1 Introduction

A lattice polytope \({\mathscr {P}}\subset {\mathbb R}^n\) is a convex polytope all of whose vertices have integer coordinates. Its normalized volume, \({{\,\mathrm{Vol}\,}}({\mathscr {P}})=\dim ({\mathscr {P}})! {{\,\mathrm{vol}\,}}({\mathscr {P}})\) where \({{\,\mathrm{vol}\,}}({\mathscr {P}})\) is the relative volume of \({\mathscr {P}}\), is always a positive integer. To compute \({{\,\mathrm{Vol}\,}}({\mathscr {P}})\) is a fundamental but hard problem in polyhedral geometry.

Let G be a simple graph on \([n]:=\{1,\ldots ,n\}\) with edge set E(G). The PV-type adjacency polytope \(\nabla ^\mathrm{PV}_G\) of G is the lattice polytope which is the convex hull of

$$\begin{aligned}\{ \pm ({\mathbf{e}}_i-{\mathbf{e}}_j) \in {\mathbb R}^n : \{i,j\} \in E(G)\},\end{aligned}$$

where \({\mathbf{e}}_i\) is the i-th unit coordinate vector in \({\mathbb R}^n\). The normalized volumes of PV-type adjacency polytopes have attracted much attention. In fact, the normalized volume of a PV-type adjacency polytope gives an upper bound on the number of possible solutions in the Kuramoto equations [5], which models the behavior of interacting oscillators [13]. For several classes of graphs, explicit formulas for the normalized volume of their PV-type adjacency polytopes have been given (e.g., [1, 7, 10]). In particular, we can compute the normalized volume of the PV-type adjacency polytope of a suspension graph by using interior polynomials [15]. Here interior polynomials are a version of the Tutte polynomials for hypergraphs introduced by Kálmán [11].

On the other hand, the PQ-type adjacency polytope \(\nabla ^PQ _G\) of G is the lattice polytope which is the convex hull of

$$\begin{aligned}\{ ({\mathbf{e}}_i,{\mathbf{e}}_j) \in {\mathbb R}^{2n} : \{i,j\} \in E(G)\; \text{ or }\; i=j\}.\end{aligned}$$

Note that an edge \(\{i,j\} \in E(G)\) results in both \(({\mathbf{e}}_i, {\mathbf{e}}_j)\) and \(({\mathbf{e}}_j, {\mathbf{e}}_i)\). There is a connection between PQ-type adjacency polytopes and the engineering problem known as power-flow study, which models the balances of electric power on a network of power generation [6]. In fact, the normalized volume of a PQ-type adjacency polytope gives an upper bound on the number of possible solutions in the algebraic power-flow equations. In the present paper, we focus on the \(h^*\)-polynomial of a PQ-type adjacency polytope. Here, the \(h^*\)-polynomial \(h^*({\mathscr {P}},x)\) of a lattice polytope \({\mathscr {P}}\) is a discrete tool to compute the normalized volume \({{\,\mathrm{Vol}\,}}({\mathscr {P}})\) (see Sect. 2).

We recall a relation between \(\nabla ^PQ _G\) and a root polytope. For a bipartite graph H on [n] with edge set E(H), the root polytope \({\mathcal Q}_H\) of H is the lattice polytope which is the convex hull of

$$\begin{aligned}\{{\mathbf{e}}_i+ {\mathbf{e}}_j \in {\mathbb R}^n:\{i,j\} \in E(H)\}.\end{aligned}$$

For a positive integer n, set \([\overline{n}]:=\{\overline{1},\ldots ,\overline{n}\}\). Define D(G) to be the bipartite graph on \([n] \cup [\overline{n}]\) with edges \(\{i, \overline{i}\}\) for each \(i \in [n]\) and \(\{i, \overline{j}\}\) and \(\{\overline{i},j\}\) for each edge \(\{i,j\}\) in G. It then follows that \(\nabla ^PQ _G\) is unimodularly equivalent to \({\mathcal Q}_{D(G)}\) [8, Lem. 2.4]. On the other hand, it is known [12] that the \(h^*\)-polynomial of the root polytope \({\mathcal Q}_H\) of a connected bipartite graph H coincides with the interior polynomial \(I_H(x)\) of the associated hypergraph of H. In particular, the normalized volume of \({\mathcal Q}_H\) is equal to \(|{{{\,\mathrm{HT}\,}}(H)}|\), where \({{\,\mathrm{HT}\,}}(H)\) denotes the set of hypertrees of an associated hypergraph of H. Therefore, we can compute the \(h^*\)-polynomial and the normalized volume of \(\nabla ^PQ _G\) of a connected graph G by using an interior polynomial and counting hypertrees. In the terminology of [17], a hypertree is called a draconian sequence [17, Defn. 9.2]. Moreover, Davis and Chen [8] have studied the normalized volume of \(\nabla ^PQ _G\) by using draconian sequences.

The main results of the present paper are formulas of the \(h^*\)-polynomial and the normalized volume of \(\nabla ^PQ _G\) of a join graph G. Let \(G_1,\ldots ,G_s\) be graphs with \(m_1,\ldots ,m_s\) vertices. Suppose that \(G_i\) and \(G_j\) have no common vertices for each \(i\ne j\). Then the join \(G_1 + \cdots + G_s\) of \(G_1,\ldots ,G_s\) is obtained from \(G_1 \cup \cdots \cup G_s\) joining each vertex of \(G_i\) to each vertex of \(G_j\) for any \(i \ne j\). Note that \(G_1 + \cdots + G_s\) \((s>1)\) is connected and hence so is \(D(G_1 + \cdots + G_s)\). For example, the complete bipartite graph \(K_{\ell ,m}\) is equal to the join \(E_\ell + E_m\) where \(E_k\) is the empty graph with k vertices. For complete graphs \(K_\ell \) and \(K_m\), one has \(K_\ell + K_m = K_{\ell +m}\). We can compute the \(h^*\)-polynomial and the normalized volume of the PQ-type adjacency polytope of a join graph by using perfectly matchable set polynomials (see Sect. 2 for the definition of perfectly matchable set polynomials).

Theorem 1.1

Let \(G_1,\ldots ,G_s\) be graphs with \(m_1,\ldots ,m_s\) vertices, respectively. Suppose that \(G_i\) and \(G_j\) have no common vertices for each \(i \ne j\). Then for the join \(G= G_1 + \cdots + G_s\) with \(m = \sum _{i=1}^s m_i\) vertices, we have

$$\begin{aligned} \begin{aligned} h^*\bigl (\nabla ^PQ _G ,x\bigr )&=\sum _{i=1}^s h^*\bigl (\nabla ^PQ _{G_i+K_{m-m_i}},x\bigr ) -(s-1)\sum _{k=0}^{m -1}\left( {\begin{array}{c}m-1\\ k\end{array}}\right) ^2x^k\\&=\sum _{i=1}^s I_{D(G_i+K_{m-m_i})}(x) -(s-1)\sum _{k=0}^{m -1}\left( {\begin{array}{c}m-1\\ k\end{array}}\right) ^2x^k\\&=\sum _{i=1}^sp(D(G_i+K_{m-m_i-1}), x)-(s-1)\sum _{k=0}^{m -1}\left( {\begin{array}{c}m-1\\ k\end{array}}\right) ^2x^k, \end{aligned} \end{aligned}$$
(1)

where \(p(H,x)\) denotes the perfectly matchable set polynomial of a graph H. In particular, one has

$$\begin{aligned} {{\,\mathrm{Vol}\,}}\bigl (\nabla ^PQ _G\bigr )&=\sum _{i=1}^s{{\,\mathrm{Vol}\,}}\bigl (\nabla ^PQ _{G_i+K_{m-m_i}}\bigr )-(s-1)\left( {\begin{array}{c}2(m-1)\\ m-1\end{array}}\right) \\&=\sum _{i=1}^s|{{{\,\mathrm{HT}\,}}(D(G_i+K_{m-m_i}))}|-(s-1)\left( {\begin{array}{c}2(m-1)\\ m-1\end{array}}\right) \\&=\sum _{i=1}^s|PM (D(G_i+K_{m-m_i-1}))|-(s-1)\left( {\begin{array}{c}2(m-1)\\ m-1\end{array}}\right) , \end{aligned}$$

where \(PM (H)\) denotes the set of perfectly matchable sets of a graph H.

By using this theorem, we give explicit formulas of the \(h^*\)-polynomial and the normalized volume of \(\nabla ^PQ _G\) when G is a complete multipartite graph (Corollary 4.4). On the other hand, Theorem 1.1 is not useful for computing the \(h^*\)-polynomial and the normalized volume of \(\nabla ^PQ _{G}\) when G is a wheel graph \(W_n\), that is, G is the join of a cycle \(C_n\) and \(K_1\). We give explicit formulas for the \(h^*\)-polynomial and the normalized volume of \(\nabla _{W_n}^PQ \) and prove the conjecture [8, Conj. 4.4] on the normalized volume of \(\nabla _{W_n}^PQ \) (Theorem 5.1) by computing the perfectly matchable set polynomial of \(D(C_n)\).

The paper is organized as follows: After reviewing the definitions and properties of the \(h^*\)-polynomials of lattice polytopes and the interior polynomials of connected bipartite graphs in Sect. 2, we give a proof of Theorem 1.1 in Sect. 3. By using Theorem 1.1, explicit formulas of the \(h^*\)-polynomial and the normalized volume of the PQ-type adjacency polytope of a complete multipartite graph are presented in Sect. 4. Finally, we compute the \(h^*\)-polynomial and the normalized volume of the PQ-type adjacency polytope of a wheel graph in Sect. 5.

