1 Introduction

Let \(Q_n\) denote the \(n\)-dimensional hypercube with vertex set \(V_n=\{0,1\}^n\). A \(0/1\)-polytope of \(Q_n\) is defined to be the convex hull of a subset of \(V_n\). The study of \(0/1\)-polytopes has received much attention, see, for example [6, 7, 1113, 15, 18, 19].

In this paper, we are concerned with the problem of determining the number of equivalence classes of \(n\)-dimensional \(0/1\)-polytopes of \(Q_n\) under the symmetries of \(Q_n\), which has been considered as a difficult problem, see Ziegler [18]. It is also listed by Zong [19, Problem 5.1] as one of the fundamental problems concerning \(0/1\)-polytopes.

An \(n\)-dimensional \(0/1\)-polytope of \(Q_n\) is also called a full-dimensional \(0/1\)-polytope of \(Q_n\). Two \(0/1\)-polytopes are said to be equivalent if one can be transformed to the other by a symmetry of \(Q_n\). Such an equivalence relation is called the \(0/1\)-equivalence relation. For example, Fig. 1 gives the representatives of \(0/1\)-equivalence classes of \(Q_2\), among which (d) and (e) are full-dimensional.

Fig. 1
figure 1

Representatives of 0/1-equivalence classes of \(Q_2\)

As the first nontrivial case, full-dimensional 0/1-equivalence classes of \(Q_4\) were counted by Below, see Ziegler [18]. With the aid of a computer program, Aichholzer [1] completed the enumeration of full-dimensional \(0/1\)-equivalence classes of \(Q_5\), and those of \(Q_6\) up to \(12\) vertices, see also Aichholzer [3] and Ziegler [18]. The 5-dimensional hypercube \(Q_5\) has been considered as the last case that one can hope for a complete solution to the enumeration of full-dimensional \(0/1\)-equivalence classes.

Let \(F_n(k)\) denote the number of full-dimensional \(0/1\)-equivalence classes of \(Q_n\). The objective of this paper is to present a method to compute \(F_n(k)\) for \(k>2^{n-2}\). We also find a way to compute \(F_n(k)\) when \(k\) is close to \(2^{n-2}\). Using our approach, we can determine \(F_6(k)\) for \(k> 12\). Combining the computation of Aichholzer [1], we reach a complete solution for the case of \(Q_6\).

To describe our approach, let \(A_n(k)\) denote the number of \(0/1\)-equivalence classes of \(Q_n\) with \(k\) vertices, and let \(H_n(k)\) denote the number of \(0/1\)-equivalence classes of \(Q_n\) with \(k\) vertices that are not full-dimensional. So we have

$$\begin{aligned} A_n(k)=F_n(k)+H_n(k). \end{aligned}$$
(1.1)

It is clear that \(F_n(k)=0\) for \(1\le k\le n\) since any full-dimensional \(0/1\)-polytope of \(Q_n\) has at least \(n+1\) vertices. As will be seen in Sect. 2, the values \(A_n(k)\) for any \(k\) can be computed from the cycle index of the hyperoctahedral group \(B_n\). Hence \(F_n(k)\) can be determined by \(H_n(k)\).

To compute \(H_n(k)\), we need a relation between the dimension of a \(0/1\)-polytope and the number of vertices. Let \(P\) be a \(0/1\)-polytope of \(Q_n\), and let \({\mathrm {dim}}(P)\) denote the dimension of \(P\). It is known that \(P\) is affinely equivalent to a full-dimensional \(0/1\)-polytope of \(Q_{d}\) for some \(d\le n\), see Ziegler [18]. Thus we have the following consequence.

Theorem 1.1

Let \(P\) be a \(0/1\)-polytope of \(Q_n\) with more than \(2^{m}\) vertices, where \(1\le m<n\). Then we have

$$\begin{aligned} \mathrm {dim}(P)\ge m+1. \end{aligned}$$

From Theorem 1.1, we see that if a \(0/1\)-polytope \(P\) of \(Q_n\) has more than \(2^{n-1}\) vertices, then \(P\) has dimension \(n\). Thus, for \(k>2^{n-1}\), we have \(F_n(k)=A_n(k)\).

Based on Theorem 1.1, we show that the computation of \(H_n(k)\) for \(2^{n-2}<k\le 2^{n-1}\) can be carried out by determining the number of equivalence classes of \(0/1\)-polytopes with \(k\) vertices that are contained in every hyperplane spanned by vertices of \(Q_n\). When \(2^{n-2}<k\le 2^{n-1}\), we can apply Pólya’s theorem to count equivalence classes of \(0/1\)-polytopes with \(k\) vertices that are contained in a hyperplane spanned by vertices of \(Q_n\). In particular, when \(n=6\), we obtain \(F_6(k)\) for \(16<k\le 32\).

We also find a way to compute \(H_n(k)\) when \(k\) is close to \(2^{n-2}\). In particular, when \(n=6\), we obtain \(F_6(k)\) for \(13\le k \le 16\).

This paper is organized as follows. In Sect. 2, we recall a method introduced by Chen [9] to determine the cycle structure of a symmetry \(w\) in the hyperoctahedral group \(B_n\) in terms of the number of vertices of \(Q_n\) fixed by \(w\). Sections 36 are devoted to the computation of \(H_n(k)\) for \(2^{n-2}<k\le 2^{n-1}\). In Sect. 7, we provide a way to compute \(H_n(k)\) when \(k\) is close to \(2^{n-2}\). This enables us to determine \(H_n(k)\) for \(n=6\) and \(13\le k\le 16\).

2 The Cycle Index of the Hyperoctahedral Group

The group of symmetries of \(Q_n\) is known as the hyperoctahedral group \(B_n\). In this section, we give an overview of a method introduced by Chen [9] to compute the cycle index of \(B_n\), which will be used in the determination of the cycle index of the subgroup consisting of symmetries that fix a hyperplane spanned by vertices of \(Q_n\).

We proceed with a brief review of the cycle index of a finite group acting on a finite set, see, for example, Brualdi [8]. Let \(G\) be a finite group that acts on a finite set \(X\). Then each element \(g\in G\) induces a permutation on \(X\). The cycle type of a permutation is defined to be a multiset \(\{1^{k_1},2^{k_2},\ldots \}\), where \(k_i\) is the number of cycles of length \(i\) that appear in the cycle decomposition of the permutation. For \(g\in G\), let \(c(g)\) denote the cycle type of the permutation on \(X\) induced by \(g\). Let \(z=(z_1,z_2,\ldots )\) be a sequence of indeterminants, and let

$$\begin{aligned} z^{c(g)}=z_1^{k_1}z_2^{k_2}\cdots . \end{aligned}$$

The cycle index of \(G\) is defined as

$$\begin{aligned} Z_G(z)=Z_G(z_1,z_2,\ldots )=\frac{1}{|G|}\sum _{g\in G}z^{c(g)}. \end{aligned}$$
(2.1)

Pólya’s enumeration theorem shows that the cycle index in (2.1) can be applied to count nonisomorphic colorings of \(X\) by using a given number of colors. To be more specific, let us color the elements of \(X\) by using \(m\) colors, say \(c_1,c_2,\ldots ,c_m\). Let \(C_G(u_1,\ldots ,u_m)\) be the polynomial obtained from the cycle index \(Z_G(z)\) by substituting \(z_i\) with \(u_1^i+\cdots +u_m^i\). Pólya’s enumeration theorem states that the number of nonisomorphic colorings of \(X\) by using the \(m\) colors \(c_1, \ldots ,c_m\) such that \(a_i\) elements of \(X\) receive the color \(c_i\) equals

$$\begin{aligned} \left[ u_1^{a_1}\cdots u_m^{a_m}\right] C_G(u_1,\ldots ,u_m), \end{aligned}$$

where \(\left[ u_1^{a_1}\cdots u_m^{a_m}\right] C_G(u_1,\ldots ,u_m)\) is the coefficient of \(u_1^{a_1}\cdots u_m^{a_m}\) in \(C_G(u_1,\ldots ,u_m)\).

For a coloring of \(Q_n\) with two colors, say, black and white, the black vertices can be considered as vertices of a \(0/1\)-polytope of \(Q_n\). This establishes a one-to-one correspondence between equivalence classes of colorings and \(0/1\)-equivalence classes of \(Q_n\). Let \(Z_n(z)\) denote the cycle index of \(B_n\) acting on the vertex set \(V_n\), and let \(C_n(u_1,u_2)\) be the polynomial obtained from \(Z_n(z)\) by substituting \(z_i\) with \(u_1^i+u_2^i\). By Pólya’s theorem, we have

$$\begin{aligned} A_n(k)=\big [u_1^ku_2^{2^n-k}\big ]C_n(u_1,u_2). \end{aligned}$$
(2.2)

The computation of \(Z_n(z)\) has been studied by Pólya [16] and Harrison and High [14]. Explicit expressions of \(Z_n(z)\) for \(n\le 6\) are given by Aguila [5], which are listed below.

$$\begin{aligned} Z_1(z)\!&= \!z_1,\\ Z_2(z)\!&= \!\frac{1}{8} \Big (z_1^4+2z_1^2z_2+3z_2^2+2z_4\Big ),\\ Z_3(z)\!&= \!\frac{1}{48} \Big (z_1^8+6z_1^4z_2^2+13z_2^4+8z_1^2z_3^2+12z_4^2+8z_2z_6\Big ),\\ Z_4(z)\!&= \!\frac{1}{384} \left( \begin{array}{l} z_1^{16}+12z_1^8z_2^4+12z_1^4z_2^6+51z_2^8+48z_8^2\\ +\,48z_1^2z_2z_4^3+84z_4^4+96z_2^2z_6^2+32z_1^4z_3^4 \end{array}\right) ,\\ Z_5(z)\!&= \!\frac{1}{3840} \left( \begin{array}{l} z_1^{32}+20z_1^{16}z_2^8+ 60z_1^8z_2^{12}+231z_2^{16}+ 80z_1^8z_3^8+240z_1^4z_2^2z_4^6\\ +\,240z_2^4z_4^6+520z_4^8+384z_1^2z_5^6+160z_1^4z_2^2z_3^4z_6^2+ 720z_2^4z_6^4\\ +\,480z_8^4+384z_2z_{10}^3+320z_4^2z_{12}^2 \end{array}\right) ,\\ Z_6(z)\!&= \!\frac{1}{46080} \left( \begin{array}{l} z_1^{64}\!+\!30z_1^{32}z_2^{16}\!+\!180z_1^{16}z_2^{24}\!+\!120z_1^8z_2^{28}\!+\! 1053z_2^{32}\!+\!160z_1^{16}z_3^{16}\\ \!+\,640z_1^4z_3^{20}\!+\!720z_1^8z_2^4z_4^{12}\!+\! 1440z_1^4z_2^6z_4^{12}\!+\!2160z_2^8z_4^{12}\!+\!4920z_4^{16}\\ \!+\,2304z_1^4z_5^{12}\!+\! 960z_1^8z_2^4z_3^8z_6^4\!+\!5280z_2^8z_6^8\!+\!3840z_1^2z_2z_3^2z_6^9\!+\!5760z_8^8\\ \!+\,1920z_2^2z_6^{10}+6912z_2^2z_{10}^6+3840z_4^4z_{12}^4+ 3840z_4z_{12}^5 \end{array}\right) \!. \end{aligned}$$

For \(k>2^{n-1}\), we have shown that \(F_n(k)=A_n(k)\). Thus, by (2.2) we obtain that for \(k>2^{n-1}\),

$$\begin{aligned} F_n(k)=\big [u_1^ku_2^{2^n-k}\big ]C_n(u_1,u_2). \end{aligned}$$

For \(n=4,5\) and \(6\), the values of \(F_n(k)\) for \(k>2^{n-1}\) are given in Tables 1, 2 and 3.

Table 1 \(F_4(k)\) for \(k>8\)
Table 2 \(F_5(k)\) for \(k>16\)
Table 3 \(F_6(k)\) for \(k>32\)

We next recall the method of Chen [9] for computing the cycle index of \(B_n\). A symmetry of \(Q_n\) can be represented as a signed permutation on \(\{1,2,\ldots ,n\}\), which is a permutation on \(\{1,2,\ldots ,n\}\) with a plus or a minus sign attached to each element. Following the notation in Chen and Stanley [10] or Chen [9], we may write a signed permutation as the form of the cycle decomposition and ignore the plus signs. For example, \((\overline{2}4\overline{5})(3)(1\overline{6})\) represents a signed permutation, where \((245)(3)(16)\) is its underlying permutation. The action of a signed permutation \(w\in B_n\) on the vertices of \(Q_n\) is defined as follows. For a vertex \((x_1, x_2,\ldots , x_n)\) of \(Q_n\), we define \(w(x_1,x_2,\ldots ,x_n)\) to be the vertex \((y_1,y_2,\ldots ,y_n)\) of \(Q_n\) as given by

$$\begin{aligned} y_i= \left\{ \begin{array}{ll} x_{\pi (i)} \qquad &{}\hbox {if}\; i\hbox { is associated with a plus sign},\\ 1-x_{\pi (i)} \qquad &{}\hbox {if}\; i\hbox { is associated with a minus sign}, \end{array}\right. \end{aligned}$$
(2.3)

where \(\pi \) is the underlying permutation of \(w\).

We end this section with the following formula of Chen [9], which will be used in Sect. 5 to compute the cycle structure of a symmetry that fixes a hyperplane spanned by vertices of \(Q_n\).

Let \(n\) be a positive integer, and let \(p_1^{n_1}\ldots p_r^{n_r}\) be the prime factorization of \(n\). Let \(\mu (n)\) be the classical number-theoretic Möbius function, that is, \(\mu (n)=(-1)^r\) if \(n_1=\cdots =n_r=1\), and \(\mu (n)=0\) otherwise.

Theorem 2.1

Let \(G\) be a group that acts on a finite set \(X\). For any \(g\in G\), the number of \(i\)-cycles of the permutation on \(X\) induced by \(g\) is given by

$$\begin{aligned} \frac{1}{i}\sum _{j|i}\mu (i/j)\psi (g^j), \end{aligned}$$

where \(\psi (g^j)\) is the number of fixed points of \(g^j\) acting on \(X\).

3 \(H_n(k)\) for \(2^{n-2}<k\le 2^{n-1}\)

Recall that \(H_n(k)\) is the number of \(0/1\)-equivalence classes of \(Q_n\) with \(k\) vertices that are not full-dimensional. In this section, we show that for \(2^{n-2}<k\le 2^{n-1}\), the number \(H_n(k)\) is determined by the number of equivalence classes of \(0/1\)-polytopes with \(k\) vertices that are contained in every hyperplane spanned by vertices of \(Q_n\). For this reason, it is necessary to consider all possible hyperplanes spanned by vertices of \(Q_n\).

