Hostname: page-component-68c7f8b79f-r4j94 Total loading time: 0 Render date: 2026-01-16T13:05:04.483Z Has data issue: false hasContentIssue false

Koszul binomial edge ideals

Published online by Cambridge University Press:  09 January 2026

Adam LaClair
Affiliation:
University of Nebraska-Lincoln , United States; E-mail: alaclair2@unl.edu.
Matthew Mastroeni
Affiliation:
SUNY Polytechnic Institute , United States; E-mail: mastromn@sunypoly.edu.
Jason McCullough*
Affiliation:
Iowa State University , United States
Irena Peeva
Affiliation:
Cornell University , United States; E-mail: ivp1@cornell.edu.
*
E-mail: jmccullo@iastate.edu (Corresponding author).

Abstract

The study of Koszul binomial edge ideals was initiated by V. Ene, J. Herzog, and T. Hibi in 2014, who found necessary conditions for Koszulness. The binomial edge ideal $J_G$ associated to a finite simple graph G is always generated by quadrics. It has a quadratic Gröbner basis if and only if the graph G is closed. However, there are many known nonclosed graphs G where $J_G$ is Koszul. We characterize the Koszul binomial edge ideals by a simple combinatorial property of the graph G.

Information

Type
Algebra
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2026. Published by Cambridge University Press

1 Introduction

Throughout the paper, $\Bbbk $ stands for a field. Consider a polynomial ring $S=\Bbbk [x_1,\dots ,x_n]$ standard graded by $\deg (x_i)=1$ for each i, and let I be a graded ideal. The graded $\Bbbk $ -algebra $Q=S/I$ is Koszul if $\Bbbk $ has a linear free resolution over Q, that is, the entries in the matrices of the differential in the resolution are linear forms; in this case, we also say that the ideal I is Koszul . Koszul algebras were formally defined by Priddy [Reference Priddy37] (who considered a more general situation which includes the noncommutative case as well). They appear in many areas of algebra, geometry, and topology. Two of their exceptionally nice properties are that the minimal free resolution of $\Bbbk $ can be explicitly described as the generalized Koszul complex (see [Reference McCullough and Peeva28, 8.12]), and there is an elegant formula relating the Hilbert function of Q and the Poincaré series of $\Bbbk $ (which is the generating function of the minimal free resolution of $\Bbbk $ over Q), see [Reference McCullough and Peeva28, 8.5]. Another important property of commutative Koszul algebras is that they are the algebras over which every finitely generated module has finite regularity [Reference Avramov and Eisenbud1, Reference Avramov and Peeva2]. The papers [Reference Conca14], [Reference Conca, De Negri and Evelina Rossi12], [Reference Fröberg23], and [Reference Polishchuk and Positselski36] provide excellent overviews of Koszul algebras. It is well known that if I is Koszul then it is generated by quadrics; in this case, we say that Q and I are quadratic .

The Koszul property has been studied for quadratic graded $\Bbbk $ -algebras coming from graphs. Let G be a graph on vertex set $[n] = \{1, 2, \dots , n\}$ . Throughout this paper, all graphs are assumed to be simple, that is, with no loops and no double edges. There are various constructions that produce a quadratic graded $\Bbbk $ -algebra starting from G; we discuss some of them: The monomial edge ideal $N_G=\big (x_ix_j\,|\, \{i,j\} \ \text { is an edge in } G\,\big )$ is generated by quadratic monomials, and it is always Koszul. One may consider the toric edge ring $A_G = \Bbbk [x_ix_j\,|\, \{i,j\} \ \text { is an edge in } G\,]$ , for which Hibi and Ohsugi provide a characterization when $A_G$ is quadratic. Furthermore, if the graph G is bipartite, they prove in [Reference Ohsugi and Hibi32] that $A_G$ is Koszul if and only if it is quadratic, if and only if there is a quadratic Gröbner basis.

We focus on the concept of binomial edge ideal associated with a graph G, which generalizes the classical concept of determinantal ideal generated by the $2$ -minors of a $(2\times n)$ -matrix of indeterminates.

Definition 1.1. The binomial edge ideal $J_G$ is the ideal

$$\begin{align*}J_G=\Big(\,x_iy_j-y_ix_j\,\Big|\, \,\{i,j\}\ \text{ is an edge in } G\,\Big)\, \end{align*}$$

in the polynomial ring

$$\begin{align*}S_G =\Bbbk[x_1,\dots ,x_n,y_1,\dots ,y_n]\,. \end{align*}$$

We set $R_G = S_G/J_G$ . We will use this notation throughout the paper.

Binomial edge ideals were first introduced in [Reference Herzog, Hibi, Hreinsdóttir, Kahle and Rauh25] and independently in [Reference Ohtani30] as a generalization of the ideals of minors that appear in algebraic statistics; see [Reference Herzog, Hibi and Ohsugi26, Section 4]. Binomial edge ideals also generalize ladder determinantal ideals and ideals of adjacent minors studied in [Reference Conca13]. They have been studied extensively, and the book [Reference Herzog, Hibi and Ohsugi26] provides an overview of their properties.

It turns out that the Koszul property of $J_G$ is related to three special graphs, defined in Figure 1: the tent, the claw, and the net.

Figure 1 The tent (left), claw (center), and net (right) graphs.

The study of the Koszul property of binomial edge ideals was launched by Ene, Herzog, and Hibi in 2014 in [Reference Ene, Herzog and Hibi21]. They showed that if the binomial edge ideal $J_G$ is Koszul, then the graph G is chordal [Reference Ene, Herzog and Hibi21, Reference Herzog, Hibi and Ohsugi26], that is, G does not contain an induced cycle of length greater than three. Furthermore, they found that the binomial edge ideal of the claw and the binomial edge ideal of the tent are not Koszul. The ideal $J_G$ is multigraded, and this implies that if $J_G$ is Koszul, then $J_F$ is Koszul for any induced subgraph F of G, by [Reference Herzog, Hibi and Ohsugi26, 7.46]. So, if $J_G$ is Koszul, then G is chordal, claw-free, and tent-free. The rather sparse known results on Koszulness of binomial edge ideals are surveyed in the book [Reference Herzog, Hibi and Ohsugi26].

The main result in our paper is that, in Section 2, we prove the following characterization of Koszulness:

Theorem 1.2. The binomial edge ideal $J_G$ is Koszul if and only if the graph G is chordal, claw-free, and tent-free.

We also show in Corollary 2.4 that the above conditions are equivalent to G being strongly chordal and claw-free.

The usual technique for establishing that an ideal is Koszul is to show that it has a quadratic Gröbner basis. Proving Koszulness in the absence of a quadratic Gröbner basis is very challenging, and there are very few such results known. For example, Caviglia and Conca study the Koszul property of projections of the Veronese cubic surface in [Reference Caviglia and Conca11], and Hibi and Ohsugi provide such toric rings in [Reference Ohsugi and Hibi33]. It is known that the binomial edge ideal $J_G$ has a quadratic Gröbner basis if and only if the graph G is closed by [Reference Herzog, Hibi, Hreinsdóttir, Kahle and Rauh25, 1.1] and [Reference Crupi and Rinaldo16, 3.4]. In this case, $J_G$ has the very rare property that its graded Betti numbers coincide with those of its lex initial ideal [Reference Peeva35]. By [Reference Herzog, Hibi and Ohsugi26, 7.10], G is closed if and only if it is chordal, claw-free, tent-free, and net-free. Thus, any Koszul binomial edge ideal for which G contains a net as an induced subgraph does not have a quadratic Gröbner basis. There are many such examples, starting with the net itself. This is in stark contrast with the binomial edge ideals of a pair of graphs, where the two notions are equivalent by [Reference Baskoroputro, Ene and Ion3].

