This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Invariants of Binomial Edge Ideals via Linear Programs

Adam LaClair Adam LaClair
Department of Mathematics
Purdue University
West Lafayette
IN 47907
USA
[email protected]
Abstract.

We associate to every graph a linear program for packings of vertex disjoint paths. We show that the optimal primal and dual values of the corresponding integer program are the binomial grade and height of the binomial edge ideal of the graph. We deduce from this a new combinatorial characterization of graphs of König type and use it to show that all trees are of König type.

The log canonical threshold and the F-threshold are important invariants associated to the singularities of a variety in characteristic 0 and characteristic pp. We show that the optimal value of the linear program (computed over the rationals) agrees with both the F-threshold and the log canonical threshold of the binomial edge ideal if the graph is a block graph or of König type. We conjecture that this linear program computes the log canonical threshold of the binomial edge ideal of any graph.

Our results resemble theorems on monomial ideals arising from hypergraphs due to Howald and others.

The author was partially supported by NSF grant DMS-2100288 and by Simons Foundation Collaboration Grant for Mathematicians #580839

1. Introduction

The theory of matchings of a graph has a long and rich history within graph theory with important applications in industry and also to other branches of mathematics. There are classical theorems relating the combinatorics of hypergraphs, the linear program realizing a maximal matching in a hypergraph, and algebraic invariants of the edge ideal associated to a hypergraph. For a hypergraph \mathcal{H} let MM,\operatorname{MM}_{\mathcal{H},\mathbb{Z}} (respectively MM,\operatorname{MM}_{\mathcal{H},\mathbb{Q}}) denote the optimal value of the linear program associated to \mathcal{H} realizing a maximal matching of \mathcal{H} over \mathbb{Z} (respectively \mathbb{Q}), and MM,\operatorname{MM}_{\mathcal{H},\mathbb{Z}}^{\ast} denote the optimal value of the dual linear program over \mathbb{Z}. For background on hypergraphs and the edge ideal of a hypergraph see [MV12]. For background on (fractional) maximal matchings of a hypergraph or the associated linear program MM,\operatorname{MM}_{\mathcal{H},-} see [LP86].

Theorem A (classical).

Let \mathcal{H} be a hypergraph, I()I(\mathcal{H}) the associated edge ideal, and MM\operatorname{MM} as above.

  1. (1)

    The monomial grade of I()I(\mathcal{H}) is equal to MM,\operatorname{MM}_{\mathcal{H},\mathbb{Z}}, and also to the size of a maximal matching of \mathcal{H}. (folklore)

  2. (2)

    The following quanitites are equal:

    • For any prime pp the F-threshold of I()I(\mathcal{H}),

    • The log canonical threshold of I()I(\mathcal{H}),

    • MM,\operatorname{MM}_{\mathcal{H},\mathbb{Q}},

    • The size of a maximal fractional matching of \mathcal{H}.

    This was proven by [How01], but see also [Her16], [Pag18].

  3. (3)

    The height of I()I(\mathcal{H}) is equal to MM,\operatorname{MM}^{\ast}_{\mathcal{H},\mathbb{Z}}, and also to the size of a minimal vertex cover of \mathcal{H} (see for instance [HH11, Lemma 9.1.4] and [LP86]).

Theorem A can be formulated for an arbitrary monomial ideal if one is willing to sacrifice combinatorial notions (see [How01], [Her16], [Pag18], [HH11, Lemma 9.1.4]).

Another classical theorem in matching theory is König’s theorem.

Theorem B (König).

Let GG be a bipartite graph, then the size of a maximal matching is equal to the size of a minimal vertex cover.

Finally, we recall results on the coordinates of the vertices of the fractional matching polytope.

Theorem C.

For a graph GG each coordinate of a vertex of the fractional matching polytope is either 0, 11, or 12\frac{1}{2}. (Balinski, see [LP86, Theorem 7.5.1]. If GG is a bipartite graph, then each coordinate of the vertex of the fractional matching polytope is either 0 or 11. (see [LP86, Theorem 7.1.2].

In this paper we introduce a new linear program associated to a graph. We prove analogues of the above statements relating the linear program, the combinatorics of the graph, and algebraic properties of the binomial edge ideal.

The class of binomial edge ideals was introduced in [HHH+10] and independently in [Oht11]. They are defined as

Definition 1.1 ([HHH+10]).

Let G=(V,E)G=(V,E) be a finite simple graph with vertex set V={1,,n}V=\{1,\ldots,n\} and edge set EE. Fix a field 𝕂\mathbb{K}. Consider the polynomial ring R:=𝕂[X1,,Xn,Y1,,Yn]R:=\mathbb{K}[X_{1},\ldots,X_{n},Y_{1},\ldots,Y_{n}], and for each edge {i,j}E\{i,j\}\in E with i<ji<j define fij:=XiYjXjYiRf_{ij}:=X_{i}Y_{j}-X_{j}Y_{i}\in R. Define the binomial edge ideal of GG, denoted JGJ_{G}, to be the ideal

(1) JG:=({fij{i,j}E}).J_{G}:=(\{f_{ij}\mid\{i,j\}\in E\}).

Numerous papers have studied algebraic invariants of a binomial edge ideal in terms of combinatorial properties of the underlying graph; see [SM18] for a survey.

In [TW04], the authors introduced the notion of F-pure threshold which is a characteristic pp-invariant having properties analogous to log canonical thresholds. In our situation the F-pure threshold is equivalent to the F-threshold introduced in [MTW05]. Throughout this paper pp will always denote a positive prime integer. When S=[X1,,Xn]S=\mathbb{Z}[X_{1},\ldots,X_{n}] and II is an SS-ideal, we denote by II_{\mathbb{C}}, respectively IpI_{p}, the image of II in SS\otimes_{\mathbb{Z}}\mathbb{C}, respectively S/pS\otimes_{\mathbb{Z}}\mathbb{Z}/p\mathbb{Z}. We denote by lct(I)\operatorname{lct}(I) the log canonical threshold of II_{\mathbb{C}} around the origin which is an important invariant appearing in birational geometry (see [Laz04]) and by ft(Ip)\operatorname{ft}(I_{p}) the F-threshold of IpI_{p}. Throughout this paper we will often write ft(JG)\operatorname{ft}(J_{G}) instead of ft(JG,p)\operatorname{ft}(J_{G,p}) since our computations will be independent of characteristic. It is known that:

Theorem 1.2 ([MTW05, Theorem 3.4]).

In the above setup,

(2) ft(Ip)lct(I)\operatorname{ft}(I_{p})\leq\operatorname{lct}(I)

for all p0p\gg 0 and

(3) lct(I)=limpft(Ip).\operatorname{lct}(I)=\lim_{p\rightarrow\infty}\operatorname{ft}(I_{p}).

In practice it is difficult to give explicit formulas for the log canonical threshold or F-threshold of an ideal, apart from special cases. Relevant for us are the following computations. The F-threshold and log canonical threshold of monomial ideals can be computed via a linear program (see [How01], [Her16], [Pag18]). Shibuta and Takagi [ST09] used this linear program to compute the log canonical threshold of certain classes of binomial ideals as the optimal value of a linear programming problem over \mathbb{Q}. Hernández [Her14] expands on their ideas to show how the optimal value of this linear program over \mathbb{Q} can be used to compute the F-threshold of a binomial hypersurface. In [MSV14], the authors compute the F-threshold of determinantal ideals. This computation was later generalized by Henriques and Varbaro [HV16] where they compute the multiplier and test ideals of GL-equivariant ideals, and as a consequence they obtain a formula for the F-threshold of GL-equivariant ideals. In [BE18], the authors give an algorithm to compute the log canonical threshold of any binomial ideal. Velez [Vel19] computes the F-threshold of the ideal of adjacent maximal minors which in the case of a 2×n2\times n matrix corresponds to the binomial edge ideal of a path. González-Martínez computes the F-threshold of the graded maximal ideal in the quotient ring R/JGR/J_{G} for any graph GG and uses this computation to characterize which binomial edge ideals are Gorenstein [GM21]. The main goal of this paper is to investigate the computation of the log canonical threshold and F-threshold of a binomial edge ideal in terms of combinatorial properties of the underlying graph. To this end we introduce the following linear programming problem which is a modification of the linear programming problem considered by Shibuta and Takagi (see [ST09]) and Hernández (see [Her14]).

For UV(G)U\subseteq V(G), define G[U]G[U] to be the induced graph on UU. Let 𝒜:={UV(G)|U|2}\mathcal{A}:=\{U\subseteq V(G)\mid\lvert U\rvert\geq 2\}. Consider the following linear program associated to the graph GG.

(4) maximizeZPP,G\displaystyle\operatorname*{maximize}\,Z_{\operatorname{PP},G} :=eE(G)XPP,e\displaystyle:=\phantom{a}\sum_{e\in E(G)}X_{\operatorname{PP},e}
subjectto:\displaystyle\operatorname{subject\,to:}
XPP,e\displaystyle X_{\operatorname{PP},e} 𝕜0 for all eE(G)\displaystyle\in\mathbbm{k}_{\geq 0}\text{ for all }e\in E(G)
CPP,U\displaystyle C_{\operatorname{PP},U} |U|1 for all U𝒜\displaystyle\leq\lvert U\rvert-1\text{ for all }U\in\mathcal{A}
CPP,v\displaystyle C_{\operatorname{PP},v} 2 for all vV(G)\displaystyle\leq 2\text{ for all }v\in V(G)

where for each U𝒜U\in\mathcal{A} and vVv\in V, CPP,UC_{\operatorname{PP},U} and CPP,vC_{\operatorname{PP},v} are the functions

CPP,U\displaystyle C_{\operatorname{PP},U} :=eE(G[U])XPP,e\displaystyle:=\phantom{}\sum_{e\in E(G[U])}X_{\operatorname{PP},e}
CPP,v\displaystyle C_{\operatorname{PP},v} :=v is incident with eXPP,e.\displaystyle:=\sum_{\begin{subarray}{c}v\text{ is incident }\\ \text{with }e\end{subarray}}X_{\operatorname{PP},e}.

Throughout this paper let 𝕜\mathbbm{k} denote a choice of ring \mathbb{Z}, \mathbb{Q}, or \mathbb{R}. Use (4) on different choices of 𝕜\mathbbm{k}, and denote the optimal value of (4) with respect to a choice of 𝕜\mathbbm{k} by PPG,𝕜\operatorname{PP}_{G,\mathbbm{k}}.

This paper studies PPG,𝕜\operatorname{PP}_{G,\mathbbm{k}} as it relates to the combinatorial properties of GG and the algebraic properties of JGJ_{G}. In Section 2 we introduce the relevant background. In Section 3 we show

Theorem D.

With the notation as above,

  1. (1)

    PPG,\operatorname{PP}_{G,\mathbb{Z}} equals the binomial grade of JGJ_{G} (which is equal to the maximal number of edges in a semi-path of GG) (Proposition 3.1 and Remark 3.2),

  2. (2)

    The optimal value of the dual program over \mathbb{Z}, PPG,\operatorname{PP}^{\ast}_{G,\mathbb{Z}}, equals htJG\operatorname{ht}J_{G} (Proposition 3.4),

  3. (3)

    ft(JG,p)PPG,\operatorname{ft}(J_{G,p})\leq\operatorname{PP}_{G,\mathbb{Q}} for any prime number pp (Proposition 3.12).

This establishes an analogue of Theorem A parts (1) and (3).

In Section 4 we prove the main result of this paper.

Theorem E (Theorem 4.14, Corollary 4.18).

Let GG be a block graph, then

ft(JG,p)=lct(JG)=PPG,[12]\displaystyle\operatorname{ft}(J_{G,p})=\operatorname{lct}(J_{G})=\operatorname{PP}_{G,\mathbb{Q}}\in\mathbb{Z}[\tfrac{1}{2}]

for any prime number pp.

This result establishes a special case of an analogue of Theorem A part (2).

Motivated by Theorem E, Corollary 5.2, and computational experiments we formulate

Conjecture 1.3.

Let GG be a graph, then

  1. (1)

    lct(JG)=PPG,\operatorname{lct}(J_{G})=\operatorname{PP}_{G,\mathbb{Q}}

  2. (2)

    PPG,[12]\operatorname{PP}_{G,\mathbb{Q}}\in\mathbb{Z}[\frac{1}{2}].

We prove this conjecture when GG is a block graph or GG is of König type. This conjecture suggests that a partial analogue of Theorem A item (2) may hold.

In Section 5 we give a new characterization of graphs of König type, Lemma 5.3. This enables us to show that every tree is of König type, Theorem 5.9. This establishes an analogue of Theorem B. In Section 6 we investigate for a connected graph GG the relationship between the equality PPG,=|V|1\operatorname{PP}_{G,\mathbb{Q}}=\lvert V\rvert-1 and GG being traceable.

In upcoming work [LaC], the author shows when the graph is a tree that the polytope determined by the constraints of Algorithm 4 (i.e. fractional path packing polytope) has integral vertices, i.e. each coordinate of such a vertex is either 0 or 11. This result gives a partial analogue of Theorem C. The author also investigates the question whether for a general graph each coordinate of a vertex of the fractional path packing is either 0, 11, or 12\frac{1}{2} [LaC].

Acknowledgments

The author found computations in Macaulay 2 [GS], Python [Pyt], Sage [S+22], and Online Linear Optimization Solver [Zwo] immensely helpful. The author would like to thank Alex Black, Jean-Philippe Labbe, and Vic Reiner for helpful conversations, Santiago Encinas for providing source code for the algorithm described in his paper [BE18], and Uli Walther for many helpful discussions and suggestions.

2. Background

2.1. Linear Programs

For further background on linear programs and for proofs not contained here see [Dan63]. A linear programming problem refers to the computation of the extremal value of a linear function over a convex polytope, i.e. any region defined by linear half-spaces. Any linear programming problem can be put into standard form which in this paper we take to be the following setup.

Let AA be an n×mn\times m matrix having entries in 𝕜\mathbbm{k}. Let 𝐜𝕜m\mathbf{c}\in\mathbbm{k}^{m} and 𝐛𝕜n\mathbf{b}\in\mathbbm{k}^{n}. Then, the linear programming problem in standard form refers to

(5) maximize𝐱𝕜m𝐜T𝐱subjectto:A𝐱𝐛𝐱0\begin{split}\operatorname*{maximize}\limits_{\mathbf{x}\in\mathbbm{k}^{m}}\quad&\mathbf{c}^{T}\mathbf{x}\\ \operatorname{subject\,to:}\hskip 6.99997pt&\phantom{}\\ &A\mathbf{x}\leq\mathbf{b}\\ &\phantom{A}\mathbf{x}\geq 0\end{split}

A linear programming problem written in this form will be called primal, and it can be shown that every linear programming problem is equivalent to a primal linear programming problem.

Given a primal problem we can associate the following linear programming problem

(6) minimize𝐲𝕜n𝐛T𝐲subjectto:AT𝐲𝐜𝐲0\begin{split}\operatorname*{minimize}_{\mathbf{y}\in\mathbbm{k}^{n}}\quad&\mathbf{b}^{T}\mathbf{y}\\ \operatorname{subject\,to:}\hskip 5.0pt&\phantom{}\\ &A^{T}\mathbf{y}\geq\mathbf{c}\\ &\phantom{A^{T}}\mathbf{y}\geq 0\end{split}

which is called the dual linear programming problem of the primal.

Given a linear programming problem, a vector is called feasible if it satisfies the constraints of the problem. If such a vector exists the linear programming problem is called feasible. A linear programming problem is called bounded if the optimal value of the linear programming problem is finite. When a linear programming problem is bounded feasible we call the value attained by the linear program to be the optimal value, and we call a vector which realizes the optimal value and satisfies the constraints of the linear programming problem an optimal solution.

Theorem 2.1 ([Dan63, p.125 Duality Theorem] Strong duality of linear programs).

Given a primal (resp. dual) feasible, bounded linear programming problem, then the dual (resp. primal) programming problem is feasible, bounded and the optimal values of the primal (resp. dual) agrees with the optimal value of the dual (resp. primal) provided that 𝕜\mathbb{Q}\subseteq\mathbbm{k}\subseteq\mathbb{R}.

Remark 2.2.

When 𝕜=\mathbbm{k}=\mathbb{Z}, then optimal value of a bounded feasible primal can be strictly smaller than the optimal value of the dual.

In the above setup of the Equation (5), let AiA_{i} denote the ii-th column of AA and ajTa_{j}^{T} denote the jj-th row of AA.

Theorem 2.3 ( [Dan63, p.136 Theorem 4] Complementary Slackness).

Let 𝐱\mathbf{x} and 𝐲\mathbf{y} be primal (respectively dual) optimal solutions of (5) (respectively (6)) and suppose that 𝐜T𝐱=𝐛T𝐲\mathbf{c}^{T}\mathbf{x}=\mathbf{b}^{T}\mathbf{y}, then

  1. (1)

    for 1im1\leq i\leq m, (𝐜i𝐲TAi)xi=0(\mathbf{c}_{i}-\mathbf{y}^{T}A_{i})x_{i}=0.

  2. (2)

    for 1jn1\leq j\leq n, (ajT𝐱𝐛j)yj=0(a_{j}^{T}\mathbf{x}-\mathbf{b}_{j})y_{j}=0.

Sketch of Proof.

It is always true that

𝐜T𝐱𝐲TA𝐱𝐲T𝐛.\displaystyle\mathbf{c}^{T}\mathbf{x}\leq\mathbf{y}^{T}A\mathbf{x}\leq\mathbf{y}^{T}\mathbf{b}.

Equality of the outer two terms implies equality throughout and complementary slackness follows by distributing terms appropriately.

2.2. Binomial Edge Ideals

Throughout this paper we will only consider finite simple graphs, i.e. graphs without multiple edges or loops and on a finite number of vertices. Given a graph GG we denote its set of vertices by V(G)V(G), or VV when the context is clear, and denote its set of edges by E(G)E(G), or EE when the context is clear. Throughout this paper we will assume that V(G)={1,,n}V(G)=\{1,\ldots,n\} for some positive integer nn. Given a subset UV(G)U\subseteq V(G), we denote by G[U]G[U] the subgraph of GG induced on the vertices UU.

By a path in GG we mean an alternating sequence of distinct vertices and edges of GG, v1,e1,v2,,e1,vv_{1},e_{1},v_{2},\ldots,e_{\ell-1},v_{\ell} so that ei={vi,vi+1}E(G)e_{i}=\{v_{i},v_{i+1}\}\in E(G) for 1i11\leq i\leq\ell-1. By a cycle in GG we mean an alternating sequence of distinct vertices and edges of GG, v1,e1,v2,,e1,v,ev_{1},e_{1},v_{2},\ldots,e_{\ell-1},v_{\ell},e_{\ell} so that ei={vi,vi+1}E(G)e_{i}=\{v_{i},v_{i+1}\}\in E(G) for 1i11\leq i\leq\ell-1 and e={v1,v}E(G)e_{\ell}=\{v_{1},v_{\ell}\}\in E(G). By a semi-path we mean a subgraph PP of GG such that each connected component of GG is a path. Let G^\hat{G} be the induced graph on V(G)(G)V(G)\smallsetminus\mathcal{I}(G) where (G)\mathcal{I}(G) denotes the isolated vertices of GG.

