1 Introduction

Let \(S = K[x_{1},\ldots , x_{n}] \) be the polynomial ring in n variables over a field K and I be a squarefree monomial ideal in S Fröberg [5, Theorem 1] proved that the edge ideal of a finite graph G has a 2-linear resolution if and only if the complement graph \(\overline{G}\) of G is a chordal graph. Emtander [4], Woodroofe [12], Van Tuyl and Villarreal [11] have partially generalized the Fröberg’s Theorem. It is known that monomial ideals with 2-linear resolution have linear quotients [7]. We denote by \(\varDelta ^{\vee }\) the Alexander dual of \(\varDelta \). Let I be the Stanley–Reisner ideal of \(\varDelta ^{\vee }\) which is generated in degree 2 and has a linear resolution. By a result of Eagon–Reiner [3], \(\varDelta \) is a Cohen–Macaulay of codimension 2. In [1], it is shown that if \(\varDelta \) is a Cohen–Macaulay simplicial complex of codimension 2, then \(\varDelta \) is vertex decomposable. Hence, by a result of [10], \(I=I_{\varDelta ^{\vee }}\) is a variable-decomposable monomial ideal generated in degree 2. Therefore, if I is a monomial ideal generated in degree 2, then

\((*)\; \; I\) is a variable-decomposable ideal\(\ \Leftrightarrow I \) has linear quotients \(\Leftrightarrow I \) has a linear resolution.

In [8], some other classes of squarefree monomial ideals with property \((*)\) were introduced. These three concepts are not equivalent in general, but we have the following implications:

I is a variable-decomposable ideal \(\Rightarrow I\) has linear quotients \(\Rightarrow I\) has a linear resolution.

In the present paper, we consider squarefree monomial ideals generated in degree d and try to find some other classes of squarefree monomial ideals with \((*)\) property. We show that if \(d=2\), \(d=n-2\), \(n\le 5\) or \(\overline{G_I}\) is a chordal graph, then the property \((*)\) holds for I. We also study squarefree monomial ideals with linear quotients and show that if I has linear relations and \(\overline{G_I} \) is chordal, then I has linear quotients.

The paper proceeds as follows. In Sect. 2, we recall some definitions and some known results which will be needed later. Let \(I=I(G)\) be an edge ideal of a simple graph G. In Sect. 3, we present a different proof of the well-known fact that property \((*)\) holds for I. Also we show that the concepts of linear relations, linear resolution, linear quotients and variable–decomposability are equivalent when I is a squarefree monomial ideal generated in degree \(n-2\). Consequently, if \(I\subset K[x_1,\ldots ,x_5]\) is a squarefree monomial ideal, then I has property \((*)\), see Corollary 1. We consider monomial ideal I, where \(\overline{G_I}\) is a chordal graph, in Sect. 4. If \(\overline{G_I}\) is a chordal graph and I has linear relations, then I has linear quotients as will be proved in Theorem 3. Hence, if \(\overline{G_I}\) is a chordal graph, then I has property \((*)\), see Corollary 2. In addition, another property of \(\overline{G_I}\) which implies that I has linear quotients is given in Theorem 4.

2 Preliminary

For a monomial ideal I in S, there is a minimal graded free resolution of the form

$$\begin{aligned} F_{\bullet }: \; 0 \longrightarrow F_p \longrightarrow \cdots \longrightarrow F_2 \longrightarrow F_1 \longrightarrow I \longrightarrow 0 \end{aligned}$$

where \(F_i=\bigoplus _{j}S(-j)^{\beta _{i j}} \) and \( S(-j)\) denotes the free S-module obtained by shifting the degrees of S by j. The numbers \(\beta _{i j}=\beta _{i j} (I) \) are called the graded Betti numbers of I. Recall that I has a d-linear resolution over K if \(\beta _{i j} (I) = 0 \) for all \(j\ne i + d\). It follows that I is generated in degree d. Let \(\varphi : F_1 \longrightarrow I\) be the maps which send the basis element \(e_i\)s of \(F_1\) to the generators \(u_i\) of I. Recall that I has linear relations if the kernel of \(\varphi \) is generated by linear forms. Note that if the ideal I has linear relations, then all of its generators have the same degree.