We prove the “if” implication of Theorem 2 in Section 2. The key idea in our proof is that such graphs are strongly chordal and claw-free, which yields orders on both the vertices and edges with special properties. Our proof is an inductive argument based on the properties of this class of graphs. Although the “only if” implication was already established by Ene, Herzog, and Hibi [Reference Ene, Herzog and Hibi21], [Reference Herzog, Hibi and Ohsugi26, Section 7.4.1], we give a self-contained, characteristic-free proof that does not rely on computer algebra computations in Section 4. In Section 5, we characterize strongly chordal, claw-free graphs with small clique number and show that this characterization does not generalize to graphs with higher clique numbers.

2 Koszul binomial edge ideals

The aim of this section is to prove the sufficient condition in our main result Theorem 1.2:

Theorem 2.1. If a simple graph G is chordal, claw-free, and tent-free, then the binomial edge ideal $J_G$ is Koszul.

We first develop the graph machinery necessary. We refer the reader to [Reference West39] and [Reference Brandstädt, Bang Le and Spinrad9] for more on graph theory. Throughout, we assume that G is a finite simple graph. Let $G = (V(G),E(G))$ have a set of vertices $V(G)$ , and a set of edges $E(G)$ . For simplicity, we may write $\{i,j\}\in G$ meaning $\{i,j\}\in E(G)$ , or may write $ij$ to denote the edge $\{i,j\}$ . The set of neighbors of a vertex i in G is the set $N(i)$ of all vertices j such that $\{i, j\}$ is an edge of G. The set $N[i] = N(i) \cup \{i\}$ is called the closed neighborhood of i. A subgraph H of G is induced if for any $v_i,v_j \in V(H)$ , if $v_iv_j \in E(G)$ then $v_iv_j \in E(H)$ . The complete graph contains all possible edges between distinct vertices. A cycle graph $C_n$ of length n has n vertices $v_1,\ldots ,v_n$ and edges $v_iv_{i+1}$ for $1 \le i \le n$ , where we identify $v_{n+1}$ with $v_1$ .

A chordal graph is one with no induced cycles of length at least four. There are several characterizations of chordal graphs; we highlight two characterizations involving vertex orderings and special types of vertices.

A perfect elimination order for a graph G is an ordering $v_1, \dots , v_n$ of its vertices such that whenever $v_iv_j, v_iv_k \in E(G)$ and $i < j, k$ , we have that $v_jv_k \in E(G)$ . We say that v is a simplicial vertex of G if $N[v]$ is a clique, that is, the induced subgraph of G on the set $N[v]$ is a complete graph. In that case, $N[v]$ is a maximal clique of G. If G has a perfect elimination order $v_1, \dots , v_n$ , it is easily seen that $v_i$ is a simplicial vertex of the induced subgraph $G_i = G[v_i, \dots , v_n]$ for all i.

Theorem 2.2 [Reference Dirac19].

For a graph G, the following are equivalent:

  1. (a) G is chordal.

  2. (b) G has a perfect elimination order.

  3. (c) Every induced subgraph of G has a simplicial vertex.

We will use the stronger concept of strongly chordal graphs. A strong elimination order on the set of vertices of G is a perfect elimination order $v_1, \dots , v_n$ such that whenever $v_iv_k, v_kv_j, v_iv_\ell \in E(G)$ with $i < k < \ell $ and $i < j$ , then we also have $v_jv_\ell \in E(G)$ . The graph G is called strongly chordal if it has a strong elimination order.

Similar to the case of chordal graphs, strongly chordal graphs can also be characterized by special types of vertices. A vertex v of G is called simple if the collection of distinct sets in $\{N[w] \mid w \in N[v]\}$ is totally ordered by inclusion. In particular, every simple vertex is also simplicial. If $v_1, \dots , v_n$ is a strong elimination order for G, it is easily seen that the vertex $v_i$ is a simple vertex of the induced subgraph $G_i := G[v_i, \dots , v_n]$ for all i.

Strongly chordal graphs can also be characterized by forbidden induced subgraphs. A graph T is an n -trampoline if it has vertices $v_1,\dots , v_n, w_1, \dots , w_n$ for some $n \geq 3$ and edges $v_iv_j$ for all $i \neq j$ , $w_iv_i$ for all i, $w_iv_{i+1}$ for all $i < n$ , and $w_nv_1$ . We note that the n-trampoline is sometimes referred to as the n-sun graph, and the 3-trampoline is also called the tent in the literature on binomial edge ideals. The $4$ -trampoline is pictured in Figure 2. Setting $w_{i + n} = w_i$ , we note that $w_i \in N[v_{i+1}] \smallsetminus N[v_{i+2}]$ and $w_{i+2} \in N[v_{i+2}] \smallsetminus N[v_{i+1}]$ for all i since $n \geq 3$ . Consequently, the sets $N[v_{i+1}]$ and $N[v_{i+2}]$ are incomparable for all i. Hence, no vertex of the n-trampoline is simple, and so no trampoline graph is strongly chordal.

Figure 2 The 4-trampoline.

Theorem 2.3 [Reference Farber22, 3.3, 4.1].

For a graph G, the following are equivalent:

  1. (a) G is strongly chordal.

  2. (b) Every induced subgraph of G has a simple vertex.

  3. (c) G is chordal and n-trampoline-free for every $n \geq 3$ .

Applying the above theorem to the graphs that we consider in Theorem 2, we get:

Corollary 2.4. A graph G is chordal, claw-free, and tent-free if and only if it is strongly chordal and claw-free.

Proof . Note that the 3-trampoline is the tent. Hence, if G is strongly chordal and claw-free, then G is clearly chordal, claw-free, and tent-free. Conversely, if G is chordal, claw-free, and tent-free, then G is n-trampoline-free since every n-trampoline with $n> 3$ contains an induced claw, and so, G is strongly chordal.

In the rest of this section, we assume that G is a simple graph with vertex set $[n] = \{1, 2, \dots , n\}$ .

If i is a simplicial vertex of G, then $N[i]$ is the unique maximal clique containing i. This motivates the following definition.

Definition 2.5. Let G be a chordal graph.

  • A simplicial edge of a graph G is an edge e that is contained in exactly one maximal clique of G and $G \smallsetminus e$ is chordal.

  • A simplicial edge is claw-avoiding if it does not belong to an induced claw of G.

  • A claw-avoiding perfect edge elimination order is an ordering $e_1, \dots , e_m$ of the edges of G such that $e_i$ is a claw-avoiding simplicial edge of $G \smallsetminus \{e_1, \dots , e_{i-1}\}$ for every i.

Remark 2.6. We caution the reader that there is at least one other notion of a simplicial edge in the literature. Dragan [Reference Dragan20] defines the class of strongly orderable graphs as a common generalization of both strongly chordal and chordal bipartite graphs and characterizes them in terms of edges $e = \{i, j\}$ such that any distinct vertices $k \in N(i)$ and $l \in N(j)$ are adjacent; he calls such an edge a simplicial edge. For chordal graphs, every simplicial edge in Dragan’s sense is a simplicial edge as defined above. However, our notion of simplicial edge is strictly weaker than Dragan’s since the middle edge of a path of length 3 is simplicial in our sense but not Dragan’s.

Lemma 2.7. Let G be a strongly chordal graph with at least one edge. Then G has a simplicial edge e such that $G \smallsetminus e$ is strongly chordal. Moreover, one can choose the simplicial edge e so that, if G is claw-free then so is $G \smallsetminus e$ .

Proof . We may assume that the vertices of G have been labeled so that $1, 2, \dots , n$ is a strong elimination order. Then $1$ is a simple vertex of G, and we can choose j so that $N[j]$ is maximal in the totally ordered set $\{ N[i] \mid i \in N[1]\}$ . We also note that $N[1]$ is minimal in $\{ N[i] \mid i \in N[1]\}$ . Since $1$ is in particular a simplicial vertex of G, it follows that $N[1]$ is the unique maximal clique of G containing $1$ .

