On the Ehrhart Theory of Generalized Symmetric Edge Polytopes
Abstract.
The symmetric edge polytope (SEP) of a (finite, undirected) graph is a centrally symmetric lattice polytope whose vertices are defined by the edges of the graph. SEPs have been studied extensively in the past twenty years. Recently, Tóthmérész and, independently, D’Alí, Juhnke-Kubitzke, and Koch generalized the definition of an SEP to regular matroids, as these are the matroids that can be characterized by totally unimodular matrices. Generalized SEPs are known to have symmetric Ehrhart -polynomials, and Ohsugi and Tsuchiya conjectured that (ordinary) SEPs have nonnegative -vectors.
In this article, we use combinatorial and Gröbner basis techniques to extend additional known properties of SEPs to generalized SEPs. Along the way, we show that generalized SEPs are not necessarily -nonnegative by providing explicit examples. We prove these polytopes to be “nearly” -nonnegative in the sense that, by deleting exactly two elements from the matroid, one obtains SEPs for graphs that are -nonnegative. This provides further evidence that Ohsugi and Tsuchiya’s conjecture holds in the ordinary case.
1. Introduction
A lattice polytope in is the convex hull of finitely many points in . That is, is a lattice polytope if and only if there are some such that
The main objects of study in this paper will be lattice polytopes arising from (finite, undirected) graphs and matroids. First, given an undirected graph where , the symmetric edge polytope (or SEP) associated to is
This polytope was introduced in [11] and has been the object of much study for its algebraic and combinatorial properties [6, 9, 12, 13], as well as its applications to problems arising from engineering [3, 4, 5].
SEPs have recently been generalized to regular matroids in two ways. Suppose is a totally unimodular matrix representing a matroid . Tóthmérész [17], working within the context of oriented matroids, defined the root polytope of to be the convex hull of the columns of . Independently, D’Alì, Juhnke-Kubitzke, and Koch [7] defined the generalized symmetric edge polytope (or generalized SEP) associated to as
Thus, may be considered the root polytope of the matroid represented by the matrix , since this is a totally unimodular matrix. Because we will not be working with oriented matroids, and because are focused on the polytopes in this work, we will follow the terminology and notation in [7].
Although the definition of depends on choice of , the polytope is well-defined up to unimodular equivalence ([17, Proposition 3.2] and [7, Theorem 3.2]). In [7], the authors proved that many properties held by ordinary SEPs also hold for generalized SEPs. Further, this more general setting allowed the authors to show that two SEPs are unimodularly equivalent if and only if the underlying matroids are isomorphic [7, Theorem 4.6]. In addition to these aspects of SEPs, much attention has been given to their Ehrhart theory, which we briefly introduce here.
The function , defined on the positive integers, agrees with a polynomial of degree , called the Ehrhart polynomial of . Consequently, the Ehrhart series of , which is the (ordinary) generating function for the sequence , may be written as the rational function
for some . We denote the numerator by and refer to it as the -polynomial of . The -vector of , , is the list of coefficients of .
The -vectors of polytopes have been the focus of much study in recent decades, as it lies within the intersection of polyhedral geometry, commutative algebra, algebraic geometry, combinatorics, and much more (see, e.g., [15, 16]). For example, when the -vector is symmetric, meaning for each , then the associated semigroup algebra
over a field , is Gorenstein. In fact, the converse holds as well. Determining when is Gorenstein can be detected entirely using polyhedral geometry, so, from this perspective, there is a complete characterization of when is symmetric.
On the other hand, a more mysterious property held by some -vectors is unimodality. We call unimodal if there is some index for which and . Unlike symmetry, there is no complete characterization of when an -vector is unimodal. There are various sets of sufficient conditions that ensure a polytope has a unimodal -vector and various sets of necessary conditions [2], but no set of conditions capturing both simultaneously.
When an -vector is already known to be symmetric, new avenues for proving -unimodality appear. For instance, in this setting, the -polynomial may be expressed
for certain choices of . The polynomial on the right is the -polynomial of and is often denoted , and the vector is the -vector of . Note that if for each , then for each as well. This is not necessary, though: the -vector of , realized by the lattice triangle with vertices , , and , is unimodal but has a -vector of . Nevertheless, -vectors remain the topic of intense study, due in part to wide-open conjectures surrounding their nonnegativity. The most famous of these is Gal’s Conjecture: that the -vector of a flag homology sphere is -nonnegative. To an Ehrhart theorist, the conjecture states that the -polynomial of a lattice polytope with a regular, unimodular, flag triangulation is -nonnegative.
