovo-Mumford regularity of the toric rings coming from the associated graphs by computing the length of the Artinian rings
from Definition 3.7. We know by Corollary 3.12 that is a linear regular sequence for , which allows us to compute the multiplicity of the rings . As a corollary of Theorem 3.16, which establishes the Hilbert function for , we obtain the multiplicity and regularity of . We also develop an alternate graph-theoretic proof for the regularity of , which is included at the end of this section.
We begin with a lemma establishing a vector space basis for , which we use extensively for our results.
Lemma 3.13.
The image of all squarefree monomials with only odd indices whose indices are at least four apart, together with the image of , forms a vector space basis for over .
Proof.
We recall for the reader the definition of and then find the initial ideal of and use Macaulay’s Basis Theorem to show that the desired representatives form a basis for as a vector space over .
From Definition 3.7, we have
where
By Definition 3.8, for the ideal is generated by
where , , and for ,
We first show that the image of the monomials with the desired property is a basis in the quotient of by the initial ideal . By Macaulay’s Basis Theorem, which is Theorem 1.5.7 in [21], the image of these monomials in is also a basis.
To find the initial ideal of , we establish that the given generators are a Gröbner basis for with respect to the degree reverse lexicographic order . This is a relatively straightforward computation using Buchberger’s Criterion, with separate cases for when .
If we adopt -polynomial notation for the -polynomial of and , then the cases to consider are
To give a flavor of the computation involved, we show the case for , and leave the remaining cases to the reader. We show in each subcase that is equal to zero or to a sum of basis elements with coefficients in , so that the reduced form of is zero in each subcase. We have
Case 1: If , then and , so we have
We note that if , then .
Case 2: If , then and , so we have
Case 2.1: If in addition , then and , so we have
We note that if , then .
Case 2.2: If in addition , then and , so we have
This concludes the case for . The remaining cases are similar.
Then the given generators are a Gröbner Basis for , so that the initial ideal is
in the ring . Since consists precisely of all squares of variables in and all degree two products of variables whose indices differ by exactly two, it follows that the image of the squarefree monomials whose indices are at least four apart, together with the image of , forms a basis for . By Macaulay’s Basis Theorem, the image of these monomials in is also a basis.
∎
We use the lemma above to establish facts about the vector space dimensions of degree pieces of , which are applied further below to establish length and multiplicity.
Notation 3.14.
Throughout this section, we use for the vector space dimension of the degree piece of , that is, for the th coefficient in the Hilbert series of . By Lemma 3.13, these are independent of .
We establish a recursive relationship between these dimensions by introducing a short exact sequence of vector spaces.
Lemma 3.15.
For and , the vector space dimension satisfies the recursive relationship
Proof.
We use the vector space basis defined in Lemma 3.13. We note that the basis elements described are actually monomial representatives (which do not depend on ) of equivalence classes (which do depend on ), but we suppress this and speak as if they are monomials, not depending on . We then take the liberty of suppressing in what follows, for convenience. We recall for the reader that
Let be multiplication by , and let
be defined for a basis element by
We note that these vector space maps are well-defined, since or a squarefree monomial with odd indices at least four apart has an output of 0, , or a monomial with the same properties. The following sequence of vector spaces is exact
so that
∎
Applying Lemma 3.15 and induction, we achieve the following closed formula for the coefficients of the Hilbert series of .
Theorem 3.16.
If and , we have
In particular, when .
Proof.
We begin with the proof of the last statement, and note that throughout the following proof we use Notation 3.14. When is even, , and , we have a factor of zero, and when is odd, , and , we have a factor of zero. Thus when .
Now we establish the base cases , then proceed by induction. It is clear that , generated by . By Lemma 3.13 and by the fact that is a graded quotient, every nonzero element of positive degree can be represented uniquely as a sum of degree squarefree monomials with odd indices whose indices are at least four apart. Then is generated by the images of all the odd variables
in , so that
matches the given formula.
Now we establish the base cases and for all . We recognize that the first monomial of degree two with odd indices at least four apart is , which does not exist until , so we have
and
which match the given formula.