We claim that the edge $e = \{1, j\}$ is simplicial. We have already shown that $N[1]$ is the unique maximal clique of G containing e, so it remains to show that $G \smallsetminus e$ is chordal. If $G \smallsetminus e$ is not chordal, then it contains an induced cycle C of length at least 4. As G is chordal, e must be the unique chord of C, and so C has length 4. Hence, G has an induced subgraph of the form

which is impossible as $i, k \in N[1]$ and 1 is a simplicial vertex. Hence, e is a simplicial edge of G as claimed.

To see that $G \smallsetminus e$ is still strongly chordal, we note that $N_{G \smallsetminus e}[1] = N_G[1] \smallsetminus \{j\}$ and that $N_{G \smallsetminus e}[i] = N_G[i]$ for all $i \neq 1, j$ , so the set

$$\begin{align*}\{N_{G \smallsetminus e}[i] \mid i \in N_{G \smallsetminus e}[1]\} = \{N_G[1] \smallsetminus \{j\}\} \cup \{N_G[i] \mid i \in N_{G}[1] \smallsetminus \{j\} \} \end{align*}$$

is still totally ordered by inclusion. Hence, $1$ is still a simple vertex of $G \smallsetminus e$ , and the graph $(G \smallsetminus e) \smallsetminus \{1\} = G \smallsetminus \{1\}$ has $2, \dots , n$ as a simple elimination order. Thus, $1, 2, \dots , n$ is a simple elimination order for $G \smallsetminus e$ , so $G \smallsetminus e$ is strongly chordal by Theorem 2.3.

Suppose in addition we know that G is claw-free. If $G \smallsetminus e$ has an induced claw, then G has an induced subgraph of the form

But this is impossible since $k \in N[1]$ implies $l \in N[k] \subseteq N[j]$ .

The following is a refinement of [Reference Mohammadi and Sharifan29, 3.7].

Proposition 2.8. Let G be a graph and $e = \{i,j\} \in E(G)$ . Denote by $G_{e}$ the graph on $[n]=\{1,\dots ,n\}$ having edge set

$$\begin{align*}E(G_{e}) = E(G \smallsetminus e) \cup \binom{N_{G \smallsetminus e}(i)}{2} \cup \binom{N_{G \smallsetminus e}(j)}{2}. \end{align*}$$

For any cycle $C: i = i_0, i_1, \dots , i_s, i_{s+1}= j$ of G containing e and $0 \leq t \leq s$ , set

$$\begin{align*}g_{C,t} = y_{i_1}\cdots y_{i_t}x_{i_{t+1}} \cdots x_{i_s}. \end{align*}$$

Then

$$\begin{align*}J_{G \smallsetminus e} : f_e = J_{G_e} + (g_{C, t} : e \subseteq C \text{ an induced cycle of } G, 0 \leq t \leq s). \end{align*}$$

We observe that the graph $G_e$ of the theorem is the graph obtained from $G \smallsetminus e$ by adding edges to turn the ends of e into simplicial vertices of $G \smallsetminus e$ .

Proof . The statement of the proposition is precisely [Reference Mohammadi and Sharifan29, 3.7] if we omit the condition that the cycles C are induced. In the statement of that theorem, the monomials are indexed instead by paths from i to j in $G \smallsetminus e$ , but such paths correspond bijectively with cycles of G containing e. We will show that each monomial $g_{C, t}$ is divisible by a suitable monomial $g_{C', t'}$ whenever C admits a chord.

Let $C: i = i_0, i_1, \dots , i_s, i_{s+1} = j$ be any cycle of G containing e. If C has a chord $\{i_p, i_q\}$ for some $p < q$ , then we have a cycle $C': i = i_0, i_1, \dots , i_p, i_q, i_{q+1}, \dots , i_{s+1} = j$ so that

$$\begin{align*}g_{C', t} = y_{i_1} \cdots y_{i_t}x_{i_{t+1}} \cdots x_{i_p}x_{i_q} \cdots x_{i_s} \end{align*}$$

divides $g_{C, t}$ for all $t < p$ ,

$$\begin{align*}g_{C', p} = y_{i_1} \cdots y_{i_p}x_{i_q} \cdots x_{i_s} \end{align*}$$

divides $g_{C, t}$ for all $p \leq t < q$ , and

$$\begin{align*}g_{C', p + t - q + 1} = y_{i_1} \cdots y_{i_p}y_{i_q} \cdots y_{i_t}x_{i_{t+1}} \cdots x_{i_s} \end{align*}$$

divides $g_{C,t}$ for all $t \geq q$ .

Corollary 2.9. Let G be a chordal graph, and let $e \in E(G)$ be a claw-avoiding simplicial edge. Let U be the unique maximal clique of G containing e. Then

$$ \begin{align*} J_{G\smallsetminus e} : f_e = J_{G \smallsetminus e} + (x_k, y_k \mid k \in U \smallsetminus e). \end{align*} $$

Proof . By the preceding proposition, we know that $J_{G \smallsetminus e}: f_e = J_{G_e} + L$ , where

$$\begin{align*}L = (g_{C,t} : C \text{ is an induced cycle of } G, e \subseteq C, \text{ and } 0 \leq t \leq s).\end{align*}$$

If $e = \{i,j\}$ , then for every $k \in U \smallsetminus e$ , we have an induced $3$ -cycle $i,j,k$ so that

$$\begin{align*}(x_k,y_k \mid k \in U\smallsetminus e) \subseteq L. \end{align*}$$

On the other hand, every induced cycle of G has length 3 since G is chordal so that $L = (x_k,y_k \mid k \in U\smallsetminus e)$ .

Next, we show that $J_{G_e} + L = J_{G \smallsetminus e} + L$ . Since $G \smallsetminus e$ is a subgraph of $G_e$ , it is clear that $J_{G \smallsetminus e} + L \subseteq J_{G_e} + L$ . Every additional generator of $J_{G_e}$ but not $J_{G \smallsetminus e}$ is of the form $f_{k, l}$ for some $k,l \in N_{G \smallsetminus e}(i)$ or $k, l \in N_{G \smallsetminus e}(j)$ . If $k, l \in N_{G \smallsetminus e}(i)$ , the induced subgraph on $\{i,j,k,l\}$ has as an edge $\{j,k\}$ , $\{j,l\}$ , or $\{k,l\}$ because e does not belong to an induced claw of G. If $\{k,l\}$ is an edge, then $f_{k, l} \in J_{G \smallsetminus e}$ . On the other hand, if $\{j,l\}$ or $\{j,k\}$ is an edge of G, then either $\{i, j, l\}$ or $\{i, j, k\}$ is a clique containing e, and so, either l or k belongs to U as e is simplicial, in which case it is clear that $f_{k,l} \in L$ . Either way, we see that $f_{k, l} \in J_{G \smallsetminus e} + L$ , and the argument is completely analogous if $k, l \in N_{G \smallsetminus e}(j)$ . Thus, we have $J_{G_e} + L = J_{G \smallsetminus e} + L$ as claimed.

We are now ready to prove our main theorem. Recall the notation in Definition 1.1.

Proof of Theorem 2.1.

We argue by induction on the number of edges of G. If G has no edges, then $J_G = 0$ and the statement holds trivially. So, we may assume that G has at least one edge and that the theorem holds for all claw-free strongly chordal graphs with fewer edges. In that case, since G is strongly chordal and claw-free, it has a simplicial edge e such that $G \smallsetminus e$ is strongly chordal and claw-free by Lemma 2.7. Since $G \smallsetminus e$ is a claw-free strongly chordal graph with fewer edges than G, we know that $R_{G \smallsetminus e}$ is Koszul by induction. We will show that $R_G$ is Koszul as well.

