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

On Simplicial Complexes with Extremal Total Betti number and Total Bigraded Betti Number

Pimeng Dai*,   Li Yu *School of Mathematics, Nanjing University, Nanjing, 210093, P.R.China [email protected] **School of Mathematics, Nanjing University, Nanjing, 210093, P.R.China [email protected]
Abstract.

We determine which simplicial complexes have the maximum or minimum sum of Betti numbers and sum of bigraded Betti numbers with a given number of vertices in each dimension.

Key words and phrases:
Total Betti number, bigraded Betti number, Sperner family
2020 Mathematics Subject Classification:
05E45, 13F55

1. Introduction

For a finite CW-complex XX and a field 𝔽\mathbb{F}, let

βi(X;𝔽):=dim𝔽Hi(X;𝔽),β~i(X;𝔽):=dim𝔽H~i(X;𝔽)\beta_{i}(X;\mathbb{F}):=\dim_{\mathbb{F}}H_{i}(X;\mathbb{F}),\ \ \ \widetilde{\beta}_{i}(X;\mathbb{F}):=\dim_{\mathbb{F}}\widetilde{H}_{i}(X;\mathbb{F})

where Hi(X;𝔽)H_{i}(X;\mathbb{F}) and H~i(X;𝔽)\widetilde{H}_{i}(X;\mathbb{F}) are the singular and reduced singular homology group of XX with 𝔽\mathbb{F}-coefficients, respectively. Moreover, define

tb(X;𝔽):=iβi(X;𝔽),tb~(X;𝔽):=iβ~i(X;𝔽).tb(X;\mathbb{F}):=\sum_{i}\beta_{i}(X;\mathbb{F}),\ \ \ \widetilde{tb}(X;\mathbb{F}):=\sum_{i}\widetilde{\beta}_{i}(X;\mathbb{F}).

We call tb(X;𝔽)tb(X;\mathbb{F}) and tb~(X;𝔽)\widetilde{tb}(X;\mathbb{F}) the total Betti number and reduced total Betti number of XX with 𝔽\mathbb{F}-coefficients, respectively. The difference between tb(X;𝔽)tb(X;\mathbb{F}) and tb~(X;𝔽)\widetilde{tb}(X;\mathbb{F}) is just 11. In addition, we call

χ~(X):=i(1)iβ~i(X;𝔽)\widetilde{\chi}(X):=\sum_{i}(-1)^{i}\widetilde{\beta}_{i}(X;\mathbb{F})

the reduced Euler characteristic of XX, which is independent on the coefficient 𝔽\mathbb{F}.

Convention: The coefficient field 𝔽\mathbb{F} will be omitted when there is no ambiguity in the context or the coefficients are not essential. In fact, all the results obtained in this paper are independent on the coefficients.

The total Betti number of a topological space plays an important role in many theories in mathematics. For example, in the Arnold conjecture in symplectic geometry, the total Betti number of a symplectic manifold MM serves as the lower bound of the number of fixed points of any Hamiltonian symplectomorphism of MM; see Abbondandolo [1, Chapter 6] for more information. Another example is from the Smith theory of cyclic group actions where the Smith inequality says that for a prime pp and a p\mathbb{Z}_{p}-action on a finite CW-complex XX, the fixed point set XpX^{\mathbb{Z}_{p}} must satisfy tb(Xp;p)tb(X;p)tb(X^{\mathbb{Z}_{p}};\mathbb{Z}_{p})\leqslant tb(X;\mathbb{Z}_{p}); see Allday and Puppe [2, Chapter 1]. In addition, an important conjecture in equivariant topology and rational homotopy theory due to Halperin [12] states that if a kk-dimensional torus Tk=(S1)kT^{k}=(S^{1})^{k} can act almost freely on a finite-dimensional topological space XX, then tb(X;)2ktb(X;\mathbb{Q})\geqslant 2^{k}. A parallel conjecture due to Carlsson [7] states that for a prime pp, if a pp-torus (p)k(\mathbb{Z}_{p})^{k} can act freely on a finite CW-complex XX, then tb(X;p)2ktb(X;\mathbb{Z}_{p})\geqslant 2^{k}.

In this paper, we focus our study on the total Betti number and total bigraded Betti number (see Definition 1.4) of a simplicial complex with a fixed number of vertices. Let KK be a simplicial complex whose vertex set is

Ver(K)=[m]={1,2,,m}.\mathrm{Ver}(K)=[m]=\{1,2,\cdots,m\}.

Each simplex σ\sigma of KK is considered as a subset of [m][m]. Let StarKσ\mathrm{Star}_{K}\sigma and LinkKσ\mathrm{Link}_{K}\sigma denote the star and the link of σ\sigma in KK, respectively. For any subset J[m]J\subseteq[m], let

K|J=the full subcomplex of K obtained by restricting toJ.K|_{J}=\ \text{the \emph{full subcomplex} of $K$ obtained by restricting to}\ J.

In particular, when J=J=\varnothing, K|J=K|_{J}=\varnothing and define

βi()=0,i0;β~i()={1,if i=1 ;0,otherwise.\beta_{i}(\varnothing)=0,\ \forall i\geqslant 0;\ \ \ \widetilde{\beta}_{i}(\varnothing)=\begin{cases}1,&\text{if $i=-1$ };\\ 0,&\text{otherwise}.\end{cases}

Question 1: For a positive integer mm, which simplicial complexes have the maximum (reduced) total Betti number among all the simplicial complexes with mm vertices? When m=1,2,3m=1,2,3, the answer is just the discrete mm points. When m=4m=4, the answer is either the discrete 44 points or the complete graph on 44 vertices. A complete answer to Question 1 is contained in the following theorem.

Theorem 1.1 (Björner and Kalai [4, Theorem 1.4]).

Suppose KK is a simplicial complex with at most n+1n+1 vertices. Then |χ~(K)|tb~(K)(n[n/2])|\widetilde{\chi}(K)|\leqslant\widetilde{tb}(K)\leqslant\binom{n}{[n/2]}. Moreover, the following conditions are equivalent:

  • (i)

    |χ~(K)|=(n[n/2])|\widetilde{\chi}(K)|=\binom{n}{[n/2]},

  • (ii)

    tb~(K)=(n[n/2])\widetilde{tb}(K)=\binom{n}{[n/2]},

  • (iii)

    KK is the kk-skeleton of an nn-simplex, where k=n/21k=n/2-1 if nn is even and k=(n1)/2k=(n-1)/2 or k=(n3)/2k=(n-3)/2 if nn is odd.

Note that Theorem 1.1 not only gives the answer to Question 1, but also reveal the connection between tb~(K)\widetilde{tb}(K) and the Euler characteristic of KK. The proof of Theorem 1.1 in [4] uses a nontrivial operation called algebraic shifting of a simplicial complex and a well-known theorem of Sperner in [18]. But if we only want to answer Question 1, we can use some easier argument as shown in Section 2.

Moreover, we can ask a more subtle question than Question 1 as follows.

Question 2: For each 0d<m0\leqslant d<m, which dd-dimensional simplicial complexes with mm vertices have the maximum (reduced) total Betti number among all the dd-dimensional simplicial complexes with mm vertices?

For a pair of integers (m,d)(m,d), 0d<m0\leqslant d<m, we introduce the following notations:

  • Let Σ(m)\Sigma(m) denote the set of all simplicial complexes with vertex set [m][m].

  • Let Σ(m,d)\Sigma(m,d) denote the set of all dd-dimensional simplicial complexes with vertex set [m][m].

Moreover, define (with respect to a fixed coefficient field 𝔽\mathbb{F})

Σtb(m)={KΣ(m)tb~(K)=maxLΣ(m)tb~(L)}Σ(m).\Sigma^{tb}(m)=\Big{\{}K\in\Sigma(m)\mid\widetilde{tb}(K)=\max_{L\in\Sigma(m)}\widetilde{tb}(L)\Big{\}}\subseteq\Sigma(m).
Σtb(m,d)={KΣ(m,d)tb~(K)=maxLΣ(m,d)tb~(L)}Σ(m,d).\Sigma^{tb}(m,d)=\Big{\{}K\in\Sigma(m,d)\mid\widetilde{tb}(K)=\max_{L\in\Sigma(m,d)}\widetilde{tb}(L)\Big{\}}\subseteq\Sigma(m,d).

We consider Σ(m)\Sigma(m), Σ(m,d)\Sigma(m,d), Σtb(m)\Sigma^{tb}(m) and Σtb(m,d)\Sigma^{tb}(m,d) as partial ordered sets with respect to the inclusion of simplicial complexes. Clearly, if KK is an (maximal or minimal) element of Σtb(m)\Sigma^{tb}(m), then KK is also an (maximal or minimal) element of Σtb(m,dim(K))\Sigma^{tb}(m,\dim(K)).

In addition, for any m1m\geqslant 1, we use Δ[m]\Delta^{[m]} to denote the (m1)(m-1)-dimensional simplex with vertex set [m][m]. So the boundary Δ[m]\partial\Delta^{[m]} of Δ[m]\Delta^{[m]} is a simplicial sphere of dimension m2m-2. Moreover, for any 0k<d<m0\leqslant k<d<m,

  • let Δ(k)[m]\Delta^{[m]}_{(k)} denote the kk-skeleton of Δ[m]\Delta^{[m]};

  • let Δ(k)[m]d\Delta_{(k)}^{[m]}\langle d\rangle denote the minimal dd-dimensional subcomplex of Δ[m]\Delta^{[m]} that contains Δ(k)[m]\Delta_{(k)}^{[m]}, which is unique up to simplicial isomorphism. Indeed, we can think of Δ(k)[m]d\Delta_{(k)}^{[m]}\langle d\rangle as the union of Δ(k)[m]\Delta_{(k)}^{[m]} with the dd-simplex Δ[d+1]\Delta^{[d+1]} in Δ[m]\Delta^{[m]} where [d+1]={1,,d+1}[m][d+1]=\{1,\cdots,d+1\}\subseteq[m].