A hyperplane spanned by vertices of \(Q_n\) is also called a spanned hyperplane of \(Q_n\). In other words, a spanned hyperplane of \(Q_n\) is a hyperplane in \(\mathbb {R}^n\) such that the affine space spanned by the vertices of \(Q_n\) contained in this hyperplane is of dimension \(n-1\). Let

$$\begin{aligned} H:a_1x_1+a_2x_2+\cdots +a_nx_n=b \end{aligned}$$

be a spanned hyperplane of \(Q_n\), where \(a_1,\ldots ,a_n\) and \(b\) are integers. For \(n\le 8\), all spanned hyperplanes of \(Q_n\) have been found by Aichholzer and Aurenhammer [4].

As will be seen, in order to compute \(H_n(k)\) for \(2^{n-2}<k\le 2^{n-1}\), we need to consider equivalence classes of spanned hyperplanes of \(Q_n\) under the symmetries of \(Q_n\). Note that the symmetries of \(Q_n\) can be expressed by permuting the coordinates and changing \(x_i\) to \(1-x_i\) for some indices \(i\). Therefore, for each equivalence class of spanned hyperplanes of \(Q_n\), we can choose a representative of the form

$$\begin{aligned} a_1x_1+a_2x_2+\cdots +a_tx_t=b, \end{aligned}$$
(3.1)

where \(t\le n\) and \(0<a_1\le a_2\le \cdots \le a_t\).

A complete list of spanned hyperplanes of \(Q_n\) for \(n\le 6\) can be found in Aichholzer [2]. The following hyperplanes are representatives of equivalence classes of spanned hyperplanes of \(Q_4\):

$$\begin{aligned}&x_1=0,\\&x_1+x_2=1,\\&x_1+x_2+x_3=1,\\&x_1+x_2+x_3+x_4=\hbox {1 or 2},\\&x_1+x_2+x_3+2x_4=2. \end{aligned}$$

In addition to the above hyperplanes, which can also be viewed as spanned hyperplanes of \(Q_5\), we have the following representatives of equivalence classes of spanned hyperplanes of \(Q_5\):

$$\begin{aligned}&x_1+x_2+x_3+x_4+x_5=\hbox {1 or 2},\\&x_1+x_2+x_3+x_4+2x_5=\hbox {2 or 3},\\&x_1+x_2+x_3+2x_4+2x_5=\hbox {2 or 3},\\&x_1+x_2+2x_3+2x_4+2x_5=\hbox {3 or 4},\\&x_1+x_2+x_3+x_4+3x_5=3,\\&x_1+x_2+x_3+2x_4+3x_5=3,\\&x_1+x_2+2x_3+2x_4+3x_5=4. \end{aligned}$$

When \(n=6\), for the purpose of computing \(F_6(k)\) for \(16<k\le 32\), we need the representatives of equivalence classes of spanned hyperplanes of \(Q_6\) containing more than 16 vertices. There are 6 such representatives:

$$\begin{aligned}&x_1=0,\\&x_1+x_2=1,\\&x_1+x_2+x_3=1,\\&x_1+x_2+x_3+x_4=2,\\&x_1+x_2+x_3+x_4+x_5=2,\\&x_1+x_2+x_3+x_4+x_5+x_6=3. \end{aligned}$$

Note that two equivalent spanned hyperplanes of \(Q_n\) contain the same number of vertices of \(Q_n\) because the symmetry of \(Q_n\) preserves the number of vertices. So we may say that an equivalence class of spanned hyperplanes of \(Q_n\) contains \(k\) vertices, by which we mean that every spanned hyperplane in this class contains \(k\) vertices of \(Q_n\).

To state the main result of this section, we need to define an equivalence relation on \(0/1\)-polytopes contained in a set of points in \(\mathbb {R}^n\). Given a set \(S\subset \mathbb {R}^n\), consider the set of \(0/1\)-polytopes of \(Q_n\) that are contained in \(S\). Restricting the \(0/1\)-equivalence relation to this set induces an equivalence relation. More precisely, two \(0/1\)-polytopes in the set of \(0/1\)-polytopes of \(Q_n\) contained in \(S\) are equivalent if one can be transformed to the other by a symmetry of \(Q_n\). Such an equivalence class is called a partial \(0/1\)-equivalence class of \(S\). Denote by \(\mathcal {P}(S,k)\) the set of partial \(0/1\)-equivalence classes of \(S\) with \(k\) vertices. The cardinality of \(\mathcal {P}(S,k)\) is denoted by \(N_S(k)\).

Let \(h(n,k)\) denote the number of equivalence classes of spanned hyperplanes of \(Q_n\) that contain at least \(k\) vertices. Assume that \(H_1,H_1,\ldots ,H_{h(n,k)}\) are the representatives of equivalence classes of spanned hyperplanes of \(Q_n\) containing at least \(k\) vertices. We use \(\mathcal {H}_n(k)\) to denote the set of \(0/1\)-equivalence classes of \(Q_n\) with \(k\) vertices that are not full-dimensional. We shall define a map, denoted \(\Phi \), from the (disjoint) union of \(\mathcal {P}(H_i,k)\), where \(1\le i\le h(n,k)\), to \(\mathcal {H}_n(k)\). Given a partial \(0/1\)-equivalence class \(\mathcal {P}\in \mathcal {P}(H_i,k)\), we define \(\varPhi (\mathcal {P})\) to be the unique \(0/1\)-equivalence class in \(\mathcal {H}_n(k)\) containing \(\mathcal {P}\). Then we have the following theorem.

Theorem 3.1

For \(2^{n-2}<k\le 2^{n-1}\), the map \(\varPhi \) is a bijection.

Proof

We first show that \(\varPhi \) is injective. Let \(\mathcal {P}_1\) and \(\mathcal {P}_2\) be two distinct partial \(0/1\)-equivalence classes with \(k\) vertices, which are contained in the spanned hyperplanes \(H_i\) and \(H_j\) of \(Q_n\), respectively. Let \(P_1\) be a \(0/1\)-polytope in \(\mathcal {P}_1\), and \(P_2\) be a \(0/1\)-polytope in \(\mathcal {P}_2\). To prove that \(\varPhi \) is an injection, it suffices to show that \(P_1\) and \(P_2\) are not equivalent. This is clear when \(i=j\). We now consider the case \(i\ne j\). Suppose to the contrary that \(P_1\) and \(P_2\) are equivalent. So there exists a symmetry \(w\in B_n\) such that \(w(P_1)=P_2\). Since \(2^{n-2}<k\le 2^{n-1}\), by Theorem 1.1 we see that \(P_1\) and \(P_2\) are of dimension \(n-1\). For a spanned hyperplane \(H\) of \(Q_n\), we use \(w(H)\) to denote the hyperplane obtained from \(H\) under the action of \(w\). So we have \(w(H_i)=H_j\), contradicting the fact that the spanned hyperplanes \(H_i\) and \(H_j\) are not equivalent. Consequently, the 0/1-polytopes \(P_1\) and \(P_2\) are not equivalent.

It remains to show that \(\varPhi \) is surjective. For any \(\mathcal {C}\in \mathcal {H}_n(k)\), we aim to find a partial \(0/1\)-equivalence class \(\mathcal {P}\) such that \(\varPhi (\mathcal {P})=\mathcal {C}\). Let \(P\) be any \(0/1\)-polytope in \(\mathcal {C}\). Since \(P\) is not full-dimensional, there exists a spanned hyperplane \(H\) of \(Q_n\) such that \(P\) is contained in \(H\). It follows that \(H\) contains at leat \(k\) vertices. Thus there exists a representative \(H_j\) (\(1\le j\le h(n,k)\)) such that \(H\) is in the equivalence class of \(H_j\). Assume that \(w(H)=H_j\) for some \(w\in B_n\). So \(w(P)\) is contained in \(H_j\). Let \(\mathcal {P}\) be the partial \(0/1\)-equivalence class of \(H_j\) containing \(w(P)\). Clearly, we have \(\varPhi (\mathcal {P})=\mathcal {C}\). This completes the proof. \(\square \)

It should also be noted that in the proof of Theorem 3.1, the condition \(2^{n-2}<k\le 2^{n-1}\) is required. When \(k\le 2^{n-2}\), the map \(\varPhi \) is not necessarily an injection while is always a surjection. For a \(0/1\)-polytope \(P\) with \(k\le 2^{n-2}\) vertices contained in a spanned hyperplane of \(Q_n\), it is not always true that \(\mathrm {dim}(P)=n-1\). So there may exist equivalent \(0/1\)-polytopes \(P\) and \(P'\) with \(k\) vertices and nonequivalent spanned hyperplanes \(H\) and \(H'\) such that \(P\) is contained in \(H\) and \(P'\) is contained in \(H'\). If this is the case, then \(\varPhi \) maps these two partial 0/1-equivalence classes containing \(P\) and \(P'\) to the same 0/1-equivalence class in \(\mathcal {H}_n(k)\).

As a consequence of Theorem 3.1, we obtain the following formula.

Corollary 3.2

For \(2^{n-2}< k\le 2^{n-1}\),

$$\begin{aligned} H_n(k)=\sum _{i=1}^{h(n,k)}N_{H_i}(k). \end{aligned}$$
(3.2)

By Corollary 3.2, the computation of \(H_n(k)\) for \(2^{n-2}<k\le 2^{n-1}\) is carried out by determining the number of partial \(0/1\)-equivalence classes of every spanned hyperplane of \(Q_n\). We shall explain how to compute the latter in the rest of this section.

For \(2^{n-2}<k\le 2^{n-1}\), let \(H\) be a spanned hyperplane of \(Q_n\) containing at least \(k\) vertices. Let \(P\) and \(P'\) be two distinct 0/1-polytopes of \(Q_n\) with \(k\) vertices that are contained in \(H\). Assume that \(P\) and \(P'\) belong to the same partial \(0/1\)-equivalence class of \(H\). Then there exists a symmetry \(w\in B_n\) such that \(w(P)=P'\). By Theorem 1.1, both \(P\) and \(P'\) have dimension \(n-1\). Hence we have \(w(H)=H\).

Let

$$\begin{aligned} F(H)=\{w\in B_n\,|\, w(H)=H\} \end{aligned}$$

be the stabilizer subgroup of \(H\), namely, the subgroup of \(B_n\) that fixes \(H\). By the above argument, we see that \(P\) and \(P'\) belong to the same partial \(0/1\)-equivalence class of \(H\) if and only if one can be transformed to the other by a symmetry in \(F(H)\). So, for \(2^{n-2}<k\le 2^{n-1}\), we can use Pólya’s theorem to compute the number \(N_H(k)\) of partial \(0/1\)-equivalence classes of \(H\) with \(k\) vertices.

Denote by \(V_n(H)\) the set of vertices of \(Q_n\) that are contained in \(H\). Consider the action of \(F(H)\) on \(V_n(H)\). Assume that each vertex in \(V_n(H)\) is assigned one of the two colors, say, black and white. For such a coloring of the vertices in \(V_n(H)\), assume that the black vertices are vertices of a \(0/1\)-polytope contained in \(H\). Clearly, for \(2^{n-2}< k\le 2^{n-1}\), this leads to a one-to-one correspondence between partial \(0/1\)-equivalence classes of \(H\) with \(k\) vertices and equivalence classes of colorings of the vertices in \(V_n(H)\) with \(k\) black vertices.

Write \(Z_H(z)\) for the cycle index of \(F(H)\), and let \(C_H(u_1,u_2)\) denote the polynomial obtained from \(Z_H(z)\) by substituting \(z_i\) with \(u_1^i+u_2^i\).

Theorem 3.3

Assume that \(2^{n-2}<k\le 2^{n-1}\), and let \(H\) be a spanned hyperplane of \(Q_n\) containing at least \(k\) vertices of \(Q_n\). Then we have

$$\begin{aligned} N_H(k)=\big [u_1^ku_2^{|V_n(H)|-k}\big ]C_H(u_1,u_2). \end{aligned}$$

We shall compute the cycle index \(Z_H(z)\) in Sects. 4 and 5. Section 4 is devoted to a characterization of the stabilizer group \(F(H)\). In Sect. 5, we will give an explicit expression for \(Z_H(z)\).

4 The Structure of the Stabilizer \(F(H)\)

In this section, we aim to characterize the stabilizer \(F(H)\) for a given spanned hyperplane \(H\) of \(Q_n\).

As mentioned in Sect. 3, for every equivalence class of spanned hyperplanes of \(Q_n\), we can choose a representative of the form

$$\begin{aligned} H:\, a_1x_1+a_2x_2+\cdots +a_tx_t=b, \end{aligned}$$
(4.1)

where the coefficients \(a_i\) are positive integers with \(a_1\le a_2\le \cdots \le a_t\), and \(b\) is a nonnegative integer.

From now on, we shall restrict our attention only to spanned hyperplanes of \(Q_n\) in the form of (4.1). We define the type of the spanned hyperplane \(H\) in (4.1) to be a vector \(\alpha =(\alpha _1,\alpha _2,\ldots ,\alpha _\ell )\), where \(\alpha _i\) is the multiplicity of \(i\) occurring in the set \(\{a_1,a_2,\ldots ,a_t\}\). For example, let

$$\begin{aligned} H:x_1+x_2+2x_3+2x_4+3x_5=4 \end{aligned}$$

be a spanned hyperplane of \(Q_5\). Then the type of \(H\) is \(\alpha =(\alpha _1,\alpha _2,\alpha _3)=(2,2,1)\).

For positive integers \(i\) and \(j\) with \(i\le j\), let \([i,j]\) denote the interval \(\{i,i+1,\ldots ,j\}\). Let \(\alpha =(\alpha _1,\alpha _2,\ldots ,\alpha _\ell )\) be the type of a spanned hyperplane. For \(i=1,2,\ldots ,\ell \), let \(S_{\alpha _i}\) be the group of permutations on the interval

$$\begin{aligned} \big [\alpha _1+\cdots +\alpha _{i-1}+1,\alpha _1+\cdots + \alpha _{i-1}+\alpha _i\big ], \end{aligned}$$
(4.2)

where we assume that \(\alpha _0=0\). We define

$$\begin{aligned} S_\alpha =S_{\alpha _1}\times S_{\alpha _2}\times \cdots \times S_{\alpha _\ell }, \end{aligned}$$
(4.3)

where \(\times \) denotes the direct product of groups. We also define

$$\begin{aligned} \overline{S}_\alpha =\overline{S}_{\alpha _1}\times \overline{S}_{\alpha _2}\times \cdots \times \overline{S}_{\alpha _\ell }, \end{aligned}$$
(4.4)

where \(\overline{S}_{\alpha _i}\) is the set of signed permutations on the interval (4.2) for which every element is associated with the minus sign.

Let

$$\begin{aligned} P(H)=\left\{ \begin{array}{lll} S_\alpha &{} \ \ \mathrm {if} \ \sum _{i=1}^t a_i\ne 2b,\\ S_\alpha \cup \overline{S}_\alpha &{} \ \ \mathrm {if}\ \sum _{i=1}^t a_i=2b. \end{array}\right. \end{aligned}$$
(4.5)

We have the following characterization of the stabilizer of a spanned hyperplane of \(Q_n\).

