Invariants of Binomial Edge Ideals via Linear Programs
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 and characteristic . 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.
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 let (respectively ) denote the optimal value of the linear program associated to realizing a maximal matching of over (respectively ), and denote the optimal value of the dual linear program over . 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 see [LP86].
Theorem A (classical).
Let be a hypergraph, the associated edge ideal, and as above.
-
(1)
The monomial grade of is equal to , and also to the size of a maximal matching of . (folklore)
- (2)
- (3)
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 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.
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 be a finite simple graph with vertex set and edge set . Fix a field . Consider the polynomial ring , and for each edge with define . Define the binomial edge ideal of , denoted , to be the ideal
(1) |
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 -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 will always denote a positive prime integer. When and is an -ideal, we denote by , respectively , the image of in , respectively . We denote by the log canonical threshold of around the origin which is an important invariant appearing in birational geometry (see [Laz04]) and by the F-threshold of . Throughout this paper we will often write instead of since our computations will be independent of characteristic. It is known that:
Theorem 1.2 ([MTW05, Theorem 3.4]).
In the above setup,
(2) |
for all and
(3) |
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 . Hernández [Her14] expands on their ideas to show how the optimal value of this linear program over 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 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 for any graph 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 , define to be the induced graph on . Let . Consider the following linear program associated to the graph .
(4) | ||||
where for each and , and are the functions
Throughout this paper let denote a choice of ring , , or . Use (4) on different choices of , and denote the optimal value of (4) with respect to a choice of by .
This paper studies as it relates to the combinatorial properties of and the algebraic properties of . In Section 2 we introduce the relevant background. In Section 3 we show
Theorem D.
In Section 4 we prove the main result of this paper.
Conjecture 1.3.
Let be a graph, then
-
(1)
-
(2)
.
We prove this conjecture when is a block graph or 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 the relationship between the equality and 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 or . 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 , , or [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 be an matrix having entries in . Let and . Then, the linear programming problem in standard form refers to
(5) |
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) |
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 .
Remark 2.2.
When , 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 denote the -th column of and denote the -th row of .
Theorem 2.3 ( [Dan63, p.136 Theorem 4] Complementary Slackness).
Sketch of Proof.
It is always true that
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 we denote its set of vertices by , or when the context is clear, and denote its set of edges by , or when the context is clear. Throughout this paper we will assume that for some positive integer . Given a subset , we denote by the subgraph of induced on the vertices .
By a path in we mean an alternating sequence of distinct vertices and edges of , so that for . By a cycle in we mean an alternating sequence of distinct vertices and edges of , so that for and . By a semi-path we mean a subgraph of such that each connected component of is a path. Let be the induced graph on where denotes the isolated vertices of .
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 be a graph on vertices. If , we will denote the induced subgraph on the vertices by . Let denote the number of connected components of , and let denote the distinct connected components of . Let denote the complete graph on the vertices . Put
where is as defined in Definition 1.1. Then, is a minimal prime ideal of and
Moreover, every minimal prime ideal of is of the form where , or and for each one has that .
Definition 2.5.
For a graph define the invariant
We call a disjoint union of paths realizing a path packing of .
Definition 2.6.
A graph is called traceable if has a Hamiltonian path.
Remark 2.7.
is traceable if and only if .
2.3. F-Thresholds
Definition 2.8 ([MTW05, Lemma 1.1, Remark 1.5]).
Let be a polynomial ring, , the homogeneous maximal ideal, and an -ideal with . For , set
where . Then,
exists and is called the F-threshold of .
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 , with its homogeneous maximal ideal, with its homogeneous maximal ideal, with its homogeneous maximal ideal. Then,
In particular, if is a graph and are its connected components, then
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 be ideals in . Then,
In particular, if is a subgraph of , then
Proof.
It follows from the definition that for all . ∎
The final lemma says that the F-threshold can only decrease under taking initial ideals.
Lemma 2.11 ([TW04, Proposition 4.5]).
Let be ideal in and a term order on the polynomial ring. Then, .
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 , and let be an matrix of indeterminates over a field of prime characteristic. Let be the polynomial ring , and the ideal generated by the size minors of .
The F-threshold of is
Remark 2.13.
In the above theorem, when , then corresponds to where denotes the complete graph on vertices, and thus
In the master’s thesis of Velez [Vel19], he proves
Proposition 2.14 ([Vel19, Theorem 3.1.6]).