The unique minimal monomial set of generators of a monomial ideal I is denoted by G(I). Monomial ideal I has linear quotients if there exists an order \(\sigma = u_1, \ldots , u_m\) of G(I) such that the colon ideal \(\langle u_1, \ldots , u_{i-1}\rangle : u_i\) is generated by a subset of variables for \(i = 2, \ldots , m\). Any order of the generators for which we have linear quotients will be called an admissible order. For a monomial \(u\in S\), set \( F(u_i) :={\,}\)Supp\((u)=\{i\; :\; x_i\mid u\} \). It is easy to see that if I is a squarefree monomial ideal, then I has linear quotients if and only if for all i and all \(j < i\) there exist \(l \in F(u_j) {\setminus } F(u_i)\) and an integer \(k < i\) such that \(F(u_k) {\setminus } F(u_i) =\{l\}\) (see [6, Corollary 8.2.4]).

Let \(u = x_1^{a_1} \ldots x_n^{a_n}\) be a monomial in S. For another monomial v, set \([u, v] = 1\) if \(x_i^{a_i} \not \mid v \) for all \(i \in F(u)\). Otherwise, set \([u, v] \ne 1\). For a monomial \(u \in S\) and a monomial ideal \(I \subseteq S\), set \(I_u=\langle u_i \in G(I) :\ [u, u_i] = 1\rangle \) and \(I^u = \langle u_j \in G(I) :\ [u, u_j] \ne 1\rangle>\).

In this paper, by the notation \(u_j:u_i=x_l\) we mean \(<u_j>:<u_i>=<x_l>.\)

Definition 1

Let I be a monomial ideal with the minimal system of generators \(\{u_1, \ldots , u_m \}\). A monomial \(u =x_1^{a_1} \cdots x_n^{a_n}\) is called shedding if \(I_u \ne 0\), and for each \(u_i \in G(I_u)\) and \(l \in F(u)\), there exists \(u_j \in G(I^u)\) such that \(u_j:u_i=x_l.\)

Definition 2

Let I be a monomial ideal minimally generated by \(\{u_1, \ldots , u_m \}\). Monomial ideal I is called a k-decomposable ideal if \(m= 1\) or else has a shedding monomial u with \(\mid \)Supp\((u)\mid \le k + 1\) such that the ideals \(I_u\) and \(I^u\) are k-decomposable. (Note that since the number of minimal generators of I is finite, the recursion procedure will stop.)

A monomial ideal is decomposable if it is k-decomposable for some \(k \ge 0\) . Also, a 0-decomposable ideal is called variable-decomposable. The concepts of shedding monomial and k-decomposable monomial ideals were first introduced by Rahmati and Yassemi in [10].

Let \(I\subset S\) be a squarefree monomial ideal which is generated in one degree. We associate with I a simple graph \(G_{I}\), called the first syzygies graph of I, whose vertices are the elements of G(I). Two vertices \(u_i\) and \(u_j\) are adjacent if there exist variables x and y such that \(xu_i=yu_j\). Since the generators of a monomial ideal form a Gröbner basis, hence the S-polynomials of each pair of its generators would give the generators of the first syzygies of the monomial ideal, that is, \(\frac{lcm(u_i,u_j)}{ u_i} e_i- \frac{lcm(u_i,u_j)}{ u_j} e_j\). So one can easily find nonlinear syzygies or, equivalently, linear syzygies.

Remark 1

Let I be a squarefree monomial ideal . If \(G_I\) is a complete graph, then I is a variable-decomposable ideal.

Let G be a finite graph on vertex set \(V=\{1,\ldots , n\}\). The edge ideal of G is defined by \(I(G) = \langle x_ix_j {:} \{i, j\} \text { is an edge in}\; G\rangle \subset S\). The induced subgraph \(G_W\) is the graph whose vertex set is \(W \subseteq V\) and the edge set consists of all the edges of G whose both endpoints are in W.