Theorem 4.1

Let \(H:a_1x_1+a_2x_2+\cdots +a_tx_t=b\) be a spanned hyperplane of \(Q_n\). Then

$$\begin{aligned} F(H)=P(H)\times B_{n,t}, \end{aligned}$$

where \(B_{n,t}\) is the group of signed permutations on the interval \([t+1,n]\).

To give a proof of Theorem 4.1, we need to describe the action of a symmetry of \(Q_n\) on a hyperplane in \(\mathbb {R}^n\). Let \(H:a_1x_1+a_2x_2+\cdots +a_nx_n=b\) be a hyperplane in \(\mathbb {R}^n\), and \(w\) be a symmetry in \(B_n\). Recall that \(w(H)\) is the hyperplane obtained from \(H\) under the action of \(w\). Let \(s(w)\) be the set of entries of \(w\) that are assigned the minus sign. In view of (2.3), we see that \(w(H)\) is of the form

$$\begin{aligned} \sum _{i\,\notin s(w)}a_{\pi (i)}x_i+ \sum _{j\,\in s(w)}a_{\pi (j)}(1-x_j)=b, \end{aligned}$$
(4.6)

where \(\pi \) is the underlying permutation of \(w\). For \(1\le j\le n\), let

$$\begin{aligned} s(w,j)= \left\{ \begin{array}{l@{\quad }l} -1 \qquad &{}\text{ if } j\in s(w),\\ 1 \qquad &{}\text{ otherwise }.\end{array}\right. \end{aligned}$$

Then (4.6) can be rewritten as

$$\begin{aligned} s(w,1)\cdot a_{\pi (1)}x_1+s(w,2)\cdot a_{\pi (2)} x_2+\cdots +s(w,n)\cdot a_{\pi (n)}x_n=b-\sum _{j\in s(w)}a_{\pi (j)}. \end{aligned}$$
(4.7)

For example, let

$$\begin{aligned} H:x_1-x_2-x_3+2x_4=1 \end{aligned}$$

be a hyperplane in \(\mathbb {R}^4\), and let \(w=(1)(\bar{2}\bar{3})(4)\in B_4\). Then \(w(H)\) is the following hyperplane:

$$\begin{aligned} x_1+x_2+x_3+2x_4=3. \end{aligned}$$

We are now in a position to prove Theorem 4.1.

Proof

Assume that \(w\in F(H)\) and \(\pi \) is the underlying permutation of \(w\). We aim to show that \(w\in P(H)\times B_{n,t}\). Notice that \(w(H)\) can be expressed in the form of (4.7). Since \(H=w(H)\), it follows that for \(1\le j\le t\), \(s(w,j)\) are either all positive or all negative. So we have the following two cases.

Case 1: \(s(w,j)\) is positive for \(1\le j\le t\). In this case, it is clear that \(w(H)\) is of the following form:

$$\begin{aligned} a_{\pi (1)}x_1+a_{\pi (2)}x_2+\cdots +a_{\pi (t)}x_t=b, \end{aligned}$$

where \(a_{\pi (j)}=a_j\) for \(1\le j\le t\). So we deduce that, for any \(1\le j\le t\), \(\pi (j)\) is in the interval \(\big [\alpha _1+\cdots +\alpha _{i-1}+1,\alpha _1+\cdots + \alpha _{i-1}+\alpha _i\big ]\) that contains the element \(j\). This implies that \(w\in S_\alpha \times B_{n,t}\).

Case 2: \(s(w,j)\) is negative for \(1\le j\le t\). Then \(w(H)\) is of the following form:

$$\begin{aligned} -a_{\pi (1)}x_1-a_{\pi (2)}x_2-\cdots -a_{\pi (t)}x_t=b-(a_1+\cdots +a_t). \end{aligned}$$

Since \(w(H)=H\), we have \(a_{\pi (j)}=a_j\) for \(1\le j\le t\) and \(b-(a_1+\cdots +a_t)=-b\). This yields that \(w\in \overline{S}_\alpha \times B_{n,t}\).

Combining the above two cases, we deduce that \(w\in P(H)\times B_{n,t}\). It remains to show that if \(w\) belongs to \(P(H)\times B_{n,t}\), then \(w\) fixes \(H\). Write \(w=\pi \, \sigma \), where \(\pi \in P(H)\) and \(\sigma \in B_{n,t}\). We have the following two cases.

Case 1: \(\pi \in S_{\alpha }\). By (4.7), the hyperplane \(w(H)\) is of the following form:

$$\begin{aligned} a_{\pi (1)}x_1+\cdots +a_{\pi (t)}x_t=b. \end{aligned}$$

By the definition of \(S_\alpha \), we see that \(a_{\pi (i)}=a_i\) for \(1\le i\le t\). So we have \(w(H)=H\).

Case 2: \(\pi \in \overline{S}_{\alpha }\). Let \(\pi _0\) be the underlying permutation of \(\pi \). By (4.7), the hyperplane \(w(H)\) can be expressed as

$$\begin{aligned} -a_{\pi _0(1)}x_1-\cdots -a_{\pi _0(t)}x_t=b-(a_1+\cdots +a_t). \end{aligned}$$

By the definition of \(\overline{S}_\alpha \), we see that \(a_{\pi _0(i)}=a_i\) for \(1\le i\le t\), which, together with the following relation:

$$\begin{aligned} 2b=a_1+\cdots +a_t, \end{aligned}$$

implies that \(w(H)=H\). This completes the proof. \(\square \)

We conclude this section with a sufficient condition to determine whether two elements in the subgroup \(P(H)\) are in the same conjugacy class. Recall that for a group \(G\), two elements \(g_1\) and \(g_2\) are in the same conjugacy class of \(G\) if there exists an element \(g\in G\) such that \(g_1=gg_2 g^{-1}\). This condition will be used in Sect. 5 for the purpose of computing the cycle index of the stabilizer group of a spanned hyperplane \(H\).

Let \(H\) be a spanned hyperplane of \(Q_n\) of type \(\alpha =(\alpha _1,\ldots ,\alpha _\ell )\). Recall that each element \(\pi \) in the subgroup \(P(H)\) is either in \(S_\alpha \) or in \(\overline{S}_{\alpha }\). Hence \(\pi \) can be expressed as a product \(\pi =\pi _1\pi _2\cdots \pi _\ell \), where, for \(1\le i\le \ell \), \(\pi _i\) belongs to \( S_{\alpha _i}\) if \(\pi \in S_\alpha \), and \(\pi _i\) belongs to \( \overline{S}_{\alpha _i}\) if \(\pi \in \overline{S}_\alpha \).

Theorem 4.2

Let \(\pi =\pi _1\pi _2\cdots \pi _\ell \) and \(\pi '=\pi _1'\pi _2'\cdots \pi _\ell '\) be two elements in \(P(H)\) such that \(\pi \) and \(\pi '\) are both in \( S_{\alpha }\), or \(\pi \) and \(\pi '\) are both in \(\overline{S}_{\alpha }\). If the underlying permutations of \(\pi _i\) and \(\pi _i'\) have the same cycle type for any \(1\le i\le \ell \), then \(\pi \) and \(\pi '\) are in the same conjugacy class of \(P(H)\).

Proof

We first consider the case when both \(\pi \) and \(\pi '\) are in \(S_\alpha \). Since \(\pi _i\) and \(\pi _i'\) are permutations of the same cycle type, they are in the same conjugacy class. So there is a permutation \(w_i\in S_{\alpha _i}\) such that \(\pi _i=w_i\pi _i'w_i^{-1}\). It follows that \(\pi =(w_1\pi _1'w_1^{-1})\cdots (w_\ell \pi _\ell 'w_\ell ^{-1})=w\pi 'w^{-1}\), where \(w=w_1\cdots w_\ell \in S_\alpha \). This shows that \(\pi \) and \(\pi '\) are in the same conjugacy class.

It remains to consider the case when both \(\pi \) and \(\pi '\) are in \(\overline{S}_\alpha \). Let \(\pi _0\) (resp., \(\pi '_0\)) be the underlying permutation of \(\pi \) (resp., \(\pi '\)). Then there is a symmetry \(w\in S_\alpha \) such that \(\pi _0=w\pi '_0w^{-1}\). We claim that \(\pi =w\pi 'w^{-1}\). Indeed, it is enough to show that \(\pi (x_1,x_2,\ldots ,x_t) =w\pi 'w^{-1}(x_1,x_2,\ldots ,x_t)\) for any point \((x_1,x_2,\ldots ,x_t)\) in \(\mathbb {R}^t\). Assume that \(\pi (x_1,x_2,\ldots ,x_t)=(y_1,y_2,\ldots ,y_t)\) and \(w\pi 'w^{-1}(x_1,x_2,\ldots ,x_t)=(z_1,z_2,\ldots ,z_t)\). Since every element of \(\pi \) is associated with the minus sign, by (2.3) we find that \(y_i=1-x_{\pi _0(i)}\) for \(1\le i\le t\). On the other hand, using (2.3), it is easy to check that \(z_i=1-x_{w^{-1}\pi '_0w(i)}\) for \(1\le i\le t\). Since \(\pi _0=w\pi '_0w^{-1}\), we deduce that \(\pi _0(i)=w^{-1}\pi '_0w(i)\). Therefore, we have \(y_i=z_i\) for \(1\le i\le t\). So the claim is justified. This completes the proof. \(\square \)

5 The Computation of \(Z_H(z)\)

In this section, we obtain a formula for the cycle index \(Z_H(z)\) of the stabilizer group \(F(H)\) of a spanned hyperplane \(H\) of \(Q_n\).

Let

$$\begin{aligned} H:\, a_1x_1+a_2x_2+\cdots +a_tx_t=b \end{aligned}$$
(5.1)

be a spanned hyperplane of \(Q_n\). Recall that \(V_n(H)\) is the set of vertices of \(Q_n\) contained in \(H\). To compute the cycle index \(Z_H(z)\), we need to determine the cycle structures of permutations on \(V_n(H)\) induced by the symmetries in \(F(H)\). By Theorem 4.1, each symmetry in \(F(H)\) can be written uniquely as a product \(\pi w\), where \(\pi \in P(H)\) and \(w\in B_{n,t}\). We shall define two group actions for the subgroups \(P(H)\) and \(B_{n,t}\), and we derive an expression for the cycle type of the permutation on \(V_n(H)\) induced by \(\pi w\) in terms of the cycle types of the permutations induced by \(\pi \) and \(w\).

Let \(H\) be a spanned hyperplane of \(Q_n\) as given in (5.1). To define the action of \(P(H)\), we should consider \(H\) as a hyperplane in \(\mathbb {R}^t\). Clearly, if \(H\) is regarded as a hyperplane in \(\mathbb {R}^t\), it is a spanned hyperplane of \(Q_t\). Denote by \(V_t(H)\) the set of vertices of \(Q_t\) that are contained in \(H\), namely,

$$\begin{aligned} V_t(H)=\{(x_1,x_2,\ldots ,x_t)\in V_t\,|\, a_1x_1+a_2x_2+\cdots +a_tx_t=b\}. \end{aligned}$$

Since \(P(H)\) stabilizes the set \(V_t(H)\), we get an action of the group \(P(H)\) on \(V_t(H)\).

We also need to describe the action of a symmetry in the group \(B_{n,t}\) on the set of vertices of \(Q_{n-t}\). Assume that \(w\in B_{n,t}\), namely, \(w\) is a signed permutation on the interval \([t+1,n]\). Subtracting each element of \(w\) by \(t\), we get a signed permutation on \([1,n-t]\). In this way, each signed permutation in \(B_{n,t}\) corresponds to a symmetry of \(Q_{n-t}\). Hence, \(B_{n,t}\) is isomorphic to the group \(B_{n-t}\) of symmetries of \(Q_{n-t}\). This leads to an action on \(V_{n-t}\).

Let \(\pi w\) be a symmetry in \(F(H)\), where \(\pi \in P(H)\) and \(w\in B_{n,t}\). The following lemma shows that the cycle type of the permutation on \(V_n(H)\) induced by \(\pi w\) is determined by the cycle types of the permutations on \(V_t(H)\) and \(V_{n-t}\) induced by \(\pi \) and \(w\). For an element \(g\) in a group \(G\) acting on a finite set \(X\), we use \(c(g)\) to denote the cycle type of the permutation on \(X\) induced by \(g\), which is written as a multiset \(\{1^{c_1}, 2^{c_2}, \ldots \}\).

Lemma 5.1

Let \(H:\, a_1x_1+a_2x_2+\cdots +a_tx_t=b\) be a spanned hyperplane of \(Q_n\), and \(\pi w\) be a symmetry in \(F(H)\), where \(\pi \in P(H)\) and \(w\in B_{n,t}\). Assume that \(c(\pi )=\{1^{m_1},2^{m_2},\ldots \}\) and \(c(w)=\{1^{k_1},2^{k_2},\ldots \}\). Then we have

$$\begin{aligned} c(\pi w)=\bigcup _{i\ge 1}\bigcup _{j\ge 1} \left\{ \Big (\mathrm {lcm}(i,j)\right) ^{\frac{ijm_ik_j}{\mathrm {lcm}(i,j)}}\Big \}, \end{aligned}$$
(5.2)

where \(\bigcup \) denotes the disjoint union of multisets, and \(\mathrm {lcm}(i,j)\) denotes the least common multiple of \(i\) and \(j\).

Proof

Clearly, each vertex in \(V_n(H)\) can be expressed as a vector of the following form

$$\begin{aligned} (x_1,\ldots ,x_{t},y_1,\ldots ,y_{n-t}), \end{aligned}$$

where \((x_1,\ldots ,x_{t})\) is a vertex in \(V_t(H)\) and \((y_1,\ldots ,y_{n-t})\) is a vertex of \(Q_{n-t}\). Assume that \(|V_t(H)|=m\). Let \(V_t(H)=\{u_1,u_2, \ldots ,u_{m}\}\) and \(V_{n-t}=\{v_1,v_2,\ldots ,v_{2^{n-t}}\}\). Then each vertex in \(V_n(H)\) can be expressed as an ordered pair \((u_i,v_j)\), where \(1\le i\le m\) and \(1\le j\le 2^{n-t}\).