We will prove the following theorem in Section 2 which answers Question 2.

Theorem 1.2 (see Theorem 2.12).

The sets Σtb(m,d)\Sigma^{tb}(m,d) are classified as follows:

  • (i)

    If d[m2]1d\leqslant\left[\frac{m}{2}\right]-1 or d=m1d=m-1, then Σtb(m,d)={Δ(d)[m]}\Sigma^{tb}(m,d)=\Big{\{}\Delta_{(d)}^{[m]}\Big{\}};

  • (ii)

    If [m2]dm3\left[\frac{m}{2}\right]\leqslant d\leqslant m-3, then Σtb(m,d)={Δ([m2]1)[m]d}\Sigma^{tb}(m,d)=\Big{\{}\Delta_{\left(\left[\frac{m}{2}\right]-1\right)}^{[m]}\langle d\rangle\Big{\}};

  • (iii)

    If d=m2d=m-2,

    • when mm is odd, Σtb(m,d)={Δ([m2]1)[m]d,Δ([m2])[m]d}\Sigma^{tb}(m,d)=\Big{\{}\Delta_{\left(\left[\frac{m}{2}\right]-1\right)}^{[m]}\langle d\rangle,\Delta_{\left(\left[\frac{m}{2}\right]\right)}^{[m]}\langle d\rangle\Big{\}};

    • when mm is even, Σtb(m,d)={Δ([m2]1)[m]d}\Sigma^{tb}(m,d)=\Big{\{}\Delta_{\left(\left[\frac{m}{2}\right]-1\right)}^{[m]}\langle d\rangle\Big{\}}.

By the above Theorem, the set Σtb(m,d)\Sigma^{tb}(m,d) is independent on the coefficients 𝔽\mathbb{F}.

Remark 1.3.

It is also meaningful to ask what kind of simplicial complexes with mm vertices have the minimum total Betti number. But the answer is just all the acyclic simplicial complexes which are numerous; see Kalai [16] for more discussion on this subject.

There is another family of integers associated to a simplicial complex KK called bigraded Betti numbers, which are derived from the Stanley-Reisner ring of KK. Recall the Stanley-Reisner ring of KK over 𝔽\mathbb{F} (see Stanley [19]) is

𝔽[K]=𝔽[v1,,vm]/K\mathbb{F}[K]=\mathbb{F}[v_{1},\cdots,v_{m}]/\mathcal{I}_{K}

where K\mathcal{I}_{K} is the ideal in the polynomial ring 𝔽[v1,,vm]\mathbb{F}[v_{1},\cdots,v_{m}] generated by all the square-free monomials vi1visv_{i_{1}}\cdots v_{i_{s}} where {i1,,is}\{i_{1},\cdots,i_{s}\} is not a simplex of KK. Since 𝔽[K]\mathbb{F}[K] is naturally a module over 𝔽[v1,,vm]\mathbb{F}[v_{1},\cdots,v_{m}], by the standard construction in homological algebra, we obtain a canonical algebra Tor𝔽[v1,,vm](𝔽[K],𝔽)\mathrm{Tor}_{\mathbb{F}[v_{1},\cdots,v_{m}]}(\mathbb{F}[K],\mathbb{F}) from 𝔽[K]\mathbb{F}[K], where 𝔽\mathbb{F} is considered as the trivial 𝔽[v1,,vm]\mathbb{F}[v_{1},\cdots,v_{m}]-module. Moreover, there is a bigraded 𝔽[v1,,vm]\mathbb{F}[v_{1},\cdots,v_{m}]-module structure on Tor𝔽[v1,,vm](𝔽[K],𝔽)\mathrm{Tor}_{\mathbb{F}[v_{1},\cdots,v_{m}]}(\mathbb{F}[K],\mathbb{F}) (see [19]):

Tor𝔽[v1,,vm](𝔽[K],𝔽)=i,j0Tor𝔽[v1,,vm]i,2j(𝔽[K],𝔽)\mathrm{Tor}_{\mathbb{F}[v_{1},\cdots,v_{m}]}(\mathbb{F}[K],\mathbb{F})=\bigoplus_{i,j\geqslant 0}\mathrm{Tor}^{-i,2j}_{\mathbb{F}[v_{1},\cdots,v_{m}]}(\mathbb{F}[K],\mathbb{F})

where deg(vi)=2\mathrm{deg}(v_{i})=2 for each 1im1\leqslant i\leqslant m. The integers

βi,2j(𝔽(K)):=dim𝔽Tor𝔽[v1,,vm]i,2j(𝔽[K],𝔽)\beta^{-i,2j}(\mathbb{F}(K)):=\dim_{\mathbb{F}}\mathrm{Tor}^{-i,2j}_{\mathbb{F}[v_{1},\cdots,v_{m}]}(\mathbb{F}[K],\mathbb{F})

are called the bigraded Betti numbers of KK with 𝔽\mathbb{F}-coefficients. Note that unlike Betti numbers of KK, bigraded Betti numbers are not topological invariants, but only combinatorial invariants of KK in general.

Definition 1.4.

The total bigraded Betti number of KK with 𝔽\mathbb{F}-coefficients is

D~(K;𝔽)=i,jβi,2j(𝔽(K))=dim𝔽Tor𝔽[v1,,vm](𝔽[K],𝔽).\widetilde{D}(K;\mathbb{F})=\sum_{i,j}\beta^{-i,2j}(\mathbb{F}(K))=\dim_{\mathbb{F}}\mathrm{Tor}_{\mathbb{F}[v_{1},\cdots,v_{m}]}(\mathbb{F}[K],\mathbb{F}).

By Hochster’s formula (see Hochster [14] or [19, Theorem 4.8]), the bigraded Betti numbers of KK can also be computed from the homology groups of the full subcomplexes of KK:

(1) βi,2j(𝔽(K))=J[m],|J|=jdim𝔽H~ji1(K|J;𝔽).\beta^{-i,2j}(\mathbb{F}(K))=\sum_{J\subseteq[m],|J|=j}\dim_{\mathbb{F}}\widetilde{H}_{j-i-1}(K|_{J};\mathbb{F}).

So we can also express D~(K;𝔽)\widetilde{D}(K;\mathbb{F}) as

(2) D~(K;𝔽)=J[m]tb~(K|J;𝔽).\widetilde{D}(K;\mathbb{F})=\sum_{J\subseteq[m]}\widetilde{tb}(K|_{J};\mathbb{F}).

This explains why we put “~\,\widetilde{\ }\,” in the notation D~(K;𝔽)\widetilde{D}(K;\mathbb{F}).

Question 3: For each 0d<m0\leqslant d<m, which dd-dimensional simplicial complexes with mm vertices have the minimum total bigraded Betti number over a field 𝔽\mathbb{F} among all the dd-dimensional simplicial complexes with mm vertices? We readily call such kind of simplicial complexes D~\widetilde{D}-minimal (over 𝔽\mathbb{F}).

It follows from Cao and Lü [6, Theorem 1.4] or Ustinovsky [20, Theorem 3.2] that there is a universal lower bound of D~(K;𝔽)\widetilde{D}(K;\mathbb{F}) (see (11)):

D~(K;𝔽)2md1,KΣ(m,d).\widetilde{D}(K;\mathbb{F})\geqslant 2^{m-d-1},\,K\in\Sigma(m,d).
Definition 1.5 (Tight Simplicial Complex).

Let KK be a dd-dimensional simplicial complex with mm vertices. We call KK tight (over 𝔽\mathbb{F}) if D~(K;𝔽)=2md1\widetilde{D}(K;\mathbb{F})=2^{m-d-1}.

We classify all the tight simplicial complexes in the following theorem.

Theorem 1.6 (see Theorem 3.10).

A finite simplicial complex KK is tight if and only if KK is of the form Δ[n1]Δ[nk]\partial\Delta^{[n_{1}]}*\cdots*\partial\Delta^{[n_{k}]} or Δ[r]Δ[n1]Δ[nk]\Delta^{[r]}*\partial\Delta^{[n_{1}]}*\cdots*\partial\Delta^{[n_{k}]} for some positive integers n1,,nkn_{1},\cdots,n_{k} and rr.

It is a convention to let Δ[1]=\partial\Delta^{[1]}=\varnothing and K=KK*\varnothing=K. Theorem 1.6 implies that the tightness of a simplicial complex is independent on the coefficient field 𝔽\mathbb{F}.

From Theorem 1.6, we can easily deduce that if KΣ(m,d)K\in\Sigma(m,d) is tight, it is necessary that [m12]d\left[\frac{m-1}{2}\right]\leqslant d. In particular, the equality [m12]=d\left[\frac{m-1}{2}\right]=d is achieved by Δ[2]Δ[2]Δ[2]\partial\Delta^{[2]}*\partial\Delta^{[2]}*\cdots*\partial\Delta^{[2]} when mm is even and by Δ[1]Δ[2]Δ[2]Δ[2]\Delta^{[1]}*\partial\Delta^{[2]}*\partial\Delta^{[2]}*\cdots*\partial\Delta^{[2]} when mm is odd. Conversely, for any (m,d)(m,d) with [m12]dm1\left[\frac{m-1}{2}\right]\leqslant d\leqslant m-1, there always exists a tight simplicial complex KΣ(m,d)K\in\Sigma(m,d). So if [m12]dm1\left[\frac{m-1}{2}\right]\leqslant d\leqslant m-1, the D~\widetilde{D}-minimal simplicial complexes in Σ(m,d)\Sigma(m,d) are exactly all the tight simplicial complexes.