In this article, we use combinatorial and Gröbner basis techniques to extend additional known properties of SEPs to generalized SEPs. Along the way, we show that generalized SEPs are not -nonnegative by providing an infinite class of explicit examples (Theorem 4.2). Computational evidence strongly suggests that these examples are “nearly” -nonnegative in the sense that, by deleting exactly two elements from the matroid, we obtain SEPs for graphs that appear to be -nonnegative. We present and prove a formula for their -polynomials (Theorem 4.3), deduce the normalized volumes of these SEPs (Corollary 4.6), and compute their -polynomials (Theorem 4.4). This provides further evidence that Ohsugi and Tsuchiya’s conjecture holds in the ordinary case.
A brief structure of the article is as follows. In Section 2 we discuss the necessary background for matroids and the relationships among Ehrhart theory, triangulations of polytopes, and toric ideals. Section 3 discusses how various important results for ordinary SEPs extend to generalized SEPs. In Section 4 we present explicit examples of generalized SEPs whose -polynomials are not -nonnegative (Theorem 4.2) and examine the consequences of making minor variations to them. We also state two of the main results of this article: Theorem 4.3 and Theorem 4.4. In Section 5 we both prepare for and present the full proof of Theorem 4.3, which follows a combinatorial argument. Finally, in Section 6, we use a generating function argument to prove Theorem 4.4.
2. Background
To keep this article as self-contained as possible, this section provides the necessary background on matroids, Ehrhart theory, and generalized SEPs. For a comprehensive reference on the first two topics, we refer the reader to [14] for matroids and [1] for Ehrhart theory. Throughout this article we use the convention of writing and for the union or deletion, respectively, of by a one-element set .
2.1. Matroids
A matroid consists of a pair where is finite and satisfies the following three properties:
-
(i)
;
-
(ii)
if and , then ; and
-
(iii)
if and , then there is some such that .
The set is called the ground set of and the elements of are called the independent sets of .
Matroids are useful for unifying various notions of “independence” in mathematics. For example, if is a (finite) graph, then there is a corresponding matroid whose ground set is , the edge set of , and whose independent sets are the sets of edges that form acyclic subgraphs of . This matroid is called the cycle matroid of , and any matroid that is isomorphic to for some is a graphic matroid.
Another fundamental example of a matroid uses the columns of a matrix as the ground set. Here, the independent sets are the columns of that form linearly independent sets. This matroid, denoted , is the vector matroid of , and any matroid isomorphic to for some matrix is called a representable matroid. More precisely, for a field , the matrix is called -representable if is isomorphic to a vector matroid whose ground set consists of vectors in . Hence, to say a matroid is representable is to say it is -representable for some field . On the other hand, a matroid is regular if it is -representable over every field .
The rank of a matroid is defined as
When is graphic, is the size of a maximal spanning forest for any satisfying . When is representable, for any matrix satisfying . With these definitions in hand, we can now describe some of the many benefits of working with regular matroids.
Theorem 2.1 ([7, Definition/Theorem 2.3]).
A positive-rank matroid is regular if and only if any of the following hold:
-
(i)
is representable over every field.
-
(ii)
for some matrix in which every minor is or .
-
(iii)
for some full-rank matrix in which every maximal minor is or .
Note that a matrix satisfying the minor condition in (ii) is called totally unimodular and a matrix satisfying the minor condition in (iii) is called weakly unimodular. Borrowing more terminology from linear algebra, an element of of cardinality is called a basis of . The set of bases, together with the ground set set , can be used to define a matroid without explicit mention of independent sets; when this is done, we usually write .
Now, from a matroid we may produce its dual matroid by setting
Many properties of can be easily deduced from . For example, is representable if and only if is representable. This does not hold for graphic matroids, though: if is graphic, then is graphic if and only if for some planar graph .