The following proposition gives a combinatorial description of the minimal primes of a binomial edge ideal and a formula for the height of prime ideals.

Proposition 2.4 ([HHH+10, Lemma 3.1, Corollary 3.9]).

Let GG be a graph on nn vertices. If T[n]:={1,,n}T\subseteq[n]:=\{1,\ldots,n\}, we will denote the induced subgraph on the vertices GTG\smallsetminus T by GTG_{T}. Let cG(T)c_{G}(T) denote the number of connected components of GTG_{T}, and let G1,,GcG(T)G_{1},\ldots,G_{c_{G}(T)} denote the distinct connected components of GTG_{T}. Let G~i\tilde{G}_{i} denote the complete graph on the vertices V(Gi)V(G_{i}). Put

PT(G):=(iT{Xi,Yi},JG~1,,JG~cG(T))\displaystyle P_{T}(G):=\big{(}\bigcup_{i\in T}\{X_{i},Y_{i}\},J_{\tilde{G}_{1}},\ldots,J_{\tilde{G}_{c_{G}(T)}}\big{)}

where JG~J_{\tilde{G}_{\ast}} is as defined in Definition 1.1. Then, PT(G)P_{T}(G) is a minimal prime ideal of JGJ_{G} and

htPT(G)=2|T|+i=1cG(T)(|V(Gi)|1)=ncG(T)+|T|.\displaystyle\operatorname{ht}P_{T}(G)=2\lvert T\rvert+\sum_{i=1}^{c_{G}(T)}\left(\lvert V(G_{i})\rvert-1\right)=n-c_{G}(T)+\lvert T\rvert.

Moreover, every minimal prime ideal of JGJ_{G} is of the form PT(G)P_{T}(G) where T=T=\varnothing, or TT\neq\varnothing and for each iTi\in T one has that cG(Ti)<cG(T)c_{G}(T\smallsetminus i)<c_{G}(T).

Definition 2.5.

For a graph GG define the invariant

r(G):=sup{|E(P)||PG is a semi-path}.\displaystyle r(G):=\sup\left\{\lvert E(P)\rvert\mathrel{}\middle|\mathrel{}P\subseteq G\text{ is a semi-path}\right\}.

We call a disjoint union of paths realizing r(G)r(G) a path packing of GG.

Definition 2.6.

A graph GG is called traceable if GG has a Hamiltonian path.

Remark 2.7.

GG is traceable if and only if r(G)=|V|1r(G)=\lvert V\rvert-1.

2.3. F-Thresholds

Definition 2.8 ([MTW05, Lemma 1.1, Remark 1.5]).

Let S=𝕂[X1,,Xn]S=\mathbb{K}[X_{1},\ldots,X_{n}] be a polynomial ring, char(𝕂)=p>0\operatorname{char}(\mathbb{K})=p>0, 𝔪=(X1,,Xn)\mathfrak{m}=(X_{1},\ldots,X_{n}) the homogeneous maximal ideal, and II an SS-ideal with I𝔪I\subseteq\mathfrak{m}. For ee\in\mathbb{N}, set

νe(I):=max{rIr𝔪[pe]}\displaystyle\nu_{e}(I):=\max\{r\mid I^{r}\not\subseteq\mathfrak{m}^{[p^{e}]}\}

where 𝔪[pe]:=(X1pe,,Xnpe)\mathfrak{m}^{[p^{e}]}:=(X_{1}^{p^{e}},\ldots,X_{n}^{p^{e}}). Then,

ft(I):=limeνe(I)pe\displaystyle\operatorname{ft}(I):=\lim_{e\rightarrow\infty}\frac{\nu_{e}(I)}{p^{e}}

exists and is called the F-threshold of II.

We review three lemmas well-known to experts. The first lemma says that the computation of F-threshold is additive over ideals in disjoint variables and enables us to reduce the computation of the F-threshold of a binomial edge ideal to the computation of the F-threshold of the binomial edge ideal of its connected components.

Lemma 2.9.

Let char𝕂=p>0\operatorname{char}\mathbb{K}=p>0, IS:=𝕂[X1,,Xn]I\subseteq S:=\mathbb{K}[X_{1},\ldots,X_{n}] with 𝔪\mathfrak{m} its homogeneous maximal ideal, JS:=𝕂[Y1,,Ym]J\subseteq S^{{}^{\prime}}:=\mathbb{K}[Y_{1},\ldots,Y_{m}] with 𝔫\mathfrak{n} its homogeneous maximal ideal, T:=S𝕂ST:=S\otimes_{\mathbb{K}}S^{{}^{\prime}} with 𝔬\mathfrak{o} its homogeneous maximal ideal. Then,

ftT(IT+JT)=ftS(I)+ftS(J).\displaystyle\operatorname{ft}_{T}(I\cdot T+J\cdot T)=\operatorname{ft}_{S}(I)+\operatorname{ft}_{S^{{}^{\prime}}}(J).

In particular, if GG is a graph and G1,,GcG_{1},\ldots,G_{c} are its connected components, then

ft(JG)=i=1cft(JGi).\displaystyle\operatorname{ft}(J_{G})=\sum_{i=1}^{c}\operatorname{ft}(J_{G_{i}}).

The next lemma says that the F-threshold is monotonic along ideal inclusion which allows us to obtain lower bounds on a graph via its subgraphs.

Lemma 2.10.

Let IJ𝔪I\subseteq J\subseteq\mathfrak{m} be ideals in 𝕂[X1,,Xn]\mathbb{K}[X_{1},\ldots,X_{n}]. Then,

ft(I)ft(J).\displaystyle\operatorname{ft}(I)\leq\operatorname{ft}(J).

In particular, if HH is a subgraph of GG, then

ft(JH)ft(JG).\displaystyle\operatorname{ft}(J_{H})\leq\operatorname{ft}(J_{G}).
Proof.

It follows from the definition that νe(I)νe(J)\nu_{e}(I)\leq\nu_{e}(J) for all ee\in\mathbb{N}. ∎

The final lemma says that the F-threshold can only decrease under taking initial ideals.

Lemma 2.11 ([TW04, Proposition 4.5]).

Let I𝔪I\subseteq\mathfrak{m} be ideal in 𝕂[X1,,Xn]\mathbb{K}[X_{1},\ldots,X_{n}] and << a term order on the polynomial ring. Then, ft(in<(I))ft(I)\operatorname{ft}(\operatorname{in}_{<}(I))\leq\operatorname{ft}(I).

Miller, Singh, and Varbaro [MSV14] computed the F-threshold of a determinantal ideal of a generic matrix as follows.

Theorem 2.12 ([MSV14, Theorem 1.2]).

Fix positive integers tmnt\leq m\leq n, and let XX be an m×nm\times n matrix of indeterminates over a field 𝔽\mathbb{F} of prime characteristic. Let SS be the polynomial ring 𝔽[X]\mathbb{F}[X], and ItI_{t} the ideal generated by the size tt minors of XX.

The F-threshold of ItI_{t} is

min{(mk)(nk)tkk=0,,t1}.\displaystyle\min\bigg{\{}\frac{(m-k)(n-k)}{t-k}\mid k=0,\ldots,t-1\bigg{\}}.
Remark 2.13.

In the above theorem, when t=m=2nt=m=2\leq n, then It(X)I_{t}(X) corresponds to JKnJ_{K_{n}} where KnK_{n} denotes the complete graph on nn vertices, and thus

ft(JKn)=n1.\displaystyle\operatorname{ft}(J_{K_{n}})=n-1.

In the master’s thesis of Velez [Vel19], he proves

Proposition 2.14 ([Vel19, Theorem 3.1.6]).

Let MM be a generic m×nm\times n matrix with mnm\leq n. Let δi\delta_{i} denote the determinant of the submatrix given by the columns {i,i+1,,i+m1}\{i,i+1,\ldots,i+m-1\} for 1inm+11\leq i\leq n-m+1. Let J:=(δi1inm+1)J:=(\delta_{i}\mid 1\leq i\leq n-m+1). Then, ft(J)=nm+1\operatorname{ft}(J)=n-m+1.

Sketch of Proof.

By Lemma 2.11 and Theorem 2.12,

nm+1=ft(in<(JG))ft(JG)nm+1.\displaystyle n-m+1=\operatorname{ft}(\operatorname{in}_{<}(J_{G}))\leq\operatorname{ft}(J_{G})\leq n-m+1.
Remark 2.15.

When t=m=2t=m=2 in Proposition 2.14 the ideal JGJ_{G} corresponds the binomial edge ideal of a path on nn vertices. I.e., if G=PnG=P_{n} is a path on nn vertices, then ft(JG)=n1\operatorname{ft}(J_{G})=n-1.

Theorem 2.16 ([Luc78] Lucas’s theorem).

Let pp be a prime integer and mm and nn positive integers. Choose a positive integer ee so that we can write m=i=0emipim=\sum_{i=0}^{e}m_{i}p^{i} and n=i=0enipin=\sum_{i=0}^{e}n_{i}p^{i} where 0mi,ni<p0\leq m_{i},n_{i}<p for 0ie0\leq i\leq e . Then,

(nm)i=1e(nimi)modp\displaystyle\binom{n}{m}\equiv\prod_{i=1}^{e}\binom{n_{i}}{m_{i}}\mod{p}

with the convention that (ab)=0\binom{a}{b}=0 if a<ba<b.

3. Linear Programs and Binomial Edge Ideals

3.1. Binomial Grade, Height, and Linear Programs

Recall that r(G)r(G) is the maximal length of a path packing (Definition 2.5), and PPG,\operatorname{PP}_{G,\mathbb{Z}} is the optimal value of an integer program (discussion surrounding (4)).

Proposition 3.1.

Given a graph GG, r(G)=PPG,r(G)=\operatorname{PP}_{G,\mathbb{Z}}.

Proof.

For vVv\in V, the constraint CPP,v2C_{\operatorname{PP},v}\leq 2 forces that a feasible solution has at most two edges incident to vv. For U𝒜U\in\mathcal{A}, the constraint CPP,U|U|1C_{\operatorname{PP},U}\leq\lvert U\rvert-1 forces that a feasible solution does not induce a Hamiltonian cycle on G[U]G[U]. Thus, an optimal solution to (4) consists of disjoint paths. Conversely, a path packing is a feasible solution of (4). ∎

Remark 3.2.

The feasible solutions of (4) are precisely the semi-paths of GG. Moreover, in [HHM22, Lemma 3.2, Lemma 3.3] the authors show that r(G)r(G) is equal to the binomial grade of JGJ_{G}, i.e. the length of a longest regular sequence in JGJ_{G} consisting of binomials which form part of a minimal generating set of JGJ_{G}.

The dual linear program of (4) solved over 𝕜\mathbbm{k} is given by

(7) minimizeZPP,G:=vV(G)2WPP,v+U𝒜(|U|1)WPP,Usubjectto:WPP,v𝕜0 for all vV(G)WPP,U𝕜0 for all U𝒜DPP,e:=WPP,i+WPP,j+U𝒜eG[U]WPP,U1 for all e={i,j}E(G)\begin{split}\operatorname*{minimize}\quad&Z_{\operatorname{PP^{\ast}},G}:=\sum_{v\in V(G)}2\cdot W_{\operatorname{PP^{\ast}},v}+\sum_{U\in\mathcal{A}}(\lvert U\rvert-1)\cdot W_{\operatorname{PP^{\ast}},U}\\ \operatorname{subject\,to:}\hskip 5.0pt&\phantom{}\\ &W_{\operatorname{PP^{\ast}},v}\in\mathbbm{k}_{\geq 0}\text{ for all }v\in V(G)\\ &W_{\operatorname{PP^{\ast}},U}\in\mathbbm{k}_{\geq 0}\text{ for all }U\in\mathcal{A}\\ &D_{\operatorname{PP^{\ast}},e}:=W_{\operatorname{PP^{\ast}},i}+W_{\operatorname{PP^{\ast}},j}+\sum_{\begin{subarray}{c}U\in\mathcal{A}\\ e\in G[U]\end{subarray}}W_{\operatorname{PP^{\ast}},U}\geq 1\text{ for all }e=\{i,j\}\in E(G)\end{split}

Denote the optimal value of (7) over 𝕜\mathbbm{k} by PPG,𝕜\operatorname{PP}_{G,\mathbbm{k}}^{\ast}.

From a combinatorial perspective, the integral linear program (7) computes a minimal weighted edge covering of the graph by stars and induced subgraphs. From a commutative algebra perspective, (7) computes the height of JGJ_{G}.

Lemma 3.3.

Fix a graph GG and set 𝕜=\mathbbm{k}=\mathbb{Z}. There exists an optimal solution of the linear program (7) of the form ({bv}vV,{cU}U𝒜)(\{b_{v}\}_{v\in V},\{c_{U}\}_{U\in\mathcal{A}}) satisfying

  1. (1)

    bv{0,1}b_{v}\in\{0,1\} for all vVv\in V and cU{0,1}c_{U}\in\{0,1\} for all U𝒜U\in\mathcal{A},

  2. (2)

    if UU and UU^{{}^{\prime}} are elements of 𝒜\mathcal{A} with cU=cU=1c_{U}=c_{U^{{}^{\prime}}}=1, then UU=U\cap U^{{}^{\prime}}=\varnothing,

  3. (3)

    if bv=1b_{v}=1 for some vV(G)v\in V(G), then vUv\notin U for all cU=1c_{U}=1,

  4. (4)

    if cU=1c_{U}=1 for some U𝒜U\in\mathcal{A}, then UU is connected.

For such a solution, let T:={vVcv=1}T:=\{v\in V\mid c_{v}=1\} and U1,,UcU_{1},\ldots,U_{c} to be the elements of 𝒜\mathcal{A} which satisfy cUi=1c_{U_{i}}=1. Then, the UiU_{i} are the connected components of GT^\widehat{G\smallsetminus T}.

Proof.

Let ({bv}vV,{cU}U𝒜)(\{b_{v}\}_{v\in V},\{c_{U}\}_{U\in\mathcal{A}}) be any optimal solution to (7).

  1. (1)

    We may assume that bv,cU{0,1}b_{v},c_{U}\in\{0,1\} for all vVv\in V and U𝒜U\in\mathcal{A}. Otherwise, for each non-zero bvb_{v} or non-zero cUc_{U} reassign it the value 11.

  2. (2)

    We may suppose that if UU and UU^{{}^{\prime}} are elements of 𝒜\mathcal{A} with cU=cU=1c_{U}=c_{U^{{}^{\prime}}}=1, then UU=U\cap U^{{}^{\prime}}=\varnothing. Otherwise, redefine cU=cU=0c_{U}=c_{U^{{}^{\prime}}}=0 and cUU=1c_{U\cup U^{{}^{\prime}}}=1.

  3. (3)

    We may suppose that if bv=1b_{v}=1 for some vV(G)v\in V(G), then vUv\notin U for all cU=1c_{U}=1. Otherwise, if |U|>2\lvert U\rvert>2 put cU=0c_{U}=0 and cU{v}=1c_{U\smallsetminus\{v\}}=1, and if |U|=2\lvert U\rvert=2 put cU=0c_{U}=0.

  4. (4)

    We may suppose that if cU=1c_{U}=1 for some U𝒜U\in\mathcal{A}, then UU is connected. Otherwise, write U=U1U2U=U_{1}\cup U_{2} with U1U2=U_{1}\cup U_{2}=\varnothing. Put cU=0c_{U}=0 and cUi=1c_{U_{i}}=1 if Ui𝒜U_{i}\in\mathcal{A} for i=1,2i=1,2.

After each application of one of the above steps this defines a new feasible solution of (7) which does not increase the cost function.

After these reductions, define T:={vVav=1}T:=\{v\in V\mid a_{v}=1\} and U1,,UcU_{1},\ldots,U_{c} to be elements of 𝒜\mathcal{A} which satisfy cUi=1c_{U_{i}}=1. The above reductions together with the constraints DPP,e1D_{\operatorname{PP^{\ast}},e}\geq 1 for every eE(G)e\in E(G) implies that the UiU_{i} are the connected components of GT^\widehat{G\smallsetminus T}. ∎

Proposition 3.4.

Let GG be a graph, then PPG,=htJG\operatorname{PP}_{G,\mathbb{Z}}^{\ast}=\operatorname{ht}J_{G}.

Proof.

We first show PPG,htJG\operatorname{PP}_{G,\mathbb{Z}}^{\ast}\leq\operatorname{ht}J_{G}. Let TV(G)T\subseteq V(G) be chosen so that htJG=htPT(G)=2|T|+i=1cT(G)(|V(Gi)|1)\operatorname{ht}J_{G}=\operatorname{ht}P_{T}(G)=2\lvert T\rvert+\sum_{i=1}^{c_{T}(G)}\big{(}\lvert V(G_{i})\rvert-1\big{)} where the notation is as in Proposition 2.4. For vVv\in V let

bv:={1,vT0,else\displaystyle b_{v}:=\begin{cases}1,&v\in T\\ 0,&\text{else}\end{cases}

and for U𝒜U\in\mathcal{A} let

cU={1,U=V(Gi) for some 1icT(G)0,else.\displaystyle c_{U}=\begin{cases}1,&U=V(G_{i})\text{ for some }1\leq i\leq c_{T}(G)\\ 0,&\text{else}\end{cases}.

Then the tuple ({bv}vV,{cU}U𝒜)\big{(}\{b_{v}\}_{v\in V},\{c_{U}\}_{U\in\mathcal{A}}\big{)} is a feasible solution to (7), and moreover

PPG,ZPP,G({bv}vV,{cU}U𝒜)=2|T|+i=1cT(G)(|V(Gi)|1)=htJG.\displaystyle\operatorname{PP}_{G,\mathbb{Z}}^{\ast}\leq Z_{\operatorname{PP^{\ast}},G}\big{(}\{b_{v}\}_{v\in V},\{c_{U}\}_{U\in\mathcal{A}}\big{)}=2\lvert T\rvert+\sum_{i=1}^{c_{T}(G)}\big{(}\lvert V(G_{i})\rvert-1\big{)}=\operatorname{ht}J_{G}.

Next, we show that htJGPPG,\operatorname{ht}J_{G}\leq\operatorname{PP}_{G,\mathbb{Z}}^{\ast}. Let ({bv}vV,{cU}U𝒜)\big{(}\{b_{v}\}_{v\in V},\{c_{U}\}_{U\in\mathcal{A}}\big{)} be an optimal solution of (7) when 𝕜=\mathbbm{k}=\mathbb{Z} of the form guaranteed by Lemma 3.3. Let TT and U1,,UcU_{1},\ldots,U_{c} be as constructed by Lemma 3.3. The cost of this optimal tuple is precisely 2|T|+i=1c(|Ui|1)2\lvert T\rvert+\sum_{i=1}^{c}(\lvert U_{i}\rvert-1) which is equal to htPT(G)\operatorname{ht}P_{T}(G). This proves that

htJGhtPT(G)=PPG,\displaystyle\operatorname{ht}J_{G}\leq\operatorname{ht}P_{T}(G)=\operatorname{PP}_{G,\mathbb{Z}}^{\ast}

which completes the proof. ∎

3.2. Thresholds and Linear Programs

Next, we study the LP relaxation of the above linear programs, i.e. we allow the variables to take values in 0\mathbb{R}_{\geq 0}, and how the optimal solution relates to commutative algebra invariants.