Let K be the unique maximal clique of G containing e. Corollary 2.9 yields the exact sequence of $R_{G \smallsetminus e}$ -modules

$$\begin{align*}0 \to R_H(-2) \stackrel{f_e}{\to} R_{G \smallsetminus e} \to R_G \to 0, \end{align*}$$

where H is the induced subgraph of $G \smallsetminus e$ obtained by deleting the vertices of $K \smallsetminus e$ . Since H is an induced subgraph of $G \smallsetminus e$ , we have that $R_H$ is an algebra retract of $R_{G \smallsetminus e}$ ; this means that we have the homogeneous inclusion $i: \ R_H\subset R_{G \smallsetminus e}$ and homogeneous surjection $\varphi : R_{G \smallsetminus e} \longrightarrow R_H$ so that $\varphi i=id$ (see [Reference Römer and Saeedi Madani38] for other applications of algebra retracts in an algebra-graph setting). Combining this with the assumption that $R_{G \smallsetminus e}$ is Koszul yields that $R_H$ has a linear free resolution as a module over $R_{G \smallsetminus e}$ by [Reference Ohsugi, Herzog and Hibi31, 1.4]. Now, combine this with the above short exact sequence, and conclude that $\operatorname {\mathrm {reg}}_{R_{G \smallsetminus e}} R_G \leq 1$ . It follows that $R_G$ is Koszul by [Reference Conca, De Negri and Evelina Rossi12, Theorem 2].

Corollary 2.10. Let G be a chordal graph and $e \in E(G)$ be a claw-avoiding simplicial edge of G. If $G \smallsetminus e$ is strongly chordal and claw-free, then so is G.

Proof . By our main theorem 1.2, the ring $R_{G \smallsetminus e}$ is Koszul. Then the proof above implies that $R_G$ is also Koszul. Applying our main theorem once more yields that G is strongly chordal and claw-free.

Corollary 2.11. A chordal graph is strongly chordal and claw-free if and only if it has a claw-avoiding perfect edge elimination order.

Proof . ( $\Rightarrow $ ): We argue by induction on the number of edges of G. The statement holds trivially if G has no edges, so we may assume that G has at least one edge and that all claw-free strongly chordal graphs with fewer edges have a claw-avoiding perfect edge elimination order. By Lemma 2.7, there is a simplicial edge $e_1$ of G such that $G \smallsetminus e_1$ is strongly chordal and claw-free. Note that $e_1$ is claw-avoiding as G is claw-free. By induction $G \smallsetminus e_1$ has a claw-avoiding perfect edge elimination order $e_2, \dots , e_m$ , and so $e_1, e_2, \dots , e_m$ is a claw-avoiding perfect edge elimination order for G.

( $\Leftarrow $ ): Let $e_1, \dots , e_m$ be a claw-avoiding perfect edge elimination order for G. Again, we argue by induction on m. The case $m = 0$ being trivial, we may assume that $m \geq 1$ and that the corollary holds for all chordal graphs having a claw-avoiding perfect edge elimination order with fewer edges. Since $e_1$ is a claw-avoiding simplicial edge of G, we know in particular that $G \smallsetminus e_1$ is chordal, and as $G \smallsetminus e_1$ has claw-avoiding perfect edge elimination order $e_2, \dots , e_m$ , it follows by induction that $G \smallsetminus e_1$ is strongly chordal and claw-free. Thus, Corollary 2.10 implies that G is also strongly chordal and claw-free.

3 Closed graphs and quadratic Gröbner bases

In the next section, we will make use of the concept of closed graphs, whose properties are reviewed here.

A closed order for a graph G is an ordering $v_1, \dots , v_n$ of its vertices such that whenever $v_iv_j, v_iv_k \in E(G)$ and either $i < j, k$ or $i> j, k$ , then $v_jv_k \in E(G)$ . The graph G is a closed graph if it has a closed order. It is straightforward to check that every closed order is a strong elimination order. (There are two cases to consider depending on whether $k < j$ or $j < k$ .) Thus, every closed graph is strongly chordal.

The significance of closed graphs is that they characterize when binomial edge ideals have a quadratic Gröbner basis. Hence, the binomial edge ideal of any closed graph is Koszul.

Theorem 3.1 [Reference Herzog, Hibi, Hreinsdóttir, Kahle and Rauh25, 1.1][Reference Crupi and Rinaldo16, 3.4].

For a graph G, the following are equivalent:

  1. (a) $J_G$ has a quadratic Gröbner basis with respect to some monomial order.

  2. (b) $J_G$ has a quadratic Gröbner basis with respect to some lex order.

  3. (c) G is a closed graph.

Denoting by

$$\begin{align*}N^{\geq}[v_i] = \{v_j \in N[v_i] \mid j \geq i\} \qquad N^{\leq}[v_i] = \{v_j \in N[v_i] \mid j \leq i\} \end{align*}$$

the sets of neighbors of the vertex $v_i$ that come after or before i respectively in the given order, we can characterize when G has a closed order as follows.

Proposition 3.2 [Reference Crupi and Rinaldo16, 4.8][Reference Cox and Erskine15, 2.2].

For a graph G, the following are equivalent:

  1. (a) $v_1, \dots , v_n$ is closed order for G.

  2. (b) The sets $N^{\leq }[v_i]$ and $N^{\geq }[v_i]$ are both cliques for all i.

  3. (c) The set $N^{\geq }[v_i]$ is a clique and $N^{\geq }[v_i] = [v_i, v_{i+r}]$ , where $r = \left \vert N^{\geq }[v_i] \right \vert $ , for all i.

A graph G is an intersection graph if there is a family of sets $\{F_i\}_{i \in V(G)}$ such that $ij \in E(G)$ if and only if $F_i \cap F_j \neq \varnothing $ . A graph G is an interval graph if it is the intersection graph of a family of open intervals of real numbers. If the family of intervals can be chosen so that no interval properly contains any other, then G is a proper interval graph . A proper interval order for a graph G is an ordering $v_1, \dots , v_n$ of its vertices such that whenever $v_iv_k \in E(G)$ and $i < j < k$ , then $v_iv_j, v_jv_k \in E(G)$ . We note that every proper interval order is a closed order and vice versa by Proposition 3.2.

Theorem 3.3 [Reference Crupi and Rinaldo17, 2.2][Reference Herzog, Hibi and Ohsugi26, 7.10].

For a graph G, the following are equivalent:

  1. (a) G is a closed graph.

  2. (b) G has a proper interval order.

  3. (c) G is a proper interval graph.

  4. (d) G is chordal, claw-free, net-free, and tent-free.

The net (see Figure 1) is the smallest example of a nonclosed graph whose binomial edge ideal is Koszul; a Koszul filtration argument is given in [Reference Herzog, Hibi and Ohsugi26, 7.56]. Our main theorem 1.2 shows that the presence of induced nets is precisely the difference between those binomial edge ideals that are Koszul and those that admit a quadratic Gröbner basis.

4 Obstructions to Koszulness

As discussed in the introduction, the “only if” direction of Theorem 1.2 was already shown by Ene, Herzog, and Hibi, namely:

Theorem 4.1 [Reference Ene, Herzog and Hibi21], [Reference Herzog, Hibi and Ohsugi26, Section 7.4.1].

If the binomial edge ideal $J_G$ is Koszul, then the graph G is chordal, claw-free, and tent-free.

Their argument easily reduces to checking that the binomial edge ideals of n-cycles for $n \ge 4$ , the claw, and the tent are not Koszul, which they verify by computer. For completeness, in the three lemmas below, we provide direct proofs that the binomial edge ideals of n-cycles for $n \ge 4$ , the claw, and the tent are not Koszul in any characteristic. The arguments for the n-cycles and the claw are fairly quick, but the reasoning for the tent is longer and more intricate.

Some care about the characteristic is warranted since D’Alì and Venturello have shown [Reference D’Alì and Venturello18, 5.15] that for any finite list of primes P, it is possible to construct ideals of polynomials with integer coefficients that fail to be Koszul modulo p exactly when $p \in P$ . Moreover, there are examples of graphs G such that the finite resolution of $J_G$ over the ambient polynomial ring can depend on the characteristic of the coefficient field [Reference Bolognini, Macchia and Strazzanti5, 7.6], [Reference Bolognini, Macchia, Strazzanti and Welker6, 5.9]. A priori, the Koszul property of $R_G$ could depend on the characteristic as well, although it follows from Theorem 2.1 that this is not the case.