Let be a generic matrix with . Let denote the determinant of the submatrix given by the columns for . Let . Then, .
Remark 2.15.
When in Proposition 2.14 the ideal corresponds the binomial edge ideal of a path on vertices. I.e., if is a path on vertices, then .
Theorem 2.16 ([Luc78] Lucas’s theorem).
Let be a prime integer and and positive integers. Choose a positive integer so that we can write and where for . Then,
with the convention that if .
3. Linear Programs and Binomial Edge Ideals
3.1. Binomial Grade, Height, and Linear Programs
Recall that is the maximal length of a path packing (Definition 2.5), and is the optimal value of an integer program (discussion surrounding (4)).
Proposition 3.1.
Given a graph , .
Proof.
For , the constraint forces that a feasible solution has at most two edges incident to . For , the constraint forces that a feasible solution does not induce a Hamiltonian cycle on . Thus, an optimal solution to (4) consists of disjoint paths. Conversely, a path packing is a feasible solution of (4). ∎
Remark 3.2.
The dual linear program of (4) solved over is given by
(7) |
Denote the optimal value of (7) over by .
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 .
Lemma 3.3.
Fix a graph and set . There exists an optimal solution of the linear program (7) of the form satisfying
-
(1)
for all and for all ,
-
(2)
if and are elements of with , then ,
-
(3)
if for some , then for all ,
-
(4)
if for some , then is connected.
For such a solution, let and to be the elements of which satisfy . Then, the are the connected components of .
Proof.
Let be any optimal solution to (7).
-
(1)
We may assume that for all and . Otherwise, for each non-zero or non-zero reassign it the value .
-
(2)
We may suppose that if and are elements of with , then . Otherwise, redefine and .
-
(3)
We may suppose that if for some , then for all . Otherwise, if put and , and if put .
-
(4)
We may suppose that if for some , then is connected. Otherwise, write with . Put and if for .
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 and to be elements of which satisfy . The above reductions together with the constraints for every implies that the are the connected components of . ∎
Proposition 3.4.
Let be a graph, then .
Proof.
We first show . Let be chosen so that where the notation is as in Proposition 2.4. For let
and for let
Then the tuple is a feasible solution to (7), and moreover
Next, we show that . Let be an optimal solution of (7) when of the form guaranteed by Lemma 3.3. Let and be as constructed by Lemma 3.3. The cost of this optimal tuple is precisely which is equal to . This proves that
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 , and how the optimal solution relates to commutative algebra invariants.
For a graph denote by the optimal value of the following linear program over .
(8) |
Remark 3.5.
Let be a graph and the connected components of . Then,
The following proposition says that the constraint in (8) can be strengthened by requiring that in addition is biconnected. Recall that a connected graph is biconnected if is connected for every . In the following linear program we differentiate the intdeterminates and constraints from those appearing in (8) by utilizing the superscript “”.
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) |
where means that the label on the vertex is smaller than the label on the vertex .
Proof.
It suffices to show that feasible solutions of (9) are feasible solutions of (8). It suffices to restrict to the case of is connected. Write as a union of its maximal biconnected components. Let denote a feasible solution of (8). Observe that
The above assumptions on together with induction on shows that
∎
Remark 3.7.
Linear program (9) will have constraints for a complete graph on vertices.
In the linear program (10) below, whenever we write for an edge of a graph we will in addition assume that . The dual of the linear program (8) is given by
(10) |
We denote by the optimal value of (10).
Proposition 3.9.
Let be a labeled graph. Then, and .
Proof.
Let (respectively ) denote the set of feasible solutions of (4) (respectively of the feasible solutions of (8)). If is a field, then there are linear bijections
If , then there are a linear bijections
The constraint for shows that the map from is well-defined.
The dual case is proved analogously. ∎
Proposition 3.10.
Proof.
Let denote the lex monomial order on with . Then,
∎
Lemma 3.11 ([MSV14, Lemma 1.2]).
Fix a generic matrix and . There exists a non-negative integer , so that for every where , we have .
Proof.
Follows from the proof of their Lemma 1.2 by taking, in their notation, and . ∎
Proposition 3.12.
Let be a labeled graph, then
Proof.
Fix notation of as a generic matrix, . For we consider and the restriction of and to the submatrix (respectively subring) of variables corresponding to the vertices of . We take where is as defined in Lemma 3.11.