For a graph GG denote by LPG,𝕜\operatorname{LP}_{G,\mathbbm{k}} the optimal value of the following linear program over (𝕜0)2|E|(\mathbbm{k}_{\geq 0})^{2\lvert E\rvert}.

(8) maximizeZLP,G:=eE(G)(XLP,e+YLP,e)subjectto:XLP,e𝕜0 for all eE(G)YLP,e𝕜0 for all eE(G)CLP,U:=eE(G[U])(XLP,e+YLP,e)|U|1 for all U𝒜CLP,v,x:=e={v,w}v<wXLP,e+e={w,v}w<vYLP,e1 for all vV(G)CLP,v,y:=e={v,w}v<wYLP,e+e={w,v}w<vXLP,e1 for all vV(G)\begin{split}\operatorname*{maximize}\quad&Z_{\operatorname{LP},G}:=\sum_{e\in E(G)}(X_{\operatorname{LP},e}+Y_{\operatorname{LP},e})\\ \operatorname{subject\,to:}\hskip 6.99997pt&\phantom{}\\ &X_{\operatorname{LP},e}\in\mathbbm{k}_{\geq 0}\text{ for all }e\in E(G)\\ &Y_{\operatorname{LP},e}\in\mathbbm{k}_{\geq 0}\text{ for all }e\in E(G)\\ &C_{\operatorname{LP},U}:=\sum_{e\in E(G[U])}(X_{\operatorname{LP},e}+Y_{\operatorname{LP},e})\leq\lvert U\rvert-1\text{ for all }U\in\mathcal{A}\\ &C_{\operatorname{LP},v,x}:=\sum_{\begin{subarray}{c}e=\{v,w\}\\ v<w\end{subarray}}X_{\operatorname{LP},e}+\sum_{\begin{subarray}{c}e=\{w,v\}\\ w<v\end{subarray}}Y_{\operatorname{LP},e}\leq 1\text{ for all }v\in V(G)\\ &C_{\operatorname{LP},v,y}:=\sum_{\begin{subarray}{c}e=\{v,w\}\\ v<w\end{subarray}}Y_{\operatorname{LP},e}+\sum_{\begin{subarray}{c}e=\{w,v\}\\ w<v\end{subarray}}X_{\operatorname{LP},e}\leq 1\text{ for all }v\in V(G)\\ \end{split}
Remark 3.5.

Let GG be a graph and G1,,GcG_{1},\ldots,G_{c} the connected components of GG. Then,

LPG,=i=1cLPGi,.\displaystyle\operatorname{LP}_{G,\mathbb{Q}}=\sum_{i=1}^{c}\operatorname{LP}_{G_{i},\mathbb{Q}}.

The following proposition says that the constraint CLP,U|U|1C_{\operatorname{LP},U}\leq\lvert U\rvert-1 in (8) can be strengthened by requiring that in addition UU is biconnected. Recall that a connected graph GG is biconnected if G{v}G\smallsetminus\{v\} is connected for every vV(G)v\in V(G). In the following linear program we differentiate the intdeterminates and constraints from those appearing in (8) by utilizing the superscript “bi\operatorname{bi}”.

Proposition 3.6.

The set of feasible solutions of (8) is the same as the set of feasible solutions of the following linear program

(9) maximizeZLP,Gbi:=eE(G)(XLP,ebi+YLP,ebi)subjectto:XLP,ebi𝕜0 for all eE(G)YLP,ebi𝕜0 for all eE(G)CLP,Ubi:=eE(G[U])(XLP,ebi+YLP,ebi)|U|1 for all U𝒜,G[U] biconnectedCLP,v,xbi:=e={v,w}v<wXLP,ebi+e={w,v}w<vYLP,ebi1 for all vV(G)CLP,v,ybi:=e={v,w}v<wYLP,ebi+e={w,v}w<vXLP,ebi1 for all vV(G)\begin{split}\operatorname*{maximize}\quad&Z_{\operatorname{LP},G}^{\operatorname{bi}}:=\sum_{e\in E(G)}(X_{\operatorname{LP},e}^{\operatorname{bi}}+Y_{\operatorname{LP},e}^{\operatorname{bi}})\\ \operatorname{subject\,to:}\hskip 6.99997pt&\phantom{}\\ &X_{\operatorname{LP},e}^{\operatorname{bi}}\in\mathbbm{k}_{\geq 0}\text{ for all }e\in E(G)\\ &Y_{\operatorname{LP},e}^{\operatorname{bi}}\in\mathbbm{k}_{\geq 0}\text{ for all }e\in E(G)\\ &C_{\operatorname{LP},U}^{\operatorname{bi}}:=\sum_{e\in E(G[U])}(X_{\operatorname{LP},e}^{\operatorname{bi}}+Y_{\operatorname{LP},e}^{\operatorname{bi}})\leq\lvert U\rvert-1\text{ for all }U\in\mathcal{A},\phantom{}G[U]\text{ biconnected}\\ &C_{\operatorname{LP},v,x}^{\operatorname{bi}}:=\sum_{\begin{subarray}{c}e=\{v,w\}\\ v<w\end{subarray}}X_{\operatorname{LP},e}^{\operatorname{bi}}+\sum_{\begin{subarray}{c}e=\{w,v\}\\ w<v\end{subarray}}Y_{\operatorname{LP},e}^{\operatorname{bi}}\leq 1\text{ for all }v\in V(G)\\ &C_{\operatorname{LP},v,y}^{\operatorname{bi}}:=\sum_{\begin{subarray}{c}e=\{v,w\}\\ v<w\end{subarray}}Y_{\operatorname{LP},e}^{\operatorname{bi}}+\sum_{\begin{subarray}{c}e=\{w,v\}\\ w<v\end{subarray}}X_{\operatorname{LP},e}^{\operatorname{bi}}\leq 1\text{ for all }v\in V(G)\\ \end{split}

where v<wv<w means that the label on the vertex vv is smaller than the label on the vertex ww.

Proof.

It suffices to show that feasible solutions of (9) are feasible solutions of (8). It suffices to restrict to the case of U𝒜U\in\mathcal{A} is connected. Write U=U1UcU=U_{1}\cup\cdots\cup U_{c} as a union of its maximal biconnected components. Let 𝐚\mathbf{{a}} denote a feasible solution of (8). Observe that

CU(𝐚)=i=1cCUi(𝐚)=i=1cCUibi(𝐚)i=1c(|Ui|1).\displaystyle C_{U}(\mathbf{{a}})=\sum_{i=1}^{c}C_{U_{i}}(\mathbf{{a}})=\sum_{i=1}^{c}C_{U_{i}}^{\operatorname{bi}}(\mathbf{{a}})\leq\sum_{i=1}^{c}(\lvert U_{i}\rvert-1).

The above assumptions on UU together with induction on cc shows that

i=1c(|Ui|1)|U|1.\displaystyle\sum_{i=1}^{c}(\lvert U_{i}\rvert-1)\leq\lvert U\rvert-1.

Remark 3.7.

Linear program (9) will have O(2n)O(2^{n}) constraints for a complete graph on nn vertices.

Example 3.8.

Let GG be the graph depicted in Figure 1.

112244663355acfedb
Figure 1. Net

We label the edges of the graph with the letters a through f and the vertices with the numbers 11 through 66 as in Figure 1. Then, the constraints in (9) are

CLP,{1,2}bi\displaystyle C_{\operatorname{LP},\{1,2\}}^{\operatorname{bi}} =Xabi+Yabi1\displaystyle=X_{a}^{\operatorname{bi}}+Y_{a}^{\operatorname{bi}}\leq 1 CLP,4,xbi\displaystyle C_{\operatorname{LP},4,x}^{\operatorname{bi}} =Xfbi+Ycbi+Ydbi1\displaystyle=X_{f}^{\operatorname{bi}}+Y_{c}^{\operatorname{bi}}+Y_{d}^{\operatorname{bi}}\leq 1 CLP,1,xbi\displaystyle C_{\operatorname{LP},1,x}^{\operatorname{bi}} =Xabi1\displaystyle=X_{a}^{\operatorname{bi}}\leq 1
CLP,{2,3}bi\displaystyle C_{\operatorname{LP},\{2,3\}}^{\operatorname{bi}} =Xbbi+Ybbi1\displaystyle=X_{b}^{\operatorname{bi}}+Y_{b}^{\operatorname{bi}}\leq 1 CLP,4,ybi\displaystyle C_{\operatorname{LP},4,y}^{\operatorname{bi}} =Yfbi+Xcbi+Xdbi1\displaystyle=Y_{f}^{\operatorname{bi}}+X_{c}^{\operatorname{bi}}+X_{d}^{\operatorname{bi}}\leq 1 CLP,1,ybi\displaystyle C_{\operatorname{LP},1,y}^{\operatorname{bi}} =Yabi1\displaystyle=Y_{a}^{\operatorname{bi}}\leq 1
CLP,{2,4}bi\displaystyle C_{\operatorname{LP},\{2,4\}}^{\operatorname{bi}} =Xcbi+Ycbi1\displaystyle=X_{c}^{\operatorname{bi}}+Y_{c}^{\operatorname{bi}}\leq 1 CLP,2,xbi\displaystyle C_{\operatorname{LP},2,x}^{\operatorname{bi}} =Yabi+Xbbi+Xcbi1\displaystyle=Y_{a}^{\operatorname{bi}}+X_{b}^{\operatorname{bi}}+X_{c}^{\operatorname{bi}}\leq 1 CLP,5,xbi\displaystyle C_{\operatorname{LP},5,x}^{\operatorname{bi}} =Yebi1\displaystyle=Y_{e}^{\operatorname{bi}}\leq 1
CLP,{3,4}bi\displaystyle C_{\operatorname{LP},\{3,4\}}^{\operatorname{bi}} =Xdbi+Ydbi1\displaystyle=X_{d}^{\operatorname{bi}}+Y_{d}^{\operatorname{bi}}\leq 1 CLP,2,ybi\displaystyle C_{\operatorname{LP},2,y}^{\operatorname{bi}} =Xabi+Ybbi+Ycbi1\displaystyle=X_{a}^{\operatorname{bi}}+Y_{b}^{\operatorname{bi}}+Y_{c}^{\operatorname{bi}}\leq 1 CLP,5,ybi\displaystyle C_{\operatorname{LP},5,y}^{\operatorname{bi}} =Xebi1\displaystyle=X_{e}^{\operatorname{bi}}\leq 1
CLP,{3,5}bi\displaystyle C_{\operatorname{LP},\{3,5\}}^{\operatorname{bi}} =Xebi+Yebi1\displaystyle=X_{e}^{\operatorname{bi}}+Y_{e}^{\operatorname{bi}}\leq 1 CLP,3,xbi\displaystyle C_{\operatorname{LP},3,x}^{\operatorname{bi}} =Xdbi+Xebi+Ybbi1\displaystyle=X_{d}^{\operatorname{bi}}+X_{e}^{\operatorname{bi}}+Y_{b}^{\operatorname{bi}}\leq 1 CLP,6,xbi\displaystyle C_{\operatorname{LP},6,x}^{\operatorname{bi}} =Yfbi1\displaystyle=Y_{f}^{\operatorname{bi}}\leq 1
CLP,{4,6}bi\displaystyle C_{\operatorname{LP},\{4,6\}}^{\operatorname{bi}} =Xfbi+Yfbi1\displaystyle=X_{f}^{\operatorname{bi}}+Y_{f}^{\operatorname{bi}}\leq 1 CLP,3,ybi\displaystyle C_{\operatorname{LP},3,y}^{\operatorname{bi}} =Ydbi+Yebi+Xbbi1\displaystyle=Y_{d}^{\operatorname{bi}}+Y_{e}^{\operatorname{bi}}+X_{b}^{\operatorname{bi}}\leq 1 CLP,6,ybi\displaystyle C_{\operatorname{LP},6,y}^{\operatorname{bi}} =Xfbi1\displaystyle=X_{f}^{\operatorname{bi}}\leq 1

and C{2,3,4}bi=Xbbi+Ybbi+Xcbi+Ycbi+Xdbi+Ydbi2C_{\{2,3,4\}}^{\operatorname{bi}}=X_{b}^{\operatorname{bi}}+Y_{b}^{\operatorname{bi}}+X_{c}^{\operatorname{bi}}+Y_{c}^{\operatorname{bi}}+X_{d}^{\operatorname{bi}}+Y_{d}^{\operatorname{bi}}\leq 2.

In the linear program (10) below, whenever we write {i,j}\{i,j\} for an edge of a graph we will in addition assume that i<ji<j. The dual of the linear program (8) is given by

(10) minimizeZLP,G:=vV(G)(WLP,v,x+WLP,v,y)+U𝒜(|U|1)WLP,Usubjectto:WLP,v,x𝕜0 for all vV(G)WLP,v,y𝕜0 for all vV(G)WLP,U𝕜0 for all UV(G) with |U|2DLP,e,x:=eE(G[U])U𝒜WLP,U+WLP,i,x+WLP,j,y1 for all e={i,j}E(G)DLP,e,y:=eE(G[U])U𝒜WLP,U+WLP,j,x+WLP,i,y1 for all e={i,j}E(G)\begin{split}\operatorname*{minimize}\hskip 8.50006pt&Z_{\operatorname{LP^{\ast}},G}:=\sum_{v\in V(G)}(W_{\operatorname{LP^{\ast}},v,x}+W_{\operatorname{LP^{\ast}},v,y})+\sum_{U\in\mathcal{A}}(\lvert U\rvert-1)W_{\operatorname{LP^{\ast}},U}\\ \operatorname{subject\,to:}\hskip 2.5pt&\phantom{}\\ &W_{\operatorname{LP^{\ast}},v,x}\in\mathbbm{k}_{\geq 0}\text{ for all }v\in V(G)\\ &W_{\operatorname{LP^{\ast}},v,y}\in\mathbbm{k}_{\geq 0}\text{ for all }v\in V(G)\\ &W_{\operatorname{LP^{\ast}},U}\in\mathbbm{k}_{\geq 0}\text{ for all }U\subseteq V(G)\text{ with }\lvert U\rvert\geq 2\\ &D_{\operatorname{LP^{\ast}},e,x}:=\sum_{\begin{subarray}{c}e\in E(G[U])\\ U\in\mathcal{A}\end{subarray}}W_{\operatorname{LP^{\ast}},U}+W_{\operatorname{LP^{\ast}},i,x}+W_{\operatorname{LP^{\ast}},j,y}\geq 1\text{ for all }e=\{i,j\}\in E(G)\\ &D_{\operatorname{LP^{\ast}},e,y}:=\sum_{\begin{subarray}{c}e\in E(G[U])\\ U\in\mathcal{A}\end{subarray}}W_{\operatorname{LP^{\ast}},U}+W_{\operatorname{LP^{\ast}},j,x}+W_{\operatorname{LP^{\ast}},i,y}\geq 1\text{ for all }e=\{i,j\}\in E(G)\end{split}

We denote by LPG,𝕜\operatorname{LP}_{G,\mathbbm{k}}^{\ast} the optimal value of (10).

Proposition 3.9.

Let GG be a labeled graph. Then, PPG,𝕜=LPG,𝕜\operatorname{PP}_{G,\mathbbm{k}}=\operatorname{LP}_{G,\mathbbm{k}} and PPG,𝕜=LPG,𝕜\operatorname{PP}_{G,\mathbbm{k}}^{\ast}=\operatorname{LP}_{G,\mathbbm{k}}^{\ast}.

Proof.

Let Σ\Sigma (respectively Σ\Sigma^{{}^{\prime}}) denote the set of feasible solutions of (4) (respectively of the feasible solutions of (8)). If 𝕜\mathbbm{k} is a field, then there are linear bijections

Σ\displaystyle\Sigma Σ\displaystyle\leftrightarrow\Sigma^{{}^{\prime}}
(ae)eE(G)\displaystyle(a_{e})_{e\in E(G)} ({ae2}eE(G),{ae2)}eE(G))\displaystyle\mapsto\big{(}\{\frac{a_{e}}{2}\}_{e\in E(G)},\{\frac{a_{e}}{2})\}_{e\in E(G)}\big{)}
(ae+be)eE(G)\displaystyle(a_{e}+b_{e})_{e\in E(G)} ({ae}eE(G),{be}eE(G))\displaystyle\mapsfrom\big{(}\{a_{e}\}_{e\in E(G)},\{b_{e}\}_{e\in E(G)}\big{)}

If 𝕜=\mathbbm{k}=\mathbb{Z}, then there are a linear bijections

Σ\displaystyle\Sigma Σ\displaystyle\leftrightarrow\Sigma^{{}^{\prime}}
(ae)eE(G)\displaystyle(a_{e})_{e\in E(G)} ({ae}eE(G),{0}eE(G))\displaystyle\mapsto\big{(}\{a_{e}\}_{e\in E(G)},\{0\}_{e\in E(G)}\big{)}
(ae+be)eE(G)\displaystyle(a_{e}+b_{e})_{e\in E(G)} (ae}eE(G),{be}eE(G))\displaystyle\mapsfrom\big{(}a_{e}\}_{e\in E(G)},\{b_{e}\}_{e\in E(G)}\big{)}

The constraint CU1C_{U}\leq 1 for U=eU=e shows that the map from ΣΣ\Sigma^{{}^{\prime}}\rightarrow\Sigma is well-defined.

The dual case is proved analogously. ∎

Proposition 3.10.

Let GG be a graph, then

r(G)ft(JG)\displaystyle r(G)\leq\operatorname{ft}(J_{G})

where r(G)r(G) is as found in Definition 2.5.

Proof.

Let << denote the lex monomial order on RR with X1>>Xn>Y1>>YnX_{1}>\cdots>X_{n}>Y_{1}>\cdots>Y_{n}. Then,

r(G)ft(in<JG)ft(JG).\displaystyle r(G)\leq\operatorname{ft}(\operatorname{in}_{<}J_{G})\leq\operatorname{ft}(J_{G}).

Lemma 3.11 ([MSV14, Lemma 1.2]).

Fix a 2×n2\times n generic matrix MM and S=𝔽p[M]S=\mathbb{F}_{p}[M]. There exists a non-negative integer NMN_{M}, so that for every q=peq=p^{e} where e0e\in\mathbb{Z}_{\geq 0}, we have νq(I2(M))NM+(q1)(n1)\nu_{q}(I_{2}(M))\leq N_{M}+(q-1)(n-1).

Proof.

Follows from the proof of their Lemma 1.2 by taking, in their notation, m=t=2m=t=2 and k=1k=1. ∎

Proposition 3.12.

Let GG be a labeled graph, then

ft(JG)LPG,\displaystyle\operatorname{ft}(J_{G})\leq\operatorname{LP}_{G,\mathbb{Q}}
Proof.

Fix notation of MM as a 2×|V(G)|2\times\lvert V(G)\rvert generic matrix, S=𝔽p[M]S=\mathbb{F}_{p}[M]. For UV(G)U\subseteq V(G) we consider MUM_{U} and SUS_{U} the restriction of MM and SS to the submatrix (respectively subring) of variables corresponding to the vertices of UU. We take N:=max{|V(G)|,{NMUU𝒜}}N:=\max\big{\{}\lvert V(G)\rvert,\{N_{M_{U}}\mid U\in\mathcal{A}\}\big{\}} where NMUN_{M_{U}} is as defined in Lemma 3.11.