But when [m12]>d\left[\frac{m-1}{2}\right]>d, a D~\widetilde{D}-minimal simplicial complex in Σ(m,d)\Sigma(m,d) is never tight, and it seems to us that there is no very good way to describe D~\widetilde{D}-minimal simplicial complexes in general. One reason is that the full subcomplexes of a D~\widetilde{D}-minimal simplicial complex may not be D~\widetilde{D}-minimal. For example, by exhausting all the 3333 members of Σ(5,1)\Sigma(5,1), we find that all the D~\widetilde{D}-minimal simplicial complexes in Σ(5,1)\Sigma(5,1) are K2,3K_{2,3} and C5C_{5} (see Figure 1) whose D~\widetilde{D}-value is 1212. Note that none of the full subcomplexes of C5C_{5} on four vertices are D~\widetilde{D}-minimal.

Refer to caption
Figure 1. D~\widetilde{D}-minimal 11-dimensional simplicial complexes with 55 vertices

This example suggests that we may not be able to inductively construct all the D~\widetilde{D}-minimal simplicial complexes with mm vertices from the D~\widetilde{D}-minimal ones with m1m-1 vertices. In addition, it is not clear how to compute the minimal value of D~()\widetilde{D}(\cdot) on Σ(m,d)\Sigma(m,d) when [m12]>d\left[\frac{m-1}{2}\right]>d except exhausting all the members.

Moreover, we can ask the following question parallel to Question 1 for bigraded Betti numbers.

Question 4: For a positive integer mm, which simplicial complexes with mm vertices have the maximum total bigraded Betti numbers among all the simplicial complexes with mm vertices?

In Section 4, we prove the following theorem which answers Question 4. Let

(3) g(m,d)=j=d+1m(mj)(j1d), 0d<m.g(m,d)=\sum_{j=d+1}^{m}\binom{m}{j}\binom{j-1}{d},\ 0\leqslant d<m.
Theorem 1.7 (Theorem 4.2).

If KK is a simplicial complex with mm vertices, then

D~(K;𝔽)g(m,[m13])+1\widetilde{D}(K;\mathbb{F})\leqslant g\left(m,\left[\frac{m-1}{3}\right]\right)+1

for any field 𝔽\mathbb{F}, where the equality holds if and only if K=Δ([m13]1)[m]K=\Delta_{\left(\left[\frac{m-1}{3}\right]-1\right)}^{[m]}.

Remark 1.8.

Parallelly to Question 2, we can also ask for each 0d<m0\leqslant d<m, which simplicial complexes KΣ(m,d)K\in\Sigma(m,d) have the maximum total bigraded Betti number over a field 𝔽\mathbb{F} among all the members of Σ(m,d)\Sigma(m,d). But we do not know the complete answer of this question yet. The question can be turned into a very complicated combinatorial extremum problem.

The paper is organized as follows. In Section 2, we first write an easy argument using Mayer-Vietoris sequence to answer Question 1. Then we use the theory of algebraic shifting of simplicial complexes from [4] and some results on Sperner families from [17] and [11] to give a complete answer to Question 2. In Section 3, we classify all the tight simplicial complexes using some results from [20] and [21]. In Section 4, we give a complete answer to Question 4 via some combinatorial inequality proved in the appendix.


2. Simplicial complexes with the maximum total Betti number

In this section, we first give an alternative proof of Theorem 2.2 which answers Question 1. Our approach is different from [4].

Lemma 2.1.

For any two subcomplexes K,LK,L of Δ[m]\Delta^{[m]}, one always have

(4) tb~(K)+tb~(L)tb~(KL)+tb~(KL).\widetilde{tb}(K)+\widetilde{tb}(L)\leqslant\widetilde{tb}(K\cap L)+\widetilde{tb}(K\cup L).
Proof.

By the following Mayer-Vietoris sequence for (K,L)(K,L) :

H~j+1(KL)dj+1H~j(KL)H~j(K)H~j(L)H~j(KL)djH~j1(KL),\cdots\widetilde{H}_{j+1}(K\cup L)\overset{d_{j+1}}{\longrightarrow}\widetilde{H}_{j}(K\cap L)\longrightarrow\widetilde{H}_{j}(K)\oplus\widetilde{H}_{j}(L)\longrightarrow\widetilde{H}_{j}(K\cup L)\overset{d_{j}}{\longrightarrow}\widetilde{H}_{j-1}(K\cap L)\cdots,

we obtain:

β~j(K)+β~j(L)β~j(KL)+β~j(KL),j0.\widetilde{\beta}_{j}(K)+\widetilde{\beta}_{j}(L)\leqslant\widetilde{\beta}_{j}(K\cap L)+\widetilde{\beta}_{j}(K\cup L),\ \forall j\geqslant 0.

Then sum it over all j0j\geqslant 0, we get the desired inequality. ∎

Note that the equality in (4) holds if and only if Im(dj)=0\mathrm{Im}(d_{j})=0 for all j0j\geqslant 0 in the above Mayer-Vietoris sequence.

Theorem 2.2.

Let KK be a simplicial complex with mm vertices. Then

tb~(K)(m1[m12]).\widetilde{tb}(K)\leqslant\binom{m-1}{\left[\frac{m-1}{2}\right]}.

Moreover, tb~(K)=(m1[m12])\widetilde{tb}(K)=\binom{m-1}{\left[\frac{m-1}{2}\right]} if and only if KK is isomorphic to the kk-skeleton of an (m1)(m-1)-dimensional simplex, where k=m32k=\frac{m-3}{2} if mm is odd and k=m21k=\frac{m}{2}-1 or k=m22k=\frac{m}{2}-2 if mm is even.

Proof.

Suppose the vertex set of KK is [m]={1,,m}[m]=\{1,\cdots,m\} and consider KK as a subcomplex of Δ[m]\Delta^{[m]}. In the rest of the proof, we assume that m3m\geqslant 3.

Note that each permutation ϕ\phi of [m][m] determines a simplicial isomorphism Δ[m]Δ[m]\Delta^{[m]}\rightarrow\Delta^{[m]}, still denoted by ϕ\phi. Let ϕK\phi K denote the image of KK under ϕ\phi. Then by Lemma 2.1,

(5) tb~(K)+tb~(ϕK)tb~(KϕK)+tb~(KϕK).\widetilde{tb}(K)+\widetilde{tb}(\phi K)\leqslant\widetilde{tb}(K\cap\phi K)+\widetilde{tb}(K\cup\phi K).

Since ϕK\phi K is homeomorphic to KK, if KK is an element of Σtb(m)\Sigma^{tb}(m), then both KϕKK\cap\phi K and KϕKK\cup\phi K must also belong to Σtb(m)\Sigma^{tb}(m) since their vertex sets are also [m][m]. So in particular, if KK is a maximal (or minimal) element of Σtb(m)\Sigma^{tb}(m), we must have K=KϕKK=K\cup\phi K (or K=KϕKK=K\cap\phi K) for every permutation ϕ\phi of [m][m], which implies K=ϕKK=\phi K. Then KK is the skeleton Δ(dim(K))[m]\Delta^{[m]}_{(\dim(K))}. An easy calculation shows

tb~(Δ(d)[m])=(m1d+1).\widetilde{tb}\big{(}\Delta^{[m]}_{(d)}\big{)}=\binom{m-1}{d+1}.

So tb~(Δ(d)[m])\widetilde{tb}\big{(}\Delta^{[m]}_{(d)}\big{)} reaches the maximum (m1[m12])\binom{m-1}{\left[\frac{m-1}{2}\right]} when and only when

d={m32,if m is odd;m21orm22,if m is even.d=\begin{cases}\frac{m-3}{2},&\text{if $m$ is odd};\\ \frac{m}{2}-1\ \text{or}\ \frac{m}{2}-2,&\text{if $m$ is even}.\end{cases}

It remains to prove that there are no other elements in Σtb(m)\Sigma^{tb}(m) except the skeleta of Δ[m]\Delta^{[m]} described in the theorem. Let KK be an arbitrary element of Σtb(m)\Sigma^{tb}(m).

When mm is odd, since both the maximal and the minimal elements of Σtb(m)\Sigma^{tb}(m) are Δ(m32)[m]\Delta^{[m]}_{(\frac{m-3}{2})}, KK can only be Δ(m32)[m]\Delta^{[m]}_{(\frac{m-3}{2})}.

When mm is even, the maximal element of Σtb(m)\Sigma^{tb}(m) is Δ(m21)[m]\Delta^{[m]}_{(\frac{m}{2}-1)} while the minimal element is Δ(m22)[m]\Delta^{[m]}_{(\frac{m}{2}-2)}. So dim(K)\dim(K) can only be m21\frac{m}{2}-1 or m22\frac{m}{2}-2. If dim(K)=m22\dim(K)=\frac{m}{2}-2, then Δ(m22)[m]K\Delta^{[m]}_{(\frac{m}{2}-2)}\subseteq K implies K=Δ(m22)[m]K=\Delta^{[m]}_{(\frac{m}{2}-2)}. But if dim(K)=m21\dim(K)=\frac{m}{2}-1, the situation is a bit complicated.

In the following, suppose KK is a minimal element of Σtb(m,m21)\Sigma^{tb}(m,\frac{m}{2}-1). If KK has only one (m21)(\frac{m}{2}-1)-simplex, then clearly tb~(K)\widetilde{tb}(K) is less than tb~(Δ(m22)[m])\widetilde{tb}\big{(}\Delta^{[m]}_{(\frac{m}{2}-2)}\big{)}, which contradicts our assumption that KΣtb(m)K\in\Sigma^{tb}(m). So KK must have at least two (m21)(\frac{m}{2}-1)-simplices. We have the following two cases.