Let \(C_i=({s_1},\ldots ,{s_i})\) be an \(i\)-cycle of the permutation on \(V_t(H)\) induced by \(\pi \), that is, \(C_i\) maps the vertex \(u_{s_p}\) to the vertex \(u_{s_{p+1}}\) for \(1\le p\le i-1\), and to the vertex \(u_{s_1}\) for \(p=i\). Similarly, let \(C_j=({t_1},\ldots ,t_j)\) be a \(j\)-cycle of the permutation on \(V_{n-t}\) induced by \(w\), that is, \(C_j\) maps the vertex \(v_{t_q}\) to the vertex \(v_{t_{q+1}}\) for \(1\le q \le j-1\), and to the vertex \(v_{t_1}\) for \(q=j\). Define \(C_{i,j}\) to be the permutation on the subset \(\{(u_{s_p},v_{t_q})\,|\, 1\le p\le i,\ 1\le q\le j\}\) of \(V_n(H)\) such that

$$\begin{aligned} C_{i,j}(u_{s_p},v_{t_q})=(C_i(u_{s_p}),C_j(v_{t_q})). \end{aligned}$$

It is easily seen that the induced permutation of \(\pi w\) on \(V_n(H)\) is the direct product of \(C_{i,j}\), where \(C_i\) (resp., \(C_j\)) runs over the cycles of the permutation on \(V_t(H)\) (resp., \(V_{n-t}\)) induced by \(\pi \) (resp., \(w\)).

It can be verified that the cycle type of \(C_{i,j}\) is

$$\begin{aligned} \Big \{\left( \mathrm {lcm}(i,j)\right) ^{\frac{ij}{\mathrm {lcm} (i,j)}}\Big \}. \end{aligned}$$

Thus the cycle type of the induced permutation of \(\pi w\) on \(V_n(H)\) is given by (5.2). This completes the proof. \(\square \)

For convenience, we introduce the following notation. Let \(\pi \) be a symmetry in \(P(H)\). Assume that the cycle type of the permutation on \(V_t(H)\) induced by \(\pi \) is

$$\begin{aligned} c(\pi )=\{1^{m_1},2^{m_2}, \ldots \}. \end{aligned}$$

For \(j\ge 1\), we define

$$\begin{aligned} f_{\pi ,j}(z)=\prod _{i\ge 1}(z_{\mathrm {lcm}(i, j)})^{\frac{ijm_i}{\mathrm {lcm}(i,j)}}. \end{aligned}$$
(5.3)

We have the following proposition.

Proposition 5.2

Let \(H\) be a spanned hyperplane of \(Q_n\) of type \(\alpha \). Assume that \(\pi =\pi _1\pi _2\cdots \pi _\ell \) and \(\pi '=\pi _1'\pi _2' \cdots \pi _\ell '\) are two symmetries in \(P(H)\) such that \(\pi \) and \(\pi '\) are both in \( S_{\alpha }\), or \(\pi \) and \(\pi '\) are both in \(\overline{S}_{\alpha }\). If the underlying permutations of \(\pi _i\) and \(\pi _i'\) have the same cycle type for \(1\le i\le \ell \), then, for \(j\ge 1\),

$$\begin{aligned} f_{\pi ,j}(z)=f_{\pi ',j}(z). \end{aligned}$$
(5.4)

Proof

It follows from Theorem 4.2 that \(\pi \) and \(\pi '\) are in the same conjugacy class of \(P(H)\). Hence the permutations on \(V_t(H)\) induced by \(\pi \) and \(\pi '\) are in the same conjugacy class, that is, \(c(\pi )=c(\pi ')\). Since \(f_{\pi ,j}(z)\) depends only on the cycle type \(c(\pi )\), we deduce that \(f_{\pi ,j}(z)=f_{\pi ',j}(z)\). This completes the proof. \(\square \)

To compute the cycle index \(Z_H(z)\), we recall some notation and terminology on integer partitions. A partition \(\lambda \) of a positive integer \(n\), denoted \(\lambda \vdash n\), will be expressed in the multiset form, that is, \(\lambda =\{1^{m_1},2^{m_2},\ldots \}\), where \(m_i\) is the number of occurrences of \(i\) in \(\lambda \). Denote by \(\ell (\lambda )\) the number of parts of \(\lambda \), that is, \(\ell (\lambda )=m_1+m_2+\cdots \). For a partition \(\lambda =\{1^{m_1},2^{m_2},\ldots \}\), let

$$\begin{aligned} m_\lambda =1^{m_1} m_1! 2^{m_2}m_2!\cdots . \end{aligned}$$

For two partitions \(\lambda \) and \(\mu \), define \(\lambda \cup \mu \) to be the partition obtained by putting the parts of \(\lambda \) and \(\mu \) together. For example, for \(\lambda =\{1,2\}\) and \(\mu =\{1^2,3\}\), we have \(\lambda \cup \mu =\{1^3,2,3\}\).

Let \(H\) be a spanned hyperplane of \(Q_n\) of type \(\alpha =(\alpha _1,\alpha _2,\ldots ,\alpha _\ell )\). For \(1\le i\le \ell \), let \(\mu ^i\) be a partition of \(\alpha _i\), and let \(\mu =\mu ^1\cup \cdots \cup \mu ^\ell \). Assume that \(\pi =\pi _1\pi _2\cdots \pi _\ell \) (resp., \(\pi '=\pi _1'\pi _2'\cdots \pi _\ell '\)) is a symmetry in \(S_\alpha \) (resp., \(\overline{S}_\alpha \)) such that the underlying permutation of \(\pi _i\) (resp., \(\pi _i'\)) has cycle type \(\mu ^i\) for \(1\le i\le \ell \). For \(j\ge 1\), define

$$\begin{aligned} g_{\mu ,j}(z)=f_{\pi ,j}(z) \end{aligned}$$

and

$$\begin{aligned} \overline{g}_{\mu ,j}(z)=f_{\pi ',j}(z). \end{aligned}$$

By Proposition 5.2, the functions \(g_{\mu ,j}(z)\) and \(\overline{g}_{\mu ,j}(z)\) are well defined.

Let

$$\begin{aligned} g_\mu (z)=(g_{\mu ,1}(z),g_{\mu ,2}(z),\ldots ) \end{aligned}$$

and

$$\begin{aligned} \overline{g}_\mu (z)=(\overline{g}_{\mu ,1}(z),\overline{g}_{\mu , 2}(z),\ldots ). \end{aligned}$$

In the above notation, we obtain the following formula for the cycle index \(Z_H(z)\).

Theorem 5.3

Let \(H:a_1x_1+a_2x_2+\cdots +a_tx_t=b\) be a spanned hyperplane of \(Q_n\). Assume that \(H\) is of type \(\alpha =(\alpha _1,\alpha _2,\ldots ,\alpha _\ell )\). Then we have

$$\begin{aligned} Z_H(z)=\frac{1}{2^{\delta (H)}}\sum _{(\mu ^1,\ldots ,\mu ^\ell )} \prod _{i=1}^\ell m_{\mu ^i}^{-1}\left( Z_{n-t}(g_{\mu }(z))+ \delta (H)Z_{n-t}(\overline{g}_\mu (z))\right) , \end{aligned}$$
(5.5)

where \(\mu ^i\vdash \alpha _i\), \(\mu =\mu ^1\cup \cdots \cup \mu ^\ell \), \(\delta (H)=1\) if \(\sum _{i=1}^t a_i= 2b\) and \(\delta (H)=0\) otherwise.

Proof

Let \(\pi \in P(H)\) and \(w\in B_{n,t}\), and let

$$\begin{aligned} c(w)=\{1^{k_1},2^{k_2},\ldots \} \end{aligned}$$

be the cycle type of the permutation on \(V_{n-t}\) induced by \(w\). In view of Lemma 5.1, we have

$$\begin{aligned} {z}^{c(\pi w)}=f_{\pi ,1}(z)^{k_1}f_{\pi ,2}(z)^{k_2}\cdots \end{aligned}$$
(5.6)

Summing over signed permutations \(w\) in \(B_{n,t}\) and using (2.1) and (5.6), we deduce that

$$\begin{aligned} \sum _{\pi w}{z}^{c(\pi w)}&=\sum _{w}f_{\pi ,1}(z)^{k_1}f_{\pi ,2}(z)^{k_2}\cdots \\&=(n-t)!2^{n-t}Z_{n-t}(f_{\pi ,1}(z),f_{\pi ,2}(z),\ldots )\\&=(n-t)!2^{n-t}Z_{n-t}(f_\pi ({z})), \end{aligned}$$

where

$$\begin{aligned} f_{\pi }(z)=(f_{\pi ,1}(z),f_{\pi ,1}(z),\ldots ). \end{aligned}$$

Thus,

$$\begin{aligned} Z_H({z})&=\frac{1}{|F(H)|}\sum _{\pi w\in F(H)}{z}^{c(\pi w)}\nonumber \\&=\frac{1}{|F(H)|}\sum _{\pi \in P(H)}(n-t)!2^{n-t}Z_{n-t}(f_\pi ({z})) \nonumber \\&=\frac{(n-t)!2^{n-t}}{|F(H)|}\Bigg (\sum _{\pi \in S_\alpha }Z_{n-t}(f_\pi ({z}))+ \delta (H)\sum _{\pi '\in \overline{S}_\alpha }Z_{n-t}(f_{\pi '}({z}))\Bigg ), \end{aligned}$$
(5.7)

where \(\delta (H)=1\) if \(\sum _{i=1}^t a_i= 2b\) and \(\delta (H)=0\) otherwise.

For a partition \(\lambda \vdash n\), there are \(\frac{n!}{m_\lambda }\) permutations on \(\{1,2,\ldots ,n\}\) that are of type \(\lambda \), see, for example, Stanley [17, Proposition 1.3.2]. So the number of symmetries \(\pi =\pi _1\pi _2\ldots \pi _\ell \) in \(S_\alpha \) (or, \(\overline{S}_\alpha \)) such that for \(i=1,2,\ldots ,\ell \), the underlying permutation of \(\pi _i\) is of type \(\mu ^i\) equals

$$\begin{aligned} \prod _{i=1}^\ell \frac{\alpha _i!}{m_{\mu ^i}}. \end{aligned}$$
(5.8)

Combining (5.7), (5.8) and Proposition 5.2, we obtain that

$$\begin{aligned} Z_H({z}) =\frac{(n-t)!2^{n-t}}{|F(H)|}\sum _{(\mu ^1,\ldots ,\mu ^\ell )} \prod _{i=1}^\ell \frac{\alpha _i!}{m_{\mu ^i}}\big (Z_{n-t}(g_\mu ({z}))+ \delta (H)Z_{n-t}(\overline{g}_\mu ({z}))\big ),\nonumber \\ \end{aligned}$$
(5.9)

where \(\mu ^i\vdash \alpha _i\) and \(\mu =\mu ^1\cup \cdots \cup \mu ^\ell \).

It is easily seen that

$$\begin{aligned} |F(H)|=(n-t)!2^{n-t+\delta (H)}\prod _{i=1}^\ell \alpha _i!. \end{aligned}$$
(5.10)

Substituting (5.10) into (5.9), we are led to (5.5). \(\square \)

By Theorem 5.3, to compute the cycle index \(Z_H({z})\), it suffices to determine the cycle type \(c(\pi )\) of the permutation on \(V_t(H)\) induced by \(\pi \in P(H)\). Let \(c(\pi )=\{1^{m_1},2^{m_2},\ldots \}\). By Theorem 2.1, we have

$$\begin{aligned} m_i=\frac{1}{i}\sum _{j|i}\mu (i/j)\psi (\pi ^j), \end{aligned}$$
(5.11)

where \(\psi (\pi ^j)\) is the number of vertices in \(V_t(H)\) that are fixed by \(\pi ^j\). The following theorem gives a formula for \(\psi (\pi )\), leading to a formula for \(\psi (\pi ^j)\) .

Theorem 5.4

Let \(H:a_1x_1+a_2x_2+\cdots +a_tx_t=b\) be a spanned hyperplane of \(Q_n\). Assume that \(\pi =\pi _1\pi _2\cdots \pi _\ell \) is a symmetry in \(P(H)\) such that the underlying permutation of \(\pi _i\) is of type \(\mu ^i=\{1^{m_{i1}},2^{m_{i2}},\ldots \}\) for \(i=1,2,\ldots ,\ell \). Then

$$\begin{aligned} \psi (\pi )= \left\{ \begin{array}{lll} \left[ x^b\right] \prod _{i=1}^\ell \prod _{j\ge 1}(1+x^{ij})^{m_{ij}} &{} \ \ \mathrm {if} \ \ \pi \in S_\alpha ,\\ \chi (\mu )2^{\ell (\mu )} &{} \ \ \mathrm {if}\ \ \pi \in \overline{S}_\alpha , \end{array}\right. \end{aligned}$$
(5.12)

where \(\mu =\mu ^1\cup \cdots \cup \mu ^\ell \), \(\chi (\mu )=1\) if \(\mu \) has no odd parts and \(\chi (\mu )=0\) otherwise.

Proof

We first consider the case when \(\pi \) is in \(S_\alpha \). Observe that, a vertex \(v=(x_1,x_2,\ldots ,x_t)\) of \(Q_t\) is both fixed by \(\pi \) and contained in \(V_t(H)\) if and only if

  1. (1)

    For \(1\le i\le \ell \) and each \(k\)-cycle \(({j_1},{j_2},\ldots ,{j_k})\) of \(\pi _i\), we have

    $$\begin{aligned} x_{{j_1}}=x_{{j_2}}=\cdots =x_{{j_k}}. \end{aligned}$$
  2. (2)

    \(a_1x_1+a_2x_2+\cdots +a_tx_t=b\), or equivalently,

    $$\begin{aligned} b_1+2b_2+\cdots +\ell b_\ell =b, \end{aligned}$$

    where \(b_i\) (\(1\le i\le \ell \)) is the sum of the entries of \(v\) equal to 1.

It can be easily deduced that the number of vertices of \(Q_t\) satisfying the above conditions is given by

$$\begin{aligned} \big [x^b\big ]\prod _{i=1}^\ell \prod _{j\ge 1}\big (1+ x^{ij}\big )^{m_{ij}}. \end{aligned}$$

This proves (5.12) for the case when \(\pi \in S_\alpha \).

We now consider the case when \(\pi \) is in \(\overline{S}_\alpha \). Notice that a vertex \(v=(x_1,x_2,\ldots ,x_t)\) of \(Q_t\) is fixed by \(\pi \) if and only if, for any \(k\)-cycle \((\overline{j_1},\overline{j_2},\ldots ,\overline{j_k})\) of \(\pi \), we have

$$\begin{aligned} (x_{{j_1}},x_{{j_2}},\ldots ,x_{{j_k}}) =(1-x_{{j_2}},1-x_{{j_3}},\ldots ,1-x_{{j_1}}). \end{aligned}$$
(5.13)

Consequently, if a vertex \(v=(x_1,x_2,\ldots ,x_t)\) of \(Q_t\) is fixed by \(\pi \), then, for any \(k\)-cycle \((\overline{j_1},\overline{j_2},\ldots ,\overline{j_k})\) of \(\pi \), the vector \((x_{j_1},x_{j_2},\ldots , x_{j_k})\) is either \((0,1,\ldots ,0,1)\) or \((1,0,\ldots ,1,0)\). This implies that \(k\) is even. Thus \(\pi \) does not have any fixed points if \(\pi \) contains an odd cycle.