Definition 3

Let G be a graph on vertex set V and \(u \in V\). The set of all neighborhoods of u is denoted by N(u).

Definition 4

A cut vertex is a vertex that when removed (with its boundary edges) from a graph creates more components than previously in the graph.

The following fact will be used later.

Proposition 1

[2, Corollary 5.6] Every non-trivial connected graph contains at least two vertices that are not cut-vertices.

A finite graph G is called chordal if each of whose cycles of length \(> 3\) has a chord. It is clear that every induced subgraph of a chordal graph is again chordal. A perfect elimination ordering of G is an ordering \( i_{n}, \ldots , i_{2}, i_{1} \) of the vertices \(1, 2, \ldots , n\) of G such that, for each \( 1 < j \le n,\) the induced graph on \(C_{i_{j}}=\{i_j\} \cup \lbrace i_{k} \in [n] : 1 \le k < j, \lbrace i_{k}, i_{j} \rbrace \in E(G) \rbrace \) is a clique of G. In 1961,  G. A. Dirac proved that a graph G is chordal if and only if it has a perfect elimination ordering.

Remark 2

Let \(\overline{G} \) be the complementary graph of G and let the ordering \( n, n-1, \ldots ,1 \) be a perfect elimination ordering of \( \overline{G} \). If \(k> i,j\) and \(\{ i , k \} , \{j , k \} \in E( \overline{G}) \), then \( \{ i , j \} \in E( \overline{G}) \). In other words, if \(\{ i, k \}, \{ j, k \} \notin E( G) \), then \( \{ i, j \} \notin E(G) \). Hence, if \( \{ i, j \} \in E(G)\) and \(k> i,j \), then \( \{ i, k\} \in E(G) \) or \( \{j , k \} \in E(G) \).

3 Resolution of Monomial Ideals of Degree 2 and \(n-2\)

Let G be a finite graph and I be the edge ideal of G. Fröberg [5] showed that I has a linear resolution if and only if \(\overline{G}\) is a chordal graph. As it is mentioned in the introduction, for monomial ideals generated in degree 2 the following conditions are equivalent:

  1. (a)

    I has a linear resolution;

  2. (b)

    I has linear quotients;

  3. (c)

    I is a variable-decomposable ideal.

We could not find a direct proof for equivalence of these three statements, but we conclude it from known results in [1, 3, 7, 10]. For convenience of reader, a direct proof of this fact is provided in the following.

Theorem 1

Let I be the edge ideal of a finite graph G. Then the following conditions are equivalent:

  1. (a)

    The complementary graph \(\overline{G}\) of G has a perfect elimination ordering;

  2. (b)

    I is variable-decomposable ideal;

  3. (c)

    I has a linear resolution;

  4. (d)

    I has linear quotients.

Proof

\((a) \Rightarrow (b)\): We proceed by induction on \(\mid V(\overline{G})\mid \). The assertion is trivial if \( \mid V(\overline{G})\mid =2\). If \(\mid V(\overline{G})\mid >2\), then without loss of generality, we may assume that \(n, n-1, \ldots , 1 \) is a perfect elimination ordering of \( \overline{G} \) and \(I_{x_n}=I(G_1)\). It is clear that \(\overline{G_1}\) is an induced subgraph of \(\overline{G}\), so \(\overline{G_1}\) is a chordal graph. Since \(\mid V(\overline{G_1})\mid <\mid V(\overline{G}) \mid \), by induction hypothesis, \(I_{x_n}\) is a variable-decomposable ideal. It is not difficult to see that \(G_{I^{x_n}}\) is a complete graph; hence, \(I^{x_n}\) is a variable-decomposable ideal. So it is enough to show that \(x_n\) is a shedding variable. If \(u=x_ix_j \in I_{x_n}\), then by Remark 2, \(x_ix_n \in G(I)\) or \(x_jx_n \in G(I) \). Therefore, \(x_n\) is a shedding monomial.