  • Case 1:

    If all the (m21)(\frac{m}{2}-1)-simplices of KK are disjoint, then KK could only have two (m21)(\frac{m}{2}-1)-simplices. But it is easy to check that tb~(K)\widetilde{tb}(K) is less than tb~(Δ(m22)[m])\widetilde{tb}\big{(}\Delta^{[m]}_{(\frac{m}{2}-2)}\big{)}, which again contradicts the assumption KΣtb(m)K\in\Sigma^{tb}(m).

  • Case 2:

    Suppose σ\sigma and τ\tau are two (m21)(\frac{m}{2}-1)-simplices in KK which share exactly ss vertices where 1sm211\leqslant s\leqslant\frac{m}{2}-1. Without loss of generality, we can assume     σ={1,,m2},τ={m2s+1,,m2,,ms}\sigma=\{1,\cdots,\frac{m}{2}\},\ \tau=\{\frac{m}{2}-s+1,\cdots,\frac{m}{2},\cdots,m-s\}.

    Let ϕ¯\overline{\phi} be an arbitrary permutation of [m][m] that preserves σ\sigma. Then Kϕ¯KK\cap\overline{\phi}K contains σ\sigma, and hence dim(Kϕ¯K)=m21\dim(K\cap\overline{\phi}K)=\frac{m}{2}-1. But by the minimality of KK in Σtb(m,m21)\Sigma^{tb}(m,\frac{m}{2}-1), we must have ϕ¯K=K\overline{\phi}K=K. In particular, ϕ¯(τ)\overline{\phi}(\tau) belongs to KK. This implies that any (m21)(\frac{m}{2}-1)-simplex in Δ[m]\Delta^{[m]} that shares exactly ss vertices with σ\sigma must also belong to KK. This is because the permutation ϕ¯\overline{\phi} can send the face στ\sigma\cap\tau to any other face of σ\sigma with ss vertices while mapping the face τ\σ\tau\backslash\sigma of τ\tau to any other simplex with m2s\frac{m}{2}-s vertices in the complement of σ\sigma in Δ[m]\Delta^{[m]}. So ϕ¯(τ)\overline{\phi}(\tau) can exhaust all the (m21)(\frac{m}{2}-1)-simplices in Δ[m]\Delta^{[m]} that share exactly ss vertices with σ\sigma. Moreover, we can similarly prove that any (m21)(\frac{m}{2}-1)-simplex in Δ[m]\Delta^{[m]} that shares exactly ss vertices with τ\tau must also belong to KK.

    Next, we prove that there exists a (m21)(\frac{m}{2}-1)-simplex τ\tau^{\prime} in KK which shares a (m22)(\frac{m}{2}-2)-face with σ\sigma. Indeed, we can just trade the m2s\frac{m}{2}-s vertices τ\σ={m2+1,,ms}\tau\backslash\sigma=\{\frac{m}{2}+1,\cdots,m-s\} with {2,,m2s}{m}\{2,\cdots,\frac{m}{2}-s\}\cup\{m\} and obtain another (m21)(\frac{m}{2}-1)-simplex τ={2,,m2,m}\tau^{\prime}=\{2,\cdots,\frac{m}{2},m\} which satisfies |ττ|=s|\tau^{\prime}\cap\tau|=s. So τ\tau^{\prime} must belong to KK and we have |τσ|=m21|\tau^{\prime}\cap\sigma|=\frac{m}{2}-1.

From the above discussion, we see that if mm is even and KK is a minimal element in Σtb(m,m21)\Sigma^{tb}(m,\frac{m}{2}-1), we can always find a pair of (m21)(\frac{m}{2}-1)-simplices σ\sigma and τ\tau in KK which share a (m22)(\frac{m}{2}-2)-face. Moreover, by the argument in Case 2, any (m21)(\frac{m}{2}-1)-simplex of Δ[m]\Delta^{[m]} that intersects σ\sigma or τ\tau at a (m22)(\frac{m}{2}-2)-face must belong to KK. So without loss of generality, assume            σ={1,,m2},τ={2,,m2+1}\sigma=\{1,\cdots,\frac{m}{2}\},\ \tau=\{2,\cdots,\frac{m}{2}+1\}.

Claim: Every (m21)(\frac{m}{2}-1)-simplex of Δ[m]\Delta^{[m]} must belong to KK, i.e. K=Δ(m21)[m]K=\Delta^{[m]}_{(\frac{m}{2}-1)}.

If a (m21)(\frac{m}{2}-1)-simplex ξ\xi of Δ[m]\Delta^{[m]} is disjoint from σ\sigma, then ξ={m2+1,,m}\xi=\{\frac{m}{2}+1,\cdots,m\}. We can construct a sequence of (m21)(\frac{m}{2}-1)-simplices σ=ξ0\sigma=\xi_{0}, ξ1,,ξm2=ξ\xi_{1},\cdots,\xi_{\frac{m}{2}}=\xi as follows:

ξi={i+1,,m2+i}\xi_{i}=\{i+1,\cdots,\frac{m}{2}+i\}, 1im21\leqslant i\leqslant\frac{m}{2}.

Since ξiξi+1\xi_{i}\cap\xi_{i+1} is a (m22)(\frac{m}{2}-2)-simplex for every 1i<m21\leqslant i<\frac{m}{2}, we can prove that ξ1,,ξn=ξ\xi_{1},\cdots,\xi_{n}=\xi must all belong to KK by iteratively using the argument in Case 2.

Similarly, if ξσ\xi\cap\sigma\neq\varnothing, we can trade the vertices in ξ\σ\xi\backslash\sigma with the vertices in σ\ξ\sigma\backslash\xi one at a time to turn σ\sigma to ξ\xi. This proves the claim.

By the above claim, the minimal element KK of Σtb(m,m21)\Sigma^{tb}(m,\frac{m}{2}-1) is Δ(m21)[m]\Delta^{[m]}_{(\frac{m}{2}-1)} which is also maximal. So Σtb(m,m21)\Sigma^{tb}(m,\frac{m}{2}-1) consists only of Δ(m21)[m]\Delta^{[m]}_{(\frac{m}{2}-1)}. Therefore, when mm is even, Σtb(m)\Sigma^{tb}(m) must either be Δ(m22)[m]\Delta^{[m]}_{(\frac{m}{2}-2)} or Δ(m21)[m]\Delta^{[m]}_{(\frac{m}{2}-1)}. The theorem is proved. ∎

For any finite simplicial complex KK of dimension dd and a field 𝔽\mathbb{F}, one can associate two integer vectors f=(f0,f1,,fd)f=(f_{0},f_{1},\cdots,f_{d}) and β=(β~0,β~1,,β~d)\beta=(\widetilde{\beta}_{0},\widetilde{\beta}_{1},\cdots,\widetilde{\beta}_{d}), where fif_{i} is the number of ii-simplices of KK and β~i\widetilde{\beta}_{i} is the ii-th reduced Betti number (over 𝔽\mathbb{F}) of KK (i=0,1,,d)i=0,1,\cdots,d). A remarkable result proved in [4] tells us that a collection of nonlinear relations along with the linear relation given by the Euler-Poincaré formula completely characterize the integer vectors which can arise as ff and β\beta for a simplicial complex. But from the description in [4] of the range of the ff-vectors and β\beta-vectors of all the simplicial complexes with mm vertices, it is still not clear what the answer of Question 2 should be. Indeed, when we fix the dimension dd of the simplicial complex KK, the upper bound of tb~(K)\widetilde{tb}(K) will depend on dd and become a little complicated (see Corollary 2.13). So to obtain the answer of Question 2, we also need to use the machinery of shifting of simplicial complexes and the theory of Sperner family just as [4] did.

Now, let us recall some basic definitions that are needed for our argument.

Definition 2.3 (Sperner Family).

Let XX be a finite set. A Sperner family of XX is a set \mathcal{F} of subsets of XX that satisfies ABA\nsubseteq B for distinct members of \mathcal{F}. Given a subset YXY\subseteq X, a Sperner family of XX over YY is a Sperner family \mathcal{F} of XX where every member of \mathcal{F} has nonempty intersection with YY.

The following is a fundamental result of Sperner on the size of a Sperner family (see Sperner [18] or Anderson [3, Theorem 1.2.2]).

Theorem 2.4 (Sperner’s Theorem).

Let \mathcal{F} be a Sperner family of subsets of a finite set XX where |X|=n|X|=n. Then ||(n[n/2])|\mathcal{F}|\leqslant\binom{n}{[n/2]}. If nn is even, the only Sperner family consisting of (n[n/2])\binom{n}{[n/2]} subsets of XX is made up of all the n2\frac{n}{2}-subsets of XX. If nn is odd, a Sperner family of size (n[n/2])\binom{n}{[n/2]} consists of either all the 12(n1)\frac{1}{2}(n-1)-subsets or all the 12(n+1)\frac{1}{2}(n+1)-subsets of XX.

A kk-subset of XX means a subset of order kk. Another very useful construction in the study of combinatorics of simplicial complexes is the shifting operation. A shifting operation is a map which assigns to every simplicial complex KK a shifted simplicial complex Δ(K)\Delta(K) with the same ff-vector.

Definition 2.5 (Shifted Complex).