We now assume that \(\pi \) has only even cycles. In this case, the number of vertices of \(Q_t\) fixed by \(\pi \) equals \(2^{\ell (\mu )}\). To prove \(\psi (\pi )=2^{\ell (\mu )}\), we need to demonstrate that any vertex of \(Q_t\) fixed by \(\pi \) is in \(V_t(H)\). Let \(v=(x_1,x_2,\ldots ,x_t)\) be a vertex of \(Q_t\) fixed by \(\pi \). Since, for each cycle \((\overline{j_1},\overline{j_2},\ldots ,\overline{j_k})\) of \(\pi \), the vector \((x_{{j_1}},x_{{j_2}},\ldots ,x_{{j_k}})\) is either \((0,1,\ldots ,0,1)\) or \((1,0,\ldots ,1,0)\), we deduce that \(a_1x_1+a_2x_2+\cdots +a_tx_t=b\) by applying the relation \( a_1+\cdots +a_t=2b.\) Hence the vertex \(v\) is in \(V_t(H)\). This completes the proof. \(\square \)

Based on Theorem 5.4, we can compute \(\psi (\pi ^j)\) since the cycle structure of \(\pi ^j\) is easily determined by the cycle structure of \(\pi \). Let \(\pi =\pi _1\pi _2\ldots \pi _\ell \) be a symmetry in \(P(H)\) such that for \(1\le i \le \ell \), the underlying permutation of \(\pi _i\) is of type \(\mu ^i=\{1^{m_{i1}},2^{m_{i2}},\ldots \}\). Clearly, we have \(\pi ^j=\pi _1^j\pi _2^j\ldots \pi _\ell ^j\). Moreover, we see that \(\pi ^j\) belongs to \(S_\alpha \) if \(\pi \) is in \(S_\alpha \) or \(\pi \) is in \(\overline{S}_\alpha \) and \(j\) is even, and \(\pi ^j\) belongs to \(\overline{S}_\alpha \) otherwise. Let \(\mathrm {gcd}(i,j)\) denote the greatest common divisor of \(i\) and \(j\). Then the cycle type of the underlying permutation of \(\pi _i^j\) is given by

$$\begin{aligned} \Big \{1^{m_{i1}}, \mathrm {gcd}(2,j)^{\frac{2m_{i2}}{\mathrm {gcd} (2,j)}}, \mathrm {gcd}(3,j)^{\frac{3m_{i3}}{\mathrm {gcd} (3,j)}}, \ldots \Big \}. \end{aligned}$$

6 \(F_n(k)\) for \(n=4,5,6\) and \(2^{n-2}< k\le 2^{n-1}\)

This section is devoted to the computation of \(F_n(k)\) for \(n=4,5,6\) and \(2^{n-2}< k\le 2^{n-1}\). This requires the cycle index \(Z_H(z)\) for every spanned hyperplane \(H\) of \(Q_n\) for \(n=4,5,6\) that contains more than \(2^{n-2}\) vertices of \(Q_n\).

Recall that \(h(n,k)\) denotes the number of equivalence classes of spanned hyperplanes of \(Q_n\) containing at least \(k\) vertices. Let \(H_1,H_2,\ldots ,H_{h(n,k)}\) be the representatives of these equivalence classes. When \(2^{n-2}< k\le 2^{n-1}\), combining relation (1.1), Corollary 3.2 and Theorem 3.3, we deduce that

$$\begin{aligned} F_n(k)&=A_n(k)-H_n(k) \nonumber \\&=A_n(k)-\sum _{i=1}^{h(n,k)}N_{H_i}(k)\nonumber \\&=A_n(k)-\sum _{i=1}^{h(n,k)}\big [u_1^{k}u_2^{|V_n(H_i)| -k}\big ]C_{H_i}(z_1,z_2). \end{aligned}$$
(6.1)

Using formula (6.1), we proceed to compute \(F_n(k)\) for \(n=4,5,6\) and \(2^{n-2}< k\le 2^{n-1}\). We start with the computation of \(F_4(k)\) for \(4<k\le 8\). For \(t\le n\), we use \(H_n^t\) to denote the following hyperplane in \(\mathbb {R}^n\)

$$\begin{aligned} x_1+x_2+\cdots +x_t=\left\lfloor t/2\right\rfloor . \end{aligned}$$

In this notation, representatives of equivalence classes of spanned hyperplanes of \(Q_4\) containing more than \(4\) vertices are as follows:

$$\begin{aligned}&H_4^1:x_1=0,\\&H_4^2:x_1+x_2=1,\\&H_4^3:x_1+x_2+x_3=1,\\&H_4^4:x_1+x_2+x_3+x_4=2. \end{aligned}$$

Employing the techniques in Sect. 5, we obtain the cycle indices \(Z_{H_4^1}(z)\) and \(Z_{H_4^2}(z)\) as given below.

$$\begin{aligned} Z_{H_4^1}(z)&=Z_3(z),\\ Z_{H_4^2}(z)&= \frac{1}{16} \big (9z_2^4+4z_4^2+2z_1^4z_2^2+z_1^8\big ). \end{aligned}$$

For the remaining two hyperplanes \(H=H_4^3\) and \(H_4^4\), it is easily checked that \(N_H(k)=1\) for \(k=5,6\), and \(N_H(k)=0\) for \(k=7,8\). Thus, applying (6.1) we can determine \(F_4(k)\) for \(k=5,6,7,8\). These values are given in Table 4, which agree with the computation of Aichholzer [1].

Table 4 \(F_4(k)\) for \(k=5,6,7,8\)

Observing that \(F_4(k)=0\) for \(k\le 4\), thus we have completed the enumeration of full-dimensional \(0/1\)-equivalence classes of \(Q_4\).

We now compute \(F_5(k)\) for \(8< k\le 16\). Representatives of equivalence classes of spanned hyperplanes of \(Q_5\) containing more than \(8\) vertices are \(H_5^1,H_5^2,H_5^3,H_5^4,H_5^5\). By utilizing the techniques in Sect. 5, we obtain that

$$\begin{aligned} Z_{H_5^1}({z})&=Z_4({z}),\\ Z_{H_5^2}({z})&=\frac{1}{96} \big (z_1^{16}+6z_1^8z_2^4+33z_2^8+8z_1^4z_3^4+24z_4^4+ 24z_2^2z_6^2\big ),\\ Z_{H_5^3}({z})&=\frac{1}{48} \big (12z_2^6+8z_4^3+2z_1^6z_2^3+ z_1^{12}+6z_1^2z_2^5+3z_1^4z_2^4+ 6z_6^2+4z_{12}+4z_3^2z_6\\&\qquad \qquad +2z_3^4\big ),\\ Z_{H_5^4}({z})&=\frac{1}{96}\big (z_1^{12}+27z_2^6+9z_1^4z_2^4+8z_3^4 +24z_6^2+18z_2^2z_4^2+6z_1^4z_4^2+3z_1^8z_2^2\big ),\\ Z_{H_5^5}({z})&=\frac{1}{120} \big (24z_5^2+30z_2z_4^2+20z_1 z_3z_6+20z_1z_3^3+15z_1^2z_2^4+10z_1^4z_2^3+z_1^{10}\big ). \end{aligned}$$

Consequently, the values \(F_5(k)\) for \(8<k\le 16\) can be derived from (6.1), and they agree with the computation of Aichholzer [1], see Table 5.

Table 5 \(F_5(k)\) for \(8< k\le 16\)

The main objective of this section is to compute \(F_6(k)\) for \(16<k\le 32\). As mentioned in Sect. 4, there are 6 representatives of equivalence classes of spanned hyperplanes of \(Q_6\) containing more than 16 vertices, namely, \(H_6^1, H_6^2, H_6^3, H_6^4, H_6^5, H_6^6\). Again, by applying the techniques in Sect. 5, we obtain that

$$\begin{aligned} Z_{H_6^1}({z})&=Z_5({z}),\\ Z_{H_6^2}({z})&=\frac{1}{768} \left( \begin{array}{l} z_1^{32}+12z_1^{16}z_2^8+12z_1^8z_2^{12}+127z_2^{16}+32z_1^8z_3^8\\ +\,\, 48z_1^4z_2^2z_4^6+168z_4^8+224z_2^4z_6^4+96z_8^4+48z_2^4z_4^6 \end{array}\right) ,\\ Z_{H_6^3}({z})&=\frac{1}{288} \left( \begin{array}{l} z_1^{24}+6z_1^{12}z_2^6+52z_2^{12}+18z_3^8+ 48z_4^6+32z_2^3z_6^3+3z_1^8z_2^8\\ +\,\, 18z_1^4z_2^{10}+24z_1^2z_3^2z_2^2z_6^2+8z_1^6z_3^6+12z_3^4 z_6^2+42z_6^4+24z_{12}^2 \end{array}\right) ,\\ Z_{H_6^4}({z})&=\frac{1}{384} \left( \begin{array}{l} z_1^{24}+81z_2^{12}++2z_1^{12}z_2^6+18z_1^4z_2^{10}+ 15z_1^8z_2^8+72z_6^4+32z_{12}^2\\ +\,\, 64z_4^6+16z_3^4z_6^2+8z_3^8 +54z_2^4z_4^4+12z_1^4z_2^2z_4^4+6z_1^8z_4^4+3z_1^{16}z_2^4 \end{array}\right) ,\\ Z_{H_6^5}({z})&=\frac{1}{240} \left( \begin{array}{l} z_1^{20}+24z_{10}^2+60z_2^2z_4^4+26z_2^{10}+20z_1^2z_3^2z_6^2\\ +\,\, 20z_1^2z_3^6+15z_1^4z_2^8+10z_1^8z_2^6+40z_2z_6^3+24z_5^4 \end{array}\right) ,\\ Z_{H_6^6}({z})&=\frac{1}{1440}\left( \begin{array}{l} z_1^{20}+144z_5^4+144z_{10}^2+320z_2z_6^3+270z_2^2z_4^4 +76z_2^{10}\\ +\,\, 90z_1^4z_4^4+30z_1^8z_2^6+45z_1^4z_2^8+240z_1^2z_3^2z_6^2+80 z_1^2z_3^6 \end{array}\right) . \end{aligned}$$

Using (6.1), we can compute \(F_6(k)\) for \(16<k\le 32\). These values are listed in Table 6.

Table 6 \(F_6(k)\) for \(16<k\le 32\)

7 \(H_6(k)\) for \(k=13,14, 15,16\)

In this section, we compute \(H_6(k)\) for \(k=13,14, 15,16\). Together with the computation of Aichholzer for \(n=6\) and \(k\le 12\), we complete the enumeration of full-dimensional \(0/1\)-equivalence classes of the 6-dimensional hypercube. In fact, we can compute \(H_n(k)\) when \(n>4\) and \(k\) is close to \(2^{n-2}\).

Let us recall the map \(\varPhi \) defined in Sect. 3. Let \(H_1,H_2,\ldots ,H_{h(n,k)}\) be the representatives of equivalence classes of spanned hyperplanes of \(Q_n\) containing at least \(k\) vertices. As before, we use \(\mathcal {P}(H_i,k)\) to denote the set of partial \(0/1\)-equivalence classes of \(H_i\) with \(k\) vertices, and use \(N_{H_i}(k)\) to denote the cardinality of \(\mathcal {P}(H_i,k)\). Let \(\mathcal {P}\) be a partial \(0/1\)-equivalence class in the (disjoint) union of \(\mathcal {P}(H_i,k)\) where \(1\le i\le h(n,k)\). Then \(\varPhi \) maps \(\mathcal {P}\) to the unique \(0/1\)-equivalence class in \(\mathcal {H}_n(k)\) that contains \(\mathcal {P}\).

When \(k\le 2^{n-2}\), it is possible that there exist equivalent \(0/1\)-polytopes \(P\) and \(P'\) that are contained respectively in \(H_i\) and \(H_j\), where \(1\le i\ne j\le h_{n,k}\). Let \(\mathcal {P}\) and \(\mathcal {P'}\) be the partial \(0/1\)-equivalence classes of \(H_i\) and \(H_j\) that contain \(P\) and \(P'\) respectively. Then we have \(\varPhi (\mathcal {P})=\varPhi (\mathcal {P'})\). So \(\varPhi \) is not necessarily an injection when \(k\le 2^{n-2}\). Note that when restricted to \(\mathcal {P}(H_i,k)\), \(\varPhi \) is always an injection. Thus, in order to compute \(H_n(k)\) for \(k\le 2^{n-2}\), we need to compute the number \(N_{H_i}(k)\) of partial 0/1-equivalence classes of each spanned hyperplane \(H_i\) as well as the number of partial \(0/1\)-equivalence classes with \(k\) vertices that are contained in the intersection of distinct spanned hyperplanes.

The objective of this section is to find a way to compute \(N_{H_i}(k)\) when \(k\) is close to \(2^{n-2}\). As will be seen, when \(2^{n-3}<k\le 2^{n-2}\), to compute \(N_{H_i}(k)\) we need to consider all possible symmetries \(w\in B_n\) such that the intersections of \(H_i\) and \(w(H_i)\) contain at least \(k\) vertices. To be more specific, we need to determine the number of partial \(0/1\)-equivalence classes with \(k\) vertices that are contained in the intersection \(H_i\cap w(H_i)\). Moreover, when \(k\) is close to \(2^{n-2}\), there are only a few symmetries \(w\) such that the intersection \(H_i\cap w(H_i)\) contains at least \(k\) vertices. This makes it possible to compute \(N_{H_i}(k)\) when \(k\) is close to \(2^{n-2}\).

When \(k\) is close to \(2^{n-2}\), the same technique can be applied to determine the number of partial \(0/1\)-equivalence classes with \(k\) vertices that are contained in the intersection of distinct spanned hyperplanes.

Notice that

$$\begin{aligned} \mathcal {H}_n(k)=A_1\cup A_2\cup \cdots \cup A_{h(n,k)}, \end{aligned}$$

where

$$\begin{aligned} A_i=\varPhi (\mathcal {P}(H_i,k)). \end{aligned}$$

By the principle of inclusion–exclusion, we have the following expression for \(H_n(k)\).

Lemma 7.1

Let \(H\) be a spanned hyperplane of \(Q_n\). Then we have

$$\begin{aligned} H_n(k)=&\sum _{1\le i\le h(n,k)}|A_i|- \sum _{1\le i_1<i_2\le h(n,k)}|A_{i_1}\cap A_{i_2}|\nonumber \\&+\sum _{1\le i_1<i_2<i_3 \le h(n,k)}|A_{i_1}\cap A_{i_2}\cap A_{i_3}|-\cdots \end{aligned}$$
(7.1)

By Lemma 7.1, the computation of \(H_n(k)\) reduces to the evaluation of the cardinalities of \(A_{i_1}\cap A_{i_2}\cap \cdots \cap A_{i_m}\), where \(1\le i_1<\cdots <i_m \le h(n,k)\). Since \(\varPhi \) is an injection when restricted to \(\mathcal {P}(H_i,k)\), we have \(|A_i|=N_{H_i}(k)\). Moreover, as will be shown, when \(2^{n-3}<k\le 2^{n-2}\) and \(m\ge 2\), the computation of \(|A_{i_1}\cap A_{i_2}\cap \cdots \cap A_{i_m}|\) can be transformed to the determination of partial \(0/1\)-equivalence classes contained in the intersection of distinct spanned hyperplanes.