2 Preliminaries

As explained in the previous section, the \(h^*\)-polynomial of \(\nabla ^PQ _G\) is equal to the interior polynomial of D(G). First, we give a brief introduction of Ehrhart polynomials and \(h^*\)-polynomials. We refer the reader to [4] for the detailed information for them. Let \({\mathscr {P}}\subset {\mathbb R}^n\) be a lattice polytope of dimension d. Given a positive integer t, we define

$$\begin{aligned}L_{{\mathscr {P}}}(t)=|t {\mathscr {P}}\cap {\mathbb Z}^n|,\end{aligned}$$

where \(t{\mathscr {P}}:=\{ t {\mathbf{x}}\in {\mathbb R}^n : {\mathbf{x}}\in {\mathscr {P}}\}\). The study on \(L_{{\mathscr {P}}}(t)\) originated in Ehrhart [9] who proved that \(L_{{\mathscr {P}}}(t)\) is a polynomial in t of degree d with the constant term 1. We call \(L_{{\mathscr {P}}}(t)\) the Ehrhart polynomial of \({\mathscr {P}}\). The generating function of the lattice point enumerator, i.e., the formal power series

$$\begin{aligned}Ehr _{\mathscr {P}}(x)=1+\sum _{k=1}^{\infty }L_{{\mathscr {P}}}(k) x^k\end{aligned}$$

is called the Ehrhart series of \({\mathscr {P}}\). It is known that it can be expressed as a rational function of the form

$$\begin{aligned}Ehr _{\mathscr {P}}(x)=\frac{h^*({\mathscr {P}},x)}{(1-x)^{d+1}},\end{aligned}$$

where \(h^*({\mathscr {P}},x)\) is a polynomial in x of degree at most d with nonnegative integer coefficients called the \(h^*\)-polynomial (or the \(\delta \)-polynomial) of \({\mathscr {P}}\). Moreover,

$$\begin{aligned}h^*({\mathscr {P}},x)=\sum _{i=0}^{d} h_i^* x^i\end{aligned}$$

satisfies \(h^*_0=1\), \(h^*_1=|{\mathscr {P}}\cap {\mathbb Z}^n|-(d +1)\), and \(h^*_{d}=|int ({\mathscr {P}}) \cap {\mathbb Z}^n|\), where \(int ({\mathscr {P}})\) is the relative interior of \({\mathscr {P}}\). Furthermore, \(h^*({\mathscr {P}},1)=\sum _{i=0}^{d} h_i^*\) is equal to the normalized volume \({{\,\mathrm{Vol}\,}}({\mathscr {P}})\) of \({\mathscr {P}}\).

Next, we recall the definition of interior polynomials and their properties. A hypergraph is a pair \({\mathcal H}= (V, E)\), where \(E=\{e_1,\ldots ,e_n\}\) is a finite multiset of non-empty subsets of \(V=\{v_1,\ldots ,v_m\}\). Elements of V are called vertices and the elements of E are the hyperedges. Then we can associate \({\mathcal H}\) with a bipartite graph \({{\,\mathrm{Bip}\,}}{\mathcal H}\) on the vertex set \(E \sqcup V\) with the edge set \(\{ \{e_j, v_i\} : v_i \in e_j\}\). Assume that \({{\,\mathrm{Bip}\,}}{\mathcal H}\) is connected. A hypertree in \({\mathcal H}\) is a function \(f:E \rightarrow {\mathbb Z}_{\ge 0}\) such that there exists a spanning tree \(\Gamma \) of \({{\,\mathrm{Bip}\,}}{\mathcal H}\) whose vertices have degree \(f (e) +1\) at each \(e \in E\). Then we say that \(\Gamma \) induces f. For example, if \({\mathcal H}\) is a hypergraph with \(V=\{v_1,v_2,v_3\}\) and \(E=\{e_1=\{v_1,v_2,v_4\}, e_2=\{v_2,v_3,v_4\}\}\), then \({{\,\mathrm{Bip}\,}}{\mathcal H}\) is a bipartite graph whose edge set is \(\{ \{e_1, v_1\}, \{e_1, v_2\}, \{e_1, v_4\}, \{e_2, v_2\}, \{e_2, v_3\}, \{e_2, v_4\}\}\). A spanning tree

$$\begin{aligned}\Gamma = \{ \{e_1, v_1\}, \{e_1, v_2\}, \{e_1, v_4\}, \{e_2, v_2\}, \{e_2, v_3\}\}\end{aligned}$$

of \({{\,\mathrm{Bip}\,}}{\mathcal H}\) induces a hypertree \(f:E \rightarrow {\mathbb Z}_{\ge 0}\) with \(f(e_1) = 2\) and \(f(e_2) = 1\). Let \({{\,\mathrm{HT}\,}}({\mathcal H})\) denote the set of all hypertrees in \({\mathcal H}\). A hyperedge \(e_j \in E\) is said to be internally inactive with respect to the hypertree f if there exists \(j' < j\) such that \(g:E \rightarrow {\mathbb Z}_{\ge 0}\) defined by

$$\begin{aligned}g(e_i) = {\left\{ \begin{array}{ll}f(e_i)+1 &{}\quad \text {if}\; i=j',\\ f(e_i)-1 &{}\quad \text {if}\;i=j,\\ f(e_i) &{} \quad \text {otherwise},\end{array}\right. }\end{aligned}$$

is a hypertree. Note that it depends on the ordering \(e_1,\ldots ,e_n\) of hyperedges. In particular, \(e_1\) is not internally inactive in general. For example, with respect to the hypertree f with \(f(e_1) = 2\) and \(f(e_2) = 1\) above, there exists no internally inactive hyperedges. Let \(\overline{\iota } (f) \) be the number of internally inactive hyperedges of f. Then the interior polynomial of \({\mathcal H}\) is the generating function \(I_{\mathcal H}(x) = \sum _{f \in {{\,\mathrm{HT}\,}}({\mathcal H})} x^{ \overline{\iota } (f)}\). It is known [11, Prop. 6.1] that \(\deg I_{\mathcal H}(x)\le \min {\{|V|,|E|\}}-1\). If \(G ={{\,\mathrm{Bip}\,}}{\mathcal H}\), then we set \({{\,\mathrm{HT}\,}}(G) ={{\,\mathrm{HT}\,}}({\mathcal H})\) and \(I_G (x) = I_{\mathcal H}(x)\). The coefficients of \(I_G(x)\) are described as follows.

Proposition 2.1

[11, Thm. 3.4] Let G be a connected bipartite graph on the vertex set \(V_1 \sqcup V_2\) where \(V_1=\{v_1,\ldots ,v_p\}\) and \(|V_2|=q\). Then the coefficient of \(x^k\) in \(I_G(x)\) is the number of functions \(f:V_1 \rightarrow {\mathbb Z}_{\ge 0}\) such that

  1. (i)

    \(\displaystyle \sum _{i=1}^p f(v_i) = q -1\);

  2. (ii)

    \(\displaystyle \sum _{v \in V'} f(v) \le |\Gamma _G(V')| -1\) for all \(V' \subset V_1\), where \(\Gamma _G(S) \subset V_2\) is the set of vertices adjacent to some vertex in S;

  3. (iii)

    \(\overline{\iota } (f) =k\), i.e., \(|\eta _G(f)|=k\) where \(\eta _G(f)\) is the set of vertices \(v_j \in V_1\) satisfying the following condition: there exists \( j' < j \) such that the function \(g:V_1 \rightarrow {\mathbb Z}_{\ge 0}\) defined by

    $$\begin{aligned}g(v_i) = {\left\{ \begin{array}{ll}f(v_i)+1 &{}\quad \text {if}\; i=j',\\ f(v_i)-1 &{} \quad \text {if}\;i=j,\\ f(v_i) &{}\quad \text {otherwise},\end{array}\right. }\end{aligned}$$

    satisfies condition (ii) above.

From [8, Lem. 2.4] and [12, Thms. 1.1 and 3.10], we have the following.

Proposition 2.2

Let G be a connected graph. Then \(h^*(\nabla _G^PQ ,x) = I_{D(G)}(x)\). In particular, the normalized volume of \(\nabla _G^PQ \) is \(I_{D(G)}(1)=|{{{\,\mathrm{HT}\,}}(D(G))}|\).

Let G be a finite graph on the vertex set \(V=[n]\). A k-matching of G is a set of k pairwise non-adjacent edges of G. The matching generating polynomial of G is

$$\begin{aligned}g(G,x) =\sum _{k \ge 0} m_G(k) x^k,\end{aligned}$$

where \(m_G(k)\) is the number of k-matchings of G. A k-matching of G is said to be perfect if \(2k =n\). A subset \(S \subset V\) is called a perfectly matchable set [2] if the induced subgraph of G on the vertex set S has a perfect matching. Let \(PM (G, k)\) be the set of all perfectly matchable sets S of G with \(|S|=2k\) and \(PM (G)\) the set of all perfectly matchable sets of G. We regard \(\emptyset \) as a perfectly matchable set and we set \(PM (G,0) = \{\emptyset \}\). Note that \(|PM (G, k)| \le m_G(k)\) holds in general. We call the polynomial

$$\begin{aligned}p(G,x) = \sum _{k \ge 0}|PM (G,k)| x^k\end{aligned}$$

the perfectly matchable set polynomial (PMS polynomial) of G.

Example 2.3

For the cycle \(C_4\) of length 4, we have \(g(C_4,x) = 2 x^2 + 4x+1\) and \(p(C_4,x) = x^2 + 4x+1\). On the other hand, if a graph G has no even cycles, then we have \(g(G,x) = p(G,x)\).

Assume that G is a bipartite graph with a bipartition \([n] =V_1 \sqcup V_2\). Then let \(\widetilde{G}\) be a connected bipartite graph on \([n+2]\) whose edge set is

$$\begin{aligned}E(\widetilde{G}) = E(G) \cup \{ \{i, n+1\} : i \in V_1\} \cup \{ \{j, n+2\} : j \in V_2 \cup \{n+1\}\}.\end{aligned}$$

Proposition 2.4

[14, Prop. 3.4] Let G be a bipartite graph. Then we have \(I_{\widetilde{G}}(x)= p(G,x)\).

Although the following lemma is easy to see, it will be useful.

Lemma 2.5

Let G be a graph. Then we have \(D(G+K_1) = \widetilde{D(G)}\).