There are additional notions in graph theory that have corresponding matroidal analogues. For example, a circuit in a matroid is a minimal dependent set; this is the analogue of a cycle in a graph. Letting denote the set of circuits of a matroid, we define the matroid to be bipartite if each of the circuits has even size. This further allows one to call a matroid Eulerian if it is the dual of a bipartite matroid. These definitions reflect the fact that a connected, planar graph is Eulerian if and only if its dual is bipartite.
A loop of is a single element of that forms a circuit; equivalently, a loop is an element that is in no basis of . On the other hand, a coloop of , called by some authors an isthmus, is a single element of that is in every basis of . Now, given a matroid and an element that is not a coloop, the deletion of from is the matroid having ground set and independent sets . The contraction of by , where is not a loop, is denoted and has ground set and independent sets . Deletion and contraction are dual operations: if is neither a loop nor a coloop of , then . Many binary operations on graphs extend to matroids as well: the direct sum, or 1-sum, of matroids and having disjoint ground sets and is , where consists of independent sets of the form where and .
To define the next operation we assume that and call this element the basepoint. The parallel connection, denoted , has ground set and is most efficiently defined through its circuits: if is neither a loop nor a coloop of , then
If is a loop of , then we set
Similarly, if is a coloop of , then we set
Of course, if and have disjoint ground sets, then a point of each ground set may be selected and relabeled to fall into these definition, although the resulting parallel connection may then depend on which points were selected. An important property of this operation is that it preserves regularity: if and are both regular matroids, then so is [14, Corollary 7.1.25].
2.2. Dissecting Tree Sets and Toric Algebra
It is frequently more tractable to gather the desired information about a polytope by decomposing it into subpolytopes. A dissection, or subdivision, of a polytope is a collection of full-dimensional subpolytopes of whose union is and for any , the intersection is a face of both. A triangulation of is a dissection in which every is a simplex. There are many ways to create triangulations of a polytope; we will be concerned with one that has been recently introduced in the study of ordinary SEPs in particular.
Given an oriented subgraph of a graph on , let . Thus, if is a spanning tree of , then is a simplex contained in . In fact, is a full-dimensional simplex contained in the boundary of . Further still, is a unimodular simplex: it has normalized volume . It follows that the simplex is a full-dimensional, unimodular simplex contained in . It is desirable, then, to find a collection of oriented spanning trees of so that the set forms a triangulation of the boundary of . Such a set is called a dissecting tree set for .
Of course, if is a dissecting tree set for , then the normalized volume of is exactly . However, there is more information about to be extracted from . To describe this information, consider a fixed vertex in an oriented tree . For each edge we say that points away from in if is contained in the same component as the tail of in . Otherwise, we say points towards in .
Theorem 2.2 ([10, Theorem 1.2]).
Let be a connected graph, let be a dissecting tree set for , and let be any vertex of . Then the entry in is
In order to determine when we have a dissecting tree set, we turn to Gröbner basis techniques. Here, we will introduce the necessary algebraic background to prove our main results. We largely follow the framework given in [9]. For more details about monomial orders and their relationships to triangulations of polytopes, see [16, Chapter 8].
Recall that, given a lattice polytope and a field , we may construct a semigroup algebra
where . The toric ideal of is defined as the kernel of the map
defined by setting ; this ideal is denoted .
Consider a monomial order on and any ideal of this algebra. For each polynomial , its initial term with respect to , denoted , is the term of that is greatest with respect to . The initial ideal of with respect to is
A Gröbner basis of , which we denote by , is a finite generating set of such that . We call reduced if each element has a leading coefficient of and if for any , does not divide any term of . There are many Gröbner bases of but there is exactly one reduced Gröbner basis of .
Now suppose is an -dimensional lattice polytope and . Choose a weight vector such that the polytope
is -dimensional, i.e., does not lie in an affine hyperplane of . Some facets of will have outward-pointing normal vectors with a negative last coordinate. By projecting these facets back to , we obtain the facets of a polyhedral subdivision of . Any triangulation that can be obtained in this way for an appropriate choice of is called a regular triangulation. Importantly, regular unimodular triangulations are in one-to-one correspondence with reduced Gröbner bases whose initial terms are all squarefree [16, Corollary 8.9]. It is well-known that any monomial order obtained as a graded reverse lexicographic (grevlex) order can be represented by a weight vector.