Throughout this section, we set $f_{i,j} = x_iy_j - x_jy_i$ for any i and j (not necessarily adjacent in G); in particular, $J_G$ is generated by the $f_{i,j}$ corresponding to edges.

We will make frequent use of known restrictions on graded Betti numbers of Koszul algebras. We refer the reader to [Reference Peeva34] for background on free resolutions and graded Betti numbers; we just recall the definition of the graded Betti number $\beta _{i,j}^R(M)$ of a finitely generated graded R-module M over a graded $\Bbbk $ -algebra R:

$$ \begin{align*}\beta_{i,j}^R(M)=\dim_{\Bbbk}\,\text{Tor}^R_i(M,k))_{j}. \end{align*} $$

The graded Betti numbers may be displayed in the Betti table, which contains $\beta _{i,j}^R(M)$ in the $i,i+j$ ’th slot. We will also use the following result of Blum concerning when the symmetric algebra $\operatorname {\mathrm {Sym}}_R(M)$ is Koszul.

Proposition 4.2 [Reference Blum4, 3.1].

Let R be a standard graded $\Bbbk $ -algebra and M be a finitely generated graded R-module such that $\operatorname {\mathrm {Sym}}_R(M)$ is Koszul. Then R is also Koszul, and $\operatorname {\mathrm {Sym}}_R^i(M)$ has a linear free resolution over R for all i. In particular, M has a linear free resolution over R.

Lemma 4.3. Let $C_n$ be the n-cycle. Then the ring $R_{C_n}$ is not Koszul.

Proof . For $n \ge 5$ , see the proof in [Reference Herzog, Hibi and Ohsugi26, 7.47] where it is shown that $J_{C_n}$ has a minimal first syzygy of degree n; for $n \ge 5$ , if $J_{C_n}$ were Koszul, this contradicts [Reference Kempf27, Lemma 4]. For $n = 4$ , consider the graph G shown below

and set $e = \{2, 3\}$ . Note that $G \smallsetminus e = C_4$ . We will show that $\beta ^S_{2,4}(R_{C_4}) = 9> \binom {4}{2}$ , so $R_{C_4}$ is not Koszul by [Reference Boocher, Hamid Hassanzadeh and Iyengar8, 3.4].

Note that $S_G=\Bbbk [x_1, \dots ,x_4, y_1, \dots , y_4].$ First, we observe that the labeling of G is closed, and so $R_G$ has a quadratic Gröbner basis with respect to the lex order with $x_1> \cdots > x_4 > y_1 > \cdots > y_4$ . In particular, we have that the initial ideal is

$$\begin{align*}\operatorname{\mathrm{in}}_>(J_G) = (x_1y_2, x_1y_3, x_2y_3, x_2y_4, x_3y_4), \end{align*}$$

so it is the edge ideal of a path with 6 vertices. Using [Reference Boocher, D’Alì, Grifo, Montaño and Sammartano7, 3.4, 3.5], one can check that the Betti table of $S_G/\operatorname {\mathrm {in}}_>(J_G)$ is:

Since G is a closed graph, $\beta _{i,j}^R(R_G) = \beta _{i,j}^{S_G}(S_G/\operatorname {\mathrm {in}}_>(J_G))$ for all $i,j$ ( see [Reference Peeva35, 1.3]).

To see that $\beta ^S_{2,4}(R_{C_4}) = 9$ , we note that Theorem 2.8 yields the exact sequence

$$\begin{align*}0 \to {S_G}/(x_1, y_1, x_4, y_4)(-2) \xrightarrow{f_{2,3}} R_{C_4} \to R_G \to 0, \end{align*}$$

from which we obtain the exact sequence

$$ \begin{align*} 0 = \operatorname{\mathrm{Tor}}_3^{S_G}(R_G, \Bbbk)_4 &\to \operatorname{\mathrm{Tor}}_2^{S_G}({S_G}/(x_1,y_1, x_4, y_4), \Bbbk)_2 \to \operatorname{\mathrm{Tor}}_2^{S_G}(R_{C_4}, \Bbbk)_4 \\ &\to \operatorname{\mathrm{Tor}}_2^{S_G}(R_G, \Bbbk)_4 \to \operatorname{\mathrm{Tor}}_1^{S_G}({S_G}/(x_1,y_1, x_4, y_4), \Bbbk)_2 = 0. \end{align*} $$

Since $x_1,y_1,x_4,y_4$ is a regular sequence on ${S_G}$ , we have

$$\begin{align*}\beta_{2,4}^S(R_{C_4}) = \beta_{2,2}^{S_G}({S_G}/(x_1, y_1, x_4, y_4)) + \beta^{S_G}_{2,4}(R_G) = 6 + 3 = 9 \end{align*}$$

as desired.

Lemma 4.4. Let G be the claw. Then the ring $R_{G}$ is not Koszul.

Proof . Consider the claw G labeled as shown below

Thus, the binomial edge ideal is $J_G = (f_{1,2}, f_{1,3}, f_{1,4}) \subseteq S_G = \Bbbk [x_1, y_1, \dots , x_4, y_4]$ . If $e = \{1,4\}$ , then Theorem 2.8 shows that $J_{G \smallsetminus e} : f_e = J_{G_e} = (f_{1,2}, f_{1,3}, f_{2,3})$ , since $G \smallsetminus e$ has no induced cycles. Hence, $J_{G \smallsetminus e} : f_e = I_2(M)$ , where

$$\begin{align*}M = \begin{pmatrix} x_1 & x_2 & x_3 \\ y_1 & y_2 & y_3 \end{pmatrix}, \end{align*}$$

which has a Hilbert-Burch resolution. Applying Theorem 2.8 once more to $G' = G \smallsetminus \{4\}$ and the edge $e' = \{1, 3\}$ shows that $(f_{1,2}) : f_{1,3} = J_{G'\smallsetminus e'} : f_{e'} = J_{G^{\prime }_{e'}} = (f_{1,2})$ , so $J_{G \smallsetminus e} = (f_{1,2}, f_{1,3})$ is a complete intersection. We can construct an explicit chain map lifting the multiplication by $f_{1,4}$ map $S_G/I_2(M)(-2) \to R_{G \smallsetminus e}$ as shown below.

The first column of the middle vertical map comes from the Plücker relation $f_{1,2}f_{3,4} - f_{1,3}f_{2,4} + f_{2,3}f_{1,4} = 0$ (see [Reference Bruns and Herzog10, 7.2.3]). Hence, the mapping cone of the above chain map yields the minimal free resolution of $R_G$ , which has Betti table:

In particular, because $\beta ^S_{2,4}(R_G) = 4> \binom {3}{2}$ , we see that $R_G$ is not Koszul by [Reference Boocher, Hamid Hassanzadeh and Iyengar8, 3.4].

Lemma 4.5. Let T be the tent. Then the ring $R_T$ is not Koszul.

Proof . Consider the tent T with labeling as shown in Figure 3.

Figure 3 The tent.

Setting $G = T \smallsetminus \{6\}$ , we note that $R_T \cong \operatorname {\mathrm {Sym}}_{R_G}(M)$ , where M is the cokernel of the matrix

$$\begin{align*}\partial_1 = \begin{pmatrix} -y_2 & -y_4 \\ x_2 & x_4 \end{pmatrix}. \end{align*}$$

If $R_T$ is Koszul, then M has a linear free resolution over $R_G$ by Proposition 4.2. However, we will show that M has a minimal nonlinear fourth syzygy, and so the ring $R_T$ is not Koszul.

To see this, we first note that the labeling of G is closed, and so $J_G$ has a quadratic Gröbner basis with respect to the lex order with $x_1> \cdots > x_5 > y_1 > \cdots > y_5$ . In particular, we have that the initial ideal is

$$\begin{align*}\operatorname{\mathrm{in}}_>(J_G) = (x_1y_2, \, x_1y_3, \, x_2y_3, \, x_2y_4, \, x_3y_4, \, x_3y_5, \, x_4y_5). \end{align*}$$

Thus, the quotient $R_G$ has a monomial basis consisting of all monomials not contained in the above monomial ideal. We also note that $R_G$ admits a $\mathbb {Z}^5$ -grading via $\deg x_i = \deg y_i = \mathbf {e}_i$ , where $\mathbf {e}_i$ is the i-th standard basis vector of $\mathbb {Z}^5$ , and $\operatorname {\mathrm {Ker}} \partial _1$ is homogeneous with respect to this grading. We exploit this grading to simplify the computation of the first few terms of the linear strand of the minimal free resolution of M.