Moreover, we have

Proposition 2.6

Let G be a graph with n vertices. Then the \(h^*\)-polynomial of \(\nabla _{G+K_1}^PQ \) is \(p(D(G), x)\). In particular, the normalized volume of \(\nabla _{G+K_1}^PQ \) is \(|PM (D(G))|\).

Proof

The assertion follows from Propositions 2.2, 2.4, and Lemma 2.5. \(\square \)

Remark 2.7

Given a graph G, the number of matchings of G is called the Hosoya index of G and is denoted by Z(G). From Proposition 2.6, the normalized volume of \(\nabla _{G+K_1}^PQ \) is at most Z(D(G)).

Let G be a graph on the vertex set [n]. Given a subset \(S \subset [n]\), we associate the (0, 1)-vector \(\rho (S)=\sum _{i \in S} \mathbf{e}_i \in {\mathbb R}^n\). For example, \(\rho (\emptyset ) = \mathbf{0} \in {\mathbb R}^n\). The convex hull of

$$\begin{aligned}\{ \rho (S) : S \in PM (G,k) \text{ for } \text{ some }\; k \}\end{aligned}$$

is called a perfectly matchable subgraph polytope (PMS polytope) of G. A system of linear inequalities for a PMS polytope of G was given in [2] for bipartite graphs, and in [3] for arbitrary graphs.

Proposition 2.8

[2, Thm. 1] Let G be a bipartite graph on the vertex set \([n]=V_1 \sqcup V_2\). Then the PMS polytope of G is the set of all vectors \((x_1,\ldots ,x_n) \in {\mathbb R}^n\) such that

$$\begin{aligned} 0&\le x_i \le 1\quad \text {for each}\quad i=1,2,\ldots ,n,\\ \sum _{i \in V_1} x_i&= \sum _{j \in V_2} x_j,\\ \sum _{i \in S} x_i&\le \sum _{j \in \Gamma _G(S)}\!\! x_j \quad \text{ for } \text{ all }\quad S \subset V_1, \end{aligned}$$

where \(\Gamma _G(S) \subset V_2\) is the set of vertices adjacent to some vertex in S.

3 Join Graphs

In the present section, we give a proof of Theorem 1.1. Given a graph G and a nonnegative integer k, let \(F_G(k)\) be the set of functions satisfying conditions (i)–(iii) in Proposition 2.1 for D(G).

Lemma 3.1

Let \(G_1\) and \(G_2\) be graphs with \(m_1\) and \(m_2\) vertices, respectively. Suppose that \(G_1\) and \(G_2\) have no common vertices. Then for the join \(G= G_1 +G_2\) with \(m = m_1+m_2\) vertices, we have

$$\begin{aligned}&\bigsqcup _{k} F_G(k) = \bigsqcup _{k}\;F_{G_1+K_{m_2}}(k)\,\cap \,\bigsqcup _{k}\;F_{K_{m_1}+G_2}(k), \end{aligned}$$
(2)
$$\begin{aligned}&\bigsqcup _{k}\;F_{G_1+K_{m_2}}(k) \cup \bigsqcup _{k} \, F_{K_{m_1}+G_2}(k)=\bigsqcup _{k} F_{K_m}(k). \end{aligned}$$
(3)

Proof

Let \(V_1 \sqcup V_2\) with \(V_1=\{v_1,\ldots , v_{m}\}\) and \(V_2=\{v_1',\ldots , v_{m}'\}\) denote the common vertex set of D(G), \(D(G_1+K_{m_2})\), \(D(K_{m_1}+G_2)\), and \(D(K_m)\) \((=K_{m,m})\). In addition, let \(V_{1,1} =\{v_1,\ldots , v_{m_1}\}\), \(V_{1,2} =\{v_{m_1+1},\ldots , v_m\}\), \(V_{2,1} =\{v_1',\ldots , v_{m_1}'\}\), and \(V_{2,2} =\{v_{m_1+1}',\ldots , v_m'\}\), where \(V_{1,j} \sqcup V_{2,j}\) corresponds to the vertex set of \(D(G_j)\) for \(j =1,2\). First we will show that (2) holds. Let H be one of D(G), \(D(G_1+K_{m_2})\), \(D(K_{m_1}+G_2)\). Since \(\bigsqcup _{k} F_H(k)\) consists of functions satisfying conditions (i) and (ii) in Proposition 2.1, we study relations between conditions (i) and (ii) as follows:

Case 1  (\(H = D(G_1+K_{m_2})\)). Suppose that f satisfies condition (i). If \(v \in V' \subset V_1\) for some \(v \in V_{1,2} \), then condition (ii) holds for \(V'\) since \(\deg (v) = m\). Thus, if f satisfies condition (i), then condition (ii) in Proposition 2.1 is equivalent to the condition

$$\begin{aligned} \sum _{v \in V'} f(v) \le |\Gamma _{D(G_1)}(V')| + m_2 -1\quad \text{ for } \text{ all }\quad V' \subset V_{1,1}. \end{aligned}$$
(4)

Case 2  (\(H = D(K_{m_1}+G_2)\)). By the similar argument as in Case 1, if f satisfies condition (i), then condition (ii) in Proposition 2.1 is equivalent to the condition

$$\begin{aligned} \sum _{v \in V'} f(v) \le |\Gamma _{D(G_2)}(V')| + m_1 -1\quad \text{ for } \text{ all }\quad V' \subset V_{1,2}. \end{aligned}$$
(5)

Case 3  (\(H = D(G)\)). Suppose that f satisfies condition (i). If \(v,v' \in V' \subset V_1\) for some \(v \in V_{1,1}\) and \(v' \in V_{1,2}\), then condition (ii) holds for \(V'\) since \(\Gamma _{D(G)}(\{v,v'\}) =V_2\). Hence, if f satisfies (i), then condition (ii) in Proposition 2.1 holds if and only if both (4) and (5) hold.

Hence (2) holds.

For the graph \(D(K_m)\), since \( |\Gamma _{D(K_m)}(V')| -1 = m-1\) for all \(V' \subset V_1\), condition (i) implies condition (ii) in Proposition 2.1. Since any element in \(F_{G_1+K_{m_2}}(k) \cup F_{K_{m_1}+G_2}(k)\) satisfies condition (i), we have

$$\begin{aligned}\bigsqcup _{k}\,F_{G_1+K_{m_2}}(k) \,\cup \,\bigsqcup _{k}\,F_{K_{m_1}+G_2}(k)\,\subset \,\bigsqcup _{k}\, F_{K_m}(k).\end{aligned}$$

Suppose that f satisfies none of (4) and (5). Then

$$\begin{aligned} \sum _{v \in V'} f(v)&\ge |\Gamma _{D(G_1)}(V')| + m_2\quad \text {for some}\quad V' \subset V_{1,1},\\ \sum _{v \in V''} f(v)&\ge |\Gamma _{D(G_2)}(V'')| + m_1\quad \text {for some}\quad V'' \subset V_{1,2}. \end{aligned}$$

It then follows that \(\sum _{v \in V_1} f(v) \ge m_1 + m_2 =m\). Hence, if f satisfies condition (i), then at least one of (4) or (5) holds. Thus (3) holds. \(\square \)

Lemma 3.2

Let \(G_1\) and \(G_2\) be graphs with \(m_1\) and \(m_2\) vertices, respectively. Suppose that \(G_1\) and \(G_2\) have no common vertices. Then for the join \(G= G_1 +G_2\) with \(m = m_1+m_2\) vertices, \(F_{K_m}(k)\) is decomposed into the disjoint sets

$$\begin{aligned}F_{K_m}(k)=F_{K_{m_1}+G_2}(k) \sqcup (F_{G_1+K_{m_2}}(k) \setminus F_G(k))\end{aligned}$$

as in Fig. 1.

Fig. 1
figure 1

Decomposition of \(F_{K_m}(k)\)

Proof

Let \(V_1 \sqcup V_2\) with \(V_1=\{v_1,\ldots , v_{m}\}\) and \(V_2=\{v_1',\ldots , v_{m}'\}\) denote the common vertex set of D(G), \(D(G_1+K_{m_2})\), \(D(K_{m_1}+G_2)\), and \(D(K_m)\) \((=K_{m,m})\). In addition, let \(V_{1,1} =\{v_1,\ldots , v_{m_1}\}\), \(V_{1,2} =\{v_{m_1+1},\ldots , v_m\}\), \(V_{2,1} =\{v_1',\ldots , v_{m_1}'\}\), and \(V_{2,2} =\{v_{m_1+1}',\ldots , v_m'\}\), where \(V_{1,j} \sqcup V_{2,j}\) corresponds to the vertex set of \(D(G_j)\) for \(j =1,2\).

Claim 1

\(F_{K_{m_1}+G_2}(k) \subset F_{K_m}(k)\).

Suppose that f belongs to \(F_{K_{m_1}+G_2}(k)\). From Lemma 3.1 (3), we have \(f \in F_{K_m}(\ell )\) for some \(\ell \). Then condition (ii) in Proposition 2.1 is independent from the value \(f(v_1)\) since \(\deg (v_1) =m\). Thus one can choose \(j'=1\) for condition (iii) for any \(j >1\), and hence we have

$$\begin{aligned}k = |\{v_j \in V_1 : v_j \ne v_1, \,f(v_j) >0\}| = \ell .\end{aligned}$$

Therefore

$$\begin{aligned} F_{K_{m_1}+G_2}(k) \subset F_{K_m}(k) \end{aligned}$$
(6)

and \(F_{K_{m_1}+G_2}(k) \cap F_{K_m}(\ell ) =\emptyset \) if \(k \ne \ell \).

Claim 2

\(F_G(k) \subset F_{G_1+K_{m_2}}(k)\).

Suppose that \(f \in F_G(k)\). From (2) of Lemma 3.1 we have \(f \in F_{G_1+K_{m_2}}(\ell )\) for some \(\ell \). Note that \(k = |\eta _{D(G)} (f)|\) and \(\ell = |\eta _{D(G_1+K_{m_2})} (f)|\) (where \(\eta \) is defined in Proposition 2.1). In order to prove \(k=\ell \), we will show that \(\eta _{D(G)} (f) = \eta _{D(G_1+K_{m_2})} (f)\). Suppose that \(f(v_j) >0\) for \(v_j \in V_1\), \(j \ne 1\).