This gives us the following table of base cases for , which match the given formula:
0 | 1 | 2 | 3 | 4 | 5 | 6 | 7 | ||
0 | 1 | 1 | 0 | 0 | 0 | 0 | 0 | 0 | |
1 | 1 | 2 | 0 | 0 | 0 | 0 | 0 | 0 | |
2 | 1 | 3 | |||||||
3 | 1 | 4 | |||||||
4 | 1 | 5 | |||||||
We recall by Lemma 3.15 that we have the recursive relationship
for and . We proceed by induction. Suppose and that the dimension formula holds for all when . By our recursion and by induction, we have
as desired. ∎
Remark 3.17.
We note from the proof of the theorem above a few facts for future reference. By our base cases, we have and . Taking the Fibonacci sequence with and , we have , , and , so that
These facts become useful in Proposition 3.19.
We see in the following corollary that the regularity of is . For an alternate proof of the regularity of which uses different machinery and more graph-theoretic properties, see the end of this section.
Corollary 3.18.
For ,
Proof.
We show that is equal to the top nonzero degree of and that this value agrees with the above. Since is Artinian by Remark 3.11, it is clear that is the top nonzero degree of . By Theorem 3.16, we know the top nonzero degree is for some , so that . In fact, the top nonzero degree is , provided . The only way we have a factor of zero in
is if , which means . This can only happen when , but , so we conclude that . Since is an Artinian quotient of by a linear regular sequence, we conclude that
∎
In the following, we first compute the lengths of the dimension zero rings , and then show a closed form for the multiplicity of our original rings by using a Fibonacci relationship between the lengths of the rings and applying Binet’s formula for , the th number in the Fibonacci sequence:
In the theorem and corollaries which follow, we suppress for convenience, since the statements are independent of .
Proposition 3.19.
The lengths of the rings satisfy the recursive formula
for . Consequently, if is the Fibonacci sequence, with and , then
Proof.
Again, we use Notation 3.14. By the recursive relationship from Lemma 3.15, since in general, and since in general for by Theorem 3.16, we have for that
Now we show the second statement. For our base cases, we see from Remark 3.17 that and that .
Now suppose that
and .
Then
we have
as desired. The closed form for follows directly from Binet’s formula for the Fibonacci sequence. ∎
Corollary 3.20.
For , there is an equality of multiplicities
Proof.
We have established the length of the Artinian rings , and hence the multiplicity . ∎
Corollary 3.21.
For , there is an equality of multiplicities
In particular,
Proof.
To obtain the multiplicity of , we look at , which by Remark 3.11 and Corollary 3.12 is the Artinian quotient of by a linear regular sequence. By a standard result, we may calculate length along the obvious short exact sequences coming from multiplication by elements of our regular sequence to obtain the equality
where is the Krull dimension of . Defining multiplicity as in and preceding [24, Thm 16.7], it follows immediately that
We are done by Proposition 3.19. ∎
We reintroduce and spend the remainder of this section providing an alternate
graph-theoretic proof for the regularity of .
Alternate proof of Corollary 3.18.
We show that by proving that . We first show that
We recall by Proposition 2.13 that the graph is chordal bipartite with vertex bipartition of cardinalities
and that does not have any vertices of degree one by Remark 2.6. Then by Theorem 4.9 of [3], we have
We note that we may equivalently prove by choosing the edges whose indices are equivalent to zero modulo , one from each row of , to obtain an edge matching (different from an induced matching, below) and then applying [14, Th 1].
We now show that . Since is homogeneous and consists of squarefree monomials by Corollary 3.3, we have by Corollary 2.7 of [4] that , so it suffices to prove that . The ideal can be viewed as the edge ideal of a simple graph, a “comb” with tines, with consecutive odd variables corresponding to vertices along the spine, as pictured below:
We know from Theorem 6.5 of [13] that the regularity of an edge ideal is bounded below by the number of edges in any induced matching plus one, so we choose edges (tines) corresponding to certain odd variables that create an induced matching. By beginning with the -tine and choosing every other tine corresponding to the variables
we obtain edges that are an induced matching, so we have
as desired.
We conclude that , and hence that ∎
References
- [1] L. Ballard. Properties of the Toric Rings of a Chordal Bipartite Family of Graphs. PhD thesis, Syracuse University, 2020.
- [2] S. K. Beyarslan, H. T. Hà, and A. O’Keefe. Algebraic properties of toric rings of graphs. Communications in Algebra, 47(1):1–16, 2019.