We denote by PP_{\infty} the polytope contained in (0)2|E|(\mathbb{R}_{\geq 0})^{2\lvert E\rvert} determined by the constraints of linear program (8).

For every tt\in\mathbb{N}, let PtP_{t} be the polytope inside (0)2|E|(\mathbb{R}_{\geq 0})^{2\lvert E\rvert} determined by the following constraints

X~LP,e0 for all eE(G)\displaystyle\tilde{X}_{\operatorname{LP},e}\geq 0\text{ for all }e\in E(G)
Y~LP,e0 for all eE(G)\displaystyle\tilde{Y}_{\operatorname{LP},e}\geq 0\text{ for all }e\in E(G)
(11) C~LP,U:=eE(G[U])(X~LP,e+Y~LP,e)N+(pt1)(|U|1)pt for all U𝒜\displaystyle\tilde{C}_{\operatorname{LP},U}:=\sum_{e\in E(G[U])}(\tilde{X}_{\operatorname{LP},e}+\tilde{Y}_{\operatorname{LP},e})\leq\frac{N+(p^{t}-1)(\lvert U\rvert-1)}{p^{t}}\text{ for all }U\in\mathcal{A}
(12) C~LP,v,x:=e={v,w}v<wX~LP,e+e={w,v}w<vY~LP,e1 for all vV(G)\displaystyle\tilde{C}_{\operatorname{LP},v,x}:=\sum_{\begin{subarray}{c}e=\{v,w\}\\ v<w\end{subarray}}\tilde{X}_{\operatorname{LP},e}+\sum_{\begin{subarray}{c}e=\{w,v\}\\ w<v\end{subarray}}\tilde{Y}_{\operatorname{LP},e}\leq 1\text{ for all }v\in V(G)
(13) C~LP,v,y:=e={v,w}v<wY~LP,e+e={w,v}w<vX~LP,e1 for all vV(G)\displaystyle\tilde{C}_{\operatorname{LP},v,y}:=\sum_{\begin{subarray}{c}e=\{v,w\}\\ v<w\end{subarray}}\tilde{Y}_{\operatorname{LP},e}+\sum_{\begin{subarray}{c}e=\{w,v\}\\ w<v\end{subarray}}\tilde{X}_{\operatorname{LP},e}\leq 1\text{ for all }v\in V(G)

We denote by LPG,,t\operatorname{LP}_{G,\mathbb{Q},t} to be the maximum of the function

2|E|:({X~LP,e}eE(G),{Y~LP,e}eE(G))eE(G)(X~LP,e+Y~LP,e)\displaystyle\mathbb{R}^{2\lvert E\rvert}\rightarrow\mathbb{R}:(\{\tilde{X}_{\operatorname{LP},e}\}_{e\in E(G)},\{\tilde{Y}_{\operatorname{LP},e}\}_{e\in E(G)})\mapsto\sum_{e\in E(G)}(\tilde{X}_{\operatorname{LP},e}+\tilde{Y}_{\operatorname{LP},e})

when restricted to the region PtP_{t}.

Fix tt a positive integer. Then, there exist non-negative integers ae,bea_{e},b_{e} for eE(G)e\in E(G) satisfying eE(G)(ae+be)=νt(JG)\sum_{e\in E(G)}(a_{e}+b_{e})=\nu_{t}(J_{G}) and a monomial

μ:=e=(i,j)E(G)i<j(XiYj)ae(XjYi)be\displaystyle\mu:=\prod\limits_{\begin{subarray}{c}e=(i,j)\in E(G)\\ i<j\end{subarray}}(X_{i}Y_{j})^{a_{e}}(X_{j}Y_{i})^{b_{e}}

belonging to the monomial support of some element of JGνt(JG)m[pt]J_{G}^{\nu_{t}(J_{G})}\smallsetminus m^{[p^{t}]}.

Observe that the tuple (𝐚,𝐛)=({aept}eE(G),{bept}eE(G))Pt\left(\mathbf{{a}},\mathbf{{b}}\right)=\left(\{\frac{a_{e}}{p^{t}}\}_{e\in E(G)},\{\frac{b_{e}}{p^{t}}\}_{e\in E(G)}\right)\in P_{t}. Indeed, since μm[pt]\mu\notin m^{[p^{t}]},

C~LP,v,x(𝐚,𝐛)\displaystyle\tilde{C}_{\operatorname{LP},v,x}\big{(}\mathbf{{a}},\mathbf{{b}}\big{)} =degXvμpt\displaystyle=\deg_{X_{v}}\mu\leq p^{t}
C~LP,v,y(𝐚,𝐛)\displaystyle\tilde{C}_{\operatorname{LP},v,y}\big{(}\mathbf{{a}},\mathbf{{b}}\big{)} =degYvμpt.\displaystyle=\deg_{Y_{v}}\mu\leq p^{t}.

Indeed, by Lemma 3.11,

C~LP,U(𝐚,𝐛)N+(pt1)(|U|1).\displaystyle\tilde{C}_{\operatorname{LP},U}\big{(}\mathbf{{a}},\mathbf{{b}}\big{)}\leq N+(p^{t}-1)(\lvert U\rvert-1).

This shows that νt(JG)ptLPG,,t\frac{\nu_{t}(J_{G})}{p^{t}}\leq\operatorname{LP}_{G,\mathbb{Q},t}. Taking limit as tt\rightarrow\infty it follows by Lemma 3.13 that ft(JG)LPG,\operatorname{ft}(J_{G})\leq\operatorname{LP}_{G,\mathbb{Q}}. ∎

Lemma 3.13.

With the notation as above,

  1. (1)

    For every 0<ϵ<10<\epsilon<1, there exists tt\in\mathbb{N} so that for every 𝐱Pt\mathbf{{x}}\in P_{t} there exists 𝐲P\mathbf{{y}}\in P_{\infty} with 𝐱𝐲1<ϵ||\mathbf{{x}}-\mathbf{{y}}||_{1}<\epsilon.

  2. (2)

    limtLPG,,t=LPG,\lim_{t\rightarrow\infty}\operatorname{LP}_{G,\mathbb{Q},t}=\operatorname{LP}_{G,\mathbb{Q}}.

Proof.
  1. (1)

    Fix 0<ϵ<10<\epsilon<1. Take tt\in\mathbb{N} so that 2|E|Npt<ϵ2\lvert E\rvert\frac{N}{p^{t}}<\epsilon. For 𝐱Pt\mathbf{{x}}\in P_{t}, define 𝐲\mathbf{{y}} via yi:=ximin{xi,Npt}y_{i}:=x_{i}-\min\{x_{i},\frac{N}{p^{t}}\} for 1i2|E|1\leq i\leq 2\lvert E\rvert. We observe that

    𝐱𝐲12|E|Npt<ϵ.\displaystyle||\mathbf{{x}}-\mathbf{{y}}||_{1}\leq 2\lvert E\rvert\frac{N}{p^{t}}<\epsilon.

    Since 𝐲\mathbf{{y}} is pointwise less than 𝐱\mathbf{{x}} and 𝐱\mathbf{{x}} satisfies the constraints (12) and (13), CLP,v,x(𝐲)<1C_{\operatorname{LP},v,x}(\mathbf{{y}})<1 and CLP,v,y(𝐲)<1C_{\operatorname{LP},v,y}(\mathbf{{y}})<1 for all vVv\in V. Thus, to show that 𝐲P\mathbf{{y}}\in P_{\infty} it suffices to show that CU(𝐲)|U|1C_{U}(\mathbf{{y}})\leq\lvert U\rvert-1 for all U𝒜U\in\mathcal{A}.

    Observation 1.

    The constraint C~LP,U(𝐱)N+(pt1)(|U|1)pt\tilde{C}_{\operatorname{LP},U}(\mathbf{{x}})\leq\frac{N+(p^{t}-1)(\lvert U\rvert-1)}{p^{t}} implies that

    (14) C~LP,U(𝐱)Npt<|U|1.\displaystyle\tilde{C}_{\operatorname{LP},U}(\mathbf{{x}})-\frac{N}{p^{t}}<\lvert U\rvert-1.
    Observation 2.

    If C~LP,U(𝐱)>|U|1\tilde{C}_{\operatorname{LP},U}(\mathbf{{x}})>\lvert U\rvert-1, then xmax,UNptx_{\max,U}\geq\frac{N}{p^{t}} where xmax,Ux_{\max,U} is the largest value of 𝐱\mathbf{{x}} that appears in the support of C~LP,U(𝐱)\tilde{C}_{\operatorname{LP},U}(\mathbf{{x}}). If this is not the case, then we have that

    |U|1\displaystyle\lvert U\rvert-1 <C~LP,U(𝐱)\displaystyle<\tilde{C}_{\operatorname{LP},U}(\mathbf{{x}})
    2|E(G[U])|xmax,U\displaystyle\leq 2\lvert E(G[U])\rvert\cdot x_{\max,U}
    2|E|Npt\displaystyle\leq 2\lvert E\rvert\frac{N}{p^{t}}
    <ϵ\displaystyle<\epsilon
    <1.\displaystyle<1.

    But, then this implies that |U|<2\lvert U\rvert<2; a contradiction.

    There is nothing to prove if C~LP,U(𝐱)|U|1\tilde{C}_{\operatorname{LP},U}(\mathbf{{x}})\leq\lvert U\rvert-1. If C~LP,U>|U|1\tilde{C}_{\operatorname{LP},U}>\lvert U\rvert-1, then by Observation 2 there exists xix_{i} appearing in the support of C~LP,U(𝐱)\tilde{C}_{\operatorname{LP},U}(\mathbf{{x}}) with xiNptx_{i}\geq\frac{N}{p^{t}}. Then, by construction yi=xiNpty_{i}=x_{i}-\frac{N}{p^{t}}. Hence, by Observation 1

    CU(𝐲)CU(𝐱)Npt<|U|1.\displaystyle C_{U}(\mathbf{{y}})\leq C_{U}(\mathbf{{x}})-\frac{N}{p^{t}}<\lvert U\rvert-1.

    Thus, 𝐲P\mathbf{{y}}\in P_{\infty}.

  2. (2)

    Let 0<ϵ<10<\epsilon<1, 𝐱\mathbf{{x}} a vertex of PtP_{t} where eE(G)(Xe+Ye)\sum_{e\in E(G)}(X_{e}+Y_{e}) attains a maximum, tt and 𝐲P\mathbf{{y}}\in P_{\infty} to be as in part (1) of this Lemma. Since 𝐲1LPG,||\mathbf{{y}}||_{1}\leq\operatorname{LP}_{G,\mathbb{Q}} and 𝐱𝐲1<ϵ||\mathbf{{x}}-\mathbf{{y}}||_{1}<\epsilon, it follows that

    LPG,,t=𝐱1𝐱𝐲1+𝐲1ϵ+LPG,.\displaystyle\operatorname{LP}_{G,\mathbb{Q},t}=||\mathbf{{x}}||_{1}\leq||\mathbf{{x}}-\mathbf{{y}}||_{1}+||\mathbf{{y}}||_{1}\leq\epsilon+\operatorname{LP}_{G,\mathbb{Q}}.

    Consequently,

    LPG,,tLPG,<ϵ.\displaystyle\operatorname{LP}_{G,\mathbb{Q},t}-\operatorname{LP}_{G,\mathbb{Q}}<\epsilon.

    This proves that limtLPG,,t=LPG,\lim_{t\rightarrow\infty}\operatorname{LP}_{G,\mathbb{Q},t}=\operatorname{LP}_{G,\mathbb{Q}}.

Corollary 3.14.

Let GG be a graph, then

lct(JG)LPG,.\displaystyle\operatorname{lct}(J_{G})\leq\operatorname{LP}_{G,\mathbb{Q}}.
Proof.

Clear from Proposition 3.12 and Theorem 1.2. ∎

Remark 3.15.

One can always bound the F-threshold of an ideal in a polynomial ring with respect to the maximal ideal by the optimal value of a linear programming problem by removing the constraints of the form C~U|U|1\tilde{C}_{U}\leq\lvert U\rvert-1 from (8) (cf. [ST09] or [Her14]).

The previous results of this section can be summarized as

(15) binomial grade JG=r(G)=LPG,ft(JG)LPG,=LPG,LPG,=ht(JG).\text{binomial grade }J_{G}=r(G)=\operatorname{LP}_{G,\mathbb{Z}}\leq\operatorname{ft}(J_{G})\leq\operatorname{LP}_{G,\mathbb{Q}}=\operatorname{LP}_{G,\mathbb{Q}}^{\ast}\leq\operatorname{LP}_{G,\mathbb{Z}}^{\ast}=\operatorname{ht}(J_{G}).
Remark 3.16.

The relation ft(I)ht(I)\operatorname{ft}(I)\leq\operatorname{ht}(I) holds more generally, cf. [TW04, Proposition 2.6(1)].

4. Proof of Conjecture 1.3 for Block Graphs

Definition 4.1.

A graph GG is called biconnected if GG is connected and GvG\smallsetminus v is connected for every vertex vv of GG. A graph GG is called a block graph if every maximal biconnected component of GG is a complete graph. We call a maximal biconnected component of GG a block. A vertex vv of GG is called a cut vertex if the number of connected components of GvG\smallsetminus v is strictly larger than the number of connected components of GG.

Let GG be a block graph and BB a block of GG, then we define cut(B)\operatorname{cut}(B) to be the set of all cut vertices of GG contained in BB. We say that BB is a leaf of GG if |cut(B)|=1\lvert\operatorname{cut}(B)\rvert=1, and we say that GG has the two block property if every cut vertex of GG is adjacent to exactly two blocks of GG. An edge of GG which is also a leaf of GG is called a whisker of GG.

Throughout this section we will let nG:=|V(G)|n_{G}:=\lvert V(G)\rvert, G\ell_{G} will denote the number of leaves of GG, wGw_{G} will denote the number of connected components of GG isomorphic to a complete graph. When the graph GG is clear from context we will omit the subscript GG.

Proposition 4.2.

Let GG be a block graph. Then,

LPG,nG12w.\displaystyle\operatorname{LP}_{G,\mathbb{Q}}\leq n_{G}-\frac{1}{2}\ell-w.
Proof.

By Remark 3.5 we may assume that GG is connected. If GG is a complete graph there is nothing to prove. We reduce to the case that GG is a connected block graph which is not a complete graph.

Since GG is a connected block graph which is not a complete graph w=0w=0 and 1\ell\geq 1. Partition the blocks of GG into 1\mathcal{B}_{1}, the leaves of GG, and 2\mathcal{B}_{2}, the remaining blocks. Write 1={B1,,B}\mathcal{B}_{1}=\{B_{1},\ldots,B_{\ell}\}. Each Bi1B_{i}\in\mathcal{B}_{1} has exactly one cut vertex which we will call cic_{i} and at least one other vertex, i.e. |V(Bi)|2\lvert V(B_{i})\rvert\geq 2. Let

T:=(B2V(B))({ci1i}).\displaystyle T:=\left(\bigcup_{B\in\mathcal{B}_{2}}V(B)\right)\cup\left(\bigcup\{c_{i}\mid 1\leq i\leq\ell\}\right).

For vVv\in V, define

(16) bv,x=bv,y={1/2,vT0,vV(G)T.b_{v,x}=b_{v,y}=\begin{cases}1/2,&v\in T\\ 0,&v\in V(G)\smallsetminus T.\end{cases}

For U𝒜U\in\mathcal{A}, define

(17) cU={1/2,U11/2,U=Bi{ci} for some 1i,0, else.c_{U}=\begin{cases}1/2,&U\in\mathcal{B}_{1}\\ 1/2,&U=B_{i}\smallsetminus\{c_{i}\}\text{ for some }1\leq i\leq\ell,\\ 0,&\text{ else}.\end{cases}

Then, the tuple ({bv,x}vV,{bv,y}vV,{cU}U𝒜)(\{b_{v,x}\}_{v\in V},\{b_{v,y}\}_{v\in V},\{c_{U}\}_{U\in\mathcal{A}}) is a feasible solution to the linear program (10), and the cost of this feasible solution is

nG12.\displaystyle n_{G}-\frac{1}{2}\ell.

Thus,

LPG,nG12.\displaystyle\operatorname{LP}_{G,\mathbb{Q}}^{\ast}\leq n_{G}-\frac{1}{2}\ell.

Hence, the result follows by strong duality, Theorem 2.1. ∎

4.1. Triangle Paths and the Two Block Property

Definition 4.3.

We say that a graph GG is a triangle path if every connected component of GG which is not an isolated vertex satisfies

  1. (1)

    the only cycles in GG are K3K_{3}’s which we label as C1,,CtC_{1},\ldots,C_{t} (we will sometimes refer to the CiC_{i}’s as the triangles of G\mathit{G}),

  2. (2)

    For all vGv\in G,

    degG(v)={3,vV(Ci) for some 1it1 or 2,otherwise.\displaystyle\deg_{G}(v)=\begin{cases}3,&v\in V(C_{i})\text{ for some }1\leq i\leq t\\ 1\text{ or }2,&\text{otherwise}.\end{cases}
  3. (3)

    Every vertex of GG belongs to at most one triangle of GG.

When GG is a triangle path and GG is a subgraph of LL, then we call GG a triangle path packing of LL.

Example 4.4.

The graph in Figure 2 is an instance of a triangle path.

Figure 2. Triangle Path
Remark 4.5.

A triangle path is an example of a family of block graphs with the two block property.

Proposition 4.6.

Let GG be a triangle path. Label the K3K_{3}’s in the triangle path C1,,CtC_{1},\ldots,C_{t}. Label the connected components of Gk=1tE(Ck)G\smallsetminus\bigcup_{k=1}^{t}E(C_{k}) by P1,,PsP_{1},\ldots,P_{s}. Then,

(18) ft(JG)=i=1s|E(Pi)|+32t.\operatorname{ft}(J_{G})=\sum_{i=1}^{s}\lvert E(P_{i})\rvert+\frac{3}{2}t.

Moreover,

(19) ft(JG)=nG12GwG.\operatorname{ft}(J_{G})=n_{G}-\frac{1}{2}\ell_{G}-w_{G}.
Proof.

By induction on tt it can be shown that (18) and (19) are equivalent. By Lemma 2.9 and Remark 2.13 we may assume that GG is connected and not a complete graph, i.e. that w=0w=0. By Proposition 3.12 and Proposition 4.2, and the above mentioned equivalence we have that ft(JG)i=1s|E(Pi)|+32t\operatorname{ft}(J_{G})\leq\sum_{i=1}^{s}\lvert E(P_{i})\rvert+\frac{3}{2}t. It thus suffices to prove the reverse inequality.

Let RR and 𝔪\mathfrak{m} be as in Definition 1.1. First consider the case that char(𝕂)=p>2\operatorname{char}(\mathbb{K})=p>2. The first step in the proof is to label the vertices of the graph and to assign to every edge ee a tuple (ae,be)(a_{e},b_{e}) so that the following conditions are satisfied.

  1. (1)

    Some vertex having degree one has been assigned the label 11.

  2. (2)

    If aV(Pj)a\in V(P_{j}) and b,cV(Pi)b,c\in V(P_{i}) for some 1i,j1\leq i,j\leq\ell with the distance from bb to aa smaller than the distance from cc to aa, then either a<b<ca<b<c or c<b<ac<b<a.