In the case of for a regular matroid , we can identify with
(2.1) |
Here, we let correspond to the column of a fixed matrix realizing that corresponds to , and we will let correspond to the negative of this column. The variable , then, corresponds to the origin. So, for each , we can treat and as encoding an orientation of . For ease of notation, if we are considering with a particular orientation, then we let denote the corresponding variable and the variable corresponding to the opposite orientation.
Lemma 2.3 (cf. [9, Proposition 3.8]).
Let be a regular matroid on the ground set , let be an order on the lattice points of , and let be the grevlex order on (2.1) induced from this ordering on the variables. The collection of all binomials of the following types forms a Gröbner basis of with respect to grevlex order:
-
(i)
For every -element circuit of and choice of orientation for each , and any -element subset not containing the weakest element among those of ,
-
(ii)
For every -element circuit of and choice of orientation for each , and any -element subset ,
-
(iii)
For any , .
In each case, the binomial is written so that the initial term has positive sign.
Proof.
This proof follows the same approach as the proof to [9, Proposition 3.8]. First, recognize we need to show that for any binomial in , one of or is divisible by the leading monomial of one of the binomials described above.
Fix a totally unimodular matrix representing , and label its columns . Both monomials and correspond to a matrix so that is a column of the matrix if and only if is present in the monomial, and is a column of the matrix if and only if is present in the monomial. Let (resp. ) be the matrix corresponding to (resp. ). Of course, we may assume that neither matrix of and has both and for any , since then the monomial would be divisible by .
Consider . Since is contained in the toric ideal of , there is some minimal linearly dependent set of columns in the matrix . Denote by the set of vectors in that are columns in , and let . Similarly, denote by the set of vectors in that are columns in , and let . We may assume .
First, suppose that for some . If , consider the product of the largest variables, with respect to , among those corresponding to vectors in . The leading term of the corresponding binomial in (i) then divides . If , we may assume without loss of generality that the smallest variable corresponding to a vector in belongs to and proceed as before. If , then since , the leading term of the binomial in (ii) corresponding to vectors in divides . ∎
3. Extensions of known properties of SEPs
Before providing the extensions of known properties of ordinary SEPs to generalized SEPs, we recall the formula of the dimensions of them.
Proposition 3.1 (cf. [7, Theorem 4.3]).
Let be a regular matroid on the ground set . Then . In particular, given a connected graph with vertices, we have . Moreover, we have .
First, we extend [13, Proposition 4.4], using the same overall strategy used in its proof.
Proposition 3.2 (cf. [13, Proposition 4.4]).
Let be a bipartite regular matroid on . For any ,
Proof.
Let be a full-rank, totally unimodular matrix of rank representing , and denote its columns by so that . Also, let be the matrix representing . Since is bipartite, the Gröbner basis from Lemma 2.3 consists of the binomials of the form (i) and (iii). Let be the rank- matrix
and let be the matroid represented by . The Gröbner basis of consists of all binomials satisfying one of the following:
where is a set of columns of corresponding to a circuit in of size and , and is a -subset of such that where ;
where corresponds to a circuit in of size and is a -subset of ;
for . Thus, the initial terms in the Gröbner bases for and are the same. Following the proof strategy of [8, Theorem 3.1], it follows that
where is a standard basis vector, as claimed. ∎
Next we present an extension of [6, Theorem 44]. Its proof is entirely analogous to the proof of [6, Theorem 44], in the same way that our proof of Lemma 2.3 adapts the proof [9, Proposition 3.8], and our proof of Proposition 3.2 adapts the proof of [13, Proposition 4.4]. So, due to the length of the proof and the similarity in how we adapt it from [6, Theorem 44], we omit its details.
Theorem 3.3 (cf. [6, Theorem 44]).
Let be a regular matroid and let be a bipartite regular matroid. Then
4. A counterexample of -nonnegativity for generalized SEPs
In [13], the authors conjectured that SEPs are -nonnegative. It is natural to want to extend this conjecture to generalized SEPs, but the conjecture in this setting is false. Throughout the remainder of this section, we will be considering the matroids for .
Proposition 4.1.
Let with and let be the ground set of .
-
(i)
can be represented by , where
-
(ii)
is not graphic for any if .
Proof.
-
(i)
It is known that if a matroid is represented by a matrix , then is represented by , where denotes the transpose ([14, Theorem 2.2.8]). Hence, we prove that the matrix represents where .