We denote by the polytope contained in determined by the constraints of linear program (8).
For every , let be the polytope inside determined by the following constraints
(11) | |||
(12) | |||
(13) |
We denote by to be the maximum of the function
when restricted to the region .
Fix a positive integer. Then, there exist non-negative integers for satisfying and a monomial
belonging to the monomial support of some element of .
Lemma 3.13.
With the notation as above,
-
(1)
For every , there exists so that for every there exists with .
-
(2)
.
Proof.
-
(1)
Fix . Take so that . For , define via for . We observe that
Since is pointwise less than and satisfies the constraints (12) and (13), and for all . Thus, to show that it suffices to show that for all .
Observation 1.
The constraint implies that
(14) Observation 2.
If , then where is the largest value of that appears in the support of . If this is not the case, then we have that
But, then this implies that ; a contradiction.
-
(2)
Let , a vertex of where attains a maximum, and to be as in part (1) of this Lemma. Since and , it follows that
Consequently,
This proves that .
∎
Corollary 3.14.
Let be a graph, then
Remark 3.15.
The previous results of this section can be summarized as
(15) |
Remark 3.16.
The relation holds more generally, cf. [TW04, Proposition 2.6(1)].
4. Proof of Conjecture 1.3 for Block Graphs
Definition 4.1.
A graph is called biconnected if is connected and is connected for every vertex of . A graph is called a block graph if every maximal biconnected component of is a complete graph. We call a maximal biconnected component of a block. A vertex of is called a cut vertex if the number of connected components of is strictly larger than the number of connected components of .
Let be a block graph and a block of , then we define to be the set of all cut vertices of contained in . We say that is a leaf of if , and we say that has the two block property if every cut vertex of is adjacent to exactly two blocks of . An edge of which is also a leaf of is called a whisker of .
Throughout this section we will let , will denote the number of leaves of , will denote the number of connected components of isomorphic to a complete graph. When the graph is clear from context we will omit the subscript .
Proposition 4.2.
Let be a block graph. Then,
Proof.
By Remark 3.5 we may assume that is connected. If is a complete graph there is nothing to prove. We reduce to the case that is a connected block graph which is not a complete graph.
Since is a connected block graph which is not a complete graph and . Partition the blocks of into , the leaves of , and , the remaining blocks. Write . Each has exactly one cut vertex which we will call and at least one other vertex, i.e. . Let
For , define
(16) |
For , define
(17) |
Then, the tuple is a feasible solution to the linear program (10), and the cost of this feasible solution is
Thus,
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 is a triangle path if every connected component of which is not an isolated vertex satisfies
-
(1)
the only cycles in are ’s which we label as (we will sometimes refer to the ’s as the triangles of ),
-
(2)
For all ,
-
(3)
Every vertex of belongs to at most one triangle of .
When is a triangle path and is a subgraph of , then we call a triangle path packing of .
Remark 4.5.
A triangle path is an example of a family of block graphs with the two block property.
Proposition 4.6.
Let be a triangle path. Label the ’s in the triangle path . Label the connected components of by . Then,
(18) |
Moreover,
(19) |
Proof.
By induction on it can be shown that (18) and (19) are equivalent. By Lemma 2.9 and Remark 2.13 we may assume that is connected and not a complete graph, i.e. that . By Proposition 3.12 and Proposition 4.2, and the above mentioned equivalence we have that . It thus suffices to prove the reverse inequality.
Let and be as in Definition 1.1. First consider the case that . The first step in the proof is to label the vertices of the graph and to assign to every edge a tuple so that the following conditions are satisfied.
-
(1)
Some vertex having degree one has been assigned the label .
-
(2)
If and for some with the distance from to smaller than the distance from to , then either or .
-
(3)
equals , , or if is not an edge of any triangle of .
-
(4)
equals or if is the edge of any triangle of .
-
(5)
.
-
(6)
The monomial
(20) does not belong to the ideal and appears in the polynomial
(21) with a coefficient which is non-zero mod .
It would follow from items (5) and (21) that
and consequently that
We prove the existence of this labeling by induction on . It , then is a path. Starting at a vertex of degree one assign it the label , then label the vertices consecutively. Assign to every edge the tuple . That the above conditions are satisfied is clear.