Case 1  (\(j \le m_1+1\)). From the argument in proof of Lemma 3.1, for the graph \(D(G_1+K_{m_2})\) (resp. D(G)), g defined in condition (iii) of Proposition 2.1 satisfies condition (ii) if and only if g satisfies (4) (resp. both (4) and (5)). Moreover, since the vertex \(v_{j'}\) in condition (iii) in Proposition 2.1 should be chosen from \(V_{1,1}\), it follows that g satisfies (5) if f satisfies (5). Thus \(v_j\) belongs to \(\eta _{D(G)}(f)\) if and only if \(v_j\) belongs to \(\eta _{D(G_1+K_{m_2})} (f)\).

Case 2  (\(j \ge m_1+2\)). Since \(v_{m_1+1}\) belongs to \(V_{1,2}\) with \(m_1+1 < j\), (4) holds for g defined in condition (iii) of Proposition 2.1 where \(v_{j'}=v_{m_1+1}\). Thus \(v_{j'}=v_{m_1+1}\) satisfies condition (iii) of Proposition 2.1 for \(v_j\) in \(D(G_1+K_{m_2})\). Hence \(v_j\) belongs to \(\eta _{D(G_1+K_{m_2})}(f)\). We now show that any \(v_j\), \(j \ge m_1+2\), belongs to \(\eta _{D(G)} (f)\), i.e., there exists \(j' < j\) such that \(g:V_1 \rightarrow {\mathbb Z}_{\ge 0}\) defined by

$$\begin{aligned} g(v_i) = {\left\{ \begin{array}{ll}f(v_i)+1 &{}\quad \text {if}\;i=j', \\ f(v_i)-1 &{}\quad \text {if}\;i=j,\\ f(v_i) &{} \quad \text {otherwise},\end{array}\right. } \end{aligned}$$
(7)

is a hypertree in D(G). Let \(\Gamma \) be a spanning tree of D(G) that induces f. Suppose that \(\Gamma \) does not contain an edge \(e=\{v_k, v_1'\}\) for some \(m_1< k \le m\). Then \(\Gamma \cup \{e\}\) has a unique cycle, and the cycle contains \(v_k\). Let \(e'\) be the edge of the cycle that is adjacent to \(v_k\) but different from e. Then the spanning tree \((\Gamma \cup \{e\}) \setminus \{e'\}\) induces f since the degree of each \(v_i \in V_1\) is same for \(\Gamma \) and \((\Gamma \cup \{e\}) \setminus \{e'\}\). Thus we may assume that \(\{ \{v_k, v_1'\} : m_1 <k\le m \} \subset \Gamma \). By the same argument, we may assume that \(\{ \{v_k, v_{m_1+1}'\} : 1\le k \le m_1\}\) is a subset of \(\Gamma \). Hence

$$\begin{aligned}S=\{ \{v_1, v_{m_1+1}'\} ,\ldots , \{v_{m_1}, v_{m_1+1}'\} ,\{v_{m_1+1},v_1'\}, \ldots , \{v_{m},v_1'\} \}\end{aligned}$$

is a subset of \(\Gamma \). Since \(\Gamma \) is spanning,

$$\begin{aligned}T=\{\{v_{i_2}, v_2'\} , \ldots , \{v_{i_{m_1}}, v_{m_1}'\},\{v_{i_{m_1+2}}, v_{m_1+2}'\} , \ldots , \{v_{i_{m}}, v_m'\}\}\end{aligned}$$

is a subset of \(\Gamma \) for some \(1 \le i_2, \ldots , i_{m_1}, i_{m_1+2}, \ldots ,i_m \le m\). Then \(S\cap T = \emptyset \) and \(|S \cup T| = 2m-2\). Since \(\Gamma \) has \(2m-1\) edges, we have

$$\begin{aligned}\Gamma = S \sqcup T \sqcup \{e\}\end{aligned}$$

for some edge e of D(G). In particular, the degree of each \(v_i'\), \(i \in [m] \setminus \{1,m_1+1\}\), in \(\Gamma \) is at most 2. Since \(f(v_j) >0\), the degree of \(v_j\) in \(\Gamma \) is at least 2, and hence \(\{v_j,v_k'\}\) is an edge of \(\Gamma \) for some \(k \in [m]\setminus \{1\}\). Let \(e'=\{v_j,v_k'\}\). We now construct a spanning tree of D(G) that induces a hypertree of the form (7). Note that, for an edge \(e''\) of D(G), \(\Gamma '=(\Gamma \cup \{e''\} ) \setminus \{e'\}\) is a spanning tree of D(G) if \(e'' \notin \Gamma \) and the (unique) cycle of \(\Gamma \cup \{e''\} \) contains \(e'\).

Case 2.A  (\(k =m_1+1\)). Then \(\Gamma \) contains a path \((v_1', v_j, v_{m_1+1}', v_1)\). Since \(\Gamma \) has no cycles, \(\{v_1,v_1'\}\) is not an edge of \(\Gamma \). Hence \((\Gamma \cup \{v_1, v_1'\} ) \setminus \{e'\}\) is a spanning tree of D(G) that induces a hypertree of the form (7) where \(j'=1\).

Case 2.B  (\(k \ne m_1+1\) and \(\deg (v_k') =1\)). If \(2 \le k \le m_1\), then \((\Gamma \cup \{v_k, v_k'\} ) \setminus \{e'\}\) is a spanning tree of D(G) that induces a hypertree of the form (7) where \(j'=k\). If \(m_1+2 \le k \le m\), then \((\Gamma \cup \{v_1, v_k'\} ) \setminus \{e'\}\) is a spanning tree of D(G) that induces a hypertree of the form (7) where \(j'=1\).

Case 2.C  (\(k \ne m_1+1\) and \(\deg (v_k') =2\)). Then \(\Gamma \) contains an edge \(\{v_\ell ,v_k'\}\) for some \(\ell \ne j\). If \(m_1 < \ell \le m\), then \(\Gamma \) contains a cycle \((v_1', v_j, v_k', v_\ell , v_1')\) of length 4. This is a contradiction. Hence \(\ell \le m_1\). Then \(\Gamma \) contains a path \((v_1', v_j, v_k', v_\ell ,v_{m_1+1}', v_1)\). Since \(\Gamma \) has no cycles, \(\{v_1,v_1'\}\) is not an edge of \(\Gamma \). Hence \((\Gamma \cup \{v_1, v_1'\} ) \setminus \{e'\}\) is a spanning tree of D(G) that induces a hypertree of the form (7) where \(j'=1\).

Thus we have \(k=\ell \), and hence

$$\begin{aligned} F_G(k) \subset F_{G_1+K_{m_2}}(k). \end{aligned}$$
(8)

Claim 3

\((F_{G_1+K_{m_2}}(k) \setminus F_G(k)) \subset F_{K_m}(k)\).

Suppose that f belongs to \(F_{G_1+K_{m_2}}(k) \setminus F_G(k)\). Then f does not satisfy (5), that is, there exists \(V' \subset V_{1,2}\) such that

$$\begin{aligned}\sum _{v \in V'} f(v) \ge |\Gamma _{D(G_2)}(V')| +m_1.\end{aligned}$$

In particular, \(\sum _{v \in V_{1,2}} f(v) > m_1\). Since \(\sum _{v \in V_1} f(v) \!=\! m \!-\!1\), \(\sum _{v \!\in V_{1,1}} f(v) \!<\! m_2 \!-\! 1\). Hence

$$\begin{aligned}\sum _{v \in V''} f(v)< m_2 - 1 < |\Gamma _{D(G_1)}(V'')| + m_2 - 1\end{aligned}$$

for all \(\emptyset \ne V'' \subset V_{1,1}\). Thus, for each \(v_j \ne v_1\) with \(f(v_j) >0\), \(g:V_1 \rightarrow {\mathbb Z}_{\ge 0}\) defined by

$$\begin{aligned}g(v_i) = {\left\{ \begin{array}{ll}f(v_i)+1 &{} \quad \text {if}\; i=1,\\ f(v_i)-1 &{} \quad \text {if}\;i=j,\\ f(v_i) &{} \quad \text {otherwise},\end{array}\right. }\end{aligned}$$

satisfies

$$\begin{aligned}\sum _{v \in V''} g(v) \le 1+\sum _{v \in V''} f(v) \le |\Gamma _{D(G_1)}(V'')| + m_2 - 1\end{aligned}$$

for all \(\emptyset \ne V'' \subset V_{1,1}\). Hence g satisfies (4) in the proof of Lemma 3.1. Therefore we have

$$\begin{aligned}k = |\{v_j \in V_1 : v_j \ne v_1, \,f(v_j) >0\}|,\end{aligned}$$

and hence

$$\begin{aligned} (F_{G_1+K_{m_2}}(k) \setminus F_G(k)) \subset F_{K_m}(k). \end{aligned}$$
(9)

Claim 4

\(F_{K_m}(k)=F_{K_{m_1}+G_2}(k) \cup (F_{G_1+K_{m_2}}(k) \setminus F_G(k))\).