Let and . Fix a spanning tree consisting of the following edges:
This corresponds to the identity matrix in the representing matrix of . Our goal is to show that . By assigning the orientation of each of these edges as described (i.e., from the first vertex to the second vertex), for the remaining edges of , we see that (resp. ) with form minimal dependent sets with the opposite of , the opposite of , and . This dependent set corresponds to the -th (resp. -th) column of , as claimed.
-
(ii)
It is known that given a matroid and an element of its ground set, we have (see [14, 3.1.1]). Hence, it is enough to show that never becomes planar for any edge of This is easy to see: the edge set of is , which contains as a subgraph.
∎
Theorem 4.2.
In every dimension at least there is a generalized SEP that is not -nonnegative.
Proof.
Consider the matroid . Since has edges, we know that . From its matrix description, it is possible to show that
hence .
Now let where the are pairwise distinct elements, none of which is already in . The generalized SEP is the -fold direct sum of with copies of the interval . Consequently, ,
from which it follows that . ∎
Computational evidence suggests that the -vectors for fail to be nonnegative for all even , although we do not aim to prove so here. Instead, we examine how these matroids are situated between ordinary SEPs and generalized SEPs: by deleting exactly two elements from , the resulting matroid is now graphic. In fact, by invoking Whitney’s -isomorphism theorem [14, Theorem 5.3.1], it is a brief exercise to show that there is, up to isomorphism, a unique graph such that . We will be more precise about what this graph looks like in Section 5. For now, we simply state our main results regarding .
Theorem 4.3.
For each ,
(4.1) |
where
For example, the following are the computations of for some small values of :
Our second main result regarding is the following. For readability, we defer its proof to Section 6.
Theorem 4.4.
For each ,
In particular, is -nonnegative for all .
As a corollary of Theorem 4.3, we can obtain the normalized volume of . For its proof, as well as the proof of Theorem 4.3, we collect some well-known formulas on the summation of binomial coefficients.
Lemma 4.5.
Let . Then the following hold:
Corollary 4.6.
The normalized volume of is
5. Proof of Theorem 4.3
In this section, we put all of the necessary pieces together for the proof of Theorem 4.3. We will produce a dissecting tree set by identifying which oriented spanning trees correspond to monomials that are not divisible by an initial monomial of one of the types in Lemma 2.3. Before this, we collect several identities needed in the proof of Theorem 4.3.
Lemma 5.1.
-
(i)
For all such that ,
-
(ii)
For all such that ,
Proof.
These straightforwardly follow by applying Lemma 4.5 (i) and (ii). ∎
Now, note that is never graphic for (see Proposition 4.1). However, by an appropriate choice from , we see that the matroid may be represented by the matrix where is the matrix with the second and third columns removed. This matrix also represents the graph on the ground set with edges , , for each , as well as . See Figure 1 for examples.

Remark 5.2.
We can obtain as the graph representing in a theoretical way. Although is not planar for any edge of as described in the proof of Proposition 4.1 (ii), we can get a planar graph by the contraction with a new edge of . By taking the dual graph of (in the sense of planar graphs), a planar graph appears.
In what follows, we concentrate on the graph . Let . To describe these spanning trees, it will be helpful to refer to a specific planar embedding of . We may view as the -cycle , together with the chords for each . In particular, throughout the remainder of the article, we will use “chord” to refer to any of the edges of the form .
Referring to a specific planar embedding will allow us to construct the -polynomial by applying Theorem 2.2. To do this, we need to identify a dissecting tree set for . Let be the reduced Gröbner basis of with respect to any grevlex order in which the variable corresponding to the origin is weakest and the variables and corresponding to are the second- and third-weakest. Using the framework developed in Section 2, we can identify which oriented spanning trees contribute to our dissecting tree set. To help us describe them, we introduce some new terminology.
Observe that if is a spanning tree of , then for each of the triangles with vertices , there are exactly three possibilities of edges appearing in : contains either
-
•
both and , which we call a chordless pair,
-
•
both and or both and , each of which we call a chorded pair, or
-
•
only or only , each of which we call an unpaired edge.
For example, the spanning tree of in Figure 2 contains one chordless pair of edges, two chorded pairs of edges, and one unpaired edge.

Proposition 5.3.