We now focus on the computation of \(N_{H}(k)\), where \(H\) is a spanned hyperplane of \(Q_n\) and \(k\) is close to \(2^{n-2}\). Let \(S\subseteq H\) be a subset of \(H\). In Sect. 3, we have defined the partial \(0/1\)-equivalence relation on the set of \(0/1\)-polytopes of \(Q_n\) contained in \(S\). Here we need another equivalence relation on this set, that is, two \(0/1\)-polytopes are said to be equivalent if one can be transformed to the other by a symmetry in the stabilizer \(F(H)\) of \(H\). The associated equivalence classes are called local \(0/1\)-equivalence classes of \(S\). Since \(F(H)\) is a subgroup of \(B_n\), each local \(0/1\)-equivalence class of \(S\) is contained in a unique partial \(0/1\)-equivalence class of \(S\).

Denote by \(\mathcal {L}(S,k)\) the set of local \(0/1\)-equivalence classes of \(S\) with \(k\) vertices. To compute \(N_{H}(k)\) when \(k\) is close to \(2^{n-2}\), we need to compute the cardinality of \(\mathcal {L}(H,k)\) and the cardinality of \(\mathcal {L}(S, k)\), where \(S\) can be expressed as \(S=H\cap w(H)\) for a symmetry \(w\) in \(B_n\) satisfying certain conditions. The cardinality of \(\mathcal {L}(H,k)\) can be obtained from the cycle index \(Z_H(z)\) of the stabilizer \(F(H)\). In the following formula, \(C_H(u_1,u_2)\) denotes the polynomial obtained from \(Z_H(z)\) by substituting \(z_i\) with \(u_1^i+u_2^i\), as defined in Sect. 3.

Lemma 7.2

For any \(1\le k\le 2^{n-1}\), we have

$$\begin{aligned} |\mathcal {L}(H,k)|=\big [u_1^ku_2^{|V_n(H)-k|}\big ]C_H(u_1,u_2). \end{aligned}$$
(7.2)

In the remaining of this section, we assume that \(2^{n-3}<k\le 2^{n-2}\). Keep in mind that \(N_H(k)\) is the cardinality of the set \(\mathcal {P}(H,k)\) of partial 0/1-equivalence classes of \(H\). To compute \(|\mathcal {P}(H,k)|\), we shall define a subset \(\mathcal {L}_1(H,k)\) of \(\mathcal {L}(H,k)\) and a subset \(\mathcal {P}_1(H,k)\) of \(\mathcal {P}(H,k)\), which satisfy the following relation:

$$\begin{aligned} |\mathcal {P}(H,k)|=|\mathcal {L}(H,k)|-|\mathcal {L}_1(H,k)|+ |\mathcal {P}_{1}(H,k)|. \end{aligned}$$

We first define the subset \(\mathcal {L}_1(H,k)\), which depends on a map \(\Psi \) from the set of local 0/1-equivalence classes of certain intersections \(H\cap w(H)\) to the set \(\mathcal {L}(H,k)\). To define \(\Psi \), let \(E(H,k)\) denote the set of affine subspaces \(H\cap w(H)\), where \(w\) ranges over symmetries in \(B_n\) such that

  1. (1)

    \(H\ne w(H)\), that is, the symmetry \(w\) of \(Q_n\) does not fix \(H\);

  2. (2)

    \(H\cap w(H)\) contains at least \(k\) vertices of \(Q_n\).

Consider the equivalence classes of \(E(H,k)\) under the symmetries in \(F(H)\). This means that two elements \(H\cap w(H)\) and \(H\cap w'(H)\) in \(E(H,k)\) are equivalent if there exists a symmetry \(\sigma \in F(H)\) such that

$$\begin{aligned} H\cap w(H)=\sigma (H\cap w'(H)). \end{aligned}$$

Denote by \(h_1(H,k)\) the number of equivalence classes of \(E(H,k)\) under the symmetries in \(F(H)\). Let

$$\begin{aligned} E_1(H,k)=\{H\cap w_{i}(H)\,|\, 1\le i\le h_1(H,k)\} \end{aligned}$$

be the set of representatives of these equivalence classes of \(E(H,k)\).

The map \(\Psi \) is defined from the (disjoint) union of \(\mathcal {L}(H\cap w_i(H),k)\), where \(1\le i\le h_1(H,k)\), to \(\mathcal {L}(H,k)\). Let \(\mathcal {L}\) be a local \(0/1\)-equivalence class in \(\mathcal {L}(H\cap w_i(H),k)\). Define \(\Psi (\mathcal {L})\) to be the unique local \(0/1\)-equivalence class in \(\mathcal {L}(H,k)\) containing \(\mathcal {L}\). We have the following property.

Theorem 7.3

For \(n>4\) and \(2^{n-3}<k\le 2^{n-2}\), the map \(\Psi \) is an injection.

Proof

Let \(\mathcal {L}\) and \(\mathcal {L}'\) be two distinct local \(0/1\)-equivalence classes with \(k\) vertices. Assume that \(\mathcal {L}\) is contained in \(\mathcal {L}(H\cap w_i(H),k)\) and \(\mathcal {L}'\) is contained in \(\mathcal {L}(H\cap w_j(H),k)\)), where \(1\le i, j\le h_1(H,k)\). To prove that \(\Psi \) is an injection, we need to show that \(\Psi (\mathcal {L})\ne \Psi (\mathcal {L}')\). If \(i=j\), from the definition of the local 0/1-equivalence relation, it is clear that \(\Psi (\mathcal {L})\ne \Psi (\mathcal {L}')\).

We now consider the case \(i\ne j\). Let \(P\) and \(P'\) be two \(0/1\)-polytopes contained in \(\mathcal {L}\) and \(\mathcal {L}'\), respectively. We claim that \(\mathrm {dim}(P)=\mathrm {dim}(P')= n-2\). We only give a proof of the assertion that \(\mathrm {dim}(P)= n-2\). The relation \(\mathrm {dim}(P')= n-2\) can be justified by the same argument.

Since \(P\) has more than \(2^{n-3}\) vertices, it follows from Theorem 1.1 that \(\mathrm {dim}(P)\ge n-2\). On the other hand, since \(P\) is contained in the intersection \(H\cap w_i(H)\), we see that \(\mathrm {dim}(P)\le n-2\). Hence we have \(\mathrm {dim}(P)= n-2\).

Based on the above claim, it can be shown that \(\Psi (\mathcal {L})\ne \Psi (\mathcal {L}')\). Suppose to the contrary that \(\Psi (\mathcal {L})= \Psi (\mathcal {L}')\). Then there is a symmetry \(w\in F(H)\) such that \(P=w(P')\). Since \(\mathrm {dim}(P)=\mathrm {dim}(P')= n-2\), we deduce that \(H\cap w_i(H)= w(H\cap w_j(H))\), which contradicts the fact that \(H\cap w_i(H)\) and \(H\cap w_j(H)\) are not equivalent under the symmetries in \(F(H)\). This completes the proof. \(\square \)

We can now give the definition of the subset \(\mathcal {L}_1(H,k)\) of \(\mathcal {L}(H,k)\). Notice that for each \(1\le i\le h_1(H,k)\), \(\Psi (\mathcal {L}(H\cap w_i(H),k))\) is a subset of \(\mathcal {L}(H,k)\). By Theorem 7.3, these subsets are disjoint. We define \(\mathcal {L}_1(H,k)\) to be the union of \(\Psi (\mathcal {L}(H\cap w_i(H),k))\), where \(1\le i\le h_1(H,k)\).

We proceed to define the subset \(\mathcal {P}_1(H,k)\) of \(\mathcal {P}(H,k)\). Let \(\overline{\mathcal {L}}_1(H,k)\) be the complement of \(\mathcal {L}_1(H,k)\), that is,

$$\begin{aligned} \overline{\mathcal {L}}_1(H,k)=\mathcal {L}(H,k)\backslash \mathcal {L}_1(H,k). \end{aligned}$$
(7.3)

In the above notation, for any local \(0/1\)-equivalence class \(\mathcal {L}\in \overline{\mathcal {L}}_1(H,k)\) and any \(0/1\)-polytope \(P\in \mathcal {L}\), if \(w\in B_n\) is a symmetry such that \(w(P)\) is contained in \(H\), then \(w(H)=H\). This yields that \(\mathcal {L}\) is also a partial \(0/1\)-equivalence class of \(H\). Consequently, when \(2^{n-3}<k\le 2^{n-2}\), \(\overline{\mathcal {L}}_1(H,k)\) is a subset of \(\mathcal {P}(H,k)\). Define

$$\begin{aligned} \mathcal {P}_1(H,k)=\mathcal {P}(H,k)\backslash \overline{\mathcal {L}}_1(H,k). \end{aligned}$$
(7.4)

From (7.3) and (7.4), we see that \(N_H(k)\) can be expressed in terms of the cardinalities of \(\mathcal {L}(H,k)\), \(\mathcal {L}_1(H,k)\) and \(\mathcal {P}_1(H,k)\). More precisely,

$$\begin{aligned} N_H(k)&=|\mathcal {P}(H,k)|\nonumber \\&=|\overline{\mathcal {L}}_1(H,k)|+|\mathcal {P}_1(H,k)|\nonumber \\&=|\mathcal {L}(H,k)|-|\mathcal {L}_1(H,k)|+|\mathcal {P}_1(H,k)|. \end{aligned}$$
(7.5)

By Lemma 7.2, \(|{\mathcal {L}}(H,k)|\) can be computed from the cycle index \(Z_H(z)\). From Theorem 7.3, \(|\mathcal {L}_1(H,k)|\) can be derived from the cardinalities of \(\mathcal {L}(H\cap w(H),k)\), where \(H\cap w(H)\in E_1(H,k)\). To compute \(|\mathcal {P}_{1}(H,k)|\), we need a map \(\Gamma \) defined as follows.

Let \(h_2(H,k)\) denote the number of equivalence classes of \(E(H,k)\) under the symmetries in \(B_n\), and let

$$\begin{aligned} E_2(H,k)=\{H\cap w_{i}(H)\,|\, 1\le i\le h_2(H,k)\} \end{aligned}$$

be the set of representatives of these equivalence classes of \(E(H,k)\). We define a map \(\Gamma \) from the (disjoint) union of \(\mathcal {P}(H\cap w_i(H),k)\), where \(1\le i\le h_2(H,k)\), to \(\mathcal {P}_{1}(H,k)\). Let \(\mathcal {P}\) be a partial \(0/1\)-equivalence class in \(\mathcal {P}(H\cap w_i(H),k)\). Then \(\Gamma \) maps \(\mathcal {P}\) to the unique partial \(0/1\)-equivalence class in \(\mathcal {P}_1(H,k)\) that contains \(\mathcal {P}\).

When \(2^{n-3}<k\le 2^{n-2}\), it has been shown that each 0/1-polytope with \(k\) vertices contained in the intersection \(H\cap w_i(H)\) has dimension \(n-2\). This enables us to use the same argument as in the proof Theorem 7.3 to reach the following assertion.

Theorem 7.4

For \(n>4\) and \(2^{n-3}<k\le 2^{n-2}\), the map \(\Gamma \) is a bijection.

Combining Lemma 7.2, Theorem 7.3 and Theorem 7.4, formula (7.5) can be rewritten as

$$\begin{aligned} N_H(k)=\,\big [u_1^{k}u_2^{|V_n(H)|-k}\big ]C_H(u_1,u_2)&-\sum _{H\cap w(H)\in E_1(H,k)}|\mathcal {L}(H\cap w(H),k)|\nonumber \\&+\sum _{H\cap w(H)\in E_2(H,k)}|\mathcal {P}(H\cap w(H),k)|.\quad \end{aligned}$$
(7.6)

So, to compute \(N_H(k)\), it is enough to determine \(|\mathcal {L}(H\cap w(H),k)|\) and \(|\mathcal {P}(H\cap w(H),k)|\). We can compute \(|\mathcal {L}(H\cap w(H),k)|\) and \(|\mathcal {P}(H\cap w(H),k)|\) by applying Pólya’s theorem.

We first consider \(|\mathcal {L}(H\cap w(H),k)|\). Let \(P\) and \(P'\) be any two \(0/1\)-polytopes belonging to the same local \(0/1\)-equivalence class in \(\mathcal {L}(H\cap w(H),k)\). Then there exists a symmetry \(\sigma \) in \(F(H)\) such that \(\sigma (P)=P'\). It is clear from Theorem 1.1 that both \(P\) and \(P'\) have dimension \(n-2\). So we deduce that \(w'(H\cap w(H))=H\cap w(H)\).

Let \(F_1(H,w)\) be the subgroup of \(F(H)\) that stabilizes \(H\cap w(H)\), that is,

$$\begin{aligned} F_1(H,w)=\left\{ \sigma \in F(H)\,|\, \sigma \left( H\cap w(H)\right) =H\cap w(H)\right\} . \end{aligned}$$

Denote by \(V_n(H\cap w(H))\) the set of vertices of \(Q_n\) contained in \(H\cap w(H)\). Consider the action of \(F_1(H,w)\) on \(V_n(H\cap w(H))\). Assume that each vertex in \(V_n(H\cap w(H))\) is assigned one of the two colors, say, black and white. Clearly, when \(2^{n-3}< k\le 2^{n-2}\), this leads to a one-to-one correspondence between local 0/1-equivalence classes in \(\mathcal {L}(H\cap w(H),k)\) and equivalence classes of colorings of the vertices in \(V_n(H\cap w(H))\) with \(k\) black vertices.

Denote by \(Z_{(H,w)}({z})\) the cycle index of \(F_1\left( H,w\right) \) acting on \(V_n(H\cap w(H))\). Write \(C_{(H,w)}(u_1,u_2)\) for the polynomial obtained from \(Z_{(H,w)}({z})\) by substituting \(z_i\) with \(u_1^i+u_2^i\). For \(2^{n-3}<k\le 2^{n-2}\), we obtain that

$$\begin{aligned} |\mathcal {L}(H\cap w(H),k)|=\big [u_1^{k}u_2^{|V_n(H\cap w(H))|-k}\big ]\, C_{(H,w)}(u_1,u_2). \end{aligned}$$
(7.7)

Similarly, we can use Pólya’s theorem to compute \(|\mathcal {P}(H\cap w(H),k)|\). Let \(F_2\left( H,w \right) \) be the subgroup of \(B_n\) that stabilizes \(H\cap w(H)\), that is,

$$\begin{aligned} F_2\left( H,w \right) =\big \{\sigma \in B_n\,|\, \sigma \left( H\cap w(H)\right) =H\cap w(H)\big \}. \end{aligned}$$

Denote by \(Z_{H\cap w(H)}({z})\) the cycle index of \(F_2\left( H,w\right) \) acting on \(V_n(H\cap w(H))\). Write \(C_{H\cap w(H)}(u_1,u_2)\) for the polynomial obtained from \(Z_{H\cap w(H)}(z)\) by substituting \(z_i\) with \(u_1^i+u_2^i\). For \(2^{n-3}<k\le 2^{n-2}\), we have

$$\begin{aligned} |\mathcal {P}(H\cap w(H),k)|=\big [u_1^{k}u_2^{|V_n(H\cap w(H))|-k}\big ]C_{H\cap w(H)}(u_1,u_2). \end{aligned}$$
(7.8)

Now, plugging (7.7) and (7.8) into (7.6), we arrive at the following formula for \(N_H(k)\).