Suppose now that . There exists a triangle of , call it , so that removing the edges of from yields a graph having three connected components—two of which are paths, say and , and the other is a triangle path, call it , having one less triangle than . Let be the vertex belonging to . Let be the vertex adjacent to in . Apply the induction hypothesis to and we may assume that has not been given the label . By item 2, it follows that . Label the remaining vertices of by and . Label the vertex of adjacent to (respectively ) by (respectively ). Label the remaining vertices of so as to satisfy condition (2). Use Table 1 to assign tuples to the edges of and to the edges of and incident to the vertices of based upon the tuple assigned to the edge . Now, and each have an edge which has been assigned a tuple; extend this label to the remaining edges of and . After performing this construction the coefficient on in from item (21) is where is the number of times that the label appears on an edge. This binomial is non-zero by Lucas’s theorem, Theorem 2.16.
Suppose now that and fix . We modify the above construction by using Table 2 in the construction. In this case we have that
Dividing by and taking limit as proves the desired inequality when .
∎
Corollary 4.7.
Let be a triangle path and be subgraphs of satisfying
-
(1)
the are triangle paths,
-
(2)
,
-
(3)
every triangle of is a subgraph of some
Then,
Proof.
Follows from Proposition 4.6. ∎
Corollary 4.8.
Let be a triangle path. Suppose satisfies that is a triangle path, no two vertices of are adjacent via an edge of , and every vertex of has degree two in . Then we have
Proof.
Since and are triangle paths, we can compute their F-threshold as the number of triangles plus the number of edges not in a triangle. The later two conditions on imply that the number of triangles of and are the same and the number of edges in not belonging to any triangle path is equal to the number of edges of not belonging to any triangle path plus . ∎
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 be a graph, , and the connected components of , then
Proof.
By strong duality, Theorem 2.1, it suffices to prove the statement for the dual linear program. For , let be an optimal solution to the linear program (10) realizing . These local solutions together with the assignments , for , and for yield a feasible solution to the linear program (10) realizing . ∎
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 be a block graph with the two block property. Let denote the number of leaves of and the number of connected components of which are isomorphic to a complete graph. Then, contains a subgraph satisfying
-
(1)
is a triangle path packing,
-
(2)
,
-
(3)
for every block of , contains at most one triangle of ,
-
(4)
a block of contains a triangle of only if is odd,
-
(5)
if is a triangle of contained in a block of , then contains two cut vertices unless in which case contains three,
-
(6)
a block of contains exactly one vertex of of degree one if and only if is a leaf of ,
-
(7)
a block of contains a vertex of of degree zero if and only if is a single vertex,
-
(8)
if is a triangle of contained in , , and is the block of incident to , then the edge of incident to and not contained in is contained in ,
-
(9)
.
The proof idea is to induce on the number of blocks of . We start by choosing a block of which is not a leaf and so that every incident block with the exception of at most one is a leaf of . In Figure 3(a), blocks and are the only such blocks. We will consider block . Next, we choose two leaves of to delete (if 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 in is less than or equal to several cases must be considered to extend the triangle path packing while preserving the above properties. However, when the number of cut vertices of in is larger than then the triangle path packing can be extended by taking a Hamiltonian path on the deleted vertices.
Proof.
It is enough to show this proposition for each connected component of . If is isomorphic to a complete graph, then take to be a Hamiltonian path and the Proposition follows since and in this case. Thus, we may assume henceforth that is a connected graph and not isomorphic to a complete graph.
The proof is by induction on the number of blocks of which we denote by . Since is not a complete graph, . If , then is two complete graphs joined at a cut vertex. In this case, and has a Hamiltonian path. Take to be a Hamiltonian path of , and it is clear that satisfies the conditions listed above.
Suppose now that and that the Proposition has been shown for all block graphs having the two block property with at most blocks. Let be a block of which is not a leaf such that every block incident to with the exception of at most one is a leaf of . Let denote the number of cut vertices of . Observe that since is not a leaf.
Case : Let be a leaf of . Let be the cut vertex of which is adjacent to . If put ; otherwise, put . Let be the image of in . Since has blocks, by induction there exists a triangle path of satisfying the above properties. Since is a leaf in , has a vertex of degree one, say , by property (6). Consider now as a subgraph of , and construct a triangle path in as follows. Observe that if , then . Otherwise, adjoin an edge from to . In either case is now a subgraph of containing as a degree one vertex. Finally, adjoin to a Hamiltonian path contained in which has a terminal vertex at . Now, is a triangle path of which satisfies properties (1) through (9). Moreover, is of the desired form by Proposition 4.6, induction hypothesis, and the fact that we obtained from by adding a path.