\((b) \Rightarrow (c)\): This proof is down by induction on \(k=\mid G(I)\mid \). The assertion is obvious for \(k = 1\). Now suppose \(k>1\) and x is a shedding. Since \(\mid G(I^x)\mid \) and \(\mid G(I_x)\mid \) are less than \(\mid G(I)\mid \), by induction hypothesis, \(I^x\) and \(I_x\) have linear resolution. The short exact sequence

$$\begin{aligned} 0\rightarrow I^x \rightarrow I=I^x\oplus I_x \rightarrow I_x \rightarrow 0 \end{aligned}$$

yields the long exact sequence

$$\begin{aligned} \cdots \rightarrow Tor_i^S(I^x,K)_{i+j} \rightarrow Tor_i^S(I,K)_{i+j} \rightarrow Tor_i^S(I_x,K)_{i+j} \rightarrow \cdots \end{aligned}$$

If \(j \ne d\), then both ends in this exact sequence vanish. Hence, the middle term vanishes too.

\((c) \Rightarrow (a)\): It follows from Fröberg’s Theorem that \(\overline{G}\) is chordal. It is well known that a graph is chordal if and only if it has a perfect elimination ordering.

\((b) \Rightarrow (d)\): Let I be a variable-decomposable ideal. Without loss of generality, assume that \(x_1\) is a shedding where \(I^{x_1}=\langle u_1, \ldots , u_{t-1}\rangle \) and \(I_{x_1}=\langle u_t, \ldots , u_m\rangle \) are variable-decomposable. Arguing by induction on \(\mid G(I)\mid \) yields that \(I^{x_1}\) and \(I_{x_1}\) have linear quotients. It can be assumed that \(u_1, \ldots , u_{t-1}\) and \(u_t, \ldots , u_m\) are admissible order for \(I^{x_1}\) and \(I_{x_1}\), respectively. It will be shown that \(u_1, \ldots , u_{t-1}, u_t, \ldots , u_m\) is an admissible order for I. Let \(u_i\in I_{x_1}\) and \(u_j\in I^{x_1}\), then, by definition of shedding, there exists \(u_k\in I^{x_1}\) such that \(F(u_k) {\setminus } F(u_i)=\{1\}\). Since \(u_j\in I^{x_1}\), one has \(1 \in F(u_k){\setminus } F(u_i)\).

\((d) \Rightarrow (c)\): Follows from [6, Proposition 8.2.1]. \(\square \)

Now we consider squarefree monomial ideals generated in degree \(n-2\).

Remark 3

Let \(I \subset S\) be a squarefree monomial ideal generated in degree \( n-2 \). If there exists l, \(1\le l\le n\), such that \(x_l \not \mid u\) for all \(u\in G(I)\), then \(G_I\) is a complete graph.

Remark 4

Let \( S=K[x_{1},\ldots ,x_{n}] \) and I be a squarefree monomial ideal generated in degree \( n-2 \). If \(u_i, u_j \in G(I)\) are not adjacent in \(G_I\), then \(F(u_i) \cup F(u_j)=[n]\).

Theorem 2

Let \( S=K[x_{1},\ldots ,x_{n}] \) and \( I =\langle u_1,\ldots , u_m\rangle \) be a squarefree monomial ideal generated in degree \( n-2 \). The following are equivalent:

  1. (a)

    \( G_{I} \) is a connected graph;

  2. (b)

    I has linear relations;

  3. (c)

    I has a \((n-2)\)-linear resolution;

  4. (d)

    I has linear quotients;

  5. (e)

    I is a variable-decomposable ideal.

Proof

\((a) \Rightarrow (d)\): We order the vertices of \(G_I\) in a way that the induced subgraph on \(\{ u_1, \ldots , u_i \}\) is a connected graph for \( 2\le i\le m\). It will be shown that \(u_1, \ldots , u_m\) is an admissible order. If \( j < i\) and \(u_i, u_j\) are adjacent, take \(u_k=u_j\). Otherwise, by connectivity of the induced subgraph \(\{ u_1, \ldots , u_i \}\), there exist a vertex \(u_k\) such that \(u_i\) and \(u_k\) are adjacent. Set \(\{l\}=F(u_k) {\setminus } F(u_i)\). Then by Remark 4 one has \(l \in F(u_j) {\setminus } F(u_i)\).