Theorem 7.5

Assume that \(n>4\) and \(2^{n-3}<k\le 2^{n-2}\). Let \(H\) be a spanned hyperplane of \(Q_n\) containing at least \(k\) vertices of \(Q_n\). Let \(q(w)=|V_n(H\cap w(H))|\). Then we have

$$\begin{aligned} N_H(k)=\,\big [u_1^{k}u_2^{|V_n(H)|-k}\big ]C_H(u_1,u_2)&-\sum _{H\cap w(H)\in E_1(H,k)}\big [u_1^{k}u_2^{q(w)-k}\big ]C_{(H,w)}(u_1,u_2)\nonumber \\&+\sum _{H\cap w(H)\in E_2(H,k)}\big [u_1^{k}u_2^{q(w)- k}\big ]C_{H\cap w(H)}(u_1,u_2).\nonumber \\ \end{aligned}$$
(7.9)

For \(n=6\) and \(k=13,14,15,16\), we can use Theorem 7.5 to compute \(N_H(k)\), where \(H\) is a spanned hyperplane of \(Q_6\) containing more than \(12\) vertices. By the computation of Aichholzer [2], in addition to the spanned hyperplanes \(H_6^1, H_6^2, H_6^3, H_6^4, H_6^5, H_6^6\), there are \(8\) representatives of equivalence classes of spanned hyperplanes of \(Q_6\) containing more than \(12\) vertices, namely,

$$\begin{aligned}&H_1:x_1+x_2+x_3+2x_4=2,\\&H_2:x_1+x_2+x_3+x_4=1,\\&H_3:x_1+x_2+x_3+x_4+2x_5=3,\\&H_{4}:x_1+x_2+x_3+x_4+x_5+2x_6=3,\\&H_5:x_1+x_2+x_3+x_4+x_5+x_6=2,\\&H_6:x_1+x_2+x_3+x_4+2x_5=2,\\&H_7:x_1+x_2+x_3+2x_4+2x_5=3,\\&H_8:x_1+x_2+x_3+x_4+2x_5+2x_6=4. \end{aligned}$$

Using a Maple program, when \(k=13,14,15,16\), it is routine to check that \(E(H,k)=\emptyset \) for \(H=H_6^3,H_6^4,H_6^5,H_6^6\) and \(H=H_1,H_2,\ldots ,H_8\). Therefore, for these spanned hyperplanes, by Theorem 7.5 we obtain that

$$\begin{aligned} N_H(k)=\big [u_1^{k}u_2^{|V_n(H)|-k}\big ]\,C_H(u_1,u_2). \end{aligned}$$
(7.10)

The cycle indices \(Z_H(z)\) for \(H=H_6^3,H_6^4,H_6^5,H_6^6\) are given in Sect. 6. Using the techniques in Sect. 5, we can derive the cycle indices for \(H_1,H_2,\ldots , H_5\), which are given below.

$$\begin{aligned} Z_{H_1}({z})&=\frac{1}{48} \left( \begin{array}{l} z_1^{16}+4z_{12}z_4+4z_3^2z_6z_1^2z_2+2z_3^4z_1^4\\ +12z_2^8+8z_4^4+6z_1^4z_2^6+5z_1^8z_2^4+6z_6^2z_2^2 \end{array}\right) ,\\ Z_{H_2}({z})&=\frac{1}{192} \left( \begin{array}{l} z_1^{16}+68z_4^4+24z_6^2z_2^2+16z_{12}z_4+8z_3^4z_1^4\\ +\,\, 39z_2^8+12z_1^4z_2^6+8z_1^8z_2^4+16z_3^2z_6z_1^2z_2 \end{array}\right) ,\\ Z_{H_3}({z})&=\frac{1}{96} \left( \begin{array}{l} z_1^{16}+24z_6^2z_2^2+8z_3^4z_1^4+33z_2^8+ 6z_1^8z_2^4+24z_4^4 \end{array}\right) ,\\ Z_{H_4}({z})&=\frac{1}{120} \left( \begin{array}{l} z_1^{15}\!+\!24z_5^3\!+\!30z_2z_4^3z_1\!+\!20z_1z_3^2z_6z_2\!+\! 20z_1^3z_3^4\!+\!15z_1^3z_2^6\!+\!10z_1^7z_2^4 \end{array}\right) ,\\ Z_{H_5}({z})&=\frac{1}{720} \left( \begin{array}{l} z_1^{15}+120z_3z_6^2+144z_5^3+40z_3^5+180z_1z_2z_4^3\\ +\,\, 40z_1^3z_3^4+60z_1^3z_2^6+ 15z_1^7z_2^4+ 120z_1z_2z_3^2z_6 \end{array}\right) . \end{aligned}$$

For \(H=H_6\), \(H_7\), \(H_8\), we obtain that \(N_{H}(13)=2\), \(N_{H}(14)=1\), and \(N_{H}(15)=N_{H}(16)=0\) without computing the cycle index \(Z_H(z)\). For example, for \(H=H_6\), since \(H_6\) contains \(14\) vertices of \(Q_6\), we have \(N_{H}(14)=1\) and \(N_{H}(15)=N_{H}(16)=0\). On the other hand, there are 14 0/1-polytopes with 13 vertices contained in \(H_6\). It is easy to check that these 14 0/1-polytopes form two partial 0/1-equivalence classes. So we have \(N_{H}(13)=2\). Similarly, we get \(N_{H}(13)=2\), \(N_{H}(14)=1\), and \(N_{H}(15)=N_{H}(16)=0\) for \(H=H_7, H_8\).

It remains to compute \(N_H(k)\) for \(H=H_6^1, H_6^2\) and \(k=13,14,15,16\). We first consider \(H_6^1\). Keep in mind that \(H_6^1\) is the spanned hyperplane \(x_1=0\). Thus, for \(H_6^1\) and \(k=13,14,15,16\), it is easily seen that the intersections \(H_6^1\cap w(H_6^1)\) in \(E(H_6^1,k)\) form only one equivalence class under the symmetries in \(F(H_6^1)\) or \(B_n\). A representative of this equivalence class can be chosen as \(H_6^1\cap w(H_6^1)\), where \(w=(1,2)(3)(4)(5)(6)\). So we have

$$\begin{aligned} E_1(H_6^1,k)=E_2(H_6^1,k) =\big \{(x_1,x_2,\ldots ,x_6)\in \mathbb {R}^6\,|\, x_1=x_2=0\big \}. \end{aligned}$$

Moreover, for \(k=13,14,15,16\), it is easy to check that if two \(0/1\)-polytopes in \(H_6^1\cap w(H_6^1)\) with \(k\) vertices are equivalent under the symmetries in \(B_n\), then they are equivalent under the symmetries in \(F(H_6^1)\). This implies that each local 0/1-equivalence class of \(H_6^1\cap w(H_6^1)\) is also a partial 0/1-equivalence class of \(H_6^1\cap w(H_6^1)\) and vice versa. Hence we obtain

$$\begin{aligned} \mathcal {L}(H_6^1\cap w(H_6^1),k)=\mathcal {P}(H_6^1\cap w(H_6^1),k). \end{aligned}$$

Therefore, for \(k=13,14,15,16\), by formula (7.6) we have

$$\begin{aligned} N_{H_6^1}(k)=\big [u_1^{k}u_2^{32-k}\big ]C_{H_6^1}(u_1,u_2). \end{aligned}$$
(7.11)

We now compute \(N_{H_6^2}(k)\) for \(k=13,14,15,16\). Recall that \(H_6^2\) is the spanned hyperplane \(x_1+x_2=1\). It is not hard to check that the intersections \(H_6^2\cap w(H_6^2)\) in \(E(H_6^2,k)\) form two equivalence classes under the symmetries in \(F(H_6^2)\) or \(B_n\). Moreover, each equivalence class in \(E(H_6^2,k)\) under the symmetries in \(F(H_6^2)\) is an equivalence class in \(E(H_6^2,k)\) under the symmetries in \(B_n\) and vice versa. The representatives of these two equivalence classes can be chosen as \(H_6^2\cap w_1(H_6^2)\) and \(H_6^2\cap w_2(H_6^2)\), where \(w_1=(1,3,2)(4)(5)(6)\) and \(w_2=(1,3)(2,4)(5)(6)\). Notice that the intersections \(H_6^2\cap w_1(H_6^2)\) and \(H_6^2\cap w_2(H_6^2)\) are of the following form:

$$\begin{aligned} H_6^2\cap w_1(H_6^2)=&\, \big \{(x_1,x_2,\ldots ,x_6)\in \mathbb {R}^6\,|\, x_1+x_2=1\ \ \text {and}\ \ x_2+x_3=1 \big \}, \\ H_6^2\cap w_2(H_6^2)=&\,\big \{(x_1,x_2,\ldots ,x_6)\in \mathbb {R}^6\,|\, x_1+x_2=1\ \ \text {and}\ \ x_3+x_4=1 \big \}. \end{aligned}$$

Since the set of vertices contained in \(H_6^2\cap w_1(H_6^2)\) is

$$\begin{aligned}&\{(1,0,1,x_4,x_5,x_6),\, (0,1,0,x_4,x_5,x_6)\,|\, x_i=0\;\hbox {or 1 for}\; i=4,5,6\}, \end{aligned}$$

it is easy to check that for \(k=13,14,15,16\), if two 0/1-polytopes contained in \(H_6^2\cap w_1(H_6^2)\) with \(k\) vertices are equivalent under the symmetries in \(B_n\), then they are equivalent under the symmetries in \(F(H_6^2)\). This means that each local 0/1-equivalence class of \(H_6^2\cap w_1(H_6^2)\) is also a partial 0/1-equivalence class of \(H_6^2\cap w_1(H_6^2)\) and vice versa. So, we have

$$\begin{aligned} \mathcal {L}(H_6^2\cap w_1(H_6^2),k)=\mathcal {P}(H_6^2\cap w_1(H_6^2),k). \end{aligned}$$

Therefore, by formula (7.6) we obtain that for \(k=13,14,15,16\),

$$\begin{aligned} N_{H_6^2}(k)=\big [u_1^{k}u_2^{32-k}\big ]C_{H_6^2}(u_1,u_2)+ |\mathcal {P}(H_6^2\cap w(H_6^2),k)-|\mathcal {L}(H_6^2\cap w(H_6^2),k),\nonumber \\ \end{aligned}$$
(7.12)

where \(w=(1,3)(2,4)(5)(6)\).

Combining (7.10), (7.11) and (7.12) , for \(n=6\) and \(k=13,14,15,16\), we obtain that

$$\begin{aligned} \sum _{i=1}^{h(6,k)}|A_i|=&\sum _{i=1}^6\big [u_1^{k}u_2^{|V_6(H_6^i)| -k}\big ]C_{H_6^i}(u_1,u_2)+\sum _{i=1}^8\big [u_1^{k}u_2^{|V_6(H_i) |-k}\big ]C_{H_i}(u_1,u_2)\nonumber \\&+|\mathcal {P}(H_6^2\cap w(H_6^2),k)-|\mathcal {L}(H_6^2\cap w(H_6^2),k), \end{aligned}$$
(7.13)

where \(w=(1,3)(2,4)(5)(6)\).

By Lemma 7.1, to determine \(H_6(k)\) for \(k=13,14,15,16\), we still need to compute \(|A_{i_1}\cap A_{i_2}\cap \cdots \cap A_{i_m}|\) for \(m\ge 2\). We first consider the case \(m=2\). The computation of the general case can be carried out in the same way.

We now demonstrate how to compute \(|A_i\cap A_j|\) for \(1\le i<j\le h(n,k)\). Let \(E(H_i,H_j,k)\) be the set of affine subspaces \(H_i\cap w(H_j)\) that contain at least \(k\) vertices of \(Q_n\). Denote by \(h(H_i,H_j,k)\) the number of equivalence classes in \(E(H_i,H_j,k)\) under the symmetries in \(B_n\), and let

$$\begin{aligned} E_1(H_i,H_j,k)=\left\{ H_i\cap w_t(H_j)\,|\, 1\le t \le h(H_i,H_j,k)\right\} \end{aligned}$$

be the set of representatives of equivalence classes in \(E(H_i,H_j,k)\).

We consider the union of the sets \(\mathcal {P}(H_i\cap w_t(H_j),k)\) of partial 0/1-equivalence classes of \(H_i\cap w_t(H_j)\) with \(k\) vertices, where \(1\le t\le h(H_i,H_j,k)\), and we define a map \(\Upsilon \) from this set of partial 0/1-equivalence classes to \(A_i\cap A_j\). Let \(\mathcal {P}\) be a partial \(0/1\)-equivalence class in \(\mathcal {P}(H_i\cap w_t(H_j),k)\). Then there is a unique \(0/1\)-equivalence class \(\mathcal {P}'\) in \(A_i\cap A_j\) that contains \(\mathcal {P}\). Define \(\Upsilon (\mathcal {P})=\mathcal {P}'\). We have the following property. The proof is omitted since it is similar to that of Theorem 7.3.

Theorem 7.6

For \(n>4\) and \(2^{n-3}<k\le 2^{n-2}\), the map \(\Upsilon \) is a bijection.

As a consequence of Theorem 7.6, for \(n>4\) and \(2^{n-3}<k\le 2^{n-2}\), we have

$$\begin{aligned} |A_i\cap A_j|=\sum _{t=1}^{h(H_i,H_j,k)}|\mathcal {P}(H_i\cap w_t(H_j),k)|. \end{aligned}$$
(7.14)

The above approach can be used to determine \(|A_{i_1}\cap A_{i_2}\cap \cdots \cap A_{i_m}|\) for \(m\ge 3\). Let

$$\begin{aligned} E(H_{i_1},\ldots ,H_{i_m},k) \end{aligned}$$

be the set of affine subspaces \(H_{i_1}\cap w_2(H_{i_2})\cap \cdots \cap w_m(H_{i_m})\), where \(w_2,\ldots , w_m\) are symmetries in \(B_n\) such that \(H_{i_1}\cap w_2(H_{i_2})\cap \cdots \cap w_m(H_{i_m})\) contains at least \(k\) vertices of \(Q_n\). Denote by \(E_1(H_{i_1},\ldots ,H_{i_m},k)\) the set of representatives of equivalence classes of \(E(H_{i_1},\ldots ,H_{i_m},k)\) under the symmetries in \(B_n\).