Case : Let . Let be leaves adjacent to at for .
Sub-case : Let be the graph attained after removing , , and . Note that is adjacent to at and that may or may not be a leaf. Consider the graph which is adjoined a whisker at the vertex . Denote by and the number of vertices and leaves of , respectively. By induction hypothesis there exists a triangle path for satisfying the above properties and
First, consider the sub-sub-case that there exists a triangle in containing ; label the other vertices of this triangle as and . Construct from by deleting the whisker attached at , deleting the edge , adding the edge , and adding Hamiltonian paths in and which have a terminal vertex at and , respectively. Observe that is a triangle path, and by Proposition 4.6 we compute that
where the term is counting in 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 now follows after simplifying and using that and that .
Second, consider the sub-sub-case that has degree two in ; call the neighbor of contained in to be . Construct from by deleting the whisker, adding the edges , , and adding Hamiltonian paths in and which have a terminal vertex at and , respectively. Observe that is a triangle path, and by Proposition 4.6 we compute that
where the is the deletion of the whisker and the plus is the addition of the triangle. The formula for now follows after simplifying and using that and that .
Sub-case : Let . By induction hypothesis there exists a triangle path of satisfying the desired properties. Since is a leaf in , property (6) guarantees the existence of a vertex in having degree one. Consider as a subgraph of , and construct a triangle path of from by adding the edges , , and adding Hamiltonian paths in and which have a terminal vertex at and , respectively. Observe that is a triangle path, and by Proposition 4.6 we compute that
where the plus is the addition of the triangle. The desired formula on now follows after simplifying and using that and that .
Case : Let and be two cut vertices of which have leaves and , respectively. Let . By induction hypothesis there exists a triangle path of satisfying the desired properties. Consider as a subgraph of , and construct a triangle path of from by adding the edges and adding Hamiltonian paths in and which have a terminal vertex at and , respectively. Observe that is a triangle path, and by Proposition 4.6 we compute that
where the plus is the addition of the edge . The desired formula on now follows after simplifying and using that and that . ∎
Corollary 4.11.
Let be a block graph having the two block property. Let denote the number of leaves of and the number of connected components of isomorphic to a complete graph. Then, there exists a triangle path packing of such that
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 be a block graph. We call a two block approximation of if can be obtained via the following process
-
(1)
Enumerate the cut vertices of incident to at least three blocks of ; call this set .
-
(2)
For each , choose two blocks incident to , say and .
-
(3)
Delete all edges incident to and not contained in .
Example 4.13.
Let be the block graph in Figure 4(a). Then, the graphs in Figure 4(b) and 4(c) are two possible block approximations of . Observe that the F-threshold of the graphs in Figure 4(b) and 4(c) are and , respectively.
Our next goal is to show
Theorem 4.14.
Let be a block graph, then the following quantities are equal
-
(1)
-
(2)
-
(3)
-
(4)
-
(5)
-
(6)
.
Proof.
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 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 be a connected block graph, a two block approximation of realizing (5) in Theorem 4.14, and a path packing of given by Lemma 4.10. Take to be the set of cut vertices of incident to at least three blocks of such that in at least one of these blocks becomes a leaf or an isolated connected component. For , let denote the blocks of incident to for some . Without loss of generality, we may suppose that and that becomes a leaf or an isolated connected component of . Since is a leaf or isolated component of , Proposition 4.10 implies that there exists a vertex having degree one or zero in .
Lemma 4.16.
With the above setup suppose that . Then satisfies the conclusions of Theorem 4.14. Moreover, .
Proof.
Lemma 4.17.
With the setup from above let . Then
-
(1)
does not belong to a triangle of ,
-
(2)
each vertex of is not contained in a triangle of ,
-
(3)
, and
-
(4)
for ,
-
(5)
no two vertices of are adjacent via an edge of .
Proof.
-
(1)
Suppose by contradiction that belongs to a triangle of , i.e. that without loss of generality there exists vertices so that , , and belong to . Let be the triangle path defined from by deleting the edges and and adding the edge . Notice that no new cycles are created in after these modifications or else it would contradict being a block graph. Moreover,
because in obtaining from we delete two edges of having weight , increase the weight of one edge of from weight to weight in , and add one edge to having weight . This contradicts our choice of via Lemma 4.15.