An oriented spanning tree of is part of the dissecting tree set if and only if, for every cycle in , either contains no more than edges of that are oriented in the same direction, or the following conditions holds:
-
When is even, both and contains exactly edges of that are oriented in the same direction.
Proof.
This follows from Lemma 2.3. ∎
Corollary 5.4.
Suppose is an oriented spanning tree in the dissecting tree set. Then the edges in the chordless pair share the same tail or share the same head. Moreover, their orientations may be reversed to obtain another oriented spanning tree in the dissecting tree set.
Proof.
If the edges in the chordless pair, say, and , share neither the same tail nor the same head, then the -cycle of contains two edges that are oriented in the same direction. ∎
For a spanning tree of , let denote the number of chords of and let
Lemma 5.5.
Let be a member of our dissecting tree set.
-
(i)
If , then contains a total of chordless pairs and chorded pairs of edges as well as exactly one unpaired edge.
-
(ii)
If , then contains a total of chordless and chorded pairs of edges. Moreover, must be always even in this case.
Proof.
Note that each spanning tree of has edges. Hence, the statements on the numbers of chordless/chorded pairs and unpaired edge follow. Regarding the evenness of , on the contrary, suppose that for some . Then there are at least edges that have the same direction. On the other hand, since there are chordless pairs that consist of edges, Corollary 5.4 claims that the exactly half of these edges should have the same oriantation. Hence, we can find a cycle consisting of those chords of chorded pairs, the edges of chordless pairs and , that have length , such that there are at least edges in in the same direction, a contradiction to Proposition 5.3. ∎
At this point we will take time here to closely examine some particular cases. We do this with the goal of helping motivate and illustrate the key aspects of the full proof, which becomes somewhat technical.
: Let be a member of our dissecting tree set with . Then the (undirected) edges of consist entirely of chordless pairs. When orienting this tree, each pair of edges , has two possibilities: either both edges point toward or both edges point away from (see Corollary 5.4). In either case, there are edges that will point towards . So, by Theorem 2.2 with , each such oriented tree corresponds to a summand of in . There are valid orientations of this tree, resulting in a summand of in the -polynomial after over all of these trees.
: Let be a member of our dissecting tree with . Then and (see Lemma 5.5 (ii)). We can establish orientations on the chordless pairs as in the previous case, although there are now only of these pairs. Any cycle containing one edge of a chordless pair must also contain the other. By Corollary 5.4, these two edges must have opposite orientations. For any cycle containing , the orientations of and the unpaired edge should have the opposite direction; otherwise we can find at least edges with the same direction in -cycle of (see Corollary 5.4 and Lemma 5.5).
For to contribute to , the orientation of and the unpaired edge must be opposite. To help describe which summand of the -polynomial corresponds to , we will refer to two portions of it relative to . Given a spanning tree of that contains , the graph consists of two components. Call the component containing the left side of and denote it by . The right side of the tree, which we denote , is the subgraph of consisting of and all edges not in .
This definition makes it clear that can correspond to any of the following summands of the -polynomial, depending on its orientation and the location of the unpaired edge, :
-
•
If and are pointing away each other and , then corresponds to the summand .
-
•
If and are pointing away each other and , then corresponds to the summand .
-
•
If and are pointing towards each other and , then corresponds to the summand .
-
•
If and are pointing towards each other and , then corresponds to the summand .
Thus, by considering all orientations on the edges of a spanning tree satisfying , the tree will correspond to a summand of . Note that .
: Because of one additional subtlety that arises when , we will discuss this case as well. When this happens, there are two subcases: the first being where and and the second being where and . In the first case, we can repeat the same argument as when , but we now have only possible chordless pairs. As before, the orientations on the chordless pairs result in a factor of in the contribution makes to the -polynomial. Such a factor occurs times, since this is the number of chords that can appear as parts of chorded pairs.
Now, for each chorded pair, the orientation on the edge dictates the orientation of the other edge of the pair: they have the opposite direction. However, whether the chorded pair consists of or will make a difference in the summand of the -polynomial corresponding to . We say that a chorded pair has type if the non-chord edge is incident to the vertex that is closer to within ; otherwise, we say that a chorded pair has type . See Figure 2 for an example of each.