\((d) \Rightarrow (c)\): It follows from [6, Proposition 8.2.1].

\((c) \Rightarrow (b)\): The implication is trivial.

\((b) \Rightarrow (a)\): See [8, Lemma 1.8].

\((a)\Rightarrow (e)\): We proceed by induction on \(\mid G(I)\mid \). If \(\mid G(I) \mid =1\) the assertion is trivial; hence, assume that \(\mid G(I)\mid \ge 2\). Let \(x_i\) be an arbitrary variable. By Remark 3, all elements in \(G_{I_{x_{i}}}\) are adjacent. If \(x_i\) is not a shedding, then there exist \(u \in I_{x_i} \) such that \(v: u \ne x_i\) for all \(v \in I^{x_i} \). Hence, u and v are not adjacent for all \(v \in I^{x_i}\). Assume that \(F(u)=[n] {\setminus } \{ i,j\}\). Now, we show that \(x_j\) is a shedding. If \(w \in I_{x_j}\), then u and w are adjacent which implies \(w \in I_{x_i} \), \(F(w)=[n] {\setminus } \{ i,j\}=F(u)\) and \( I_{x_j}=\langle u\rangle \). Since \( G_I\) is a connected graph, there exists \(v \in I^{x_j}\) such that u and v are adjacent and \(v:u=x_j\).

Now, it is enough to show that \(G_{I^{x_j}}\cong G_I {\setminus } \{u\} \) is a connected graph. If \(v ,w \in G(I){\setminus } \{u\}\), then there exists a path between v and w in \(G_I\). If u is not a vertex in this path, then v and w are connected in \(G_{I^{x_j}}\). Otherwise, we have a path \(v, \ldots , z, u, q, \ldots , w\). We know zq are adjacent to u; hence, \(z, q \in I_{x_i}\) and z and q are adjacent. Therefore, v and w are connected in \(G_{I^{x_j}}\).

Now, we consider a case that every variable is a shedding for I. By Remark 3, it is enough to show that \(G_{I^{x_i}} \) is a connected graph for some \(1 \le i \le n\). Assume \(G_{I^{x_1}} \) is not a connected graph and \(C_1, \ldots , C_p\) are its connected components. Since \(G_I\) is a connected graph, \(C_i\) is connected to \(G_{ I_{x_1}}\), for \(i=1,\ldots ,p\). If \(u_1 \in C_1\) and \( l \notin F(u_1)\), then \(u_1 \in I_{x_l}\). Since \(G_{ I_{x_l}}\) is a complete graph and \(u \in C_i\), \(1 < i \le p\), is not adjacent to \(u_1\), we conclude that \(V(C_i)\subseteq G( I^{x_l})\). We claim that if \(u_t \in I_{x_1}\) and \(u_t\) is adjacent to \(u \in C_i\), for some \(1 < i \le p\), then \( u_t \in I^{x_l}\). Suppose on the contrary that \( u_t \in I_{x_l}\), then \(F(u_t)=[n] {\setminus } \{ 1, l\}\). On the other hand, \( u \in I^{x_1}\) and \( u \in I^{x_l}\). Hence, \(\{1, l\} \subseteq F(u)\); therefore, \( u_t\) and u are not adjacent, a contradiction.

Now we consider the following two cases:

Case 1 If for all \(u \in C_1\), then there exists \(u_t \in I_{x_1}\) such that \( u_t\) is adjacent to u, then \(G_{I^{x_l}} \) is a connected graph.