From (6) and (9),

$$\begin{aligned}F_{K_m}(k)\supset F_{K_{m_1}+G_2}(k) \cup (F_{G_1+K_{m_2}}(k) \setminus F_G(k)).\end{aligned}$$

Let \(f \in F_{K_m}(k) \setminus F_{K_{m_1}+G_2}(k) \). Since \(F_{K_{m_1}+G_2}(k') \cap F_{K_m}(k) =\emptyset \) for any \(k'\ne k\), it follows that \(f \notin F_{K_{m_1}+G_2}(\ell ) \) for any \(\ell \). Hence from (3), \(f \in F_{G_1+K_{m_2}}(\ell )\) for some \(\ell \). If \(f \in F_G(\ell )\), then \(f \in F_{K_{m_1}+G_2}(\ell ') \) for some \(\ell '\) by (2). This is a contradiction. Thus \(f \in F_{G_1+K_{m_2}}(\ell ) \setminus F_G(\ell )\). From (9), we have

$$\begin{aligned}f \in F_{G_1+K_{m_2}}(\ell ) \setminus F_G(\ell )\subset F_{K_m}(\ell ).\end{aligned}$$

Then \(\ell =k\). It follows that

$$\begin{aligned}F_{K_m}(k)=F_{K_{m_1}+G_2}(k) \cup (F_{G_1+K_{m_2}}(k) \setminus F_G(k)).\end{aligned}$$

Finally, we show that this is a decomposition. Suppose that \(f \in F_{G_1+K_{m_2}}(k) \cap F_{K_{m_1}+G_2}(k)\). From (2), \(f \in F_G(\ell )\) for some \(\ell \). Moreover, from (8), we have \(\ell = k\). Thus

$$\begin{aligned} F_{G_1+K_{m_2}}(k) \cap F_{K_{m_1}+G_2}(k) \subset F_G(k).\end{aligned}$$

Therefore \(F_{K_m}(k)\) is decomposed into the disjoint sets

$$\begin{aligned}F_{K_m}(k)=F_{K_{m_1}+G_2}(k) \sqcup (F_{G_1+K_{m_2}}(k) \setminus F_G(k))\end{aligned}$$

as in Fig. 1. \(\square \)

Now, we are in the position to give a proof of Theorem 1.1.

Proof of Theorem 1.1

We prove this by induction on s. First we discuss the case when \(s=2\), i.e., \(G= G_1+G_2\). It is known [12, Exam. 5.3] that

$$\begin{aligned}|F_{K_m}(k)| = \left( {\begin{array}{c}m - 1\\ k\end{array}}\right) ^{\!2}.\end{aligned}$$

From Lemma 3.2,

$$\begin{aligned}\left( {\begin{array}{c}m - 1\\ k\end{array}}\right) ^{\!2} = |F_{K_{m_1}+G_2}(k)| + (|F_{G_1+K_{m_2}}(k) | - |F_G(k)| ).\end{aligned}$$

Thus we have

$$\begin{aligned}I_{D(G)}(x)=I_{D(G_1+K_{m_2})}(x) + I_{D(G_2+K_{m_1})}(x) -\sum _{k=0}^{m -1}\left( {\begin{array}{c}m-1\\ k\end{array}}\right) ^{\!2}x^k.\end{aligned}$$

Let \(s>2\) and assume that the assertion holds for the join of at most \(s-1\) graphs. Since \(G = (G_1 + \cdots + G_{s-1}) + G_s\), we have

$$\begin{aligned} I_{D(G)}(x)&=I_{D((G_1 + \cdots + G_{s-1})+K_{m_s})}(x) + I_{D(G_s+K_{m - m_s})}(x) -\sum _{k=0}^{m -1}\left( {\begin{array}{c}m-1\\ k\end{array}}\right) ^2x^k\\&=\sum _{i=1}^{s-2} I_{D(G_i+K_{m-m_i})}(x)+I_{D((G_{s-1}+K_{m_s})+K_{m-m_{s-1}-m_s})}(x)\\&\qquad -(s-2)\sum _{k=0}^{m -1}\left( {\begin{array}{c}m-1\\ k\end{array}}\right) ^{\!2}x^k+ I_{D(G_s+K_{m - m_s})}(x) -\sum _{k=0}^{m -1}\left( {\begin{array}{c}m-1\\ k\end{array}}\right) ^2x^k\\&=\sum _{i=1}^s I_{D(G_i+K_{m-m_i})}(x) -(s-1)\sum _{k=0}^{m -1}\left( {\begin{array}{c}m-1\\ k\end{array}}\right) ^{\!2}x^k. \end{aligned}$$

From Proposition 2.4 and Lemma 2.5, this is equation (1). \(\square \)

4 Complete Multipartite Graphs

In this section, applying Theorem 1.1, we give explicit formulas for the \(h^*\)-polynomial and the normalized volume of the PQ-type adjacency polytope of a complete multipartite graph. Given positive integers \(\ell \) and m, let

$$\begin{aligned}f_{\ell ,m}(x) = \sum _{k=0}^{\ell + m-1}\sum _{\alpha =0}^{k}\left( {\begin{array}{c}\ell -1\\ k-\alpha \end{array}}\right) \left( {\begin{array}{c}m\\ \alpha \end{array}}\right) \sum _{\beta =\alpha }^{k}\left( {\begin{array}{c}\ell +\alpha -1\\ \beta \end{array}}\right) \left( {\begin{array}{c}m-\alpha \\ k-\beta \end{array}}\right) x^k.\end{aligned}$$

Since

$$\begin{aligned}&\sum _{k=0}^{\ell + m-1}\sum _{\alpha =0}^{k}\left( {\begin{array}{c}\ell -1\\ k-\alpha \end{array}}\right) \left( {\begin{array}{c}m\\ \alpha \end{array}}\right) \sum _{\beta =0}^{k}\left( {\begin{array}{c}\ell +\alpha -1\\ \beta \end{array}}\right) \left( {\begin{array}{c}m-\alpha \\ k-\beta \end{array}}\right) x^k\\&\quad =\sum _{k=0}^{\ell + m-1}\sum _{\alpha =0}^{k}\left( {\begin{array}{c}\ell -1\\ k-\alpha \end{array}}\right) \left( {\begin{array}{c}m\\ \alpha \end{array}}\right) \left( {\begin{array}{c}\ell +m-1\\ k\end{array}}\right) x^k\\&\quad =\sum _{k=0}^{\ell + m-1}\left( {\begin{array}{c}\ell +m-1\\ k\end{array}}\right) \sum _{\alpha =0}^{k}\left( {\begin{array}{c}\ell -1\\ k-\alpha \end{array}}\right) \left( {\begin{array}{c}m\\ \alpha \end{array}}\right) x^k=\sum _{k=0}^{\ell + m-1}\left( {\begin{array}{c}\ell +m-1\\ k\end{array}}\right) ^{\!2}x^k \end{aligned}$$

holds, we have

$$\begin{aligned}f_{\ell ,m}(x)= & {} \sum _{k=0}^{\ell +m -1}\left( {\begin{array}{c}\ell +m-1\\ k\end{array}}\right) ^{\!2} x^k-\sum _{k=1}^{m-1}\sum _{\alpha =1}^{k}\left( {\begin{array}{c}\ell -1\\ k-\alpha \end{array}}\right) \left( {\begin{array}{c}m\\ \alpha \end{array}}\right) \\&\quad \times \sum _{\beta =0}^{\alpha -1}\left( {\begin{array}{c}\ell +\alpha -1\\ \beta \end{array}}\right) \left( {\begin{array}{c}m-\alpha \\ k-\beta \end{array}}\right) x^k.\end{aligned}$$

The \(h^*\)-polynomial of \(\nabla _G^PQ \) for the graph \(G= K_\ell +E_m \) \((=K_{1,\ldots ,1,m})\) coincides with \(f_{\ell ,m} (x)\).

Theorem 4.1

Let \(G= K_\ell +E_m \). Then we have

$$\begin{aligned} h^*\bigl (\nabla _G^PQ ,x\bigr )&=f_{\ell ,m}(x),\\ {{\,\mathrm{Vol}\,}}\bigl (\nabla _G^PQ \bigr )&=f_{\ell ,m}(1)=\sum _{\alpha =0}^{m}\left( {\begin{array}{c}m\\ \alpha \end{array}}\right) \sum _{\beta =0}^{\ell -1}\left( {\begin{array}{c}\ell +\alpha -1\\ \beta \end{array}}\right) \left( {\begin{array}{c}\ell +m-\alpha -1\\ \beta \end{array}}\right) . \end{aligned}$$

Proof

Let \(G'=D(K_{\ell -1}+E_m)\). Since \(G =(K_{\ell -1}+E_m) +K_1\), from Proposition 2.6, we have

$$\begin{aligned}h^*\bigl (\nabla _G^PQ ,x\bigr )\,=\,p(G',x)\,=\!\sum _{k = 0}^{\ell + m-1}|{PM (G' ,k)}| x^k.\end{aligned}$$

Let \(n =\ell +m-1\) and let \([n] \cup [\overline{n}]\) be the vertex set of \(G'\). We decompose [n] into two disjoint sets \(V_{1,1} = [\ell -1]\) and \(V_{1,2} = [n]\setminus [\ell -1]\) where \(V_{1,1}\) (resp. \(V_{1,2}\)) corresponds to \(K_{\ell -1}\) (resp. \(E_m\)). Similarly, we decompose \([\overline{n}]\) into two disjoint sets \(V_{2,1}\) and \(V_{2,2}\). From Proposition 2.8, each \(|PM (G' ,k)|\) is the number of (0, 1)-vectors \((x_1,\ldots ,x_n, y_1,\ldots ,y_n) \in {\mathbb R}^{2n}\) such that

$$\begin{aligned} \sum _{i = 1}^n x_i&=\sum _{i = 1}^n y_i = k, \end{aligned}$$
(10)
$$\begin{aligned} \sum _{i \in S} x_i&\le \sum _{\overline{j} \in \Gamma _{G'}(S)} y_j \quad \text {for all}\quad S \subset [n]. \end{aligned}$$
(11)

If (10) holds and a subset \(S \subset [n]\) contains an element of \(V_{1,1}\), then we have \(\Gamma _{G'}(S)=[\overline{n}]\), and hence

$$\begin{aligned} \sum _{\overline{j} \in \Gamma _{G'}(S)} y_j=k=\sum _{i = 1}^n x_i \ge \sum _{i \in S} x_i.\end{aligned}$$

Thus (10) and (11) hold if and only if

$$\begin{aligned} \sum _{i = 1}^n x_i&=\sum _{i = 1}^n y_i = k, \end{aligned}$$
(12)
$$\begin{aligned} \sum _{i \in S} x_i&\le \sum _{\overline{j} \in \Gamma _{G'}(S)} y_j \quad \text {for all}\quad S \subset V_{1,2}. \end{aligned}$$
(13)

We count such vectors with \(|\{ i \in V_{1,2} : x_i = 1 \}|=\alpha \) for each \(\alpha = 0,1,\ldots , k\). Let \(S'=\{ i \in V_{1,2} : x_i = 1 \} \subset V_{1,2}\).