  3. (3)

    (ae,be)(a_{e},b_{e}) equals (0,1)(0,1), (1,0)(1,0), or (0.5,0.5)(0.5,0.5) if ee is not an edge of any triangle of GG.

  4. (4)

    (ae,be)(a_{e},b_{e}) equals (0,0.5)(0,0.5) or (0.5,0)(0.5,0) if ee is the edge of any triangle of GG.

  5. (5)

    eE(G)(ae+be)=i=1s|E(Pi)|+32t\sum_{e\in E(G)}\left(a_{e}+b_{e}\right)=\sum_{i=1}^{s}\lvert E(P_{i})\rvert+\frac{3}{2}t.

  6. (6)

    The monomial

    (20) m:=e={i,j}E(G)i<j(XiYj)aij(pe1)(XjYi)bij(pe1)m:=\prod_{\begin{subarray}{c}e=\{i,j\}\in E(G)\\ i<j\end{subarray}}(X_{i}Y_{j})^{a_{ij}\cdot(p^{e}-1)}\cdot(X_{j}Y_{i})^{b_{ij}\cdot(p^{e}-1)}

    does not belong to the ideal 𝔪[pe]\mathfrak{m}^{[p^{e}]} and appears in the polynomial

    (21) f:=e={i,j}E(G)i<jfij(aij+bij)(pe1).f:=\prod_{\begin{subarray}{c}e=\{i,j\}\in E(G)\\ i<j\end{subarray}}f_{ij}^{(a_{ij}+b_{ij})\cdot(p^{e}-1)}.

    with a coefficient which is non-zero mod pp.

It would follow from items (5) and (21) that

(i=1s|E(Pi)|+32t)(pe1)νe(JG),\displaystyle\bigg{(}\sum_{i=1}^{s}\lvert E(P_{i})\rvert+\frac{3}{2}t\bigg{)}(p^{e}-1)\leq\nu_{e}(J_{G}),

and consequently that

i=1s|E(Pi)|+32tft(JG).\displaystyle\sum_{i=1}^{s}\lvert E(P_{i})\rvert+\frac{3}{2}t\leq\operatorname{ft}(J_{G}).

We prove the existence of this labeling by induction on tt. It t=0t=0, then GG is a path. Starting at a vertex of degree one assign it the label 11, then label the vertices consecutively. Assign to every edge the tuple (1,0)(1,0). That the above conditions are satisfied is clear.

Suppose now that t1t\geq 1. There exists a triangle of GG, call it TT, so that removing the edges of TT from GG yields a graph having three connected components—two of which are paths, say PP and QQ, and the other is a triangle path, call it GG^{{}^{\prime}}, having one less triangle than GG. Let bb be the vertex belonging to V(T)V(G)V(T)\cap V(G^{{}^{\prime}}). Let aa be the vertex adjacent to bb in GG^{{}^{\prime}}. Apply the induction hypothesis to GG^{{}^{\prime}} and we may assume that bb has not been given the label 11. By item 2, it follows that a<ba<b. Label the remaining vertices of TT by n+1n+1 and n+2n+2. Label the vertex of V(P)V(Q)V(P)\cup V(Q) adjacent to n+1n+1 (respectively n+2n+2) by n+3n+3 (respectively n+4n+4). Label the remaining vertices of V(P)V(Q)V(P)\cup V(Q) so as to satisfy condition (2). Use Table 1 to assign tuples to the edges of TT and to the edges of PP and QQ incident to the vertices of TT based upon the tuple assigned to the edge {a,b}\{a,b\}. Now, PP and QQ each have an edge which has been assigned a tuple; extend this label to the remaining edges of PP and QQ. After performing this construction the coefficient on mm in ff from item (21) is (pe1pe12)g\binom{p^{e}-1}{\frac{p^{e}-1}{2}}^{g} where gg is the number of times that the label (0.5,0.5)(0.5,0.5) appears on an edge. This binomial is non-zero by Lucas’s theorem, Theorem 2.16.

{a,b}\{a,b\} {b,n+1}\{b,n+1\} {b,n+2}\{b,n+2\} {n+1,n+2}\{n+1,n+2\} {n+1,n+3}\{n+1,n+3\} {n+2,n+4}\{n+2,n+4\}
(0,1)(0,1) (0,0.5)(0,0.5) (0,0.5)(0,0.5) (0.5,0)(0.5,0) (0,1)(0,1) (0.5,0.5)(0.5,0.5)
(1,0)(1,0) (0.5,0)(0.5,0) (0.5,0)(0.5,0) (0,0.5)(0,0.5) (0,1)(0,1) (0.5,0.5)(0.5,0.5)
(0.5,0.5)(0.5,0.5) (0.5,0)(0.5,0) (0,0.5)(0,0.5) (0,0.5)(0,0.5) (1,0)(1,0) (0,1)(0,1)
Table 1. char(𝕂)=p>2\operatorname{char}(\mathbb{K})=p>2

Suppose now that char(𝕂)=2\operatorname{char}(\mathbb{K})=2 and fix r0r\geq 0. We modify the above construction by using Table 2 in the construction. In this case we have that

(2i=1s|E(Pi)|+3t)(pr11)νr(JG).\displaystyle\bigg{(}2\sum_{i=1}^{s}\lvert E(P_{i})\rvert+3t\bigg{)}(p^{r-1}-1)\leq\nu_{r}(J_{G}).

Dividing by prp^{r} and taking limit as rr\rightarrow\infty proves the desired inequality when char(𝕂)=p=2\operatorname{char}(\mathbb{K})=p=2.

{a,b}\{a,b\} {b,n+1}\{b,n+1\} {b,n+2}\{b,n+2\} {n+1,n+2}\{n+1,n+2\} {n+1,n+3}\{n+1,n+3\} {n+2,n+4}\{n+2,n+4\}
(0,pr1)(0,p^{r}-1) (0,pr11)(0,p^{r-1}-1) (0,pr11)(0,p^{r-1}-1) (pr11,0)(p^{r-1}-1,0) (0,pr1)(0,p^{r}-1) (pr1,pr11)(p^{r-1},p^{r-1}-1)
(pr1,0)(p^{r}-1,0) (pr11,0)(p^{r-1}-1,0) (pr11,0)(p^{r-1}-1,0) (0,pr11)(0,p^{r-1}-1) (0,pr1)(0,p^{r}-1) (pr1,pr11)(p^{r-1},p^{r-1}-1)
(pr1,pr11)(p^{r-1},p^{r-1}-1) (pr11,0)(p^{r-1}-1,0) (0,pr11)(0,p^{r-1}-1) (0,pr11)(0,p^{r-1}-1) (pr1,0)(p^{r}-1,0) (0,pr1)(0,p^{r}-1)
Table 2. char(𝕂)=p=2\operatorname{char}(\mathbb{K})=p=2

Corollary 4.7.

Let GG be a triangle path and G1,,GrG_{1},\ldots,G_{r} be subgraphs of GG satisfying

  1. (1)

    the GiG_{i} are triangle paths,

  2. (2)

    E(G)=i=1rE(Gi)E(G)=\bigsqcup\limits_{i=1}^{r}E(G_{i}),

  3. (3)

    every triangle of GG is a subgraph of some GiG_{i}

Then,

ft(JG)=i=1rft(JGi).\displaystyle\operatorname{ft}(J_{G})=\sum_{i=1}^{r}\operatorname{ft}(J_{G_{i}}).
Proof.

Follows from Proposition 4.6. ∎

Corollary 4.8.

Let HH be a triangle path. Suppose CV(H)C\subseteq V(H) satisfies that HCH\smallsetminus C is a triangle path, no two vertices of CC are adjacent via an edge of HH, and every vertex of CC has degree two in HH. Then we have

ft(JHC)=ft(JH)2|C|.\displaystyle\operatorname{ft}(J_{H\smallsetminus C})=\operatorname{ft}(J_{H})-2\lvert C\rvert.
Proof.

Since HH and HCH\smallsetminus C are triangle paths, we can compute their F-threshold as 3/23/2 the number of triangles plus the number of edges not in a triangle. The later two conditions on CC imply that the number of triangles of HH and HCH\smallsetminus C are the same and the number of edges in HH not belonging to any triangle path is equal to the number of edges of HCH\smallsetminus C not belonging to any triangle path plus 2|C|2\lvert C\rvert. ∎

We record the following lemma which we will need later which can be viewed as an analogue of Corollary 4.8.

Lemma 4.9.

Let GG be a graph, CV(G)C\subseteq V(G), and {Gi}i=1r\{G_{i}\}_{i=1}^{r} the connected components of GCG\smallsetminus C, then

LPG,i=1rLPGi,+2|C|.\displaystyle\operatorname{LP}_{G,\mathbb{Q}}\leq\sum_{i=1}^{r}\operatorname{LP}_{G_{i},\mathbb{Q}}+2\lvert C\rvert.
Proof.

By strong duality, Theorem 2.1, it suffices to prove the statement for the dual linear program. For i=1,,ri=1,\ldots,r, let ({bi,v,x}vV,{bi,v,y}vV,{ci,U}U𝒜i)(\{b_{i,v,x}\}_{v\in V},\{b_{i,v,y}\}_{v\in V},\{c_{i,U}\}_{U\in\mathcal{A}_{i}}) be an optimal solution to the linear program (10) realizing LPGi,\operatorname{LP}^{\ast}_{G_{i},\mathbb{Q}}. These local solutions together with the assignments bv,x:=1b_{v,x}:=1, bv,y:=1b_{v,y}:=1 for vCv\in C, and cU:=0c_{U}:=0 for U𝒜i=1r𝒜iU\in\mathcal{A}\smallsetminus\cup_{i=1}^{r}\mathcal{A}_{i} yield a feasible solution to the linear program (10) realizing LPG,\operatorname{LP}^{\ast}_{G,\mathbb{Q}}. ∎

The next proposition shows that the F-threshold of a block graph with the two block property can be computed by a maximal triangle path packing.

Proposition 4.10.

Let GG be a block graph with the two block property. Let \ell denote the number of leaves of GG and ww the number of connected components of GG which are isomorphic to a complete graph. Then, GG contains a subgraph HH satisfying

  1. (1)

    HH is a triangle path packing,

  2. (2)

    V(G)=V(H)V(G)=V(H),

  3. (3)

    for every block BB of GG, BB contains at most one triangle of HH,

  4. (4)

    a block BB of GG contains a triangle of HH only if |cutB|\lvert\operatorname{cut}{B}\rvert is odd,

  5. (5)

    if TT is a triangle of HH contained in a block BB of GG, then TT contains two cut vertices unless cut(B)=V(B)\operatorname{cut}(B)=V(B) in which case TT contains three,

  6. (6)

    a block BB of GG contains exactly one vertex of HH of degree one if and only if BB is a leaf of GG,

  7. (7)

    a block BB of GG contains a vertex of HH of degree zero if and only if BB is a single vertex,

  8. (8)

    if TT is a triangle of HH contained in BB, ccut(B)V(T)c\in\operatorname{cut}(B)\cap V(T), and B~B\tilde{B}\neq B is the block of GG incident to cc, then the edge of HH incident to cc and not contained in TT is contained in B~\tilde{B},

  9. (9)

    ft(JH)=n12w\operatorname{ft}(J_{H})=n-\frac{1}{2}\ell-w.

The proof idea is to induce on the number of blocks of GG. We start by choosing a block of GG which is not a leaf and so that every incident block with the exception of at most one is a leaf of GG. In Figure 3(a), blocks {12,13}\{12,13\} and {1,2,3,4,5,6}\{1,2,3,4,5,6\} are the only such blocks. We will consider block B:={1,2,3,4,5,6}B:=\{1,2,3,4,5,6\}. Next, we choose two leaves of BB to delete (if BB has only one leaf then delete that leaf), Figure 3(b). By induction we construct a triangle path packing satisfying the above conditions, Figure 3(c). Lastly, extend the triangle path packing of the subgraph to a triangle path packing of the original graph so that the new triangle path packing has the desired properties, Figure 3(d). When the number of cut vertices of GG in BB is less than or equal to 33 several cases must be considered to extend the triangle path packing while preserving the above properties. However, when the number of cut vertices of GG in BB is larger than 33 then the triangle path packing can be extended by taking a Hamiltonian path on the deleted vertices.

1122334455667788991212131314141010111115151616
(a) Block Graph
112244551212131314141010111115151616
(b) Deleting Leaves
112244551212131314141010111115151616
(c) Triangle path of subgraph
1122334455667788991212131314141010111115151616
(d) Triangle path of original graph
Figure 3. Proof Idea of Proposition 4.10
Proof.

It is enough to show this proposition for each connected component of GG. If GG is isomorphic to a complete graph, then take HH to be a Hamiltonian path and the Proposition follows since =0\ell=0 and w=1w=1 in this case. Thus, we may assume henceforth that GG is a connected graph and not isomorphic to a complete graph.

The proof is by induction on the number of blocks of GG which we denote by tt. Since GG is not a complete graph, t2t\geq 2. If t=2t=2, then GG is two complete graphs joined at a cut vertex. In this case, =2\ell=2 and GG has a Hamiltonian path. Take HH to be a Hamiltonian path of GG, and it is clear that HH satisfies the conditions listed above.

Suppose now that t3t\geq 3 and that the Proposition has been shown for all block graphs having the two block property with at most t1t-1 blocks. Let BB be a block of GG which is not a leaf such that every block incident to BB with the exception of at most one is a leaf of GG. Let rr denote the number of cut vertices of BB. Observe that r2r\geq 2 since BB is not a leaf.

Case r=2r=2: Let LL be a leaf of BB. Let c~\tilde{c} be the cut vertex of BB which is adjacent to LL. If |V(B)|>2\lvert V(B)\rvert>2 put G~:=GL\tilde{G}:=G\smallsetminus L; otherwise, put G~:=G(Lc~)\tilde{G}:=G\smallsetminus(L\smallsetminus\tilde{c}). Let B~\tilde{B} be the image of BB in G~\tilde{G}. Since G~\tilde{G} has t1t-1 blocks, by induction there exists a triangle path H~\tilde{H} of G~\tilde{G} satisfying the above properties. Since B~\tilde{B} is a leaf in G~\tilde{G}, H~B~\tilde{H}\mid_{\tilde{B}} has a vertex of degree one, say vv, by property (6). Consider now H~\tilde{H} as a subgraph of GG, and construct HH a triangle path in GG as follows. Observe that if |V(B)|=2\lvert V(B)\rvert=2, then v=c~v=\tilde{c}. Otherwise, adjoin an edge from vv to c~\tilde{c}. In either case HH is now a subgraph of GG containing c~\tilde{c} as a degree one vertex. Finally, adjoin to HH a Hamiltonian path contained in LL which has a terminal vertex at c~\tilde{c}. Now, HH is a triangle path of GG which satisfies properties (1) through (9). Moreover, ft(JH)\operatorname{ft}(J_{H}) is of the desired form by Proposition 4.6, induction hypothesis, and the fact that we obtained HH from H~\tilde{H} by adding a path.

Case r=3r=3: Let cutB={c1,c2,c3}\operatorname{cut}{B}=\{c_{1},c_{2},c_{3}\}. Let LiL_{i} be leaves adjacent to BB at cic_{i} for i{1,2}i\in\{1,2\}.

Sub-case |V(B)|=3\lvert V(B)\rvert=3: Let G~\tilde{G} be the graph attained after removing L1L_{1}, L2L_{2}, and B{c3}B\smallsetminus\{c_{3}\}. Note that G~\tilde{G} is adjacent to BB at c3c_{3} and that G~\tilde{G} may or may not be a leaf. Consider the graph GG^{{}^{\prime}} which is G~\tilde{G} adjoined a whisker at the vertex c3c_{3}. Denote by nn^{{}^{\prime}} and \ell^{{}^{\prime}} the number of vertices and leaves of GG^{{}^{\prime}}, respectively. By induction hypothesis there exists a triangle path HH^{{}^{\prime}} for GG^{{}^{\prime}} satisfying the above properties and

ft(JH)=n12.\displaystyle\operatorname{ft}(J_{H^{{}^{\prime}}})=n^{{}^{\prime}}-\frac{1}{2}\ell^{{}^{\prime}}.

First, consider the sub-sub-case that there exists a triangle in HH^{{}^{\prime}} containing c3c_{3}; label the other vertices of this triangle as aa and bb. Construct HH from HH^{{}^{\prime}} by deleting the whisker attached at c3c_{3}, deleting the edge {a,b}\{a,b\}, adding the edge {c1,c2}\{c_{1},c_{2}\}, and adding Hamiltonian paths in L1L_{1} and L2L_{2} which have a terminal vertex at c1c_{1} and c2c_{2}, respectively. Observe that HH is a triangle path, and by Proposition 4.6 we compute that

ft(JH)=ft(JH)+(112+1)+(|V(L1)|1)+1+|V(L2)|1)\displaystyle\operatorname{ft}(J_{H})=\operatorname{ft}(J_{H^{{}^{\prime}}})+(-1-\frac{1}{2}+1)+(\lvert V(L_{1})\rvert-1)+1+\lvert V(L_{2})\rvert-1)

where the term (112+1)(-1-\frac{1}{2}+1) is counting in HH^{{}^{\prime}} the deletion of a whisker, the deletion of an edge of a triangle, and that two edges of a triangle are now counted as edges belonging to a path. The desired formula on ft(JH)\operatorname{ft}(J_{H}) now follows after simplifying and using that n=n+(|V(L1)|1)+|V(L2)|n=n^{{}^{\prime}}+(\lvert V(L_{1})\rvert-1)+\lvert V(L_{2})\rvert and that =+1\ell=\ell^{{}^{\prime}}+1.

Second, consider the sub-sub-case that c3c_{3} has degree two in HH^{{}^{\prime}}; call the neighbor of c3c_{3} contained in G~\tilde{G} to be aa. Construct HH from HH^{{}^{\prime}} by deleting the whisker, adding the edges {c1,c2}\{c_{1},c_{2}\}, {c2,c3}\{c_{2},c_{3}\}, {c1,c3}\{c_{1},c_{3}\} and adding Hamiltonian paths in L1L_{1} and L2L_{2} which have a terminal vertex at c1c_{1} and c2c_{2}, respectively. Observe that HH is a triangle path, and by Proposition 4.6 we compute that

ft(JH)=(ft(JH)1)+32+(|V(L1)|1)+|V(L2)|1)\displaystyle\operatorname{ft}(J_{H})=\big{(}\operatorname{ft}(J_{H^{{}^{\prime}}})-1\big{)}+\frac{3}{2}+(\lvert V(L_{1})\rvert-1)+\lvert V(L_{2})\rvert-1)

where the 1-1 is the deletion of the whisker and the plus 32\frac{3}{2} is the addition of the triangle. The formula for ft(JH)\operatorname{ft}(J_{H}) now follows after simplifying and using that n=n+(|V(L1)|1)+|V(L2)|n=n^{{}^{\prime}}+(\lvert V(L_{1})\rvert-1)+\lvert V(L_{2})\rvert and that =+1\ell=\ell^{{}^{\prime}}+1.