Let $u = (u_1, u_2) \in \operatorname {\mathrm {Ker}} \partial _1 \subseteq R_G(-\mathbf {e}_2) \oplus R_G(-\mathbf {e}_4)$ be a nonzero homogeneous element of degree $\mathbf {a} \in \mathbb {Z}^5$ . Since $(R_G)_{\mathbf {b}} = 0$ if any component of $\mathbf {b}$ is negative, we must have either $\mathbf {e}_2 \leq \mathbf {a}$ or $\mathbf {e}_4 \leq \mathbf {a}$ , where $\mathbf {b} = (b_1, \dots , b_5) \leq (a_1, \dots , a_5) = \mathbf {a}$ means that $b_i \leq a_i$ for all i. If u is a linear syzygy, then we have $\left \vert \mathbf {a} \right \vert = \sum _i a_i = 2$ , so $\mathbf {a} = \mathbf {e}_2 + \mathbf {e}_i, \mathbf {e}_4 + \mathbf {e}_j, \mathbf {e}_2 + \mathbf {e}_4$ for some $i \neq 4$ and $j \neq 2$ . In the first case, we have $u = (\alpha x_i + \beta y_i, 0)$ for some $\alpha , \beta \in \Bbbk $ , so $\partial _1u = 0$ implies $\alpha x_2x_i + \beta x_2y_i = 0$ . As $x_2x_i$ is part of the monomial basis for $R_G$ , it follows that $\alpha = 0$ , and so $\beta x_2y_i = 0$ . If $i \neq 5$ , then $0 = \beta x_2y_i = \beta x_iy_2$ . In any case, either $x_2y_i$ or $x_iy_2$ is part of the monomial basis for $R_G$ , so $\beta = 0$ , which contradicts that u is nontrivial. A similar argument shows that there are no nontrivial elements of $\operatorname {\mathrm {Ker}} \partial _1$ in degree $\mathbf {a} = \mathbf {e}_4 + \mathbf {e}_j$ for $j \neq 2$ . However, when $\mathbf {a} = \mathbf {e}_2 + \mathbf {e}_4$ , we see that $u = (\alpha _1x_4 + \beta _1y_4, \alpha _2x_2 + \beta _2y_2)$ , so $\partial _1u = 0$ implies

$$\begin{align*}0 = \alpha_1x_2x_4 + \beta_1x_2y_4 + \alpha_2x_2x_4 + \beta_2x_4y_2 = (\alpha_1 + \alpha_2)x_2x_4 + (\beta_2 + \beta_1)x_4y_2. \end{align*}$$

As $x_2x_4$ and $x_4y_2$ are part of the monomial basis for $R_G$ , it follows that $\alpha _2 = -\alpha _1$ and $\beta _2 = -\beta _1,$ so

$$\begin{align*}u = \alpha_1\begin{pmatrix} x_4 \\ -x_2 \end{pmatrix} + \beta_1\begin{pmatrix} y_4 \\ -y_2 \end{pmatrix}, \end{align*}$$

and the columns of the matrix

$$\begin{align*}\partial_2 = \begin{pmatrix} x_4 & y_4 \\ -x_2 & -y_2 \end{pmatrix} \end{align*}$$

are easily seen to generate the linear syzygies of $\partial _1$ .

If $u = (u_1, u_2) \in \operatorname {\mathrm {Ker}} \partial _2 \subseteq R(-\mathbf {e}_2-\mathbf {e}_4)^2$ is a nonzero homogeneous element of degree $\mathbf {a}$ , then $\mathbf {a} \geq \mathbf {e}_2 + \mathbf {e}_4$ . If in addition u is a linear syzygy, then $\mathbf {a} = \mathbf {e}_2 + \mathbf {e}_4 + \mathbf {e}_i$ for some i, so $u = (\alpha _1x_i + \beta _1y_i, \alpha _2x_i + \beta _2y_i)$ . Then $\partial _2u = 0$ implies:

$$ \begin{align*} 0 &= \alpha_1x_ix_4 + \beta_1x_4y_i + \alpha_2x_iy_4 + \beta_2y_iy_4 \\ 0 &= \alpha_1x_ix_2 + \beta_1x_2y_i + \alpha_2x_iy_2 + \beta_2y_iy_2 \end{align*} $$

As $x_ix_4$ and $y_iy_4$ are part of the monomial basis for $R_G$ , it follows that $\alpha _1 = \beta _2 = 0$ , so $\beta _1x_4y_i + \alpha _2x_iy_4 = 0$ and $\beta _1x_2y_i + \alpha _2x_iy_2 = 0$ . If $i = 1, 5$ , then $\alpha _2 = \beta _1 = 0$ as well since $x_4y_1, x_1y_4, x_2y_5$ , and $x_5y_2$ are also part of the monomial basis for $R_G$ , which contradicts that u is nontrivial. Hence, we must have $i = 2, 3, 4$ , in which case we have $0 = \beta _1x_4y_i + \alpha _2x_iy_4 = (\beta _1 + \alpha _2)x_iy_4 = (\beta _1 + \alpha _2)x_4y_i$ where either $x_iy_4$ or $x_4y_i$ is part of the monomial basis for $R_G$ , so $\alpha _2 = -\beta _1$ . From this, it is easily seen that the columns of the matrix

$$\begin{align*}\partial_3 = \begin{pmatrix} y_2 & y_4 & y_3 \\ -x_2 & -x_4 & -x_3 \end{pmatrix} \end{align*}$$

generate the linear syzygies of $\partial _2$ .

Let $u = (u_1, u_2, u_3) \in \operatorname {\mathrm {Ker}} \partial _3 \subseteq R(-2\mathbf {e}_2-\mathbf {e}_4) \oplus R(-\mathbf {e}_2-2\mathbf {e}_4) \oplus R(-\mathbf {e}_2-\mathbf {e}_3-\mathbf {e}_4)$ be a nonzero linear syzygy of degree $\mathbf {a}$ . If $\mathbf {a} \ngeq \mathbf {e}_2 + \mathbf {e}_3 + \mathbf {e}_4$ , then $u_3 = 0$ , so $(u_2, u_1) \in \operatorname {\mathrm {Ker}} \partial _1$ and

$$\begin{align*}u = \alpha\begin{pmatrix} -x_2 \\ x_4 \\ 0 \end{pmatrix} + \beta\begin{pmatrix} -y_2 \\ y_4 \\ 0 \end{pmatrix} \end{align*}$$

for some $\alpha , \beta \in \Bbbk $ . So, we may assume that $\mathbf {a} \geq \mathbf {e}_2 + \mathbf {e}_3 + \mathbf {e}_4$ , in which case we have either:

  1. 1. $\mathbf {a} = \mathbf {e}_2 +\mathbf {e}_3 + \mathbf {e}_4 + \mathbf {e}_i$ for some $i \neq 2, 4$ ,

  2. 2. $\mathbf {a} = \mathbf {e}_2 +\mathbf {e}_3 + 2\mathbf {e}_4$ , or

  3. 3. $\mathbf {a} = 2\mathbf {e}_2 +\mathbf {e}_3 + \mathbf {e}_4$ .