Consider the union of the sets \(\mathcal {P}(H_{i_1}\cap w_2(H_{i_2})\cap \cdots \cap w_m(H_{i_m}),k)\) of partial 0/1-equivalence classes, where

$$\begin{aligned} H_{i_1}\cap w_2(H_{i_2})\cap \cdots \cap w_m(H_{i_m})\in E_1(H_{i_1},\ldots ,H_{i_m},k). \end{aligned}$$

We define a map \(\Omega \) from this set of partial 0/1-equivalence classes to \(A_{i_1}\cap A_{i_2}\cap \cdots \cap A_{i_m}\). Let \(\mathcal {P}\) be a partial 0/1-equivalence of \(H_{i_1}\cap w_2(H_{i_2})\cap \cdots \cap w_m(H_{i_m})\). Then \(\Omega \) maps \(\mathcal {P}\) to the unique \(0/1\)-equivalence class in \(A_{i_1}\cap A_{i_2}\cap \cdots \cap A_{i_m}\) that contains \(\mathcal {P}\). Using the same argument as in the proof of Theorem 7.3, we obtain the following property.

Theorem 7.7

For \(n>4\) and \(2^{n-3}<k\le 2^{n-2}\), the map \(\Omega \) is a bijection.

As a consequence of Theorem 7.7, we see that for \(n> 4\) and \(2^{n-3}<k\le 2^{n-2}\),

$$\begin{aligned} |A_{i_1}\cap A_{i_2}\cap \cdots \cap A_{i_m}|=\sum |\mathcal {P}(H_{i_1} \cap w_2(H_{i_2})\cap \cdots \cap w_m(H_{i_m}),k)|, \end{aligned}$$
(7.15)

where the sum ranges over the representatives \(H_{i_1}\cap w_2(H_{i_2})\cap \cdots \cap w_m(H_{i_m})\) of equivalence classes in \(E(H_{i_1},\ldots ,H_{i_m},k)\).

The following theorem shows that for \(m\ge 3\), the set \(E(H_{i_1},\ldots ,H_{i_m},k)\) is empty under certain conditions. When \(n=6\) and \(k=13,14,15,16\), this property allows us to deduce that for any \(m\ge 4\) and any spanned hyperplanes \(H_{i_1},\ldots ,H_{i_m}\), the set \(E(H_{i_1},\ldots ,H_{i_m},k)\) is empty.

Theorem 7.8

Let \(n>4\) and \(2^{n-3}<k\le 2^{n-2}\). If there exist \(1\le p< q\le m\) such that \(E(H_{i_{p}},H_{i_{q}},k)\) is empty, then \(E(H_{i_1},\ldots ,H_{i_m},k)\) is empty.

Proof

Assume that there exist \(1\le p< q\le m\) such that \(E(H_{i_{p}},H_{i_{q}},k)\) is empty. Suppose to the contrary that \(E(H_{i_1},\ldots ,H_{i_m},k)\) is nonempty. Let

$$\begin{aligned} S=H_{i_1}\cap w_2(H_{i_2})\cap \cdots \cap w_m(H_{i_m}) \end{aligned}$$

be an affine space belonging to \(E(H_{i_1},\ldots ,H_{i_m},k)\). Let \(w_1\) be the identity element \(e\) in \(B_n\). We claim that

$$\begin{aligned} S=w_p(H_{i_p})\cap w_q(H_{i_q}). \end{aligned}$$
(7.16)

Clearly, \(S\subseteq w_p(H_{i_p})\cap w_q(H_{i_q})\). Since \(\mathrm {dim}(w_p(H_{i_p})\cap w_q(H_{i_q}))=n-2\), to prove (7.16), it suffices to show that \(\mathrm {dim}(S)=n-2\). Since \(S\) contains more than \(2^{n-3}\) vertices of \(Q_n\), by Theorem 1.1, we deduce that \(\mathrm {dim}(S)\ge n-2\). But \(S\subseteq w_p(H_{i_p})\cap w_q(H_{i_q})\), so we have \(\mathrm {dim}(S)= n-2\). This proves the claim.

Let \(w=(w_{p})^{-1}\). By (7.16), we see that \(w(S)\) is an affine space in \(E(H_{i_{p}},H_{i_{q}},k)\), contradicting the assumption that \(E(H_{i_{p}},H_{i_{q}},k)\) is empty. This completes the proof. \(\square \)

Using formulas (7.14) and (7.15), we can compute \(|A_{i_1}\cap A_{i_2}\cap \cdots \cap A_{i_m}|\) for \(n=6\), \(k=13,14,15,16\) and \(m\ge 2\). We first consider the case when \(m=2\). Using a Maple program, it can be checked that there are only four pairs for which \(E(H_i,H_j,k)\) is nonempty. Recall that for \(t\le n\), \(H_n^t\) denotes the hyperplane \(x_1+\cdots +x_t= \left\lfloor t/2\right\rfloor \) in \(\mathbb {R}^n\).

Case 1: \((H_6^1, H_6^2)\). In this case, it can be easily checked that the affine subspaces in \(E(H_6^1, H_6^2,k)\) form two equivalence classes under the symmetries in \(B_n\). The representatives can be chosen as \(H_6^1\cap H_6^2\) and \(H_6^1\cap w(H_6^2)\), where \(w=(1,3,2)(4)(5)(6)\). Notice that \(w(H_6^2)\) is the hyperplane \(x_2+x_3=1\). So we have

$$\begin{aligned} E_1(H_6^1, H_6^2,k)=\big \{H_6^1\cap H_6^2,\ H_6^1\cap H_6^3\big \}. \end{aligned}$$
(7.17)

Case 2: \((H_6^1,H_6^3)\). In this case, the affine subspaces in \(E(H_6^1, H_6^3,k)\) form only one equivalence class under the symmetries in \(B_n\). The representative can be chosen as \(H_6^1\cap H_6^3\), and hence

$$\begin{aligned} E_1(H_6^1, H_6^3,k)=\big \{H_6^1\cap H_6^3\big \}. \end{aligned}$$
(7.18)

Case 3: \((H_6^2,H_6^3)\). This case is similar to Case 2. We have

$$\begin{aligned} E_1(H_6^2, H_6^3,k)=\big \{H_6^1\cap H_6^3\big \}. \end{aligned}$$
(7.19)

Case 4: \((H_6^2, H_6^4)\). In this case, it can be verified that

$$\begin{aligned} E_1(H_6^2, H_6^4,k)=\big \{H_6^2\cap H_6^4\big \}. \end{aligned}$$
(7.20)

By (7.17)–(7.20), we obtain that for \(n=6\) and \(k=13,14,15,16\),

$$\begin{aligned} \sum _{1\le i<j\le h(6,k)}|A_i\cap A_j|=|\mathcal {P}(H_6^1\cap H_6^2,k)|+3|\mathcal {P}(H_6^1\cap H_6^3,k)|+|\mathcal {P}(H_6^2\cap H_6^4,k)|. \end{aligned}$$
(7.21)

Finally, we compute \(|A_{i_1}\cap A_{i_2}\cap \cdots \cap A_{i_m}|\) for \(n=6\), \(k=13,14,15,16\) and \(m\ge 3\). We claim that \(E(H_{i_1},\ldots ,H_{i_m},k)\) is empty for any \(m\ge 4\). If this is not the case, then, by Theorem 7.8, for any \(1\le p< q\le m\), \(E(H_{i_p},H_{i_q},k)\) is nonempty. Since \(m\ge 4\), there are at least six pairs \((H_i,H_j)\) with \(1\le i<j\le h(6,k)\) for which \(E(H_i,H_j,k)\) is nonempty. However, as shown before, there are only four pairs \((H_i,H_j)\) with \(1\le i<j\le h(6,k)\) for which \(E(H_i,H_j,k)\) is nonempty, leading to a contradiction. So the claim is proved.

When \(m=3\), it is easy to check that \(E(H_{i_1},H_{i_2},H_{i_3},k)\) is nonempty if and only if

$$\begin{aligned} (H_{i_1},H_{i_2},H_{i_3})=(H_6^1, H_6^2,H_6^3). \end{aligned}$$

Moreover, we have

$$\begin{aligned} E_1(H_6^1,H_6^2,H_6^3,k)=\big \{H_6^1\cap H_6^3\big \}. \end{aligned}$$

Thus, for \(n=6\), \(k=13,14,15,16\) and \(m\ge 3\), we have

$$\begin{aligned} \sum _{1\le i_1<\cdots <i_m \le h(6,k)}|A_{i_1}\cap \cdots \cap A_{i_m}|= \Big \{\begin{array}{l@{\quad }l} |\mathcal {P}(H_6^1\cap H_6^3,k)| &{} \text{ if }\; m=3,\\ 0 &{} \text{ if }\; m>3.\end{array} \end{aligned}$$
(7.22)

By Lemma 7.1 and formulas (7.13), (7.21) and (7.22), we deduce that for \(n=6\) and \(k=13,14,15,16\),

$$\begin{aligned} H_6(k)=&\sum _{i=1}^6\big [u_1^{k}u_2^{|V_6(H_6^i)|-k}\big ]C_{H_6^i} (u_1,u_2)+\sum _{i=1}^8\big [u_1^{k}u_2^{|V_6(H_i)|-k}\big ]C_{H_i} (u_1,u_2)\nonumber \\&+|\mathcal {P}(H_6^2\cap w(H_6^2),k)|-|\mathcal {P}(H_6^1\cap H_6^2,k)|-2|\mathcal {P}(H_6^1\cap H_6^3,k)|\nonumber \\&-|\mathcal {P}(H_6^2\cap H_6^4,k)| -|\mathcal {L}(H_6^2\cap w(H_6^2),k)|, \end{aligned}$$
(7.23)

where \(w=(1,3)(2,4)(5)(6)\). Notice that for \(w=(1,3)(2,4)(5)(6)\),

$$\begin{aligned} H_6^2\cap w(H_6^2)\!=\!H_6^2\cap H_6^4\!=\!\big \{(x_1,x_2,\ldots ,x_6)\in \mathbb {R}^6\,|\, x_1 \!+\! x_2\!=\!1\ \ \text {and}\ \ x_3 \!+\! x_4 \!=\! 1 \big \}. \end{aligned}$$

Thus, (7.23) can be rewritten as

$$\begin{aligned} H_6(k)=&\sum _{i=1}^6\big [u_1^{k}u_2^{|V_6(H_6^i)|-k}\big ] C_{H_6^i}(u_1,u_2)+\sum _{i=1}^8\big [u_1^{k}u_2^{|V_6(H_i)| -k}\big ]C_{H_i}(u_1,u_2)\nonumber \\&-|\mathcal {P}(H_6^1\cap H_6^2,k)|-2|\mathcal {P}(H_6^1\cap H_6^3,k)| -|\mathcal {L}(H_6^2\cap w(H_6^2),k)|, \end{aligned}$$
(7.24)

where \(w=(1,3)(2,4)(5)(6)\).

As for \(|\mathcal {P}(H_6^1\cap H_6^2,k)|\), we notice that

$$\begin{aligned} H_6^1\cap H_6^2=\{(0,1,x_3,x_4,x_5,x_6)\,|\, x_i=0\;\text{ or } \text{1 } \text{ for }\; i=3,4,5,6\}. \end{aligned}$$

Thus the vertices of \(Q_6\) contained in \(H_6^1\cap H_6^2\) are in one-to-one correspondence with the vertices of \(Q_4\). To be more specific, given a vertex \((0,1,x_3,x_4,x_5,x_6)\) of \(Q_6\) contained in \(H_6^1\cap H_6^2\), we get a vertex \((x_3,x_4,x_5,x_6)\) of \(Q_4\) and vice versa. Moreover, the partial 0/1-equivalence classes of \(H_6^1\cap H_6^2\) are in one-to-one correspondence with the 0/1-equivalence classes of \(Q_4\). Hence, for \(n=6\) and \(k=13,14,15,16\), we have

$$\begin{aligned} |\mathcal {P}(H_6^1\cap H_6^2,k)|=\big [u_1^{k} u_2^{16-k}\big ]C_4(u_1,u_2). \end{aligned}$$
(7.25)

We now compute \(|\mathcal {P}(H_6^1\cap H_6^3,k)|\). Since

$$\begin{aligned} H_6^1\cap H_6^3=\{(0,x_2,x_3,x_4,x_5,x_6)\,|\, x_2+x_3=1\}, \end{aligned}$$

we see that each vertex \((0,x_2,x_3,x_4,x_5,x_6)\) of \(Q_6\) contained in \(H_6^1\cap H_6^3\) corresponds to a vertex \((x_2,x_3,x_4,x_5,x_6)\) of \(Q_5\) contained in the spanned hyperplane \(H_5^2\) of \(Q_5\) and vice versa. Hence the partial 0/1-equivalence classes of \(H_6^1\cap H_6^3\) are in one-to-one correspondence with the partial 0/1-equivalence classes of the spanned hyperplane \(H_5^2\) of \(Q_5\). Therefore, for \(n=6\) and \(k=13,14,15,16\), we have

$$\begin{aligned} |\mathcal {P}(H_6^1\cap H_6^3,k)|=\big [u_1^{k}u_2^{16 -k}\big ]C_{H_5^2}(u_1,u_2). \end{aligned}$$
(7.26)

Finally, we determine \(|\mathcal {L}(H_6^2\cap w(H_6^2),k)|\) for \(w=(1,3)(2,4)(5)(6)\). By (7.7), we see that \(|\mathcal {L}(H_6^2\cap w(H_6^2),k)|\) can be obtained from the cycle index \(Z_{(H_6^2,w)}({z})\). Using the technique in Sect. 5, we obtain that

$$\begin{aligned} Z_{(H_6^2,w)}(z)=\frac{1}{32} \left( z_1^{16}+21z_2^8 +8z_4^4+2z_1^8z_2^4\right) . \end{aligned}$$

Hence

$$\begin{aligned} |\mathcal {L}(H_6^2\cap H_6^4,k)|=\big [u_1^{k}u_2^{16 -k}\big ]C_{(H_6^2,w)}(u_1,u_2), \end{aligned}$$
(7.27)

where \(C_{(H_6^2,w)}(u_1,u_2)\) is the polynomial obtained from \(Z_{(H_6^2,w)}({z})\) by substituting \(z_i\) with \(u_1^i+u_2^i\).

Using (7.24)–(7.27), we can compute \(H_6(k)\) for \(k=13,14,15,16\). Since \(F_6(k)=A_6(k)-H_6(k)\), we obtain \(F_6(k)\) for \(k=13,14,15,16\) as given in Table 7.

Table 7 \(F_6(k)\) for \(k=13,14,15,16\)