When summing the monomials of the -polynomial corresponding to , if a chorded pair has type , then one of the orientations of its edges results in a factor of , while the other results in a factor of . On the other hand, if a chorded pair has type , then each of the orientations of its edges results in a factor of . So, if we consider all spanning trees of satisfying and , then we know that there are chorded pairs of type in and chorded pairs of type in for some nonnegative satisfying . This leaves chords to be part of a chorded pair of type . This is where the set in (4.1) will come from: the pair will indicate the presence of chorded pairs of type in and of them in , where the notation is explained below.
Recognize that in any oriented spanning tree of , there is at most one unpaired edge. Consequently, if an unpaired edge exists, then and the unpaired edge must have endpoints and for some and . Thus we may, without affecting the monomial to which corresponds, modify by deleting the vertex , adding a new vertex , and adding a new edge as follows:
-
•
If and are pointing away each other and , then add the directed edge .
-
•
If and are pointing away each other and , then add the directed edge .
-
•
If and are pointing towards each other and , then add the directed edge .
-
•
If and are pointing towards each other and , then add the directed edge .
See Figure 3 for illustrations of each of these cases. We call a modified (oriented) spanning tree of . Modified spanning trees allow us to consider and or as a chorded pair, so that all edges in are part of a chorded or chordless pair. Note that, in each case, the number of edges pointing towards or away from in each of and is preserved by this operation. We are now prepared to present the proof of Theorem 4.3.

Proof of Theorem 4.3.
To begin, we first show the second equation of (4.1) holds. Consider a summand
If for some , then adding this summand and the summand for becomes
Note that if , then the binomial coefficient becomes . So, there is no harm in replacing with . Further, since , the above expression may be written
Factoring appropriately results in the expression
Thus, ranging over all yields the second equation of (4.1).
Now, fix an arbitrary . We will treat the summand corresponding to in two pieces:
(5.1) |
and
(5.2) |
since these will more closely resemble how to identify the oriented spanning trees in our dissecting tree set, and their contributions to the -polynomial.
For (5.1), we interpret as the number of chords in the modified tree . Then there is exactly one such that all of , and are missing in . Namely, we treat both cases, and , and and , simultaneously. Passing to the modified spanning trees, we can say that is part of a chorded pair, and is of type or . Suppose first that is part of a pair of type in . Then of the chords may be chosen to be parts of chorded pairs of type , and these will all be part of . We may then choose of the chords to be part of chorded pairs of type , to ensure that contains of them. Of the remaining chords of the form (excluding ), there are of them, and we must choose of them to be the chorded pairs of type . All together we have constructed
trees. On the other hand, suppose is part of a chorded pair of type . Following an analogous argument, we obtain
spanning trees.
Now we may apply Lemma 5.1 (i) to obtain
spanning trees that contribute to the -polynomial when there are chords in the modified tree .
For each of the edges that are part of a chorded pair, the structure of and Proposition 5.3 imply that exactly half of them must be pointing towards and exactly half of them must be pointing away from . The contribution to the -polynomial depends on whether those chorded pairs are of type or type .
For each of these trees, we can determine a valid orientation by considering which orientations satisfy Proposition 5.3. For each spanning tree with chorded pairs of type in , we choose of them to point toward . Similarly, we choose of them from to point toward . There are chorded pairs remaining, all of type ; we must orient these so that exactly of the chords are pointing in the same direction. By selecting which pairs are oriented so that the chord is pointing counterclockwise in our planar embedding of , this must be done in ways, since of the edges have already been given counterclockwise orientations.
For each of these trees, there will be edges oriented in toward that come from the chordless pairs, and this is true for either possible orientation of the pairs. From the paragraph above, there will be edges pointing toward that come from chorded pairs of type , and edges pointing toward that come from chorded pairs of type . Summing over all and gives us a summand of
in the computation of the -polynomial. Further, because we are considering the case of modified spanning trees, we have shown that there are
of them. By multiplying the above two expressions and summing over all , we obtain (5.1).
Now, suppose . By Lemma 5.5, . We can apply an argument analogous to the one above, using Lemma 5.1 (ii), to show that there are
oriented spanning trees with and choreded pairs of type in and in , respectively. Note that the argument holds for each possibility of whether is of type or . So, we must multiply each of these by the sum of the -polynomial contributions from these chorded pairs, which is . This establishes (5.2), which completes the proof.