-
(2)
Suppose by contradiction that there exists and vertices and of with , , and edges of . Construct a triangle path from by deleting the edges and and adding the edge . Then, ; a contradiction.
-
(3)
By Item (2) is or . Suppose by contradiction that . Construct from by adding the edge . Then, ; a contradiction.
-
(4)
Suppose by contradiction that . By the previous item, and do not belong to a triangle of . If and have degree two in , then define from by adding the edge and the edge . In the case that has degree one and has degree two in , define from by deleting the edge and adding the edges and . The case that has degree one and has degree two in is handled analogously. If and both have degree one in , define from by deleting the edge and adding the edges and . Notice that in all cases that is a triangle path, a subgraph of , and that ; a contradiction.
-
(5)
Suppose by contradiction that and that . Construct a triangle path from by deleting the edge and adding the edges and . Then, ; a contradiction.
∎
Proof of Theorem 4.14.
Let the setup be as above. We induce on the number of blocks of . If the number of blocks of is one, then there is nothing to prove. Suppose that the number of blocks of is strictly larger than one. If , then we are done by Lemma 4.16. Suppose that . Let denote the connected components of . For , let denote the restriction of to . We show
Claim 1.
For ,
Proof of Claim 1.
By induction on number of blocks and the fact that the number of blocks of is less than the number of blocks of , it suffices to show that
for every . Suppose by contradiction that for some fixed choice of and a triangle path of . Let be the vertices of which are adjacent to some vertex of via an edge of ; call this vertex for . Construct from as follows.
-
(i)
Delete the edges for .
-
(ii)
Add the edges for .
-
(iii)
Replace on by .
Now, is a triangle path. Indeed parts (i) and (ii) of the construction do not create any cycles in ; otherwise, could not be a block graph. After performing parts (i) and (ii), forms its own connected component by Lemma 4.17 item (4). Thus after performing operation (iii) will remain a triangle path. Clearly, which contradicts our choice of , Lemma 4.15.
Corollary 4.18.
Conjecture 1.3 is true when is a block graph.
Corollary 4.19.
If is a tree, then
Proof.
In the next section we show the stronger statement that for a tree
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 such that
Proof.
Let be the graph below. Let and be the subgraphs of induced on the vertices and , respectively. Let be a triangle path packing of maximal wrt . Let and be the restriction of to and respectively. For , let , denote the number of triangles of and edges of not belonging to any triangle, respectively.
Claim 2.
.
Proof of Claim 2.
Without loss of generality suppose . Construct from by deleting and replicating on . Then, is a triangle path packing with ; a contradiction.
It follows from Proposition 4.6 and Claim 2 that is an integer. However, which can be realized by the following assignment
∎
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 to be of König type if it satisfies the following characterization:
Definition 5.1 ([HHM22] Theorem 3.5).
A graph is said to be of König type if .
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 is of König type, then . In particular, Conjecture 1.3 holds for every graph that is of König type.
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 be a graph. Then, is of König type if and only if there exists a semi-path and a choice of distinct vertices , possibly empty, with the property that
-
(1)
,
-
(2)
for all , has degree two in ,
-
(3)
for all , is not an edge of ,
-
(4)
the number of connected components and the number of vertices of equals the number of connected components and the number of vertices of , respectively.
Proof.
We prove the backward implication. Suppose that a semi-path of and a choice of distinct vertices have been chosen satisfying properties (1) through (4). It is enough to show that .
Let be the number of connected components of and its connected components. Let the ideal as defined in Proposition 2.4. Then,
Let denote the subgraph of given by for . Property (4) implies that is a Hamiltonian path of , and in particular that
Let denote the set of edges of which are incident to a vertex of . By assumptions (2) and (3), . The edges of can be partitioned into those which belong to some and those which belong to . We compute that
This proves that .
Next, suppose that is of König type. Let be a semi-path realizing . Consider the vector given by
Then, is an optimal solution of the linear program (4). Let be an optimal solution with values in of the linear program (7) having the form guaranteed by Lemma 3.3, and let and be as in Lemma 3.3.
Let be the below matrix, the vector of indeterminates of length , vector of ’s of length , and the vector of length realizing the linear program (4), i.e. algorithm (4) can be expressed as
(22) |
Let denote the -th column of , and let denote the -th row of .