Case 1  (\(\alpha >0\)). There are \(\left( {\begin{array}{c}m\\ \alpha \end{array}}\right) \) possibilities for the choice of the subset \(S'\). For each \(S'\), there are \(\left( {\begin{array}{c}\ell -1\\ k-\alpha \end{array}}\right) \) possibilities for the choice of the subset \(\{ i \in V_{1,1} : x_i = 1 \} \subset V_{1,1}\). Then (12) and (13) hold if and only if \(\sum _{i = 1}^n y_i = k\) and \(\alpha \le \sum _{\overline{j} \in \Gamma _{G'}(S')} y_j\). Let \(\beta =\sum _{\overline{j} \in \Gamma _{G'}(S')} y_j\). Since \(|\Gamma _{G'}(S')| = \ell + \alpha -1\), there are \(\left( {\begin{array}{c}\ell +\alpha -1\\ \beta \end{array}}\right) \) possibilities for the choice of the subset \(S''=\{ \overline{j} \in \Gamma _{G'}(S') : y_j = 1 \} \subset \Gamma _{G'}(S')\). For each \(S''\), there are \(\left( {\begin{array}{c}m-\alpha \\ k-\beta \end{array}}\right) \) possibilities for the choice of the subset \(\{ \overline{j} \in [\overline{n}] \setminus \Gamma _{G'}(S') : y_j = 1 \}\subset [\overline{n}] \setminus \Gamma _{G'}(S')\).

Case 2  (\(\alpha = 0\)). There are \(\left( {\begin{array}{c}\ell -1\\ k\end{array}}\right) \) possibilities for the choice of the subset \(\{ i \in V_{1,1} : x_i = 1 \} \subset V_{1,1}\). Let \(\beta = \sum _{\overline{j} \in V_{2,1}} y_j\). Since \(S'=\emptyset \), condition (13) always holds. Hence there are \(\left( {\begin{array}{c}\ell -1\\ \beta \end{array}}\right) \) possibilities for the choice of the subset \(\{\overline{j} \in V_{2,1} : y_j = 1 \}\), and there are \(\left( {\begin{array}{c}m\\ k-\beta \end{array}}\right) \) possibilities for the choice of the subset \(\{\overline{j} \in V_{2,2} : y_j = 1 \}\).

Thus we have \(I_{D(G)}(x)=f_{\ell ,m}(x)\). Moreover, the normalized volume of \(\nabla _G^PQ \) is equal to

$$\begin{aligned} f_{\ell ,m}(1)&=\sum _{k=0}^{\ell + m-1}\sum _{\alpha =0}^{k}\left( {\begin{array}{c}\ell -1\\ k-\alpha \end{array}}\right) \left( {\begin{array}{c}m\\ \alpha \end{array}}\right) \sum _{\beta =\alpha }^{k}\left( {\begin{array}{c}\ell +\alpha -1\\ \beta \end{array}}\right) \left( {\begin{array}{c}m-\alpha \\ k-\beta \end{array}}\right) \\&=\sum _{\alpha =0}^{m}\left( {\begin{array}{c}m\\ \alpha \end{array}}\right) \sum _{\beta =\alpha }^{\ell +\alpha -1}\left( {\begin{array}{c}\ell +\alpha -1\\ \beta \end{array}}\right) \sum _{k=\beta }^{\ell + \alpha -1}\left( {\begin{array}{c}\ell -1\\ k-\alpha \end{array}}\right) \left( {\begin{array}{c}m-\alpha \\ k-\beta \end{array}}\right) \\&=\sum _{\alpha =0}^{m}\left( {\begin{array}{c}m\\ \alpha \end{array}}\right) \sum _{\beta =\alpha }^{\ell +\alpha -1}\left( {\begin{array}{c}\ell +\alpha -1\\ \beta \end{array}}\right) \sum _{k=\beta }^{\ell + \alpha -1}\left( {\begin{array}{c}\ell -1\\ \ell -1+\alpha -\beta -(k-\beta )\end{array}}\right) \left( {\begin{array}{c}m-\alpha \\ k-\beta \end{array}}\right) \\&=\sum _{\alpha =0}^{m}\left( {\begin{array}{c}m\\ \alpha \end{array}}\right) \sum _{\beta =\alpha }^{\ell +\alpha -1}\left( {\begin{array}{c}\ell +\alpha -1\\ \beta \end{array}}\right) \left( {\begin{array}{c}\ell +m-\alpha -1\\ \ell +\alpha -\beta -1\end{array}}\right) \\&=\sum _{\alpha =0}^{m}\left( {\begin{array}{c}m\\ \alpha \end{array}}\right) \sum _{\beta '=0}^{\ell -1}\left( {\begin{array}{c}\ell +\alpha -1\\ \beta '\end{array}}\right) \left( {\begin{array}{c}\ell +m-\alpha -1\\ \beta '\end{array}}\right) . \end{aligned}$$

\(\square \)

Remark 4.2

Let \(G= K_\ell +E_m \). Since

$$\begin{aligned}\sum _{\alpha =0}^{m}\left( {\begin{array}{c}m\\ \alpha \end{array}}\right) \sum _{\beta =0}^{\ell +\alpha -1}\left( {\begin{array}{c}\ell +\alpha -1\\ \beta \end{array}}\right) \left( {\begin{array}{c}\ell +m-\alpha -1\\ \beta \end{array}}\right) =\left( {\begin{array}{c}2(\ell +m-1)\\ \ell +m-1\end{array}}\right) ,\end{aligned}$$

we have

$$\begin{aligned}{{\,\mathrm{Vol}\,}}\bigl (\nabla _G^PQ \bigr ) =\left( {\begin{array}{c}2(\ell +m-1)\\ \ell +m-1\end{array}}\right) -\sum _{\alpha =1}^{m-1}\left( {\begin{array}{c}m\\ \alpha \end{array}}\right) \sum _{\beta =\ell }^{\ell +\alpha -1}\left( {\begin{array}{c}\ell +\alpha -1\\ \beta \end{array}}\right) \left( {\begin{array}{c}\ell +m-\alpha -1\\ \beta \end{array}}\right) .\end{aligned}$$

Example 4.3

For small \(\ell \) and m, \(f_{\ell , m}(1)\) in Theorem 4.1 is

$$\begin{aligned} f_{1,m}(1)&=2^m,\\ f_{2,m}(1)&= 2^{m-2}(m^2+3 m+8),\\ f_{3,m}(1)&=\frac{2^{m-4}}{(2!)^2}(m^4+10 m^3+59 m^2+186 m+384),\\ f_{4,m}(1)&=\frac{2^{m-6}}{(3!)^2}(m^6+21 m^5+229 m^4+1563 m^3+7762 m^2+24{,}984 m+46{,}080),\\ f_{\ell ,1}(1)&=\left( {\begin{array}{c}2\ell \\ \ell \end{array}}\right) ,\\ f_{\ell ,2}(1)&=\left( {\begin{array}{c}2(\ell +1)\\ \ell +1\end{array}}\right) -2,\\ f_{\ell ,3}(1)&=\left( {\begin{array}{c}2(\ell +2)\\ \ell +2\end{array}}\right) -(6\ell +6),\\ f_{\ell ,4}(1)&=\left( {\begin{array}{c}2(\ell +3)\\ \ell +3\end{array}}\right) -(10\ell ^2+24\ell +20). \end{aligned}$$

From Theorems 1.1 and 4.1 we have the following.

Corollary 4.4

Let G be a complete multipartite graph \(K_{m_1,\ldots ,m_s}\), and let \(m= \sum _{i=1}^s m_i\). Then we have

$$\begin{aligned} h^*\bigl (\nabla _G^PQ ,x\bigr )&=\sum _{i=1}^s f_{m-m_i,m_i}(x)- (s-1)\sum _{k=0}^{m -1}\left( {\begin{array}{c}m-1\\ k\end{array}}\right) ^{\!2}x^k,\\ {{\,\mathrm{Vol}\,}}\bigl (\nabla _G^PQ \bigr )&=\sum _{i=1}^s f_{m-m_i,m_i}(1) - (s-1)\left( {\begin{array}{c}2(m-1)\\ m-1\end{array}}\right) . \end{aligned}$$

Example 4.5

From Corollary 4.4, the normalized volume of \(\nabla _{K_{\ell ,m}}^PQ \) is

$$\begin{aligned}f_{\ell ,m}(1)+f_{m,\ell }(1) -\left( {\begin{array}{c}2(\ell +m-1)\\ \ell + m-1\end{array}}\right) .\end{aligned}$$

From Example 4.3 we have

$$\begin{aligned} {{\,\mathrm{Vol}\,}}\bigl (\nabla _{K_{1,m}}^PQ \bigr )&=2^m,\\ {{\,\mathrm{Vol}\,}}\bigl (\nabla _{K_{2,m}}^PQ \bigr )&=2^{m-2}(m^2+3 m+8) -2,\\ {{\,\mathrm{Vol}\,}}\bigl (\nabla _{K_{3,m}}^PQ \bigr )&=\frac{2^{m-4}}{(2!)^2}(m^4+10 m^3+59 m^2+186 m+384)-(6m+6),\\ {{\,\mathrm{Vol}\,}}\bigl (\nabla _{K_{4,m}}^PQ \bigr )&=\frac{2^{m-6}}{(3!)^2}(m^6+21 m^5+229 m^4+1563 m^3+7762 m^2+24984 m+46080)\\&\qquad \qquad \qquad -(10 m^2+24m +20). \end{aligned}$$

The formula for \({{\,\mathrm{Vol}\,}}\bigl (\nabla _{K_{2,m}}^PQ \bigr )\) coincides with that in [8, Prop. 4.2].