A simplicial complex Γ\Gamma with vertex set [m][m] is called shifted if for every simplex σ={i1,,is}Γ\sigma=\{i_{1},\cdots,i_{s}\}\in\Gamma where i1<<isi_{1}<\cdots<i_{s}, any {i1,,is}\{i^{\prime}_{1},\cdots,i^{\prime}_{s}\} with i1i1,,isisi^{\prime}_{1}\leqslant i_{1},\cdots,i^{\prime}_{s}\leqslant i_{s} and i1<<isi^{\prime}_{1}<\cdots<i^{\prime}_{s} is also a simplex of Γ\Gamma.

A well-known (combinatorial) shifting operation, introduced by Erdös, Ko and Rado [10], has been of great use in extremal set theory. Later, another shifting operation was introduced by Kalai in [15] which preserves both the ff-vector and the β\beta-vector of a simplicial complex.

Theorem 2.6 (see [15] and [4]).

Given a simplicial complex KK on mm vertices and a field 𝔽\mathbb{F}, there exists a canonically defined shifted simplicial complex Δ=Δ(K,𝔽)\Delta=\Delta(K,\mathbb{F}) on [m][m] such that fi(Δ;𝔽)=fi(K;𝔽)f_{i}(\Delta;\mathbb{F})=f_{i}(K;\mathbb{F}) and β~i(Δ;𝔽)=β~i(K;𝔽)\widetilde{\beta}_{i}(\Delta;\mathbb{F})=\widetilde{\beta}_{i}(K;\mathbb{F}) for all i0i\geqslant 0.

Shifted complexes belong to a slightly larger class of simplicial complexes called near-cones (see [4]).

Definition 2.7 (Near-Cone).

A simplicial complex Δ\Delta with vertex set [m][m] is called a near-cone if for every simplex SΔS\in\Delta and 2jm2\leqslant j\leqslant m, if 1S1\notin S and jSj\in S, then (S\j){1}Δ(S\backslash j)\cup\{1\}\in\Delta. For a near-cone Δ\Delta, define

(6) B(Δ)={SΔS{1}Δ}.B(\Delta)=\{S\in\Delta\mid S\cup\{1\}\notin\Delta\}.

A very nice property of a near-cone Δ\Delta is that its total Betti number can be easily computed by counting the number of elements in B(Δ)B(\Delta).

Lemma 2.8 ([4, Lemma 4.2, Theorem 4.3]).

If Δ\Delta is a near-cone on [m][m], then

  • (i)

    B(Δ)B(\Delta) is a Sperner family of {2,,m}\{2,\cdots,m\},

  • (ii)

    every simplex SB(Δ)S\in B(\Delta) is maximal in Δ\Delta,

  • (iii)

    tb~(Δ)=|B(Δ)|\widetilde{tb}(\Delta)=|B(\Delta)|.

The following proposition tells one what kind of Sperner families on [m][m] can be realized by B(Δ)B(\Delta) of a near-cone Δ\Delta.

Proposition 2.9.

Let \mathcal{F} be a Sperner family of [m]\{1}={2,,m}[m]\backslash\{1\}=\{2,\cdots,m\}. Then the following statements are equivalent:

  • (i)

    there exists a dd-dimensional near-cone Δ\Delta with vertex set contained in [m][m], such that B(Δ)=;B(\Delta)=\mathcal{F};

  • (ii)

    there exists a subset {i1,,id}{2,,m}\{i_{1},\cdots,i_{d}\}\subseteq\{2,\cdots,m\} such that no subset of {i1,,id}\{i_{1},\cdots,i_{d}\} belongs to \mathcal{F} and the order of each member of \mathcal{F} is no greater than d+1d+1.

Proof.

(i)\Rightarrow(ii). First of all, since Δ\Delta is a dd-dimensional near-cone, it has at least one dd-simplex, say {1,i1,,id}\{1,i_{1},\cdots,i_{d}\}. Then by the definition of B(Δ)B(\Delta) in (6), no subset of {i1,,id}\{i_{1},\cdots,i_{d}\} belongs to =B(Δ)\mathcal{F}=B(\Delta) since {1,i1,,id}\{1,i_{1},\cdots,i_{d}\} is a face of Δ\Delta. Moreover, the order of each member of \mathcal{F} is no greater than d+1d+1 since Δ\Delta is dd-dimensional. (ii)\Rightarrow(i). From the Sperner family \mathcal{F}, we define

(7) Δ=[1(E())]Δ{1,i1,,id},\Delta=\left[1*\left(E(\mathcal{F})\setminus\mathcal{F}\right)\right]\cup\mathcal{F}\cup\Delta^{\{1,i_{1},\cdots,i_{d}\}},

where E()E(\mathcal{F}) is the simplicial complex generated by \mathcal{F}, that is, the minimal simplicial complex taking all the members of \mathcal{F} as its maximal faces. Then Δ\Delta is clearly a dd-dimensional simplicial complex whose vertex set is a subset of [m][m]. Moreover, for any FF\in\mathcal{F} and jFj\in F, we have (Fj)11(E())(F\setminus j)\cup 1\in 1*\left(E(\mathcal{F})\setminus\mathcal{F}\right), so Δ\Delta is a near-cone. Finally, it is obvious that B(Δ)=B(\Delta)=\mathcal{F}. ∎

Definition 2.10.

Let XX be a finite set with order |X|=n|X|=n. For a nonempty subset YY of XX, let C(n,X,Y)C(n,X,Y) denote the set of all subsets of XX that have nonempty intersection with YY. For any i1i\geqslant 1, let Ci(n,X,Y)C_{i}(n,X,Y) denote the collection of sets in C(n,X,Y)C(n,X,Y) of size ii.

The following lemma combines some results from Lih [17] and Griggs [11], where the \lceil\cdot\rceil is the ceiling function.

Lemma 2.11 (see [17, Theorem 2] and [11, Theorem 7.1]).

Let XX be a finite set of order nn. The maximal possible cardinality of a Sperner family \mathcal{F} of XX over a subset YXY\subseteq X with |Y|=k|Y|=k is f(n,k)=(nn/2)(nkn/2)f(n,k)=\binom{n}{\lceil n/2\rceil}-\binom{n-k}{\lceil n/2\rceil}. Moreover, ||=f(n,k)|\mathcal{F}|=f(n,k) if and only if \mathcal{F} is one of the following cases:

  1. (a)

    C12n(n,X,Y)C_{\left\lceil\frac{1}{2}n\right\rceil}(n,X,Y);

  2. (b)

    C12(n1)(n,X,Y)C_{\frac{1}{2}(n-1)}(n,X,Y), for odd nn and k12(n+3)k\geqslant\frac{1}{2}(n+3);

  3. (c)

    C12(n+2)(n,X,Y)C_{\frac{1}{2}(n+2)}(n,X,Y), for even nn and k=1k=1.

In particular, if ||=f(n,k)|\mathcal{F}|=f(n,k), then every member in \mathcal{F} has the same order.

Now, we are ready to prove the following theorem which answers Question 2.

Theorem 2.12.

The sets Σtb(m,d)\Sigma^{tb}(m,d) are classified as follows:

  • (i)

    If d[m2]1d\leqslant\left[\frac{m}{2}\right]-1 or d=m1d=m-1, then Σtb(m,d)={Δ(d)[m]}\Sigma^{tb}(m,d)=\Big{\{}\Delta_{(d)}^{[m]}\Big{\}};

  • (ii)

    If [m2]dm3\left[\frac{m}{2}\right]\leqslant d\leqslant m-3, then Σtb(m,d)={Δ([m2]1)[m]d}\Sigma^{tb}(m,d)=\Big{\{}\Delta_{\left(\left[\frac{m}{2}\right]-1\right)}^{[m]}\langle d\rangle\Big{\}};

  • (iii)

    If d=m2d=m-2,

    • when mm is odd, Σtb(m,d)={Δ([m2]1)[m]d,Δ([m2])[m]d}\Sigma^{tb}(m,d)=\Big{\{}\Delta_{\left(\left[\frac{m}{2}\right]-1\right)}^{[m]}\langle d\rangle,\Delta_{\left(\left[\frac{m}{2}\right]\right)}^{[m]}\langle d\rangle\Big{\}};

    • when mm is even, Σtb(m,d)={Δ([m2]1)[m]d}\Sigma^{tb}(m,d)=\Big{\{}\Delta_{\left(\left[\frac{m}{2}\right]-1\right)}^{[m]}\langle d\rangle\Big{\}}.

Proof.

By the algebraic shifting construction in [4, Theorem 3.1], there is a unique shifted complex Δ\Delta associated to KΣ(m,d)K\in\Sigma(m,d) where Δ\Delta has the same ff-vector and β\beta-vector as KK. Moreover, Δ\Delta is a near-cone. So by Lemma 2.8,

tb~(K)=tb~(Δ)=|B(Δ)|,\widetilde{tb}(K)=\widetilde{tb}(\Delta)=|B(\Delta)|,

where B(Δ)B(\Delta) is a Sperner family of {2,,m}\{2,\cdots,m\}. Moreover, since Δ\Delta is shifted and dim(Δ)=dim(K)=d\dim(\Delta)=\dim(K)=d, we can assume that {1,,d+1}\{1,\cdots,d+1\} is a dd-simplex of Δ\Delta.

(i)  The case d=m1d=m-1 is trivial. We do induction on d[m2]1d\leqslant\left[\frac{m}{2}\right]-1. The case d=0d=0 is clearly true. Let KK be an arbitrary element of Σtb(m,d)\Sigma^{tb}(m,d). For any permutation ϕ\phi of [m][m], since KϕKΣ(m,d)K\cup\phi K\in\Sigma(m,d), we have

tb~(K)=tb~(ϕK)tb~(KϕK).\widetilde{tb}(K)=\widetilde{tb}(\phi K)\geqslant\widetilde{tb}(K\cup\phi K).