Because 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 , equivalently, . Let denote the row of corresponding to vertex . By complementary slackness Theorem 2.3.(2) we have that
This shows that there are two edges of incident with .
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 be a tree. Let be disjoint paths in with . Put
Put and
Definition 5.4.
Let . A subtree of based at , denoted , is the subtree of constructed as follows:
Then, is the induced subgraph of on the set of vertices .
Lemma 5.5.
Let be the subtree of based at . For , .
Proof.
Clear because , and the later sets are disjoint for since 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 -alternating path with respect to (respectively a -alternating path with respect to ) is a set of distinct vertices where satisfying
-
(1)
for ,
-
(2)
for and for some ,
-
(3)
and (respectively, ).
Lemma 5.7.
Let be a graph. Then, does not have an alternating path , , with both and .
Proof.
Algorithm 5.8.
Input: Semi-path of a tree Output: satisfying conditions (2) - (3) in Lemma 5.3, and every edge of contains a vertex belonging to .
-
(1)
Initialize and for .
-
(2)
While , repeat the following steps:
-
(a)
Increment ,
-
(b)
Pick a vertex .
-
(c)
Let denote the subtree of based at , and for , let denote the subsets of as constructed in Definition 5.4.
-
(a)
-
(3)
Put .
- (4)
-
(5)
Put .
-
(6)
For , define and .
-
(7)
Put .
Theorem 5.9.
Proof.
First, we remark that each while-loop of Algorithm 5.8 terminates because and are finite, and form a strictly ascending sequence of sets each containing more elements of than their predecessors.
Claim 3.
For , .
Proof of Claim 3.
Suppose by contradiction that . Choose even and odd satisfying the following
-
(i)
there exists ,
-
(ii)
if (respectively ) is the -alternating path (respectively -alternating path), then the vertices of and intersect only at .
Let be the vertex of adjacent to . Then, there is a -augmenting path by concatenating and along the edge . Without loss of generality, suppose that , then by construction of , but this contradicts the choice of which proves the claim.
: Suppose by contradiction that there exists . Since and for some , it must be the case that by part 2 of Algorithm 5.8. This contradicts Lemma 5.7.
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 be the graph in Figure 6.
One can check that , , and . One realization of is via the following assignment
This graph is not unmixed. For example, is a cut set of and the corresponding prime ideal has height . 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. and Traceable Graphs
In this section, we consider the relationship between maximality of 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 is a traceable graph on vertices, then .
Proof.
We have the graph containments . The result follows from monotonicity of and . ∎
Remark 6.2.
The above proof shows that for any graph on vertices, .
Example 6.3 shows that there exists a graph with yet is non-traceable. Computations in Macaulay2 [GS] suggest that may agree with independent of characteristic in Example 6.3.
Example 6.3.
Throughout this example let denote the graph in Figure 7. Observe that is not traceable.
Observe that deleting edges and produces a triangle path as a subgraph and that this triangle path has F-threshold equal to by Proposition 4.6. It follows from Proposition 2.10 that . However, it can be computed that and one such realization is as follows
This bilabeling corresponds to the monomial
belonging to the polynomial
which has coefficient given by the expression
where . For small values of and it can be checked that this coefficient does not vanish mod . This suggests that for any prime .
The following proposition shows that under additional hypotheses on that a converse of Proposition 6.1 holds.
Proposition 6.4.
If is a connected graph on vertices with at least three vertices having degree one (in particular is not traceable), then .
Proof.
Observe that is a subgraph of a complete graph on vertices with three whiskers. Suppose that the whiskers of do not share a common vertex, then by Corollary 4.11, . If two of the three whiskers or all three of the whiskers of share a common vertex, then we can list the two block approximations of and show that by Theorem 4.14. Since the result follows. ∎
It can be shown that the graph in Example 6.3 is not unmixed. This raises
Question 6.5.
If is an unmixed, connected graph and , then is traceable?
Note that being unmixed and connected implies that . 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 is unmixed, connected, and , then is traceable. Question 6.5 can be rephrased as: if is unmixed, connected, and , then is 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, -pure thresholds of binomial hypersurfaces, Proc. Amer. Math. Soc. 142 (2014), no. 7, 2227–2242.
- [Her16] by same author, -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 -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.