Case 2 There is a \(u \in C_1\) such that u is not adjacent to any \(u_t \in I_{x_1}\). Assume \(F(u)=[n] {\setminus } \{ 1, r\}\). If \(G_{I^{x_r}}\) is a connected graph, we are done. Otherwise, since u is not adjacent to any \(u_t \in I_{x_1}\) and \(G_{I_{x_r}}\) is a complete graph, one has \(G( I_{x_1}) \subseteq G(I^{x_r})\). It is clear that \(G_{I_{x_1}} \) and \(C_i\), for all \(2 \le i \le p\), form a connected component of \(G_{I^{x_r}}\). Hence, \(G_{I^{x_r}}\) has two connected components and since \( u \notin C_1^{'}\), its second connected component is \(C_1^{'}\subsetneq C_1\). Now repeat the above procedure for \(C_1^{'}\). Since the graph \(G_I\) is finite, after a finite number of steps we find \({x_r} \) for which \(G_{I^{x_r}}\) is a connected graph.

\((e) \Rightarrow (a)\): We proceed by induction on \(\mid G(I)\mid \). Since I is variable-decomposable, there exists a shedding x such that \(I_x\) and \(I^x\) are variable-decomposable. By induction hypothesis, \(G_{I_x}\) and \(G_{I^x}\) are connected graphs. Since x is a shedding, one has every vertex in \(G_{I_x}\) is adjacent to some vertex in \(G_{I^x}\). \(\square \)

Corollary 1

Let \(I \subseteq K[x_1,\ldots , x_n] \) be a squarefree monomial ideal. If \(n\le 5\), then the following are equivalent:

  1. (a)

    I has a linear resolution;

  2. (b)

    I has linear quotients;

  3. (c)

    I is variable-decomposable ideal.

Proof

If I is generated in degree 2, then the assertion follows from Theorem 1. If I is generated in degree 3, then the assertion follows from Theorem 2. The assertion is trivial when I is generated in degree 1 or 4. \(\square \)

Notice that the following examples show that the condition \(n \le 5\) is the best possible.

Example 1

Consider monomial ideal \(I=\langle x_1x_2x_3, x_1x_3x_5, x_1x_5x_6, x_2x_5x_6, x_2x_3x_6,\)\( x_3x_4x_6, x_1x_4x_6, x_1x_2x_4, x_2x_4x_5, x_3x_4x_5\rangle \). This ideal is corresponding to triangulation of the projective plane and has a linear resolution in characteristics zero but has not a linear quotients (see [13]).

Example 2

Consider monomial ideal \(I=\langle x_1x_2x_3, x_1x_2x_6, x_1x_2x_5, x_2x_4x_5, x_3x_4x_5,\)\( x_1x_3x_4, x_1x_3x_6, x_2x_4x_6, x_3x_5x_6, x_4x_5x_6\rangle \). This ideal has linear quotients and is not a variable-decomposable (see [9]).

4 Linear Quotients of Monomial Ideals

It is shown in [8] that from any path between \(u_i\) and \(u_j\) in \(G_I\), one can find \(w_i\) and \(w_j\) in Mon(S) such that \(w_i e_i-w_je_j \in \ker (\varphi )\).

Remark 5

If there exist two paths \(u_i, v, u_j\) and \(u_i, w, u_j\) in \(G_I\) between \(u_i\) and \(u_j\), then the coefficients of \(e_i\) which comes from the first and the second paths are equal.

Lemma 1

Let \(I=\langle u_1, \ldots , u_m\rangle \) be a squarefree monomial ideal which has linear relations. If \(\overline{G_I}\) is a chordal graph, then there exists a t such that the monomial ideal J with \(G(J)=G(I){\setminus } \{u_t\}\) has linear relations.