Example 4.6

Let G be the complete bipartite graph \(K_{2,n-2}\). Since

$$\begin{aligned} f_{2,m}(x)&=\sum _{k=0}^{m+1}\sum _{\alpha =0}^{k}\left( {\begin{array}{c}1\\ k-\alpha \end{array}}\right) \left( {\begin{array}{c}m\\ \alpha \end{array}}\right) \sum _{\beta =\alpha }^{k}\left( {\begin{array}{c}\alpha +1\\ \beta \end{array}}\right) \left( {\begin{array}{c}m-\alpha \\ k-\beta \end{array}}\right) x^k\\&=\sum _{k=0}^{m+1}\left( {\begin{array}{c}m\\ k-1\end{array}}\right) k(m-k+1) x^k+\sum _{k=0}^{m+1}\left( {\begin{array}{c}m\\ k-1\end{array}}\right) x^k+\sum _{k=0}^{m+1}\left( {\begin{array}{c}m\\ k\end{array}}\right) (k+1) x^k\\&=\sum _{k=0}^{m}\left( {\begin{array}{c}m\\ k\end{array}}\right) k^2x^k+\sum _{k=0}^{m}\left( {\begin{array}{c}m\\ k\end{array}}\right) x^{k+1}+\sum _{k=0}^{m}\left( {\begin{array}{c}m\\ k\end{array}}\right) (k+1) x^k\\&=\sum _{k=0}^{m}\left( {\begin{array}{c}m\\ k\end{array}}\right) (x+1+k^2+k) x^k\\&=(x+1)^{m+1} + m( (m + 1) x +2) x(x+1)^{m-2} \end{aligned}$$

and

$$\begin{aligned}f_{\ell ,2}(x)=\sum _{k=0}^{\ell +1}\left( {\begin{array}{c}\ell +1\\ k\end{array}}\right) ^{\!2}x^k-2x,\end{aligned}$$

we have

$$\begin{aligned} h^*\bigl (\nabla _G^PQ ,x\bigr )&=f_{2,n-2}(x) + f_{n-2,2}(x) - \sum _{k=0}^{n -1}\left( {\begin{array}{c}n-1\\ k\end{array}}\right) ^{\!2}x^k\\&=(x+1)^{n-1} + (n-2)( (n- 1) x +2) x (x+1)^{n-4} -2x. \end{aligned}$$

5 Wheel Graphs

For \(n\ge 3\), the wheel graph \(W_n\) with \(n+1\) vertices is the join graph \(W_n = C_n + K_1\). Unfortunately, Theorem 1.1 is not useful for computing the \(h^*\)-polynomial of \(\nabla _{W_n}^PQ \). We will give an explicit formula for the \(h^*\)-polynomial of \(\nabla _{W_n}^PQ \) and prove the conjecture [8, Conj. 4.4] on the normalized volume of \(\nabla _{W_n}^PQ \) by using Proposition 2.6 on \(\nabla _{G+K_1}^PQ \). Let

$$\begin{aligned}\gamma (n,x)={\left\{ \begin{array}{ll} 1 &{} \quad \text {if}\;n=0,\\ \displaystyle \frac{(1+\sqrt{1+8x})^n + (1-\sqrt{1+8x})^n}{2^n} &{}\quad \text {if}\;n\text { is odd},\\ \displaystyle \frac{(1+\sqrt{1+8x})^n + (1-\sqrt{1+8x})^n}{2^n} -2 x^{n/2}&{}\quad \text {otherwise}. \end{array}\right. }\end{aligned}$$

For \(n\ge 3\),

$$\begin{aligned}\frac{(1+\sqrt{1+8x})^n + (1-\sqrt{1+8x})^n}{2^n} = g(C_n, 2x),\end{aligned}$$

where \(g(C_n,x)\) is the matching generating polynomial of \(C_n\) (\(=\) the “independence polynomial” of \(C_n\)). See, e.g., [18, p. 27]. Moreover, for \(n\ge 3\), it is known [16, Exam. 4.5] that, \(\gamma (n,x)\) is the \(\gamma \)-polynomial of the PV-type adjacency polytope \(\nabla ^PV _{W_n}\) of \(W_n\). (Note that \(\nabla ^\mathrm{PV}_{W_n}\) is called the symmetric edge polytope of type A of \(W_n\) in [16].) The \(h^*\)-polynomial of \(\nabla _{W_n}^PQ \) is described by this function as follows.

Theorem 5.1

The \(h^*\)-polynomial of \(\nabla _{W_n}^PQ \) is

$$\begin{aligned} \sum _{k=0}^n \left( {\begin{array}{c}n\\ k\end{array}}\right) \gamma (n-k,x) x^k&=\biggl (\frac{1+2x+\sqrt{1+8x}}{2}\biggr )^{\!n}+\biggl (\frac{1+2x-\sqrt{1+8x}}{2}\biggr )^{\!n} \\&\quad +x^n-(x+\sqrt{x})^n-(x-\sqrt{x})^n. \end{aligned}$$

Moreover, the normalized volume of \(\nabla _{W_n}^PQ \) is \(3^n-2^n+1\).

Proof

Since \(W_n = C_n + K_1\), the \(h^*\)-polynomial of \(\nabla _{W_n}^PQ \) is

$$\begin{aligned}p(D(C_n),x) =\sum _{\ell = 0}^n|PM (D(C_n),\ell )| x^\ell \end{aligned}$$

by Proposition 2.6. From Proposition 2.8, \(|PM (D(C_n),\ell )|\) is equal to the number of (0, 1)-vectors \((x_1,\ldots ,x_n, y_1,\ldots ,y_n) \in {\mathbb R}^{2n}\) such that

$$\begin{aligned} \sum _{i =1}^{n} x_i&=\sum _{j=1}^{n} y_j = \ell , \end{aligned}$$
(14)
$$\begin{aligned} \sum _{i \in S} x_i&\,\le \!\sum _{\overline{j} \in \Gamma _{D(C_n)}(S)} \!\!y_j \quad \text{ for } \text{ all }\quad S \subset [n], \end{aligned}$$
(15)

where \(V= [n] \cup [\overline{n}]\) is the set of vertices of \(D(C_n)\). Let \(C_n=(1,2,\ldots ,n,1)\). Given a subset \(T \subset [n]\) and an integer \(\ell \in [n]\), let \(PM _{T,\ell }\) denote the set of all (0, 1) vectors \((x_1,\ldots ,x_n,y_1,\ldots ,y_n) \in {\mathbb R}^{2n}\) satisfying (14), (15), and \(T = \{ i \in [n] : x_i = y_i =1 \}\). Note that \(PM _{T,\ell } = \emptyset \) if \(\ell < |T|\).

Let \(T \subset [n]\) with \(|T| =k\). We will show that \(\sum _{\ell =k}^n |PM _{T,\ell }| x^\ell = \gamma (n-k,x) x^k\).

Case 1  (\(k=n\)). It is easy to see that \(PM _{T,n}= \{(1,\ldots ,1)\}\) and \(PM _{T,\ell }=\emptyset \) if \(\ell \ne n\). Note that \( \gamma (n-k,x) x^k =x^n\) if \(k=n\). Thus \(\sum _{\ell =k}^n |PM _{T,\ell }| x^\ell = \gamma (n-k,x) x^k\).

Case 2  (\(k=n-1\)). Let \(T=[n] \setminus \{i\}\) where \(i \in [n]\). It then follows that \(PM _{T,n-1}=\{ (1,\ldots ,1) - \mathbf{e}_i - \mathbf{e}_{n+i}\}\) and \(PM _{T,\ell }=\emptyset \) if \(\ell \ne n-1\). Note that \(\gamma (n-k,x) x^k =x^{n-1} \) if \(k=n-1\). Thus \(\sum _{\ell =k}^n |PM _{T,\ell }| x^\ell = \gamma (n-k,x) x^k\).

Case 3  (\(k=n-2\)). Let \(T=[n] \setminus \{i,j\}\) where \(1 \le i<j\le n\). Since (14) holds, each element of \(PM _{T,\ell }\) is

$$\begin{aligned}\alpha _1 = (1,\ldots ,1) - \mathbf{e}_i - \mathbf{e}_j - \mathbf{e}_{n+i}- \mathbf{e}_{n+j}\end{aligned}$$

if \(\ell =n-2\), and is one of

$$\begin{aligned}\alpha _2= (1,\ldots ,1) - \mathbf{e}_i - \mathbf{e}_{n+j},\qquad \alpha _3=(1,\ldots ,1) - \mathbf{e}_j - \mathbf{e}_{n+i},\end{aligned}$$

if \(\ell = n-1\). Then each \(\alpha _i\) corresponds to a perfectly matchable set. In fact, a matching of \(D(C_n)\) which corresponds to \(\alpha _1,\alpha _2,\alpha _3\) is

$$\begin{aligned}&\{ \{s, \overline{s}\} : s \in [n] \setminus \{i,j\} \},\\&\{ \{s, \overline{s}\} : 1 \le s \le i-1\; \text {or}\; j+1 \le s \le n \}\cup \{ \{ s+1 , \overline{s}\} : i \le s \le j-1 \},\\&\{ \{s, \overline{s}\} : 1 \le s \le i-1 \;\text {or}\; j+1 \le s \le n \}\cup \{ \{ s , \overline{s+1}\} : i \le s \le j-1 \}, \end{aligned}$$

respectively. Thus

$$\begin{aligned}PM _{T,\ell }={\left\{ \begin{array}{ll}\{\alpha _1\} &{}\quad \text {if}\;\ell =n-2,\\ \{\alpha _2, \alpha _3\} &{} \quad \text {if}\;\ell =n-1,\\ \emptyset &{} \quad \text {otherwise}.\end{array}\right. }\end{aligned}$$