In case (1), we have $u = (0, 0, \alpha x_i + \beta y_i)$ for some $\alpha , \beta \in \Bbbk $ , so $\partial _3u = 0$ implies $0 = \alpha x_iy_3 + \beta y_iy_3$ and $0 = \alpha x_ix_3 + \beta y_ix_3$ . As $x_ix_3$ and $y_iy_3$ are part of the monomial basis for $R_G$ , it follows that $\alpha = \beta = 0$ , which contradicts that u is nontrivial. In case (2), we have $u = (0, \alpha _2x_3 + \beta _2 x_3, \alpha _3x_4 + \beta _3 x_4)$ for some $\alpha _i, \beta _i \in \Bbbk $ . Then $\partial _3u = 0$ implies

$$\begin{align*}0 = \alpha_2x_3y_4 + \beta_2y_3y_4 + \alpha_3x_4y_3 + \beta_3y_3y_4 = (\alpha_2 + \alpha_3)x_4y_3 + (\beta_2 + \beta_3)y_3y_4. \end{align*}$$

As $x_4y_3$ and $y_3y_4$ are part of the monomial basis for $R_G$ , we see that $\alpha _3 = -\alpha _2$ and $\beta _3 = -\beta _2$ , so

$$\begin{align*}u = \alpha_2\begin{pmatrix} 0 \\ x_3 \\ -x_4 \end{pmatrix} + \beta_2\begin{pmatrix} 0 \\ y_3 \\ -y_4 \end{pmatrix}. \end{align*}$$

By a similar argument applied to case (3), we see that the columns of the matrix

$$\begin{align*}\partial_4 = \begin{pmatrix} -x_2 & -y_2 & x_3 & y_3 & 0 & 0 \\ x_4 & y_4 & 0 & 0 & x_3 & y_3 \\ 0 & 0 & -x_2 & -y_2 & -x_4 & -y_4 \end{pmatrix} \end{align*}$$

generate the linear syzygies of $\partial _3$ .

Finally, we observe that $x_3f_{1,5} = x_1f_{3,5} + x_5f_{1,3} \in J_G$ and similarly that $y_3f_{1,5} \in J_G$ , so $q = (0, 0, f_{1,5})$ is a quadratic syzygy of $\partial _3$ . If q is a nonminimal syzygy, the above description of the linear syzygies of $\partial _3$ implies that $f_{1,5} \in J_G + (x_2, y_2, x_4, y_4)$ , which is easily seen to imply that $f_{1,5} \in J_P$ where P is the induced path of G on the vertices $\{1,3, 5\}$ . Since the vertex labeling of P is closed, we see that $\operatorname {\mathrm {in}}_>(J_P) = (x_1y_3, x_3y_5)$ , but then $\operatorname {\mathrm {in}}_>(f_{1,5}) = x_1y_5 \notin \operatorname {\mathrm {in}}_>(J_P)$ , which is a contradiction. Therefore, q is a minimal quadratic syzygy of $\partial _3$ , hence also of M. Hence, M does not have a linear resolution over $R_G$ , and $R_T$ is not Koszul.

5 Strongly chordal claw-free graphs with small clique number

In this section, we characterize strongly chordal, claw-free graphs with small clique number. The clique number $\omega (G)$ of a graph G is the maximum size of a clique of G. To state our characterization, we need a little more terminology. A cut vertex of a connected graph G is a vertex v such that $G \smallsetminus v$ is not connected. A graph G is 2-connected if it has no cut vertices. A block of a graph G is a maximal 2-connected subgraph.

As an immediate combinatorial consequence of our main theorem 1.2 and [Reference Ene, Herzog and Hibi21, 2.6], we have the following.

Corollary 5.1. Let G be a connected graph with $\omega (G) \leq 3$ . The following are equivalent:

  1. (a) G is strongly chordal and claw-free.

  2. (b) There is a tree T in which every vertex has degree at most 3 and a family of induced closed subgraphs $\{G_v\}_{v \in T}$ of G such that:

    1. (i) $V(G_v) \cap V(G_w) = \{i\}$ is a cutset of G whenever $\{v,w\}$ is an edge of T.

    2. (ii) If u and w are distinct neighbors of v in T, then $V(G_u) \cap V(G_v) \cap V(G_w) = \varnothing $ .

    3. (iii) If v is a degree 3 vertex, then $G_v$ is a clique of size 3.

  3. (c) G is claw-free, and every block of G is a closed graph.

Proof . (a) $\Leftrightarrow $ (b): This follows immediately from our main theorem 1.2 and [Reference Ene, Herzog and Hibi21, 2.6].

(b) $\Rightarrow $ (c): If G can be decomposed as described in part (b), then G is claw-free by the equivalence of (b) and (a), and the blocks of G are just the blocks of $G_v$ for all v. Since induced subgraphs of closed graphs are closed by [Reference Crupi and Rinaldo17, 2.2], it follows that every block of G is closed.

(c) $\Rightarrow $ (a): To see that G is strongly chordal when it is claw-free and all of its blocks are closed, it suffices by our main theorem 1.2 to note that G is chordal and tent-free. Since cycles and tents are 2-connected, any induced cycle or tent of G would be contained in one of its blocks. Thus, G is chordal and tent-free since the blocks of G are closed.

We note that the implication (c) $\Rightarrow $ (a) of the preceding corollary is always valid without any assumption on $\omega (G)$ . Unfortunately, the following example shows that the decomposition of the above corollary and the implication (a) $\Rightarrow $ (c) do not carry over as stated to higher clique numbers.

Example 5.2. Consider the thick net graph $TN$ shown in Figure 4. It is easily checked that $TN$ is 2-connected, but $TN$ is not a closed graph since the induced subgraph on the vertices $\{1, 2, 4, 7, 8, 9\}$ is a net. We show that $TN$ is nonetheless strongly chordal and claw-free. If we set $e_1 = \{5,9\}$ and $e_2 = \{4, 9\}$ , it is easily seen that $e_1$ is a claw-avoiding simplicial edge for $TN$ and $e_2$ is a claw-avoiding simplicial edge for $TN \smallsetminus e_1$ . Since $TN \smallsetminus \{e_1, e_2\}$ is a closed graph for the given labeling, it follows from Corollary 2.10 that $TN$ is strongly chordal and claw-free.

Figure 4 The thick net.

It is an open question to describe the structure of claw-free strongly chordal graphs with $\omega (G) \geq 4$ . In particular, if $\omega (G) \leq k$ for some $k \geq 2$ , we don’t know if being strongly chordal and claw-free is equivalent to being chordal and every maximal $(k-1)$ -connected subgraph of G is closed; note that the thick net does not rule this out, since its maximal 3-connected subgraphs are its maximal cliques.

Acknowledgments

The authors thank Arvind Kumar for some helpful conversations related to this work. Computations with Macaulay2 [Reference Grayson and Stillman24] were very useful while working on this project.

Conflict of interest

The authors have no conflicts of interest to declare.

Financial support

LaClair was partially supported by the National Science Foundation grant DMS–2100288 and by Simons Foundation Collaboration Grant for Mathematicians #580839. Mastroeni was partially supported by an AMS-Simons Travel Grant. McCullough was partially supported by National Science Foundation grants DMS–1900792 and DMS–2401256. Peeva was partially supported by the National Science Foundation grant DMS-2401238.