Proof

By [8, Lemma 1.8], \(G_I\) is a connected graph; thus, the existence of non-cut-vertex \(u_t\) follows from Proposition 1. Let \(J=\langle u_1, \ldots ,u_{t-1}, u_{t+1}, \ldots , u_m\rangle \) and \(u_i\) and \(u_j\) be two arbitrary elements in \(G(J)= G(I){\setminus }\{u_t\}\). We have

$$\begin{aligned} X_{F(u_j) {\setminus } F(u_i)}e_i-X_{F(u_i) {\setminus } F(u_j)}e_j\in \ker (F^{'}_1\longrightarrow J)\subset \ker (\varphi ). \end{aligned}$$

Since I has linear relations, there exists a path \(L_1\) between \(u_i\) and \(u_j\) in \(G_I\) such that \(X_{F(u_j) {\setminus } F(u_i)}e_i-X_{F(u_i) {\setminus } F(u_j)}e_j\) is a linear combination of linear forms which comes from \(L_1\). If \(L_1\) is a subgraph of \(G_J\), then we are done. Otherwise, there exists a path \(L_2\) of minimum length between \(u_i\) and \(u_j\) in \(G_J\). Since \( \overline{G_I}\) is chordal, it is easy to see that \(G_I\) has no cycle of length greater than 4 without chord, and so by Remark 5 the linear combinations of linear forms which come from \(L_2\) are equal to \(X_{F(u_j) {\setminus } F(u_i)}e_i-X_{F(u_i) {\setminus } F(u_j)}e_j\). \(\square \)

Theorem 3

Let \(I=<u_1, \ldots , u_m>\) be a squarefree monomial ideal. If I has linear relations and \(\overline{G_I}\) is a chordal graph, then I has linear quotients.

Proof

We proceed by induction on \(|G(I)|=m\). The assertion is trivial for \(m=1\) and \(m=2\). If \(m>2\), then by Lemma 1 there exists a t (assume \(t=m\)) such that \(J=\langle u_1, \ldots ,u_{m-1} \rangle \) has linear relations. Induction hypothesis implies that J has linear quotients. Assume \(u_1, \ldots , u_{m-1}\) is an admissible order for J. We show \(u_1, \ldots , u_{m-1}, u_m\) is an admissible order for I. Let \(j<m\), then by [8, Proposition 1.22] there exists a path \(u_m=v_1, v_2, \ldots , v_r=u_j\) in \(G_I\) such that \(F(v_i) \subseteq F(u_m) \cup F(u_j)\) for all \(1 \le i \le r\). If \(\{l\}=F(v_2) {\setminus } F(u_m)\), then \(l \in F(u_j) {\setminus } F(u_m)\). \(\square \)

As an immediate consequence of Theorem 3, we have:

Corollary 2

Let \(I=\langle u_1, \ldots , u_m \rangle \) be a squarefree monomial ideal. If \(\overline{G_I}\) is a chordal graph, then the following are equivalent:

  1. (a)

    I has linear relations;

  2. (b)

    I has a linear resolution;

  3. (c)

    I has linear quotients.

The following simple lemma is crucial for the proof of Theorem 4.

Lemma 2

Suppose I is a squarefree monomial ideal and \( \{u_{i},u_{r} \}\) and \( \{ u_{r},u_{j} \} \) are edges in \(G_{I}\). If \( \{u_{i},u_{j} \} \notin E(G_{I}) \) and \( F(u_{r}) {\setminus } F(u_{i}) =\{l\}\), then \( l \in F(u_{j}) {\setminus } F(u_{i}) \).

Proof

Since \( \{u_{i},u_{r} \} \in E(G_{I}) \), one has \( F(u_{r}) {\setminus } F(u_{i}) =\{l\}\). Assume \( F(u_{i})=A \cup \{t\} \) and \( F(u_{r})=A \cup \{l\}\). If \( l \notin F(u_{j}) \), then \( A \subseteq F(u_{j}) \). Hence, \( \{u_{i},u_{j} \} \in E(G_{I}) \), a contradiction. \(\square \)

Theorem 4

Assume that the ordering \( u_{m}, u_{m-1}, \ldots ,u_{1} \) is a perfect elimination ordering of \( \overline{G_{I}} \). Let \( u_{g_t}, u_{g_{t-1}},\ldots , u_{g_1} \) be the neighborhoods of \( u_{m}\) in \(\overline{G_I}\) and \(t>1\). Set \(V=\{u_{g_2-1},\ldots ,u_{g_1}, \ldots , u_{1} \} \). If \(N_{G^{'}}(u_{g_1}) \ne \emptyset \), where \(G^{'}\) is the induced subgraph of \(G_I\) on V, then I has linear quotients.