Sub-case |V(B)|>3\lvert V(B)\rvert>3: Let G~:=G(L1L2)\tilde{G}:=G\smallsetminus(L_{1}\cup L_{2}). By induction hypothesis there exists a triangle path H~\tilde{H} of G~\tilde{G} satisfying the desired properties. Since B~\tilde{B} is a leaf in G~\tilde{G}, property (6) guarantees the existence of a vertex vv in H~B~\tilde{H}\cap\tilde{B} having degree one. Consider H~\tilde{H} as a subgraph of GG, and construct HH a triangle path of GG from H~\tilde{H} by adding the edges {v,c1}\{v,c_{1}\}, {v,c2}\{v,c_{2}\}, {c1,c2}\{c_{1},c_{2}\} and adding Hamiltonian paths in L1L_{1} and L2L_{2} which have a terminal vertex at c1c_{1} and c2c_{2}, respectively. Observe that HH is a triangle path, and by Proposition 4.6 we compute that

ft(JH)=ft(JH~)+32+(|V(L1)|1)+(|V(L2)|1)\displaystyle\operatorname{ft}(J_{H})=\operatorname{ft}(J_{\tilde{H}})+\frac{3}{2}+(\lvert V(L_{1})\rvert-1)+(\lvert V(L_{2})\rvert-1)

where the plus 32\frac{3}{2} is the addition of the triangle. The desired formula on ft(JH)\operatorname{ft}(J_{H}) now follows after simplifying and using that n=n~+|V(L1)|+|V(L2)|n=\tilde{n}+\lvert V(L_{1})\rvert+\lvert V(L_{2})\rvert and that =~+1\ell=\tilde{\ell}+1.

Case r>3r>3: Let c1c_{1} and c2c_{2} be two cut vertices of BB which have leaves L1L_{1} and L2L_{2}, respectively. Let G~:=G(L1L2)\tilde{G}:=G\smallsetminus(L_{1}\cup L_{2}). By induction hypothesis there exists a triangle path H~\tilde{H} of G~\tilde{G} satisfying the desired properties. Consider H~\tilde{H} as a subgraph of GG, and construct HH a triangle path of GG from H~\tilde{H} by adding the edges {c1,c2}\{c_{1},c_{2}\} and adding Hamiltonian paths in L1L_{1} and L2L_{2} which have a terminal vertex at c1c_{1} and c2c_{2}, respectively. Observe that HH is a triangle path, and by Proposition 4.6 we compute that

ft(JH)=ft(JH~)+(|V(L1)|1)+1+|V(L2)|1)\displaystyle\operatorname{ft}(J_{H})=\operatorname{ft}(J_{\tilde{H}})+(\lvert V(L_{1})\rvert-1)+1+\lvert V(L_{2})\rvert-1)

where the plus 11 is the addition of the edge {c1,c2}\{c_{1},c_{2}\}. The desired formula on ft(JH)\operatorname{ft}(J_{H}) now follows after simplifying and using that n=n~+|V(L1)|+|V(L2)|n=\tilde{n}+\lvert V(L_{1})\rvert+\lvert V(L_{2})\rvert and that =~+2\ell=\tilde{\ell}+2. ∎

Corollary 4.11.

Let GG be a block graph having the two block property. Let \ell denote the number of leaves of GG and ww the number of connected components of GG isomorphic to a complete graph. Then, there exists a triangle path packing HH of GG such that

ft(JH)=ft(JG)=LPG,=n12w.\displaystyle\operatorname{ft}(J_{H})=\operatorname{ft}(J_{G})=\operatorname{LP}_{G,\mathbb{Q}}=n-\frac{1}{2}\ell-w.
Proof.

Follows from Lemma 2.10, Proposition 3.12, Proposition 4.10, and Proposition 4.2. ∎

4.2. Two Block Approximation and General Case

Next, we show how to reduce the computation of the F-threshold of a block graph to that of a block graph with the two block property.

Definition 4.12.

Let GG be a block graph. We call HH a two block approximation of GG if HH can be obtained via the following process

  1. (1)

    Enumerate the cut vertices of GG incident to at least three blocks of GG; call this set CC.

  2. (2)

    For each cCc\in C, choose two blocks incident to cc, say AcA_{c} and BcB_{c}.

  3. (3)

    Delete all edges incident to cc and not contained in E(Ac)E(Bc)E(A_{c})\cup E(B_{c}).

Example 4.13.

Let GG be the block graph in Figure 4(a). Then, the graphs in Figure 4(b) and 4(c) are two possible block approximations of GG. Observe that the F-threshold of the graphs in Figure 4(b) and 4(c) are 88 and 77, respectively.

(a) Block Graph
(b) Two Block Approximation
(c) Two Block Approximation
Figure 4. Examples of Two Block Approximation

Our next goal is to show

Theorem 4.14.

Let GG be a block graph, then the following quantities are equal

  1. (1)

    max{ftJHH is a triangle path packing of G}\max\{\operatorname{ft}J_{H}\mid\;H\text{ is a triangle path packing of }G\}

  2. (2)

    max{ftJKK is a two block approximation of G}\max\{\operatorname{ft}J_{K}\;\mid K\text{ is a two block approximation of }G\}

  3. (3)

    ft(JG)\operatorname{ft}(J_{G})

  4. (4)

    lct(JG)\operatorname{lct}(J_{G})

  5. (5)

    max{LPK,K is a two block approximation of G}\max\{\operatorname{LP}_{K,\mathbb{Q}}\mid K\text{ is a two block approximation of }G\}

  6. (6)

    LPG,\operatorname{LP}_{G,\mathbb{Q}}.

Lemma 4.15.

In Theorem 4.14 items (1), (2), and (5) are equal.

Proof.

By Corollary 4.11 it follows that (1) is greater than or equal to (2) which is equal to (5). Any triangle path of GG has the two block property by Remark 4.5, and hence can be realized as a subgraph of a two block approximation of GG. This proves that (1) is less than or equal to (2). ∎

To prove Theorem 4.14 it suffices to prove that (1) is equal to (6). Indeed, (1), (2), and (5) are equal by Lemma 4.15, and equality of (1) and (3) would follow from Lemma 2.10 and Proposition 3.12. In particular, this shows that the computation of ft(JG)\operatorname{ft}(J_{G}) is independent of characteristic, and equality of (3) and (4) would follow from Theorem 1.2 and Corollary 3.14.

For the proof of Theorem 4.14 we take the following setup.

Let GG be a connected block graph, KK a two block approximation of GG realizing (5) in Theorem 4.14, and HH a path packing of KK given by Lemma 4.10. Take CC to be the set of cut vertices of GG incident to at least three blocks of GG such that in KK at least one of these blocks becomes a leaf or an isolated connected component. For cCc\in C, let Bc,1,Bc,2,Bc,3,,Bc,mcB_{c,1},B_{c,2},B_{c,3},\ldots,B_{c,m_{c}} denote the blocks of GG incident to cc for some mc3m_{c}\geq 3. Without loss of generality, we may suppose that NH(c)V(Bc,1)V(Bc,2)N_{H}(c)\subseteq V(B_{c,1})\cup V(B_{c,2}) and that Bc,3B_{c,3} becomes a leaf or an isolated connected component of KK. Since Bc,3B_{c,3} is a leaf or isolated component of KK, Proposition 4.10 implies that there exists a vertex vcV(Bc,3)V(H)v_{c}\in V(B_{c,3})\cap V(H) having degree one or zero in HH.

Lemma 4.16.

With the above setup suppose that C=C=\varnothing. Then GG satisfies the conclusions of Theorem 4.14. Moreover, LPK,=LPG,\operatorname{LP}_{K,\mathbb{Q}}=\operatorname{LP}_{G,\mathbb{Q}}.

Proof.

Because C=C=\varnothing, K=G\ell_{K}=\ell_{G} and wK=wGw_{K}=w_{G}. Now, the result follows by Lemma 2.10, Proposition 4.2, and Corollary 4.11. ∎

Lemma 4.17.

With the setup from above let cCc\in C. Then

  1. (1)

    cc does not belong to a triangle of HH,

  2. (2)

    each vertex of NH(c)N_{H}(c) is not contained in a triangle of HH,

  3. (3)

    |NH(c)|=2\lvert N_{H}(c)\rvert=2, and

  4. (4)

    NH(c)Bc,jN_{H}(c)\not\subseteq B_{c,j} for j=1,2j=1,2,

  5. (5)

    no two vertices of CC are adjacent via an edge of HH.

Proof.
  1. (1)

    Suppose by contradiction that cc belongs to a triangle of HH, i.e. that without loss of generality there exists vertices a,bBc,1a,b\in B_{c,1} so that {a,b}\{a,b\}, {a,c}\{a,c\}, and {b,c}\{b,c\} belong to E(H)E(H). Let H~\tilde{H} be the triangle path defined from HH by deleting the edges {c,a}\{c,a\} and {c,b}\{c,b\} and adding the edge {c,vc}\{c,v_{c}\}. Notice that no new cycles are created in H~\tilde{H} after these modifications or else it would contradict GG being a block graph. Moreover,

    ftJH~=ftJH+12\displaystyle\operatorname{ft}J_{\tilde{H}}=\operatorname{ft}J_{H}+\frac{1}{2}

    because in obtaining H~\tilde{H} from HH we delete two edges of HH having weight 12\frac{1}{2}, increase the weight of one edge of HH from weight 12\frac{1}{2} to weight 11 in H~\tilde{H}, and add one edge to H~\tilde{H} having weight 11. This contradicts our choice of KK via Lemma 4.15.

  2. (2)

    Suppose by contradiction that there exists aNH(c)V(Bc,1)a\in N_{H}(c)\cap V(B_{c,1}) and vertices bb and dd of Bc,1B_{c,1} with {a,b}\{a,b\}, {a,d}\{a,d\}, and {b,d}\{b,d\} edges of HH. Construct H~\tilde{H} a triangle path from HH by deleting the edges {c,a}\{c,a\} and {b,d}\{b,d\} and adding the edge {c,vc}\{c,v_{c}\}. Then, ftJH~=12+ftJH\operatorname{ft}J_{\tilde{H}}=\frac{1}{2}+\operatorname{ft}J_{H}; a contradiction.

  3. (3)

    By Item (2) |NH(c)|\lvert N_{H}(c)\rvert is 11 or 22. Suppose by contradiction that NH(c)={a}Bi,1N_{H}(c)=\{a\}\subseteq B_{i,1}. Construct H~\tilde{H} from HH by adding the edge {c,vc}\{c,v_{c}\}. Then, ft(JH~)>ft(JH)\operatorname{ft}(J_{\tilde{H}})>\operatorname{ft}(J_{H}); a contradiction.

  4. (4)

    Suppose by contradiction that {a,b}=NH(c)Bc,1\{a,b\}=N_{H}(c)\subseteq B_{c,1}. By the previous item, aa and bb do not belong to a triangle of HH. If aa and bb have degree two in HH, then define H~\tilde{H} from HH by adding the edge {a,b}\{a,b\} and the edge {c,vc}\{c,v_{c}\}. In the case that aa has degree one and bb has degree two in HH, define H~\tilde{H} from HH by deleting the edge {c,b}\{c,b\} and adding the edges {a,b}\{a,b\} and {c,vc}\{c,v_{c}\}. The case that bb has degree one and aa has degree two in Bc,1B_{c,1} is handled analogously. If aa and bb both have degree one in HH, define H~\tilde{H} from HH by deleting the edge {a,c}\{a,c\} and adding the edges {a,b}\{a,b\} and {c,vc}\{c,v_{c}\}. Notice that in all cases that H~\tilde{H} is a triangle path, a subgraph of GG, and that ft(JH~)>ft(JH)\operatorname{ft}(J_{\tilde{H}})>\operatorname{ft}(J_{H}); a contradiction.

  5. (5)

    Suppose by contradiction that c,cCc,c^{{}^{\prime}}\in C and that {c,c}E(H)\{c,c^{{}^{\prime}}\}\in E(H). Construct H~\tilde{H} a triangle path from HH by deleting the edge {c,c}\{c,c^{{}^{\prime}}\} and adding the edges {c,vc}\{c,v_{c}\} and {c,vc}\{c^{{}^{\prime}},v_{c^{{}^{\prime}}}\}. Then, ft(JH~)>ft(JH)\operatorname{ft}(J_{\tilde{H}})>\operatorname{ft}(J_{H}); a contradiction.

Proof of Theorem 4.14.

Let the setup be as above. We induce on the number of blocks of GG. If the number of blocks of GG is one, then there is nothing to prove. Suppose that the number of blocks of GG is strictly larger than one. If C=C=\varnothing, then we are done by Lemma 4.16. Suppose that CC\neq\varnothing. Let {Gi}i=1r\{G_{i}\}_{i=1}^{r} denote the connected components of GCG\smallsetminus C. For 1ir1\leq i\leq r, let HiH_{i} denote the restriction of HH to GiG_{i}. We show

Claim 1.

For 1ir1\leq i\leq r,

ft(JHi)=LPGi,.\displaystyle\operatorname{ft}(J_{H_{i}})=\operatorname{LP}_{G_{i},\mathbb{Q}}.
Proof of Claim 1.

By induction on number of blocks and the fact that the number of blocks of GiG_{i} is less than the number of blocks of GG, it suffices to show that

ft(JHi)=max{ftJH~iH~i is a triangle path of Gi}\displaystyle\operatorname{ft}(J_{H_{i}})=\max\{\operatorname{ft}J_{\tilde{H}_{i}}\mid\;\tilde{H}_{i}\text{ is a triangle path of }G_{i}\}

for every 1ir1\leq i\leq r. Suppose by contradiction that ft(JHi)<ft(JH~i)\operatorname{ft}(J_{H_{i}})<\operatorname{ft}(J_{\tilde{H}_{i}}) for some fixed choice of 1ir1\leq i\leq r and a triangle path H~i\tilde{H}_{i} of GiG_{i}. Let c1,,ctc_{1},\ldots,c_{t} be the vertices of CC which are adjacent to some vertex of GiG_{i} via an edge of HH; call this vertex cjc_{j}^{{}^{\prime}} for 1jt1\leq j\leq t. Construct H~\tilde{H} from HH as follows.

  1. (i)

    Delete the edges {cj,cj}\{c_{j},c_{j}^{{}^{\prime}}\} for 1jt1\leq j\leq t .

  2. (ii)

    Add the edges {cj,vcj}\{c_{j},v_{c_{j}}\} for 1jt1\leq j\leq t .

  3. (iii)

    Replace HiH_{i} on GiG_{i} by H~i\tilde{H}_{i} .

Now, H~\tilde{H} is a triangle path. Indeed parts (i) and (ii) of the construction do not create any cycles in H~\tilde{H}; otherwise, GG could not be a block graph. After performing parts (i) and (ii), HiH_{i} forms its own connected component by Lemma 4.17 item (4). Thus after performing operation (iii) H~\tilde{H} will remain a triangle path. Clearly, ft(JH~)>ft(JH)\operatorname{ft}(J_{\tilde{H}})>\operatorname{ft}(J_{H}) which contradicts our choice of HH, Lemma 4.15.

Observe that

ft(JH)LPG,\displaystyle\operatorname{ft}(J_{H})\leq\operatorname{LP}_{G,\mathbb{Q}} i=1rLPGi,+2|C|\displaystyle\leq\sum_{i=1}^{r}\operatorname{LP}_{G_{i},\mathbb{Q}}+2\lvert C\rvert (Lemma 4.9)\displaystyle(\text{Lemma }\ref{lem_LP_deleting_vertices})
=i=1rft(JHi)+2|C|\displaystyle=\sum_{i=1}^{r}\operatorname{ft}(J_{H_{i}})+2\lvert C\rvert (Claim 1)\displaystyle(\text{Claim }\ref{claim_ft_H_i_equals_LP_G_i})
=ft(JH)\displaystyle=\operatorname{ft}(J_{H}) (Lemma 4.17 parts (1),(2),(3),(5)\displaystyle(\text{Lemma }\ref{lem_vertices_incident_leaves_blocks}\text{ parts }\eqref{item_vertex_not_in_triangle},\,\eqref{item_nbrs_not_in_triangle},\,\eqref{item_vertex_has_two_nbrs},\,\eqref{item_vertices_disconnected}
and Corollary 4.8).\displaystyle\text{ and Corollary }\ref{cor_ft_deleting_vertices}).

This proves equality of (1) and (6) and completes the proof. ∎

Corollary 4.18.

Conjecture 1.3 is true when GG is a block graph.

Corollary 4.19.

If GG is a tree, then

r(G)=ft(JG).\displaystyle r(G)=\operatorname{ft}(J_{G}).
Proof.

By Theorem 4.14, ft(JG)=max{ftJKK is a two block approximation of G}\operatorname{ft}(J_{G})=\max\{\operatorname{ft}J_{K}\;\mid K\text{ is a two block approximation of }G\}. Any two block approximation KK of GG is a semi-path of GG since GG is a tree, and hence ft(JG)r(G)\operatorname{ft}(J_{G})\leq r(G). The inequality r(G)ft(JG)r(G)\leq\operatorname{ft}(J_{G}) is Equation (15). ∎

In the next section we show the stronger statement that for a tree GG

r(G)=ht(JG).\displaystyle r(G)=\operatorname{ht}(J_{G}).

4.3. Remark on the Case of Chordal Graphs

The next proposition shows that if Conjecture 1.3 is true for chordal graphs, then the proof techniques used for block graphs will not suffice. We show:

Proposition 4.20.

There exists a chordal graph GG such that

max{ft(JH)HG is triangle path packing}<LPG,.\displaystyle\max\{\operatorname{ft}(J_{H})\mid H\subseteq G\text{ is triangle path packing}\}<\operatorname{LP}_{G,\mathbb{Q}}.
Proof.

Let GG be the graph below. Let G1G_{1} and G2G_{2} be the subgraphs of GG induced on the vertices {1,2,3,4,5,6,13,14}\{1,2,3,4,5,6,13,14\} and {7,8,9,10,11,12,15,16}\{7,8,9,10,11,12,15,16\}, respectively. Let HH be a triangle path packing of GG maximal wrt ft(JH)\operatorname{ft}(J_{H}). Let H1H_{1} and H2H_{2} be the restriction of HH to G1G_{1} and G2G_{2} respectively. For i=1,2i=1,2, let tit_{i}, eie_{i} denote the number of triangles of HiH_{i} and edges of HiH_{i} not belonging to any triangle, respectively.

1122334455661313141477889910101111121215151616
Figure 5. Chordal Graph
Claim 2.

e1+32t1=e2+32t2e_{1}+\frac{3}{2}t_{1}=e_{2}+\frac{3}{2}t_{2}.

Proof of Claim 2.

Without loss of generality suppose e1+32t1<e2+32t2e_{1}+\frac{3}{2}t_{1}<e_{2}+\frac{3}{2}t_{2}. Construct H~\tilde{H} from HH by deleting H1H_{1} and replicating H2H_{2} on V(G1)V(G_{1}). Then, H~\tilde{H} is a triangle path packing with ft(JH)<ft(JH~)\operatorname{ft}(J_{H})<\operatorname{ft}(J_{\tilde{H}}); a contradiction.