- [3] J. Biermann, A. O’Keefe, and A. Van Tuyl. Bounds on the regularity of toric ideals of graphs. Advances in Applied Mathematics, 85:84–102, 2017.
- [4] A. Conca and M. Varbaro. Square-free Gröbner degenerations. Inventiones Mathematicae, 221(3):713–730, 2020.
- [5] A. Corso and U. Nagel. Monomial and toric ideals associated to Ferrers graphs. Transactions of the American Mathematical Society, 361(3):1371–1395, 2009.
- [6] D. Cox, J. Little, and D. O’Shea. Ideals, varieties, and algorithms. Undergraduate Texts in Mathematics. Springer, New York, third edition, 2007. An introduction to computational algebraic geometry and commutative algebra.
- [7] A. D’Alì. Toric ideals associated with gap-free graphs. Journal of Pure and Applied Algebra, 219(9):3862–3872, 2015.
- [8] G. Favacchio, G. Keiper, and A. Van Tuyl. Regularity and h-polynomials of toric ideals of graphs. Proceedings of the American Mathematical Society, 148:4665–4677, 2020.
- [9] F. Galetto, J. Hofscheier, G. Keiper, C. Kohne, M. E. Uribe Paczka, and A. Van Tuyl. Betti numbers of toric ideals of graphs: a case study. Journal of Algebra and its Applications, 18(12):1950226, 2019.
- [10] I. Gitler and C. E. Valencia. Multiplicities of edge subrings. Discrete Mathematics, 302(1):107–123, 2005.
- [11] D. R. Grayson and M. E. Stillman. Macaulay2, a software system for research in algebraic geometry. Available at http://www.math.uiuc.edu/Macaulay2/.
- [12] Z. Greif and J. McCullough. Green-Lazarsfeld condition for toric edge ideals of bipartite graphs. Journal of Algebra, 562:1–27, 2020.
- [13] H. T. Hà and A. Van Tuyl. Monomial ideals, edge ideals of hypergraphs, and their graded Betti numbers. Journal of Algebraic Combinatorics, 27(2):215–245, 2008.
- [14] J. Herzog and T. Hibi. The regularity of edge rings and matching numbers. Mathematics, 8(1):39, 2020.
- [15] J. Herzog, T. Hibi, and H. Ohsugi. Binomial ideals, volume 279 of Graduate Texts in Mathematics. Springer, Cham, 2018.
- [16] T. Hibi, A. Higashitani, K. Kimura, and A. O’Keefe. Depth of edge rings arising from finite graphs. Proceedings of the American Mathematical Society, 139(11):3807–3813, 2011.
- [17] T. Hibi and L. Katthän. Edge rings satisfying Serre’s condition (). Proceedings of the American Mathematical Society, 142(7):2537–2541, 2014.
- [18] T. Hibi, K. Matsuda, and H. Ohsugi. Strongly Koszul edge rings. Acta Mathematica Vietnamica, 41(1):69–76, 2016.
- [19] T. Hibi, K. Matsuda, and A. Tsuchiya. Edge rings with 3-linear resolutions. Proceedings of the American Mathematical Society, 147(8):3225–3232, 2019.
- [20] T. Hibi and H. Ohsugi. Toric ideals generated by quadratic binomials. Journal of Algebra, 218(2):509–527, 1999.
- [21] M. Kreuzer and L. Robbiano. Computational Commutative Algebra 1. Springer, Berlin, Heidelberg, 2000.
- [22] K. Mori, H. Ohsugi, and A. Tsuchiya. Edge rings with -linear resolutions. Preprint, arxiv.org/abs/2010.02854.
- [23] R. Nandi and R. Nanduri. On Betti numbers of toric algebras of certain bipartite graphs. Journal of Algebra and Its Applications, 18(12):1950231, 2019.
- [24] I. Peeva. Graded Syzygies. Springer London, London, 2011.
- [25] A. Simis, W. V. Vasconcelos, and R. H. Villarreal. On the ideal theory of graphs. Journal of Algebra, 167(2):389–416, 1994.
- [26] C. Tatakis and A. Thoma. On the universal Gröbner bases of toric ideals of graphs. Journal of Combinatorial Theory, Series A, 118(5):1540–1548, 2011.
- [27] R. H. Villarreal. Rees algebras of edge ideals. Communications in Algebra, 23(9):3513–3524, 1995.