References

Avramov, L. L. and Eisenbud, D., ‘Regularity of modules over a Koszul algebra’, J. Algebra 153(1) (1992), 8590.10.1016/0021-8693(92)90149-GCrossRefGoogle Scholar
Avramov, L. L. and Peeva, I., ‘Finite regularity and Koszul algebras’, Amer. J. Math. 123(2) (2001), 275281.10.1353/ajm.2001.0008CrossRefGoogle Scholar
Baskoroputro, H., Ene, V. and Ion, C., ‘Koszul binomial edge ideals of pairs of graphs’, J. Algebra 515 (2018), 344359.10.1016/j.jalgebra.2018.08.029CrossRefGoogle Scholar
Blum, S., ‘Subalgebras of bigraded Koszul algebras’, J. Algebra 242(2) (2001), 795809.10.1006/jabr.2001.8804CrossRefGoogle Scholar
Bolognini, D., Macchia, A. and Strazzanti, F., ‘Cohen-Macaulay binomial edge ideals and accessible graphs’, J. Algebraic Combin. 55(4) (2022), 11391170.10.1007/s10801-021-01088-wCrossRefGoogle Scholar
Bolognini, D., Macchia, A., Strazzanti, F. and Welker, V., ‘Powers of monomial ideals with characteristic-dependent Betti numbers’, Res. Math. Sci. 9(2) (2022), Paper No. 26, 17.10.1007/s40687-022-00318-2CrossRefGoogle Scholar
Boocher, A., D’Alì, A., Grifo, E., Montaño, J. and Sammartano, A., ‘Edge ideals and DG algebra resolutions’, Matematiche (Catania) 70(1) (2015), 215238.Google Scholar
Boocher, A., Hamid Hassanzadeh, S. and Iyengar, S. B., ‘Koszul algebras defined by three relations’, in Homological and Computational Methods in Commutative Algebra, volume 20 of Springer INdAM Ser. (Cham, Springer, 2017), 5368.10.1007/978-3-319-61943-9_3CrossRefGoogle Scholar
Brandstädt, A., Bang Le, V. and Spinrad, J. P., Graph Classes: A Survey. SIAM Monographs on Discrete Mathematics and Applications (Philadelphia, PA, Society for Industrial and Applied Mathematics (SIAM), 1999).10.1137/1.9780898719796CrossRefGoogle Scholar
Bruns, W. and Herzog, J., Cohen-Macaulay Rings, volume 39 of Cambridge Studies in Advanced Mathematics (Cambridge, Cambridge University Press, 1993).Google Scholar
Caviglia, G. and Conca, A., ‘Koszul property of projections of the Veronese cubic surface’, Adv. Math. 234 (2013), 404413.10.1016/j.aim.2012.11.002CrossRefGoogle Scholar
Conca, A., De Negri, E. and Evelina Rossi, M., ‘Koszul algebras and regularity’, in Commutative Algebra (New York, Springer, 2013), 285315.10.1007/978-1-4614-5292-8_8CrossRefGoogle Scholar
Conca, A., ‘Ladder determinantal rings’, J. Pure Appl. Algebra 98(2) (1995), 119134.10.1016/0022-4049(94)00039-LCrossRefGoogle Scholar
Conca, A., ‘Koszul algebras and their syzygies’, in Combinatorial Algebraic Geometry, volume 2108 of Lecture Notes in Math. (Cham, Springer, 2014), 131.10.1007/978-3-319-04870-3_1CrossRefGoogle Scholar
Cox, D. A. and Erskine, A., ‘On closed graphs I’, Ars Combin. 120 (2015), 259274.Google Scholar
Crupi, M. and Rinaldo, G., ‘Binomial edge ideals with quadratic Gröbner bases’, Electron. J. Combin. 18(1) (2011), Paper 211, 13.10.37236/698CrossRefGoogle Scholar
Crupi, M. and Rinaldo, G., ‘Closed graphs are proper interval graphs’, An. Ştiinţ. Univ. “Ovidius” Constanţa Ser. Mat. 22(3) (2014), 3744.Google Scholar
D’Alì, A. and Venturello, L., ‘Koszul Gorenstein algebras from Cohen-Macaulay simplicial complexes’, Int. Math. Res. Not. IMRN 6 (2023), 49985045.10.1093/imrn/rnac003CrossRefGoogle Scholar
Dirac, G. A., ‘On rigid circuit graphs’, Abh. Math. Sem. Univ. Hamburg 25 (1961), 7176 10.1007/BF02992776CrossRefGoogle Scholar
Dragan, F. F., ‘Strongly orderable graphs: a common generalization of strongly chordal and chordal bipartite graphs’, Discrete Appl. Math. 99(1–3) (2000), 427442.10.1016/S0166-218X(99)00149-3CrossRefGoogle Scholar
Ene, V., Herzog, J. and Hibi, T., ‘Koszul binomial edge ideals’, in Bridging Algebra, Geometry, and Topology, volume 96 of Springer Proc. Math. Stat. (Cham, Springer, 2014), 125136.10.1007/978-3-319-09186-0_8CrossRefGoogle Scholar
Farber, M., ‘Characterizations of strongly chordal graphs’, Discrete Math. 43(2–3) (1983), 173189.10.1016/0012-365X(83)90154-1CrossRefGoogle Scholar
Fröberg, R., ‘Koszul algebras’, in Advances in Commutative Ring Theory (Fez, 1997), volume 205 of Lecture Notes in Pure and Appl. Math. (New York, Dekker, 1999), 337350.Google Scholar
Grayson, D. R. and Stillman, M. E.. Macaulay2, a software system for research in algebraic geometry. URL: http://www2.macaulay2.com.Google Scholar
Herzog, J., Hibi, T., Hreinsdóttir, F., Kahle, T. and Rauh, J., ‘Binomial edge ideals and conditional independence statements’, Adv. in Appl. Math. 45(3) (2010), 317333.10.1016/j.aam.2010.01.003CrossRefGoogle Scholar
Herzog, J., Hibi, T. and Ohsugi, H., Binomial Ideals, volume 279 of Graduate Texts in Mathematics (Cham, Springer, 2018).10.1007/978-3-319-95349-6CrossRefGoogle Scholar
Kempf, G. R., ‘Some wonderful rings in algebraic geometry’, J. Algebra 134(1) (1990), 222224.10.1016/0021-8693(90)90219-ECrossRefGoogle Scholar
McCullough, J. and Peeva, I., ‘Infinite graded free resolutions’, in Commutative Algebra and Noncommutative Algebraic Geometry. Vol. I, volume 67 of Math. Sci. Res. Inst. Publ. (New York, Cambridge University Press, 2015), 215257.10.1017/9781107588530.009CrossRefGoogle Scholar
Mohammadi, F. and Sharifan, L., ‘Hilbert function of binomial edge ideals’, Comm. Algebra 42(2) (2014), 688703.10.1080/00927872.2012.721037CrossRefGoogle Scholar
Ohtani, M., ‘Graphs and ideals generated by some 2-minors’, Comm. Algebra 39(3) (2011), 905917.10.1080/00927870903527584CrossRefGoogle Scholar
Ohsugi, H., Herzog, J. and Hibi, T., ‘Combinatorial pure subrings’, Osaka J. Math. 37(3) (2000), 745757.Google Scholar
Ohsugi, H. and Hibi, T., ‘Koszul bipartite graphs’, Adv. in Appl. Math. 22(1) (1999), 2528.10.1006/aama.1998.0615CrossRefGoogle Scholar
Ohsugi, H. and Hibi, T., ‘Toric ideals generated by quadratic binomials’, J. Algebra 218(2) (1999), 509527.10.1006/jabr.1999.7918CrossRefGoogle Scholar
Peeva, I., Graded Syzygies, volume 14 of Algebra and Applications (London, Springer-Verlag London, Ltd., 2011).10.1007/978-0-85729-177-6CrossRefGoogle Scholar
Peeva, I., ‘Closed binomial edge ideals’, J. Reine Angew. Math. 803 (2023), 133.Google Scholar
Polishchuk, A. and Positselski, L., Quadratic Algebras, volume 37 of University Lecture Series (Providence, RI, American Mathematical Society, 2005).Google Scholar
Priddy, S. B., ‘Koszul resolutions’, Trans. Amer. Math. Soc. 152 (1970), 3960.10.1090/S0002-9947-1970-0265437-8CrossRefGoogle Scholar
Römer, T. and Saeedi Madani, S., ‘Retracts and algebraic properties of cut algebras’, European J. Combin. 69 (2018), 214236.10.1016/j.ejc.2017.11.002CrossRefGoogle Scholar
West, D. B., Introduction to Graph Theory (Upper Saddle River, NJ, Prentice Hall, Inc., 1996).Google Scholar
Figure 0

Figure 1 The tent (left), claw (center), and net (right) graphs.

Figure 1

Figure 2 The 4-trampoline.

Figure 2

Figure 3 The tent.

Figure 3

Figure 4 The thick net.