So we can deduce from (5) that tb~(K)tb~(KϕK)\widetilde{tb}(K)\leqslant\widetilde{tb}(K\cap\phi K).

Claim: dim(KϕK)=d\dim(K\cap\phi K)=d, hence KϕKΣtb(m,d)K\cap\phi K\in\Sigma^{tb}(m,d).

Assume that s=dim(KϕK)<ds=\dim(K\cap\phi K)<d. Then since KΣtb(m,d)K\in\Sigma^{tb}(m,d), we have tb~(K)tb~(Δ(d)[m])=(m1d+1)\widetilde{tb}(K)\geqslant\widetilde{tb}(\Delta_{(d)}^{[m]})=\binom{m-1}{d+1}, and hence tb~(KϕK)(m1d+1)\widetilde{tb}(K\cap\phi K)\geqslant\binom{m-1}{d+1}. On the other hand, by our induction hypothesis tb~(KϕK)tb~(Δ(s)[m])=(m1s+1)\widetilde{tb}(K\cap\phi K)\leqslant\widetilde{tb}(\Delta_{(s)}^{[m]})=\binom{m-1}{s+1}.

  • When mm is odd, since s<d[m2]1s<d\leqslant\left[\frac{m}{2}\right]-1, (m1s+1)<(m1d+1)\binom{m-1}{s+1}<\binom{m-1}{d+1}, a contradiction. So we must have s=ds=d.

  • When mm is even and d[m2]2d\leqslant\left[\frac{m}{2}\right]-2, the same argument as the previous case applies. So the only remaining case is d=[m2]1=m21d=\left[\frac{m}{2}\right]-1=\frac{m}{2}-1. Since B(Δ)B(\Delta) is a Sperner family of {2,,m}\{2,\cdots,m\}, by Sperner’s theorem we have

    tb~(K)=tb~(Δ)=|B(Δ)|(m1[m12])=(m1m21)\widetilde{tb}(K)=\widetilde{tb}(\Delta)=|B(\Delta)|\leqslant\binom{m-1}{\left[\frac{m-1}{2}\right]}=\binom{m-1}{\frac{m}{2}-1}

    where the equality holds if and only if B(Δ)B(\Delta) consists of either all the (m21)(\frac{m}{2}-1)-subsets or all the m2\frac{m}{2}-subsets of {2,,m}\{2,\cdots,m\}. But since the dimension dim(Δ)=dim(K)=m21\dim(\Delta)=\dim(K)=\frac{m}{2}-1, B(Δ)B(\Delta) must be the later case. This implies that K=Δ(m21)[m]K=\Delta_{(\frac{m}{2}-1)}^{[m]} and so KϕK=KK\cap\phi K=K. The claim is proved.

By the above claim, if KK is a minimal element of Σtb(m,d)\Sigma^{tb}(m,d), then KϕK=KK\cap\phi K=K for every permutation ϕ\phi of [m][m], which implies that K=Δ(d)[m]K=\Delta_{(d)}^{[m]}. But Δ(d)[m]\Delta_{(d)}^{[m]} is maximal in Σ(m,d)\Sigma(m,d), so we can assert that Σtb(m,d)={Δ(d)[m]}\Sigma^{tb}(m,d)=\Big{\{}\Delta_{(d)}^{[m]}\Big{\}}.

(ii) and (iii)  When [m2]dm2\left[\frac{m}{2}\right]\leqslant d\leqslant m-2, let

X={2,,m},Y=[m]{1,d+1}={d+2,,m}.X=\{2,\cdots,m\},\ \ Y=[m]\setminus\{1,\cdots d+1\}=\{d+2,\cdots,m\}.

Observe that B(Δ)B(\Delta) is a Sperner family of XX that satisfies the condition (2) in Proposition 2.9 with {i1,,id}={2,,d+1}\{i_{1},\cdots,i_{d}\}=\{2,\cdots,d+1\}. Therefore, no subset of {2,,d+1}\{2,\cdots,d+1\} belongs to B(Δ)B(\Delta), which implies that every member of B(Δ)B(\Delta) has nonempty intersection with YY. In other words, B(Δ)B(\Delta) is a Sperner family of XX over YY. So by Lemma 2.11, the cardinality of B(Δ)B(\Delta) satisfies:

|B(Δ)|(m1m12)(dm12)=(m1[m2])(d[m2])|B(\Delta)|\leqslant\binom{m-1}{\lceil\frac{m-1}{2}\rceil}-\binom{d}{\lceil\frac{m-1}{2}\rceil}=\binom{m-1}{\left[\frac{m}{2}\right]}-\binom{d}{\left[\frac{m}{2}\right]}

where the equality holds if and only if B(Δ)B(\Delta) is one of the three types of Sperner families listed in Lemma 2.11 with n=|X|=m1n=|X|=m-1 and k=|Y|=md1k=|Y|=m-d-1. Then since |B(Δ)||B(\Delta)| computes tb~(K)\widetilde{tb}(K), the Sperner families described in Lemma 2.11 will give us all possible members of Σtb(m,d)\Sigma^{tb}(m,d).

  • In the case (a) of Lemma 2.11, the near-cone constructed from the Sperner family C12n(n,X,Y)C_{\left\lceil\frac{1}{2}n\right\rceil}(n,X,Y) in (7) in Proposition 2.9 is exactly Δ([m2]1)[m]d\Delta^{[m]}_{\left(\left[\frac{m}{2}\right]-1\right)}\langle d\rangle for every [m2]dm2\left[\frac{m}{2}\right]\leqslant d\leqslant m-2.

  • In the case (b) of Lemma 2.11, the near-cone constructed from the Sperner family C12(n1)(n,X,Y)C_{\frac{1}{2}(n-1)}(n,X,Y) in (7) has dimension nkn12(n+3)=m22n-k\leqslant n-\frac{1}{2}(n+3)=\frac{m}{2}-2 which contradicts our assumption d[m2]d\geqslant\left[\frac{m}{2}\right], hence is invalid.

  • In the case (c) of Lemma 2.11, nn is even and then mm is odd. The near-cone constructed from the Sperner family C12(n+2)(n,X,Y)C_{\frac{1}{2}(n+2)}(n,X,Y) in (7) has dimension d=nk=m2d=n-k=m-2, which gives rise to Δ([m2])[m]m2\Delta^{[m]}_{\left(\left[\frac{m}{2}\right]\right)}\langle m-2\rangle.

From the above discussion, we obtain the desired statements in (ii) and (iii). ∎

The following is an immediate corollary of the proof of Theorem 2.12.

Corollary 2.13.

For any dd-dimensional simplicial complex KK with mm vertices, the upper bound of tb~(K)\widetilde{tb}(K) is given by:

  1. (i)

    tb~(K)(m1d+1),\widetilde{tb}(K)\leqslant\binom{m-1}{d+1}, if d[m2]1d\leqslant\left[\frac{m}{2}\right]-1;

  2. (ii)

    tb~(K)(m1[m/2])(d[m/2]),\widetilde{tb}(K)\leqslant\binom{m-1}{[m/2]}-\binom{d}{[m/2]}, if [m2]dm2\left[\frac{m}{2}\right]\leqslant d\leqslant m-2;

  3. (iii)

    tb~(K)=0\widetilde{tb}(K)=0, if d=m1d=m-1.

Moreover, the equalities in (i)\mathrm{(i)} and (ii)\mathrm{(ii)} hold if and only if KK belongs to Σtb(m,d)\Sigma^{tb}(m,d) as described in Theorem 2.12.


3. Simplicial complexes with the minimum total bigraded Betti number

For any dd-dimensional simplicial complex KK with mm vertices, there is a universal lower bound of D~(K)\widetilde{D}(K) that depends only on mm and dd. This lower bound was discovered through an interesting relation between D~(K)\widetilde{D}(K) and some canonical CW-complex associated to KK called real moment-angle complex of KK (see Davis and Januszkiewicz [9, p. 428–429] or Buchstaber and Panov [5, Section 4.1]). One way to write 𝒵K\mathbb{R}\mathcal{Z}_{K} is

(8) 𝒵K=σK(iσD(i)1×iσS(i)0)i[m]D(i)1=[1,1]m,\mathbb{R}\mathcal{Z}_{K}=\bigcup_{\sigma\in K}\Big{(}\prod_{i\in\sigma}D^{1}_{(i)}\times\prod_{i\notin\sigma}S^{0}_{(i)}\Big{)}\subseteq\prod_{i\in[m]}D^{1}_{(i)}=[-1,1]^{m},

where D(i)1D^{1}_{(i)} and S(i)0S^{0}_{(i)} is a copy of D1=[1,1]D^{1}=[-1,1] and S0=D1={1,1}S^{0}=\partial D^{1}=\{-1,1\} indexed by i[m]i\in[m], and \prod denotes Cartesian product of spaces.

It is shown in [5, Section 4] (also see [6, Theorem 4.2]) that Tor𝔽[v1,,vm](𝔽[K],𝔽)\mathrm{Tor}_{\mathbb{F}[v_{1},\cdots,v_{m}]}(\mathbb{F}[K],\mathbb{F}) computes the cohomology groups of 𝒵K\mathbb{R}\mathcal{Z}_{K}, which implies

(9) D~(K;𝔽)=dim𝔽Tor𝔽[v1,,vm](𝔽[K],𝔽)=tb(𝒵K;𝔽).\widetilde{D}(K;\mathbb{F})=\dim_{\mathbb{F}}\mathrm{Tor}_{\mathbb{F}[v_{1},\cdots,v_{m}]}(\mathbb{F}[K],\mathbb{F})=tb(\mathbb{R}\mathcal{Z}_{K};\mathbb{F}).