It follows from Proposition 4.6 and Claim 2 that ft(JH)\operatorname{ft}(J_{H}) is an integer. However, LPG,=14.5\operatorname{LP}_{G,\mathbb{Q}}=14.5 which can be realized by the following assignment

(a[1,2],b[1,2])\displaystyle(a_{[1,2]},b_{[1,2]}) =(0.0,1.0)\displaystyle=(0.0,1.0) (a[9,10],b[9,10])\displaystyle(a_{[9,10]},b_{[9,10]}) =(0.0,0.5)\displaystyle=(0.0,0.5) (a[4,13],b[4,13])\displaystyle(a_{[4,13]},b_{[4,13]}) =(0.0,0.5)\displaystyle=(0.0,0.5)
(a[2,3],b[2,3])\displaystyle(a_{[2,3]},b_{[2,3]}) =(0.0,1.0)\displaystyle=(0.0,1.0) (a[10,11],b[10,11])\displaystyle(a_{[10,11]},b_{[10,11]}) =(0.0,0.5)\displaystyle=(0.0,0.5) (a[7,9],b[7,9])\displaystyle(a_{[7,9]},b_{[7,9]}) =(0.0,0.0)\displaystyle=(0.0,0.0)
(a[3,4],b[3,4])\displaystyle(a_{[3,4]},b_{[3,4]}) =(0.0,0.5)\displaystyle=(0.0,0.5) (a[11,12],b[11,12])\displaystyle(a_{[11,12]},b_{[11,12]}) =(0.5,0.5)\displaystyle=(0.5,0.5) (a[7,10],b[7,10])\displaystyle(a_{[7,10]},b_{[7,10]}) =(0.0,0.0)\displaystyle=(0.0,0.0)
(a[4,5],b[4,5])\displaystyle(a_{[4,5]},b_{[4,5]}) =(0.5,0.5)\displaystyle=(0.5,0.5) (a[1,3],b[1,3])\displaystyle(a_{[1,3]},b_{[1,3]}) =(0.0,0.0)\displaystyle=(0.0,0.0) (a[7,12],b[7,12])\displaystyle(a_{[7,12]},b_{[7,12]}) =(0.5,0.0)\displaystyle=(0.5,0.0)
(a[5,6],b[5,6])\displaystyle(a_{[5,6]},b_{[5,6]}) =(0.0,0.5)\displaystyle=(0.0,0.5) (a[1,6],b[1,6])\displaystyle(a_{[1,6]},b_{[1,6]}) =(1.0,0.0)\displaystyle=(1.0,0.0) (a[9,15],b[9,15])\displaystyle(a_{[9,15]},b_{[9,15]}) =(0.5,0.0)\displaystyle=(0.5,0.0)
(a[6,7],b[6,7])\displaystyle(a_{[6,7]},b_{[6,7]}) =(0.5,0.0)\displaystyle=(0.5,0.0) (a[3,6],b[3,6])\displaystyle(a_{[3,6]},b_{[3,6]}) =(0.0,0.0)\displaystyle=(0.0,0.0) (a[10,12],b[10,12])\displaystyle(a_{[10,12]},b_{[10,12]}) =(0.0,0.5)\displaystyle=(0.0,0.5)
(a[7,8],b[7,8])\displaystyle(a_{[7,8]},b_{[7,8]}) =(0.5,0.5)\displaystyle=(0.5,0.5) (a[3,13],b[3,13])\displaystyle(a_{[3,13]},b_{[3,13]}) =(0.0,0.5)\displaystyle=(0.0,0.5) (a[10,15],b[10,15])\displaystyle(a_{[10,15]},b_{[10,15]}) =(0.5,0.0)\displaystyle=(0.5,0.0)
(a[8,9],b[8,9])\displaystyle(a_{[8,9]},b_{[8,9]}) =(0.5,0.5)\displaystyle=(0.5,0.5) (a[4,6],b[4,6])\displaystyle(a_{[4,6]},b_{[4,6]}) =(0.0,0.0)\displaystyle=(0.0,0.0) (a[13,14],b[13,14])\displaystyle(a_{[13,14]},b_{[13,14]}) =(0.0,1.0)\displaystyle=(0.0,1.0)
(a[15,16],b[15,16])\displaystyle(a_{[15,16]},b_{[15,16]}) =(1.0,0.0)\displaystyle=(1.0,0.0)

5. Binomial Edge Ideals of König type

It is a well-known fact that bipartite graphs satisfy the König–Egervary property, i.e. the size of a minimal vertex cover equals the size of a maximal matching. For (monomial) edge ideals the König-Egervary property translates into the statement that the monomial grade of the ideal is equal to its height. In [HHM22] the authors introduce the notion of graded ideals of König type as a way to extend the König–Egervary property to other homogeneous ideals of a polynomial ring. Because in this section we are only concerned with binomial edge ideals of König type, we (by abuse of notation) define a graph GG to be of König type if it satisfies the following characterization:

Definition 5.1 ([HHM22] Theorem 3.5).

A graph GG is said to be of König type if r(G)=ht(JG)r(G)=\operatorname{ht}(J_{G}).

In this section we prove that every graph of König type satisfies Conjecture 1.3, we give a new combinatorial characterization of the König type property, and we use this characterization to prove that every tree is of König type, and hence give a second proof that trees satisfy Conjecture 1.3. Lastly, we give an example of a bipartite graph which is not of König type.

Proposition 5.2.

If GG is of König type, then ft(JG)=r(G)\operatorname{ft}(J_{G})=r(G). In particular, Conjecture 1.3 holds for every graph GG that is of König type.

Proof.

Follows immediately from Equation (15) and Definition 5.1. ∎

The following lemma gives a combinatorial characterization for a graph to be of König type. We use this characterization to show that trees are of König type.

Lemma 5.3.

Let GG be a graph. Then, GG is of König type if and only if there exists a semi-path PP and a choice of distinct vertices TV(G)T\subseteq V(G), possibly empty, with the property that

  1. (1)

    r(G)=|E(P)|r(G)=\lvert E(P)\rvert,

  2. (2)

    for all vTv\in T, vv has degree two in PP,

  3. (3)

    for all v,wTv,w\in T, {v,w}\{v,w\} is not an edge of PP,

  4. (4)

    the number of connected components and the number of vertices of GT^\widehat{G\smallsetminus T} equals the number of connected components and the number of vertices of PT^\widehat{P\smallsetminus T}, respectively.

Proof.

We prove the backward implication. Suppose that a semi-path PP of GG and a choice of distinct vertices TV(P)T\subseteq V(P) have been chosen satisfying properties (1) through (4). It is enough to show that ht(JG)r(G)\operatorname{ht}(J_{G})\leq r(G).

Let c:=cG(T)c:=c_{G}(T) be the number of connected components of GT^\widehat{G\smallsetminus T} and G1,,GcG_{1},\ldots,G_{c} its connected components. Let Q:=({Xi,Yi}iT,JG~1,,JG~c)Q:=(\{X_{i},Y_{i}\}_{i\in T},J_{\tilde{G}_{1}},\ldots,J_{\tilde{G}_{c}}) the ideal as defined in Proposition 2.4. Then,

ht(Q)\displaystyle\operatorname{ht}(Q) =2|T|+i=1c(|V(Gi)|1)\displaystyle=2\cdot\lvert T\rvert+\sum_{i=1}^{c}(\lvert V(G_{i})\rvert-1)

Let PiP_{i} denote the subgraph of GiG_{i} given by GiPG_{i}\cap P for 1ic1\leq i\leq c. Property (4) implies that PiP_{i} is a Hamiltonian path of GiG_{i}, and in particular that

|E(Pi)|=|V(Gi)|1.\displaystyle\lvert E(P_{i})\rvert=\lvert V(G_{i})\rvert-1.

Let EE^{{}^{\prime}} denote the set of edges of PP which are incident to a vertex of TT. By assumptions (2) and (3), |E|=2|T|\lvert E^{{}^{\prime}}\rvert=2\cdot\lvert T\rvert. The edges of PP can be partitioned into those which belong to some PiP_{i} and those which belong to EE^{{}^{\prime}}. We compute that

r(G)\displaystyle r(G) =|E(P)|(by (1))\displaystyle=\lvert E(P)\rvert\phantom{aaaaaaaaaaaaaaaaaaaa}(\text{by \eqref{item_konig_1}})
=|E|+i=1c|E(Pi)|\displaystyle=\lvert E^{{}^{\prime}}\rvert+\sum_{i=1}^{c}\lvert E(P_{i})\rvert
=2|T|+i=1c(|V(Gi)|1)\displaystyle=2\cdot\lvert T\rvert+\sum_{i=1}^{c}(\lvert V(G_{i})\rvert-1)
=ht(Q).\displaystyle=\operatorname{ht}(Q).

This proves that htJGhtQ=r(G)\operatorname{ht}J_{G}\leq\operatorname{ht}Q=r(G).

Next, suppose that GG is of König type. Let PP be a semi-path realizing r(G)r(G). Consider the vector 𝐚=(ae)eE(G)\mathbf{{a}}=(a_{e})_{e\in E(G)} given by

ae={1,eE(P)0,otherwise\displaystyle a_{e}=\begin{cases}1,&e\in E(P)\\ 0,&\text{otherwise}\end{cases}

Then, 𝐚:=(ae)\mathbf{{a}}:=(a_{e}) is an optimal solution of the linear program (4). Let (𝐛,𝐜):=({bv}vV,{cU}U𝒜)(\mathbf{{b}},\mathbf{{c}}):=(\{b_{v}\}_{v\in V},\{c_{U}\}_{U\in\mathcal{A}}) be an optimal solution with values in \mathbb{Z} of the linear program (7) having the form guaranteed by Lemma 3.3, and let TT and U1,,UcU_{1},\ldots,U_{c} be as in Lemma 3.3.

Let AA be the below (2|V|1)×|E|\left(2^{\lvert V\rvert}-1\right)\times\lvert E\rvert matrix, 𝐗\mathbf{X} the vector of indeterminates of length |E|\lvert E\rvert, 𝟏\mathbf{{1}} vector of 11’s of length |E|\lvert E\rvert, and 𝐝\mathbf{{d}} the vector of length 2|V|12^{\lvert V\rvert-1} realizing the linear program (4), i.e. algorithm (4) can be expressed as

(22) maximizeZPP,G:=𝟏T𝐗subjectto:𝐗𝕜0A𝐗𝐝\begin{split}\operatorname*{maximize}\hskip 3.00003pt&Z_{\operatorname{PP},G}:=\mathbf{{1}}^{\operatorname{T}}\,\mathbf{X}\\ \operatorname{subject\,to:}&\phantom{}\\ &\mathbf{X}\in\mathbbm{k}_{\geq 0}\\ &A\cdot\mathbf{{X}}\leq\mathbf{{d}}\end{split}

Let AiA_{i} denote the ii-th column of AA, and let ajTa_{j}^{\operatorname{T}} denote the jj-th row of AA.

Because GG is of König type the cost of these optimal solutions are equal, and complementary slackness, Theorem 2.3, holds between these pair of solutions.

We prove part (2). Choose vTv\in T, equivalently, bv=1b_{v}=1. Let avTa_{v}^{\operatorname{T}} denote the row of AA corresponding to vertex vv. By complementary slackness Theorem 2.3.(2) we have that

2=avT𝐚=e incident to vae.\displaystyle 2=a_{v}^{\operatorname{T}}\mathbf{{a}}=\sum_{e\text{ incident to }v}a_{e}.

This shows that there are two edges of PP incident with vv.

We prove part (3). Let vTv\in T, i.e. bv=1b_{v}=1, and e={v,w}e=\{v,w\} an edge of PP, then we want to show that wTw\notin T. Let AeA_{e} denote column of AA corresponding to the edge ee. Theorem 2.3.(1) implies that

(23) 1=(𝐛,𝐜)TAe=bv+bw+U𝒜eG[U]cU1=(\mathbf{{b}},\mathbf{{c}})^{\operatorname{T}}A_{e}=b_{v}+b_{w}+\sum_{\begin{subarray}{c}U\in\mathcal{A}\\ e\in G[U]\end{subarray}}c_{U}

Since bv=1b_{v}=1, Equation (23) implies that bw=0b_{w}=0. Hence wTw\notin T by Lemma 3.3.

We prove part (4). Fix 1ic1\leq i\leq c. Let aUiTa_{U_{i}}^{\operatorname{T}} denote the row of AA corresponding to UiU_{i}. By the construction of 𝐜\mathbf{{c}}, cUi=1c_{U_{i}}=1. Theorem 2.3.(2) implies that

|Ui|1=aUiT𝐚=|E(P)E(Ui)|.\displaystyle\lvert U_{i}\rvert-1=a_{U_{i}}^{\operatorname{T}}\cdot\mathbf{{a}}=\lvert E(P)\cap E(U_{i})\rvert.

This shows that PP restricted to UiU_{i} induces a Hamiltonian path for every 1ic1\leq i\leq c and (4) follows. ∎

5.1. Trees are of König Type

We use Lemma 5.3 to show that trees are of König type. The proof idea is to pick a maximal semi-path and to recursively delete vertices incident with edges not belonging to the semi-path so at the end we are left with no edges between connected components of the semi-path and no two deleted vertices are adjacent.

We set up notation below. Throughout this subsection let GG be a tree. Let P1,,PP_{1},\ldots,P_{\ell} be disjoint paths in GG with r(G)=i=1|E(Pi)|r(G)=\sum_{i=1}^{\ell}\lvert E(P_{i})\rvert. Put

EL:={{v,w}E(G)\displaystyle E_{L}:=\bigg{\{}\vphantom{y^{2}}\{v,w\}\in E(G)\mid vPi and wPj for some 1i<j\displaystyle v\in P_{i}\text{ and }w\in P_{j}\text{ for some }1\leq i<j\leq\ell
or vPi for some 1i and wPj for any 1j}.\displaystyle\text{ or }\,v\in P_{i}\text{ for some }1\leq i\leq\ell\text{ and }w\notin P_{j}\text{ for any }1\leq j\leq\ell\bigg{\}}.

Put VL:={V(e)eEL}V_{L}:=\{V(e)\mid e\in E_{L}\} and

Bv\displaystyle B_{v} :={vVLv is an initial or terminal vertex of some Pi or vPi for any 1i}.\displaystyle:=\{v\in V_{L}\mid v\text{ is an initial or terminal vertex of some }P_{i}\text{ or }v\notin P_{i}\text{ for any }1\leq i\leq\ell\}.
Definition 5.4.

Let vV(G)v\in V(G). A subtree of GG based at vv, denoted TvT_{v}, is the subtree of GG constructed as follows:

Tv,k\displaystyle T_{v,k} :=,k0\displaystyle:=\varnothing,k\leq 0
Tv,1\displaystyle T_{v,1} :={v}\displaystyle:=\{v\}
Tv,2k\displaystyle T_{v,2k} :={vVLwT2k1 with {v,w}EL},k1\displaystyle:=\{v\in V_{L}\mid\exists w\in T_{2k-1}\text{ with }\{v,w\}\in E_{L}\},k\geq 1
Tv,2k+1\displaystyle T_{v,2k+1} :={vV(Pi)wT2k with {v,w}E(Pi) for some 1i},k1.\displaystyle:=\{v\in V(P_{i})\mid\exists w\in T_{2k}\text{ with }\{v,w\}\in E(P_{i})\text{ for some }1\leq i\leq\ell\},k\geq 1.

Then, TvT_{v} is the induced subgraph of GG on the set of vertices kTk\bigcup_{k\in\mathbb{Z}}T_{k}.

Lemma 5.5.

Let TvT_{v} be the subtree of GG based at vv. For NNN\neq N^{{}^{\prime}}, Tv,NTv,N=T_{v,N}\cap T_{v,N^{{}^{\prime}}}=\varnothing.

Proof.

Clear because Tv,N{wV(G)d(v,w)=N}T_{v,N}\subseteq\{w\in V(G)\mid d(v,w)=N\}, and the later sets are disjoint for NNN\neq N^{{}^{\prime}} since GG is a tree. ∎

The following definition and lemma can be viewed as analogues of the the statement that there are no augmenting paths with respect to a choice of maximal matching in a bipartite graph.

Definition 5.6.

A (v,w)(v,w)-alternating path with respect to PP (respectively a vv-alternating path with respect to PP) is a set of distinct vertices v1,v2,,v2k1,v2kv_{1},v_{2},\ldots,v_{2k-1},v_{2k} where 1k1\leq k\in\mathbb{Z} satisfying

  1. (1)

    {v2i1,v2i}EL\{v_{2i-1},v_{2i}\}\in E_{L} for 1ik1\leq i\leq k,

  2. (2)

    {v2i,v2i+1}E(Pji)\{v_{2i},v_{2i+1}\}\in E(P_{j_{i}}) for 1ik11\leq i\leq k-1 and for some 1ji1\leq j_{i}\leq\ell,

  3. (3)

    v1=vv_{1}=v and v2k=wv_{2k}=w (respectively, v1=vv_{1}=v).

Lemma 5.7.

Let GG be a graph. Then, GG does not have an alternating path v1v_{1}, v2,,v_{2},\ldots, v2k1v_{2k-1}, v2kv_{2k} with both v1Bvv_{1}\in B_{v} and v2kBvv_{2k}\in B_{v}.

Proof.

If such an alternating path were to exist, then we could construct a semi-path having more edges than PP by adding to PP the edges belonging to (1) of Definition 5.6 and deleting from PP the edges belonging to (2) of Definition 5.6. ∎

Algorithm 5.8.

Input: Semi-path PP of a tree GG Output: Ti=1V(Pi)T\subseteq\bigcup_{i=1}^{\ell}V(P_{i}) satisfying conditions (2) - (3) in Lemma 5.3, and every edge of ELE_{L} contains a vertex belonging to TT.

  1. (1)

    Initialize i=0i=0 and T0,k=T_{0,k}=\varnothing for kk\in\mathbb{Z}.

  2. (2)

    While Bvn=0ik1Tn,kB_{v}\not\subseteq\bigcup\limits_{n=0}^{i}\bigcup\limits_{k\geq 1}T_{n,k}, repeat the following steps:

    1. (a)

      Increment ii,

    2. (b)

      Pick a vertex viBvn=0i1k1Tn,kv_{i}\in B_{v}\smallsetminus\bigcup\limits_{n=0}^{i-1}\bigcup\limits_{k\geq 1}T_{n,k}.

    3. (c)

      Let TiT_{i} denote the subtree of GG based at viv_{i}, and for kk\in\mathbb{Z}, let Ti,kT_{i,k} denote the subsets of TiT_{i} as constructed in Definition 5.4.

  3. (3)

    Put S:=VLn=0ik1Tn,kS:=V_{L}\smallsetminus\bigcup\limits_{n=0}^{i}\bigcup\limits_{k\geq 1}T_{n,k}.

  4. (4)

    While SS\neq\varnothing, repeat the following steps:

    1. (a)

      Increment ii,

    2. (b)

      Choose 1j1\leq j\leq\ell minimally so that there exists a vertex viV(Pj)Sv_{i}\in V(P_{j})\cap S,

    3. (c)

      Let TiT_{i} denote the subtree of GG based at viv_{i}, and for kk\in\mathbb{Z}, let Ti,kT_{i,k} denote the subsets of TiT_{i} as constructed in Definition 5.4,

    4. (d)

      Go back to part (3) of Algorithm 5.8.

  5. (5)

    Put N:=iN:=i.