Proof

Let u be a neighborhood of \( u_{g_1} \). By Remark 2, u is a neighborhood of \( u_{g_i} \) for \( 2 \le i \le t\). We order the elements of G(I) as follows: \( u_{g_1}\), neighborhoods of \(u_{g_1}\), \(u_{g_2}\), neighborhoods of \(u_{g_2}\) which are not appeared before, ..., \( u_{g_t}\), neighborhoods of \(u_{g_t}\) which are not appeared before, \( u_{m} \), neighborhoods of \( u_{m} \) which are not appeared before.

Assume that \(v_{1}, \ldots , v_{m} \) come from the above ordering. In what follows, we show that this order is an admissible order. For \(j<i\), we must find \(k<i\) such that \(F(v_k){\setminus } F(v_i)=\{l\}\) and \(l\in F(v_j)\). Consider the following four cases:

  1. (i)

    \( v_{i}=u_{g_{s}} \) and \( v_{j}=u_{g_{l}}\) with \(l<s\). Since u is a neighborhood of \( u_{g_s} \) and \( u_{g_l} \), if we set \(v_k=u\), then we are done by Lemma 2.

  2. (ii)

    \( v_{i}=u_{g_{s}} \) and \( v_{j}\) is a neighborhood of \( u_{g_{l}} \) with \(l<s\). Remark 2 implies that \( v_{j}\) is a neighborhood of \( v_{i}\). Hence, take \(v_k=v_j\).

  3. (iii)

    \( v_{i}\) is a neighborhood of \(u_{g_{s}} \) and \( v_{j}\) is a neighborhood of \( u_{g_{l}} \) with \(l \le s\). Remark 2 implies that \( v_{j}\) is a neighborhood of \( u_{g_{s}}\). If \( \{v_{i},v_{j} \} \in E(G_{I}) \), then we take \(v_k=v_j\). Otherwise, we take \(v_k= u_{g_{s}}\) and the assertion follows from Lemma 2.

  4. (iv)

    \( v_{i}\) is a neighborhood of \(u_{g_{s}} \) and \( v_{j}= u_{g_{l}} \) with \(l \le s\). If \( l=s \), then take \(v_k=v_{j}\). So assume that \( l<s \). If \( \{v_i, u \} \in E(G_I) \), then take \(v_k=u\) and the assertion follows from Lemma 2. If \( \{v_i, u \} \notin E(G_I) \), then \( F(v_{i})= A \cup \{r_1, r_2\} \) and \( F(u)= A^{'} \cup \{r_3, r_4\} \). Remark 2 implies that \( v_{i}\) and u are neighborhoods of \( u_{m}\). It is easy to see that \( A=A^{'} \subseteq F(u_{m}) \). Hence, \( F(u_{m})= A\cup \{p,q\} \) such that \( p \in \{r_1, r_2\} \) and \( q \in \{r_3, r_4\}\). Without loss of generality, we may assume \( F(u_{m})= A \cup \{r_1, r_3\} \). Since \( v_{i}\) and u are neighborhoods of \(u_{g_{s}} \) and \( \{u_{m},u_{g_{s}} \} \notin E(G_{I}) \), one has \(F(u_{g_{s}})= A \cup \{r_2, r_4\} \). Take \( v_{k} =u_{g_{s}}\), then \( F(v_{k}){\setminus } F(v_{i}) =\{r_4\}\). If \( r_4 \notin F(v_{j}) \), since \( \lbrace u, v_{j} \rbrace \in E(G_{I}) \) , one has \( A \cup \lbrace r_3 \rbrace \subseteq F(v_{j})\). Therefore, \( \lbrace u_{m}, v_{j} \rbrace \in E(G_{I}) \), a contradiction.

\(\square \)