∎
6. Proof of Theorem 4.4
To prove Theorem 4.4, we will use a generating function argument. We will make regular use of the following three well-known identities:
(6.1) |
(6.2) |
(6.3) |
The first lemma we will need is the following.
Lemma 6.1.
For all we have
Proof.
Using (6.1), (6.2) we compute the generating function for the left side of the identity:
Applying the Taylor expansion to this rational function, we find
∎
Lemma 6.2.
For all ,
Proof.
By Taylor expansion, we have
∎
The second lemma we will need is the following.
Lemma 6.3.
For ,
Proof.
We again use a generating function argument, beginning with the left side of the desired identity multiplied by .
∎
Theorem 6.4.
For all we have
Proof.
For notational convenience, set
and
We compute the generating function for as follows:
Hence it is enough to show that
Using Lemma 6.3, we have
where . It follows that , that
and that
Hence,
Thus
Since is the coefficient of in , we have
Additionally, from Lemma 6.1,
Moreover, from Lemma 6.2,
as desired. ∎
References
- [1] Matthias Beck and Sinai Robins. Computing the continuous discretely. Undergraduate Texts in Mathematics. Springer, New York, 2007. Integer-point enumeration in polyhedra.
- [2] Benjamin Braun. Unimodality problems in Ehrhart theory. In Recent trends in combinatorics, volume 159 of IMA Vol. Math. Appl., pages 687–711. Springer, 2016.
- [3] Tianran Chen. Directed acyclic decomposition of Kuramoto equations. Chaos, 29(9):093101, 12, 2019.
- [4] Tianran Chen, Robert Davis, and Evgeniia Korchevskaia. Facets and facet subgraphs of symmetric edge polytopes. Discrete Appl. Math., 328:139–153, 2023.
- [5] Tianran Chen, Robert Davis, and Dhagash Mehta. Counting equilibria of the Kuramoto model using birationally invariant intersection index. SIAM J. Appl. Algebra Geom., 2(4):489–507, 2018.
- [6] Alessio D’Alì, Emanuele Delucchi, and Mateusz Michałek. Many faces of symmetric edge polytopes. Electron. J. Combin., 29(3):Paper No. 3.24, 42, 2022.
- [7] Alessio D’Alì, Martina Juhnke-Kubitzke, and Melissa Koch. On a generalization of symmetric edge polytopes to regular matroids, 2023. arXiv:2307.04933.
- [8] Takayuki Hibi and Akiyoshi Tsuchiya. Reflexive polytopes arising from perfect graphs. Journal of Combinatorial Theory, Series A, 157:233–246, 2018.
- [9] Akihiro Higashitani, Katharina Jochemko, and Mateusz Michałek. Arithmetic aspects of symmetric edge polytopes. Mathematika, 65(3):763–784, 2019.
- [10] Tamás Kálmán and Lilla Tóthmérész. -vectors of graph polytopes using activities of dissecting spanning trees, 2023. arXiv:2203.17127.
- [11] Tetsushi Matsui, Akihiro Higashitani, Yuuki Nagazawa, Hidefumi Ohsugi, and Takayuki Hibi. Roots of Ehrhart polynomials arising from graphs. J. Algebr. Comb., 34(4):721–749, 2011.
- [12] Hidefumi Ohsugi and Takayuki Hibi. Centrally symmetric configurations of integer matrices. Nagoya Math. J., 216:153–170, 2014.
- [13] Hidefumi Ohsugi and Akiyoshi Tsuchiya. The -polynomials of locally anti-blocking lattice polytopes and their -positivity. Discrete Comput. Geom., 66(2):701–722, 2021.
- [14] James G. Oxley. Matroid theory. Oxford Science Publications. The Clarendon Press, Oxford University Press, New York, 1992.
- [15] Richard P. Stanley. Combinatorics and commutative algebra, volume 41 of Progress in Mathematics. Birkhäuser Boston, Inc., Boston, MA, second edition, 1996.
- [16] Bernd Sturmfels. Gröbner bases and convex polytopes, volume 8 of University Lecture Series. American Mathematical Society, Providence, RI, 1996.
- [17] Lilla Tóthmérész. A geometric proof for the root-independence of the greedoid polynomial of eulerian branching greedoids, 2022. arXiv:2204.12419.