Moreover, by [6, Theorem 1.4] or [20, Theorem 3.2], there is a universal lower bound of tb(𝒵K;𝔽)tb(\mathbb{R}\mathcal{Z}_{K};\mathbb{F}) for any simplicial complex KK:

(10) tb(𝒵K;𝔽)2mdim(K)1.tb(\mathbb{R}\mathcal{Z}_{K};\mathbb{F})\geqslant 2^{m-\dim(K)-1}.

So for any dd-dimensional simplicial complex KK with mm vertices, we always have

(11) D~(K;𝔽)2md1.\widetilde{D}(K;\mathbb{F})\geqslant 2^{m-d-1}.

In the following, we study those simplicial complexes that make the equality in (10) hold, i.e. tight simplicial complexes (see Definition 1.5). The coefficient 𝔽\mathbb{F} will be omitted in the rest of this section.

The following are some easy lemmas on the properties of D~(K)\widetilde{D}(K).

Lemma 3.1.

If KK^{\prime} is a full subcomplex of KK, then D~(K)D~(K)\widetilde{D}(K^{\prime})\leqslant\widetilde{D}(K).

Proof.

This follows from the formula (2) of D~(K)\widetilde{D}(K) and the simple fact that a full subcomplex of KK^{\prime} is also a full subcomplex of KK. ∎

Lemma 3.2.

Let KK be a simplicial complex with vertex set [m]={1,2,,m}[m]=\{1,2,\cdots,m\}. Then D~(K)1\widetilde{D}(K)\geqslant 1. Moreover, D~(K)=1\widetilde{D}(K)=1 if and only if KK is the simplex Δ[m]\Delta^{[m]}.

Proof.

We use the formula of D~(K)\widetilde{D}(K) in  (2). Consider a minimal subset J[m]J\subseteq[m] that does not span a simplex in KK (called a minimal non-face). If JJ is not the empty set, then K|J=ΔJK|_{J}=\partial\Delta^{J} and tb~(K|J)=1\widetilde{tb}(K|_{J})=1. This implies

D~(K)tb~(K|)+tb~(K|J)=2.\widetilde{D}(K)\geqslant\widetilde{tb}(K|_{\varnothing})+\widetilde{tb}(K|_{J})=2.

So D~(K)=1\widetilde{D}(K)=1 if and only if all the minimal non-faces of KK are empty, i.e. KK is the simplex Δ[m]\Delta^{[m]}. ∎

Lemma 3.3.

For any finite CW-complexes XX and YY,

tb(X×Y)=tb(X)tb(Y),tb~(XY)=tb~(X)tb~(Y)tb(X\times Y)=tb(X)tb(Y),\ \ \ \widetilde{tb}(X*Y)=\widetilde{tb}(X)\widetilde{tb}(Y)

where XYX*Y is the join of XX and YY.

Proof.

The equality tb(X×Y)=tb(X)tb(Y)tb(X\times Y)=tb(X)tb(Y) follows from the Künneth formula of homology groups. In addition, by the homotopy equivalence

XYΣ(XY)X*Y\simeq\Sigma(X\wedge Y)

where “\wedge” is the smash product and “Σ\Sigma” is the (reduced) suspension, we obtain (with a field coefficient) that

(12) H~n(XY)Hn1(XY)i(H~i(X)H~n1i(Y)).\widetilde{H}_{n}(X*Y)\cong H_{n-1}(X\wedge Y)\cong\bigoplus_{i}(\widetilde{H}_{i}(X)\otimes\widetilde{H}_{n-1-i}(Y)).

The second isomorphism in (12) follows from the relative version of the Künneth formula (see [13, Corollary 3.B.7]). Notice that XYX*Y is always path-connected and hence H~0(XY)=0\widetilde{H}_{0}(X*Y)=0. Then it follows that tb~(XY)=tb~(X)tb~(Y)\widetilde{tb}(X*Y)=\widetilde{tb}(X)\widetilde{tb}(Y). ∎

Lemma 3.4.

For any finite nonempty simplicial complexes KK and LL,

D~(KL)=D~(K)D~(L).\widetilde{D}(K*L)=\widetilde{D}(K)\widetilde{D}(L).
Proof.

Since 𝒵KL𝒵K×𝒵L\mathbb{R}\mathcal{Z}_{K*L}\cong\mathbb{R}\mathcal{Z}_{K}\times\mathbb{R}\mathcal{Z}_{L} (see [5, Chapter 4]), we obtain from (9) and Lemma 3.3 that

D~(KL)=tb(𝒵KL)=tb(𝒵K×𝒵L)=tb(𝒵K)tb(𝒵L)=D~(K)D~(L).\widetilde{D}(K*L)=tb(\mathbb{R}\mathcal{Z}_{K*L})=tb(\mathbb{R}\mathcal{Z}_{K}\times\mathbb{R}\mathcal{Z}_{L})=tb(\mathbb{R}\mathcal{Z}_{K})tb(\mathbb{R}\mathcal{Z}_{L})=\widetilde{D}(K)\widetilde{D}(L).

Observe that if KK is a dd-dimensional simplicial complex with mm vertices, then for any r1r\geqslant 1, Δ[r]K\Delta^{[r]}*K is a (d+r)(d+r)-dimensional simplicial complex with m+rm+r vertices and, D~(Δ[r]K)=D~(K)\widetilde{D}(\Delta^{[r]}*K)=\widetilde{D}(K) by Lemma 3.4. So we obtain the corollary immediately.

Corollary 3.5.

A finite simplicial complex KK is tight if and only if Δ[r]K\Delta^{[r]}*K is tight for all r1r\geqslant 1.

By Yu and Masuda [21, Proposition 2.1], any simplicial complex of the form Δ[n1]Δ[nk]\partial\Delta^{[n_{1}]}*\cdots*\partial\Delta^{[n_{k}]} is tight. So by Corollary 3.5, Δ[r]Δ[n1]Δ[nk]\Delta^{[r]}*\partial\Delta^{[n_{1}]}*\cdots*\partial\Delta^{[n_{k}]} is also tight for all r1r\geqslant 1. Then one may ask whether simplicial complexes of the form Δ[n1]Δ[nk]\partial\Delta^{[n_{1}]}*\cdots*\partial\Delta^{[n_{k}]} or Δ[r]Δ[n1]Δ[nk]\Delta^{[r]}*\partial\Delta^{[n_{1}]}*\cdots*\partial\Delta^{[n_{k}]} are all the tight simplicial complexes? We will see in Theorem 3.10 that the answer is yes.

For brevity, we introduce the following terms.

Definition 3.6.

For any positive integers n1,,nkn_{1},\cdots,n_{k} and rr, we call the simplicial complex Δ[n1]Δ[nk]\partial\Delta^{[n_{1}]}*\cdots*\partial\Delta^{[n_{k}]} a sphere join and call Δ[r]Δ[n1]Δ[nk]\Delta^{[r]}*\partial\Delta^{[n_{1}]}*\cdots*\partial\Delta^{[n_{k}]} a simplex-sphere join.

The following theorem proved in [21] will be useful for our proof.

Theorem 3.7 ([21, Theorem 3.1]).

Let KK be a simplicial complex of dimension n2n\geqslant 2. Suppose that KK satisfies the following two conditions:

  • (a)

    KK is an nn-dimensional pseudomanifold,

  • (b)

    the link of any vertex of KK is a sphere join of dimension n1n-1,

Then KK is a sphere join.

Recall that KK is an nn-dimensional pseudomanifold if the following conditions hold:

  • (i)

    Every simplex of KK is a face of some nn-simplex of KK (i.e. KK is pure).

  • (ii)

    Every (n1)(n-1)-simplex of KK is the face of exactly two nn-simplices of KK.

  • (iii)

    If σ\sigma and σ\sigma^{\prime} are two nn-simplices of KK, then there is a finite sequence of nn-simplices σ=σ0,σ1,,σk=σ\sigma=\sigma_{0},\sigma_{1},\ldots,\sigma_{k}=\sigma^{\prime} such that the intersection σiσi+1\sigma_{i}\cap\sigma_{i+1} is an (n1)(n-1)-simplex for all i=0,,k1i=0,\ldots,k-1.

In particular, any closed connected PL-manifold is a pseudomanifold.

In addition, we will use the following inequality proved in [20, Theorem 3.2]:

(13) D~(K)=tb(𝒵K)2mmdim(K)1,\widetilde{D}(K)=tb(\mathbb{R}\mathcal{Z}_{K})\geqslant 2^{m-\mathrm{mdim}(K)-1},

where mm is the number of vertices of KK, and

mdim(K)=the minimal dimension of the maximal simplices ofK.\mathrm{mdim}(K)=\ \text{the minimal dimension of the maximal simplices of}\ K.

The inequality (13) refines the inequality (10) since mdim(K)dim(K)\mathrm{mdim}(K)\leqslant\dim(K).

Lemma 3.8.

Let KK be a simplicial complex with mm vertices. If KK is tight, then

  • (i)

    KK is pure.

  • (ii)

    For every simplex σ\sigma of KK, LinkKσ\mathrm{Link}_{K}\sigma is tight.

Proof.

(i) By [20, Theorem 3.2], D~(K)=tb(𝒵K)2mmdim(K)12mdim(K)1\widetilde{D}(K)=tb(\mathbb{R}\mathcal{Z}_{K})\geqslant 2^{m-\mathrm{mdim}(K)-1}\geqslant 2^{m-\mathrm{dim}(K)-1}. Then since KK is tight, we must have mdim(K)=dim(K)\mathrm{mdim}(K)=\mathrm{dim}(K), which implies that every maximal simplex of KK has the same dimension as KK. So KK is pure.