  6. (6)

    For 1iN1\leq i\leq N, define Tieven:=kTi,2kT^{\operatorname{even}}_{i}:=\bigcup_{k\in\mathbb{Z}}T_{i,2k} and Tiodd:=kTi,2k+1T^{\operatorname{odd}}_{i}:=\bigcup_{k\in\mathbb{Z}}T_{i,2k+1}.

  7. (7)

    Put T:=i=1NTievenT:=\bigcup_{i=1}^{N}T_{i}^{\operatorname{even}}.

Theorem 5.9.

Algorithm 5.8 is correct, and property (4) of Lemma 5.3 is satisfied by the choice of TT. Consequently, GG is of König type.

Proof.

First, we remark that each while-loop of Algorithm 5.8 terminates because BvB_{v} and VLV_{L} are finite, and {n=0ik1Tn,k}i=1N\{\bigcup_{n=0}^{i}\bigcup_{k\geq 1}T_{n,k}\}_{i=1}^{N} form a strictly ascending sequence of sets each containing more elements of BvVLB_{v}\cup V_{L} than their predecessors.

Claim 3.

For iji\neq j, TievenTjodd=T^{\operatorname{even}}_{i}\cap T^{\operatorname{odd}}_{j}=\varnothing.

Proof of Claim 3.

Suppose by contradiction that TievenTjoddT^{\operatorname{even}}_{i}\cap T^{\operatorname{odd}}_{j}\neq\varnothing. Choose aa even and bb odd satisfying the following

  1. (i)

    there exists wTi,aTj,bw\in T_{i,a}\cap T_{j,b},

  2. (ii)

    if q1q_{1} (respectively q2q_{2}) is the (vi,w)(v_{i},w)-alternating path (respectively (vj,w)(v_{j},w)-alternating path), then the vertices of q1q_{1} and q2q_{2} intersect only at ww.

Let ww^{{}^{\prime}} be the vertex of q2q_{2} adjacent to ww. Then, there is a (vi,vj)(v_{i},v_{j})-augmenting path by concatenating q1q_{1} and q2q_{2} along the edge {w,w}\{w,w^{{}^{\prime}}\}. Without loss of generality, suppose that i<ji<j, then vjV(Ti)v_{j}\in V(T_{i}) by construction of TiT_{i}, but this contradicts the choice of vjVLn=1j1k1Tn,kv_{j}\in V_{L}\smallsetminus\bigcup_{n=1}^{j-1}\bigcup_{k\geq 1}T_{n,k} which proves the claim.

Ti=1V(Pi)T\subseteq\bigcup_{i=1}^{\ell}V(P_{i}): Suppose by contradiction that there exists wTi=1V(Pi)w\in T\smallsetminus\bigcup_{i=1}^{\ell}V(P_{i}). Since wBvw\in B_{v} and wTievenw\in T_{i}^{\operatorname{even}} for some 1iN1\leq i\leq N, it must be the case that viBvv_{i}\in B_{v} by part 2 of Algorithm 5.8. This contradicts Lemma 5.7.

TT satisfies (2) of Lemma 5.3: TT does not contain any initial nor terminal points of any PiP_{i} by a similar argument to the previous paragraph.

TT satisfies (3) of Lemma 5.3: If not, we would contradict TievenTjodd=T_{i}^{\operatorname{even}}\cap T_{j}^{\operatorname{odd}}=\varnothing.

TT satisfies (4) of Lemma 5.3: The construction of TT shows that for each edge {v,w}EL\{v,w\}\in E_{L} that

  1. (i)

    wTw\in T if vi=1Piv\notin\bigcup_{i=1}^{\ell}P_{i}, and that

  2. (ii)

    vTv\in T or wTw\in T if {v,w}EL\{v,w\}\in E_{L}.

Thus, as sets GT^\widehat{G\smallsetminus T} and PT^\widehat{P\smallsetminus T} are equal, and hence condition (4) is trivially satisfied. Hence by Lemma 5.3 GG is of König type. ∎

5.2. Example of Bipartite Graph Not of König Type

The following example shows that there exist bipartite graphs which are not of König type.

Example 5.10.

For this example let GG be the graph in Figure 6.

11223344556677889910101111121213131414
Figure 6. Non-König type Non-Unmixed Bipartite Graph

One can check that r(G)=12r(G)=12, LPG,=12.5\operatorname{LP}_{G,\mathbb{Q}}=12.5, and htJG=13\operatorname{ht}J_{G}=13. One realization of LPG,\operatorname{LP}_{G,\mathbb{Q}} is via the following assignment

(a[1,2],b[1,2])\displaystyle(a_{[1,2]},b_{[1,2]}) =(0.5,0.5)\displaystyle=(0.5,0.5) (a[6,7],b[6,7])\displaystyle(a_{[6,7]},b_{[6,7]}) =(0.0,0.5)\displaystyle=(0.0,0.5) (a[11,12],b[11,12])\displaystyle(a_{[11,12]},b_{[11,12]}) =(1.0,0.0)\displaystyle=(1.0,0.0)
(a[2,3],b[2,3])\displaystyle(a_{[2,3]},b_{[2,3]}) =(0.5,0.5)\displaystyle=(0.5,0.5) (a[7,8],b[7,8])\displaystyle(a_{[7,8]},b_{[7,8]}) =(0.5,0.0)\displaystyle=(0.5,0.0) (a[13,14],b[13,14])\displaystyle(a_{[13,14]},b_{[13,14]}) =(0.5,0.5)\displaystyle=(0.5,0.5)
(a[3,4],b[3,4])\displaystyle(a_{[3,4]},b_{[3,4]}) =(0.5,0.5)\displaystyle=(0.5,0.5) (a[8,9],b[8,9])\displaystyle(a_{[8,9]},b_{[8,9]}) =(0.5,0.5)\displaystyle=(0.5,0.5) (a[1,6],b[1,6])\displaystyle(a_{[1,6]},b_{[1,6]}) =(0.0,0.5)\displaystyle=(0.0,0.5)
(a[4,5],b[4,5])\displaystyle(a_{[4,5]},b_{[4,5]}) =(0.0,0.5)\displaystyle=(0.0,0.5) (a[9,10],b[9,10])\displaystyle(a_{[9,10]},b_{[9,10]}) =(0.5,0.5)\displaystyle=(0.5,0.5) (a[7,12],b[7,12])\displaystyle(a_{[7,12]},b_{[7,12]}) =(0.0,1.0)\displaystyle=(0.0,1.0)
(a[5,6],b[5,6])\displaystyle(a_{[5,6]},b_{[5,6]}) =(0.5,0.5)\displaystyle=(0.5,0.5) (a[10,11],b[10,11])\displaystyle(a_{[10,11]},b_{[10,11]}) =(0.5,0.0)\displaystyle=(0.5,0.0) (a[4,13],b[4,13])\displaystyle(a_{[4,13]},b_{[4,13]}) =(0.5,0.0)\displaystyle=(0.5,0.0)
(a[10,13],b[10,13])\displaystyle(a_{[10,13]},b_{[10,13]}) =(0.0,0.5)\displaystyle=(0.0,0.5)

This graph is not unmixed. For example, {1,3}\{1,3\} is a cut set of GG and the corresponding prime ideal has height 1414. This raises

Question 5.11.

Is every unmixed bipartite graph of König type?

Bipartite graphs having unmixed binomial edge ideal were studied in [BMS18]. Bipartite graphs with Cohen–Macaulay binomial edge ideal were characterized in [BMS18, Theorem 6.1]. As a consequence of their characterization, Cohen–Macaulay bipartite graphs are traceable. Herzog et al. proved that traceable implies König type [HHM22, Proposition 3.6.(a)], and thus in particular Cohen–Macaulay bipartite graphs are of König type.

6. LPG,\operatorname{LP}_{G,\mathbb{Q}} and Traceable Graphs

In this section, we consider the relationship between maximality of LPG,\operatorname{LP}_{G,\mathbb{Q}} of a graph and the graph being traceable. Recall that a graph is called traceable if it has a Hamiltonian path, Definition 2.6.

Proposition 6.1.

If GG is a traceable graph on nn vertices, then ft(JG)=LPG,=n1\operatorname{ft}(J_{G})=\operatorname{LP}_{G,\mathbb{Q}}=n-1.

Proof.

We have the graph containments PnGKnP_{n}\subseteq G\subseteq K_{n}. The result follows from monotonicity of LPG,𝕜\operatorname{LP}_{G,\mathbbm{k}} and ft(JPn)=LPKn,=n1\operatorname{ft}(J_{P_{n}})=\operatorname{LP}_{K_{n},\mathbb{Q}}=n-1. ∎

Remark 6.2.

The above proof shows that for any graph GG on nn vertices, ft(JG)n1\operatorname{ft}(J_{G})\leq n-1.

Example 6.3 shows that there exists a graph GG with LPG,=|V|1\operatorname{LP}_{G,\mathbb{Q}}=\lvert V\rvert-1 yet GG is non-traceable. Computations in Macaulay2 [GS] suggest that ft(JG)\operatorname{ft}(J_{G}) may agree with |V|1\lvert V\rvert-1 independent of characteristic in Example 6.3.

Example 6.3.

Throughout this example let GG denote the graph in Figure 7. Observe that GG is not traceable.

12468910375
Figure 7. Non-traceable graph

Observe that deleting edges {7,8}\{7,8\} and {8,9}\{8,9\} produces a triangle path as a subgraph and that this triangle path has F-threshold equal to 8.58.5 by Proposition 4.6. It follows from Proposition 2.10 that 8.5ft(JG)8.5\leq\operatorname{ft}(J_{G}). However, it can be computed that LPG,=9\operatorname{LP}_{G,\mathbb{Q}}=9 and one such realization is as follows

(a[1,2],b[1,2])\displaystyle(a_{[1,2]},b_{[1,2]}) =(1.0,0.0)\displaystyle=(1.0,0.0) (a[3,5],b[3,5])\displaystyle(a_{[3,5]},b_{[3,5]}) =(1.0,0.0)\displaystyle=(1.0,0.0) (a[7,8],b[7,8])\displaystyle(a_{[7,8]},b_{[7,8]}) =(0.5,0.0)\displaystyle=(0.5,0.0)
(a[2,3],b[2,3])\displaystyle(a_{[2,3]},b_{[2,3]}) =(0.5,0.0)\displaystyle=(0.5,0.0) (a[4,6],b[4,6])\displaystyle(a_{[4,6]},b_{[4,6]}) =(0.5,0.5)\displaystyle=(0.5,0.5) (a[7,9],b[7,9])\displaystyle(a_{[7,9]},b_{[7,9]}) =(0.5,0.0)\displaystyle=(0.5,0.0)
(a[2,4],b[2,4])\displaystyle(a_{[2,4]},b_{[2,4]}) =(0.5,0.0)\displaystyle=(0.5,0.0) (a[5,7],b[5,7])\displaystyle(a_{[5,7]},b_{[5,7]}) =(1.0,0.0)\displaystyle=(1.0,0.0) (a[8,9],b[8,9])\displaystyle(a_{[8,9]},b_{[8,9]}) =(0.5,0.0)\displaystyle=(0.5,0.0)
(a[3,4],b[3,4])\displaystyle(a_{[3,4]},b_{[3,4]}) =(0.0,0.5)\displaystyle=(0.0,0.5) (a[6,8],b[6,8])\displaystyle(a_{[6,8]},b_{[6,8]}) =(0.5,0.5)\displaystyle=(0.5,0.5) (a[9,10],b[9,10])\displaystyle(a_{[9,10]},b_{[9,10]}) =(1.0,0.0)\displaystyle=(1.0,0.0)

This bilabeling corresponds to the monomial

i=08Xii=19Yi\displaystyle\prod_{i=0}^{8}X_{i}\cdot\prod_{i=1}^{9}Y_{i}

belonging to the polynomial

(f01f131/2f121/2f231/2f24f35f46f57f671/2f681/2f781/2f89)pe1JG9(pe1)\displaystyle\big{(}f_{01}f_{13}^{1/2}f_{12}^{1/2}f_{23}^{1/2}f_{24}f_{35}f_{46}f_{57}f_{67}^{1/2}f_{68}^{1/2}f_{78}^{1/2}f_{89}\big{)}^{p^{e}-1}\in J_{G}^{9(p^{e}-1)}

which has coefficient given by the expression

a=0N/2((N/2a)(NN/2+a)(Na))2\displaystyle\sum_{a=0}^{N/2}\bigg{(}\binom{N/2}{a}\binom{N}{N/2+a}\binom{N}{a}\bigg{)}^{2}

where N:=pe1N:=p^{e}-1. For small values of pp and ee it can be checked that this coefficient does not vanish mod pp. This suggests that ft(JG)=9\operatorname{ft}(J_{G})=9 for any prime pp.

The following proposition shows that under additional hypotheses on GG that a converse of Proposition 6.1 holds.

Proposition 6.4.

If GG is a connected graph on n+3n+3 vertices with at least three vertices having degree one (in particular GG is not traceable), then LPG,n+3/2<n+2\operatorname{LP}_{G,\mathbb{Q}}\leq n+3/2<n+2.

Proof.

Observe that GG is a subgraph of a complete graph HH on nn vertices with three whiskers. Suppose that the whiskers of HH do not share a common vertex, then by Corollary 4.11, LPH,=n+32\operatorname{LP}_{H,\mathbb{Q}}=n+\frac{3}{2}. If two of the three whiskers or all three of the whiskers of HH share a common vertex, then we can list the two block approximations of HH and show that LPH,n+1\operatorname{LP}_{H,\mathbb{Q}}\leq n+1 by Theorem 4.14. Since LPG,LPH,\operatorname{LP}_{G,\mathbb{Q}}\leq\operatorname{LP}_{H,\mathbb{Q}} the result follows. ∎

It can be shown that the graph in Example 6.3 is not unmixed. This raises

Question 6.5.

If GG is an unmixed, connected graph and LPG,=|V(G)|1\operatorname{LP}_{G,\mathbb{Q}}=\lvert V(G)\rvert-1, then is GG traceable?

Note that GG being unmixed and connected implies that LPG,=n1\operatorname{LP}_{G,\mathbb{Z}}^{\ast}=n-1. If Question 6.5 were true, this would strengthen a result of [HHM22, 3.6.(b)] that an unmixed graph of König type is traceable. In terms of the notation of this paper, their statement can be rephrased as: if GG is unmixed, connected, and LPG,=LPG,\operatorname{LP}_{G,\mathbb{Z}}=\operatorname{LP}_{G,\mathbb{Z}}^{\ast}, then GG is traceable. Question 6.5 can be rephrased as: if GG is unmixed, connected, and LPG,=LPG,\operatorname{LP}_{G,\mathbb{Q}}=\operatorname{LP}_{G,\mathbb{Z}}^{\ast}, then is GG traceable?

References

  • [BE18] R. Blanco and S. Encinas, A procedure for computing the log canonical threshold of a binomial ideal, Manuscripta Math. 155 (2018), no. 1-2, 141–181.
  • [BMS18] D. Bolognini, A. Macchia, and F. Strazzanti, Binomial edge ideals of bipartite graphs, European J. Combin. 70 (2018), 1–25.
  • [Dan63] G. B. Dantzig, Linear programming and extensions, Princeton University Press, Princeton, N.J., 1963, pp. xvi+625.
  • [GM21] R. González-Martínez, Gorenstein binomial edge ideals, Math. Nachr. 294 (2021), no. 10, 1889–1898.
  • [GS] Daniel R. Grayson and Michael E. Stillman, Macaulay2, a software system for research in algebraic geometry, Available at http://www.math.uiuc.edu/Macaulay2/.
  • [Her14] D. J. Hernández, FF-pure thresholds of binomial hypersurfaces, Proc. Amer. Math. Soc. 142 (2014), no. 7, 2227–2242.
  • [Her16] by same author, FF-purity versus log canonicity for polynomials, Nagoya Math. J. 224 (2016), no. 1, 10–36.
  • [HH11] J. Herzog and T. Hibi, Monomial ideals, Springer, 2011.
  • [HHH+10] J. Herzog, T. Hibi, F. Hreinsdóttir, T. Kahle, and J. Rauh, Binomial edge ideals and conditional independence statements, Adv. in Appl. Math. 45 (2010), no. 3, 317–333.
  • [HHM22] J. Herzog, T. H., and S. Moradi, Graded ideals of König type, Trans. Amer. Math. Soc. 375 (2022), no. 1, 301–323.
  • [How01] J. A. Howald, Multiplier ideals of monomial ideals, Trans. Amer. Math. Soc. 353 (2001), no. 7, 2665–2671.
  • [HV16] I. B. D. A. Henriques and M. Varbaro, Test, multiplier and invariant ideals, Adv. Math. 287 (2016), 704–732.
  • [LaC] A. LaClair, In preparation.
  • [Laz04] R. Lazarsfeld, Positivity in algebraic geometry. II, Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics], vol. 49, Springer-Verlag, Berlin, 2004, Positivity for vector bundles, and multiplier ideals.
  • [LP86] L. Lovász and M. D. Plummer, Matching theory, North-Holland Mathematics Studies, vol. 121, North-Holland Publishing Co., Amsterdam; North-Holland Publishing Co., Amsterdam, 1986, Annals of Discrete Mathematics, 29.
  • [Luc78] E. Lucas, Theorie des Fonctions Numeriques Simplement Periodiques, Amer. J. Math. 1 (1878), no. 4, 289–321.
  • [MSV14] L. E. Miller, A. K. Singh, and M. Varbaro, The FF-pure threshold of a determinantal ideal, Bull. Braz. Math. Soc. (N.S.) 45 (2014), no. 4, 767–775.
  • [MTW05] M. Mustaţǎ, S. Takagi, and K. Watanabe, F-thresholds and Bernstein-Sato polynomials, European Congress of Mathematics, Eur. Math. Soc., Zürich, 2005, pp. 341–364.
  • [MV12] S. Morey and R. H. Villarreal, Edge ideals: algebraic and combinatorial properties, Progress in commutative algebra 1, de Gruyter, Berlin, 2012, pp. 85–126.
  • [Oht11] M. Ohtani, Graphs and ideals generated by some 2-minors, Comm. Algebra 39 (2011), no. 3, 905–917.
  • [Pag18] G. Pagi, Enhanced algorithms for F-pure threshold computation, Ph.D. thesis, 2018.
  • [Pyt] Python language reference, version 2.7, Available at http://www.python.org.
  • [S+22] W. A. Stein et al., Sage Mathematics Software (Version 9.6), The Sage Development Team, 2022, Available at http://www.sagemath.org.
  • [SM18] S. Saeedi Madani, Binomial edge ideals: a survey, Multigraded algebra and applications, Springer Proc. Math. Stat., vol. 238, Springer, Cham, 2018, pp. 83–94.
  • [ST09] T. Shibuta and S. Takagi, Log canonical thresholds of binomial ideals, Manuscripta Math. 130 (2009), no. 1, 45–61.
  • [TW04] S. Takagi and K. Watanabe, On F-pure thresholds, J. Algebra 282 (2004), no. 1, 278–297.
  • [Vel19] D. J. Velez, On F-pure thresholds: computations and relations to others invariants.
  • [Zwo] Y. Zwols, Online linear and integer optimization solver, Available at https://online-optimizer.appspot.com.