Note that \(\gamma (n-k,x) x^k = (2x+1) x^{n-2} =x^{n-2} + 2 x^{n-1}\) if \(k=n-2\). Thus

$$\begin{aligned}\sum _{\ell =k}^n |PM _{T,\ell }| x^\ell = \gamma (n-k,x) x^k.\end{aligned}$$

Case 4  (\(k\le n-3\)). Suppose that \(T'=\{p,p+1,\ldots ,q\} \subset T\) and \(p-1,q+1\notin T\). Since \(|T|\le n-3\), we have \(n-q+p-1\ge 3\). Let \(U=\{ p,p+1,\ldots ,q, \overline{p}, \overline{p+1},\ldots ,\overline{q}\}\). We will show that there exists a perfectly matchable set S of \(D(C_{n-q+p-1})\) such that

$$\begin{aligned}\rho (S) =(x_1, \ldots , x_{p-1},x_{q+1},\ldots ,x_n, y_1, \ldots , y_{p-1},y_{q+1},\ldots ,y_n)\end{aligned}$$

for any \((x_1,\ldots ,x_n,y_1,\ldots ,y_n) \in PM _{T,\ell }\). Here \(C_{n-q+p-1}=(1,\ldots ,p-1,q+1,\ldots ,n,1)\) and the vertex set of \(D(C_{n-q+p-1})\) is \(\{1,\ldots ,p-1,q+1,\ldots ,n\} \cup \{\overline{1},\ldots ,\overline{p-1},\overline{q+1},\ldots ,\overline{n}\}\). Let M be a matching of \(D(C_n)\) which corresponds to \((x_1,\ldots ,x_n,y_1,\ldots ,y_n)\).

Case 4.1 (either \(x_{p-1}=x_{q+1}=0\) or \(y_{p-1}=y_{q+1}=0\)). Exchanging [n] and \([\overline{n}]\) if needed, we may assume that \(y_{p-1}=y_{q+1}=0\). Then, for the subset \(T' \subset [n]\),

$$\begin{aligned}\sum _{i \in T'} x_i \,=\! \sum _{\overline{j} \in \Gamma _{G'}(T')}\! y_j = |T'|.\end{aligned}$$

Hence the matching M is the union of a perfect matching of the induced subgraph \(D(C_n)_U\) of \(D(C_n)\) and a matching \(M'\) of the induced subgraph \(D(C_n)_{V\setminus U}\) of \(D(C_n)\). Then one can regard \(M'\) as a matching of \(D(C_{n-q+p-1})\) since \(D(C_n)_{V\setminus U}\) is a subgraph of \(D(C_{n-q+p-1})\).

Case 4.2  (\(x_{p-1}=y_{q+1} \ne y_{p-1}=x_{q+1}\)). Exchanging [n] and \([\overline{n}]\) if needed, we may assume that \(x_{p-1}=y_{q+1} =1\) and \(y_{p-1}=x_{q+1}=0\). If \(e = \{q+2, \overline{q+1} \}\) belongs to the matching M, then \(M \setminus \{e\}\) is the union of a perfect matching of \(D(C_n)_U\) and a matching \(M'\) of \(D(C_n)_{V\setminus U}\) by the same argument in Case 4.1. Suppose that \(e = \{q, \overline{q+1} \}\) belongs to the matching M. It then follows that M is the union of \(\{ \{p-1, \overline{p} \} , \ldots , \{q-1, \overline{q} \} , \{q, \overline{q+1}\}\}\) and a matching \(M'\) of \(D(C_n)_{V\setminus U}\). Thus one can regard \(M' \cup \{\{p-1, \overline{q+1} \} \}\) as a matching of \(D(C_{n-q+p-1})\).

Thus there exists a perfectly matchable set \(S_1\) of \(D(C_{n-q+p-1})\) such that

$$\begin{aligned}\rho (S_1) =(x_1, \ldots , x_{p-1},x_{q+1},\ldots ,x_n, y_1, \ldots , y_{p-1},y_{q+1},\ldots ,y_n)\end{aligned}$$

for \((x_1,\ldots ,x_n,y_1,\ldots ,y_n)\) if \(T'=\{p,p+1,\ldots ,q\} \subset T\) and \(p-1 , q+1 \notin T\). If, in addition, \(T''=\{p',p'+1,\ldots ,q'\} \subset T\) and \(p'-1 , q'+1 \notin T\) for some \(p' > q+1\), then there exists a perfectly matchable set \(S_2\) of \(D(C_{n-(q-p+1) -(q'-p'+1)})\) such that

$$\begin{aligned}\rho (S_2)= & {} (x_1, \ldots , x_{p-1},x_{q+1},\ldots , x_{p'-1},x_{q'+1},\ldots , x_n, \\&y_1, \ldots , y_{p-1},y_{q+1},\ldots , y_{p'-1},y_{q'+1},\ldots ,y_n) \end{aligned}$$

by the same argument as above. Repeating the above argument, it follows that there exists a perfectly matchable set S of \(D(C_{n-k})\) such that \(\rho (S) =(x_{i_1}, \ldots , x_{i_{n-k}}, y_{i_1}, \ldots , y_{i_{n-k}})\) where \([n] \setminus T = \{i_1,\ldots ,i_{n-k}\}\) and \(x_{i_r} + y_{i_r} \le 1\) for all r. Then there exists a perfectly matchable set \(S'\) of \(C_{n-k}\) such that \(|S'| = 2(\ell -k)\) and \(\rho (S') = (x_{i_1} +y_{i_1}, \ldots , x_{i_{n-k}} + y_{i_{n-k}})\). The matching corresponding to \(S'\) is not unique if and only if \(n-k\) is even and \(\rho (S')=(1,\ldots ,1)\). There exist two matchings for the perfectly matchable set S of \(D(C_{n-k})\) exactly when

$$\begin{aligned}(x_{i_1}, \ldots , x_{i_{n-k}}, y_{i_1}, \ldots , y_{i_{n-k}})= & {} (1,0,\ldots ,1,0, 0,1,\ldots , 0,1), \\&(0,1,\ldots ,0,1,1,0, \ldots , 1,0).\end{aligned}$$

Conversely, for each matching M of \(C_{n-k}\), there exist \(2^{|M|}\) vectors \((x_{i_1}, \ldots , x_{i_{n-k}}, y_{i_1}, \ldots , y_{i_{n-k}})\) where \(x_{i_r} + y_{i_r} \le 1\) for all r and associated with a perfectly matchable set of \(D(C_{n-k})\) since there are two possibilities \((x_{i_p}, x_{i_{p+1}},y_{i_p}, y_{i_{p+1}}) =(1,0,0,1), (0,1,1,0)\) for each \(\{i_p, i_{p+1}\} \in M\). Hence we have

$$\begin{aligned}| \mathrm{PM}_{T,\ell } | ={\left\{ \begin{array}{ll} 2^{\ell -k} m_{C_{n-k}} (\ell -k) -2 &{}\quad \text {if}\;\ell -k = (n-k)/2,\\ 2^{\ell -k} m_{C_{n-k}} (\ell -k)&{} \quad \text {otherwise}. \end{array}\right. }\end{aligned}$$

Recall that \(m_{G} (k)\) is the number of k-matchings of a graph G. Thus

$$\begin{aligned} \sum _{\ell =k}^n |PM _{T,\ell }| x^\ell&=\sum _{\ell =k}^n2^{\ell -k} m_{C_{n-k}} (\ell -k) x^\ell -2 x^{(n+k)/2} \\&=\left( \sum _{\ell =k}^n m_{C_{n-k}} (\ell -k)(2 x)^{\ell -k} -2 x^{(n-k)/2} \right) x^k=\gamma (n-k,x) x^k \end{aligned}$$

if \(n-k\) is even, and

$$\begin{aligned} \sum _{\ell =k}^n |PM _{T,\ell }| x^\ell&=\sum _{\ell =k}^n2^{\ell -k} m_{C_{n-k}} (\ell -k) x^\ell \\&=\left( \sum _{\ell =k}^nm_{C_{n-k}} (\ell -k)(2 x)^{\ell -k}\right) x^k=\gamma (n-k,x) x^k \end{aligned}$$

if \(n-k\) is odd. Therefore the \(h^*\)-polynomial of \(\nabla _{W_n}^PQ \) is

$$\begin{aligned}\sum _{T \subset [n]}\sum _{\ell =|T|}^n |PM _{T,\ell }| x^\ell =\sum _{T \subset [n]}\gamma (n-|T|) x^{|T|}=\sum _{k=0}^n \left( {\begin{array}{c}n\\ k\end{array}}\right) \gamma (n-k,x) x^k.\end{aligned}$$

Moreover,

$$\begin{aligned}&\sum _{k=0}^n \left( {\begin{array}{c}n\\ k\end{array}}\right) \gamma (n-k,x) x^k\\&\quad =x^n+\sum _{k=0}^n \left( {\begin{array}{c}n\\ k\end{array}}\right) \left( \frac{1+\sqrt{1+8x}}{2}\right) ^{n-k} x^k+\sum _{k=0}^n \left( {\begin{array}{c}n\\ k\end{array}}\right) \biggl (\frac{1-\sqrt{1+8x}}{2}\biggr )^{\!n-k}x^k\\&\qquad - \sum _{\begin{array}{c} k\ge 0\\ k \equiv n \; (\mathrm{mod}\;2) \end{array}}\!\!\! 2 \left( {\begin{array}{c}n\\ k\end{array}}\right) \sqrt{x}^{n-k}\,x^k\\&\quad =\biggl (\frac{1+2x+\sqrt{1+8x}}{2}\biggr )^{\!n} \!+\!\biggl (\frac{1+2x-\sqrt{1+8x}}{2}\biggr )^{\!n}\!+\!x^n-(x\!+\!\sqrt{x})^n\!-\! (x-\sqrt{x})^n. \end{aligned}$$

In particular, substituting \(x=1\), we have

$$\begin{aligned}\sum _{k=0}^n \left( {\begin{array}{c}n\\ k\end{array}}\right) \gamma (n-k,1)=3^n-2^n+1. \end{aligned}$$

\(\square \)