(ii) We do induction on mm. When m=1m=1, this is trivial. For any vertex vv of KK, let mvm_{v} be the number of vertices in LinkKv\mathrm{Link}_{K}v. By the proof of [20, Theorem 3.2] (note that the argument there works for any vertex of KK), there is a subspace XvX_{v} of 𝒵K\mathbb{R}\mathcal{Z}_{K} with tb(𝒵K)tb(Xv)tb(\mathbb{R}\mathcal{Z}_{K})\geqslant tb(X_{v}), where XvX_{v} is homeomorphic to the disjoint union of 2mmv12^{m-m_{v}-1} copies of 𝒵LinkKv\mathbb{R}\mathcal{Z}_{\mathrm{Link}_{K}v}. So we have

2mdim(K)1=tb(𝒵K)2mmv1tb(𝒵LinkKv).2^{m-\mathrm{dim}(K)-1}=tb(\mathbb{R}\mathcal{Z}_{K})\geqslant 2^{m-m_{v}-1}tb(\mathbb{R}\mathcal{Z}_{\mathrm{Link}_{K}v}).

Then since dim(LinkKv)dim(K)1\dim(\mathrm{Link}_{K}v)\leqslant\dim(K)-1, we obtain

tb(𝒵LinkKv)2mvdim(K)2mvdim(LinkKv)1.tb(\mathbb{R}\mathcal{Z}_{\mathrm{Link}_{K}v})\leqslant 2^{m_{v}-\dim(K)}\leqslant 2^{m_{v}-\dim(\mathrm{Link}_{K}v)-1}.

On the other hand, by our induction we have tb(𝒵LinkKv)2mvdim(LinkKv)1tb(\mathbb{R}\mathcal{Z}_{\mathrm{Link}_{K}v})\geqslant 2^{m_{v}-\dim(\mathrm{Link}_{K}v)-1} since mv<mm_{v}<m. So LinkKv\mathrm{Link}_{K}v is tight.

Now suppose we have proved that LinkKσ\mathrm{Link}_{K}\sigma is tight for any simplex σ\sigma of KK with dimension less than jj. Let τ\tau be a jj-simplex in KK and let vv be a vertex of τ\tau. So σ=τ\{v}\sigma=\tau\backslash\{v\} is a (j1)(j-1)-simplex and it is easy to see that

(14) LinkKτ=LinkLinkKσv.\mathrm{Link}_{K}\tau=\mathrm{Link}_{\mathrm{Link}_{K}\sigma}v.

By our assumption, we know that LinkKσ\mathrm{Link}_{K}\sigma is tight. So by our preceding argument, we can assert that LinkKτ\mathrm{Link}_{K}\tau is also tight. This finishes the proof. ∎

Lemma 3.9.

If a simplicial complex KK is tight but not connected, then KK must be S0=Δ[2]S^{0}=\partial\Delta^{[2]}.

Proof.

If dim(K)=0\dim(K)=0, then KK being tight implies that KK is isomorphic either to Δ[1]\Delta^{[1]} or to Δ[2]\partial\Delta^{[2]}. If KΣ(m,d)K\in\Sigma(m,d) with dim(K)=d1\dim(K)=d\geqslant 1 and KK is not connected, let K=K1K2K=K_{1}\sqcup K_{2} where K1K_{1} and K2K_{2} are two subcomplexes of KK that are disjoint. Then we can add a 11-simplex σ={i1,i2}\sigma=\{i_{1},i_{2}\} to KK with a vertex {i1}K1\{i_{1}\}\in K_{1} and {i2}K2\{i_{2}\}\in K_{2} and obtain a new simplicial complex KΣ(m,d)K^{\prime}\in\Sigma(m,d). Observe that adding the 11-simplex σ\sigma to KK kills the generator of H~0(K|{i1,i2})=H~0(S0)\widetilde{H}_{0}(K|_{\{i_{1},i_{2}\}})=\widetilde{H}_{0}(S^{0}). So for any J[m]J\subseteq[m], it is easy to see that

tb~(K|J)={tb~(K|J)1,if J contains {i1,i2};tb~(K|J),otherwise.\widetilde{tb}(K^{\prime}|_{J})=\begin{cases}\widetilde{tb}(K|_{J})-1,&\text{if $J$ contains $\{i_{1},i_{2}\}$};\\ \widetilde{tb}(K|_{J}),&\text{otherwise}.\end{cases}

So by the formula (2) of D~(K)\widetilde{D}(K), we can deduce that D~(K)<D~(K)\widetilde{D}(K^{\prime})<\widetilde{D}(K). But since KK is tight, D~(K)<D~(K)=2md1\widetilde{D}(K^{\prime})<\widetilde{D}(K)=2^{m-d-1} contradicting the inequality in (11). ∎

Now, we are ready to prove the following theorem which answers Question 3.

Theorem 3.10.

A finite simplicial complex KK is tight if and only if KK is of the form Δ[n1]Δ[nk]\partial\Delta^{[n_{1}]}*\cdots*\partial\Delta^{[n_{k}]} or Δ[r]Δ[n1]Δ[nk]\Delta^{[r]}*\partial\Delta^{[n_{1}]}*\cdots*\partial\Delta^{[n_{k}]} for some positive integers n1,,nkn_{1},\cdots,n_{k} and rr.

Proof.

By our preceding discussion, any sphere join or simplex-sphere join is tight. Conversely, suppose KΣ(m,d)K\in\Sigma(m,d) is tight and we do induction on mm. When m2m\leqslant 2, the statement is trivial. For m3m\geqslant 3, by Lemma 3.8 and Lemma 3.9, KK is connected, pure and LinkKσ\mathrm{Link}_{K}\sigma is tight for every simplex σ\sigma of KK. Then since the number of vertices of LinkKσ\mathrm{Link}_{K}\sigma is less than mm, our induction hypothesis implies that LinkKσ\mathrm{Link}_{K}\sigma is either a sphere join or a simplex-sphere join.

Case 1: If for every vertex vv of KK, LinkKv\mathrm{Link}_{K}v is a sphere join, then by the relation in (14), we can inductively prove that LinkKσ\mathrm{Link}_{K}\sigma is also a sphere join for every simplex σ\sigma of KK. This implies that KK is a closed connected PL-manifold hence a pseudomanifold. So by Theorem 3.7, KK is sphere join.

Case 2: If there exists a vertex vKv\in K such that LinkKv\mathrm{Link}_{K}v is a simplex-sphere join, let LinkKv=Δ[r]Δ[n1]Δ[nk]\mathrm{Link}_{K}v=\Delta^{[r]}*\partial\Delta^{[n_{1}]}*\cdots*\partial\Delta^{[n_{k}]}. Since KK is a pure dd-dimensional simplicial complex, dim(LinkKv)=d1\dim(\mathrm{Link}_{K}v)=d-1. Take a vertex wΔ[r]w\in\Delta^{[r]} and consider the full subcomplex K\w:=K|[m]wK\backslash w:=K|_{[m]\setminus w} of KK. Note that

LinkK\wv=Δ[r]wΔ[n1]Δ[n2]Δ[nk],\mathrm{Link}_{K\backslash w}v=\Delta^{[r]\setminus w}*\partial\Delta^{[n_{1}]}*\partial\Delta^{[n_{2}]}*\cdots*\partial\Delta^{[n_{k}]},

which has dimension d2d-2. So the dimension of StarK\wv\mathrm{Star}_{K\backslash w}v is d1d-1, which implies

mdim(K\w)d1.\mathrm{mdim}(K\backslash w)\leqslant d-1.

But removing a vertex can reduce mdim(K)\mathrm{mdim}(K) at most by one. Then since KK is pure, mdim(K\w)mdim(K)1=dim(K)1=d1\mathrm{mdim}(K\backslash w)\geqslant\mathrm{mdim}(K)-1=\dim(K)-1=d-1. So mdim(K\w)=d1\mathrm{mdim}(K\backslash w)=d-1. Then by (13), we obtain

D~(K\w)2m1mdim(Kw)1=2md1.\widetilde{D}(K\backslash w)\geqslant 2^{m-1-\mathrm{mdim}(K\setminus w)-1}=2^{m-d-1}.

But by Lemma 3.1, D~(K\w)D~(K)=2md1\widetilde{D}(K\backslash w)\leqslant\widetilde{D}(K)=2^{m-d-1}. So we have

(15) D~(K\w)=2md1=D~(K).\widetilde{D}(K\backslash w)=2^{m-d-1}=\widetilde{D}(K).

Moreover, by the formula (2) of D~(K)\widetilde{D}(K), we obtain

D~(K)D~(K\w)=wJ[m]tb~(K|J).\widetilde{D}(K)-\widetilde{D}(K\backslash w)=\sum_{w\in J\subseteq[m]}\widetilde{tb}(K|_{J}).

So the equality (15) implies that tb~(K|J)=0\widetilde{tb}(K|_{J})=0 for every J[m]J\subseteq[m] that contains ww.

Claim: For any simplex σ\sigma of K\wK\backslash w, {w}σ\{w\}\cup\sigma is a simplex of KK.

We prove the claim by induction on |σ||\sigma|. When |σ|=1|\sigma|=1, i.e. σ\sigma is a vertex, {w}σK\{w\}\cup\sigma\in K since tb~(K|{w}σ)=0\widetilde{tb}(K|_{\{w\}\cup\sigma})=0. Assume that the claim is true when |σ|<s|\sigma|<s. If |σ|=s|\sigma|=s, then by induction K|{w}τK|_{\{w\}\cup\tau} is a simplex for every τσ\tau\subsetneq\sigma. If {w}σ\{w\}\cup\sigma is not a simplex of KK, then K|{w}σK|_{\{w\}\cup\sigma} is isomorphic to the boundary of an ss-dimensional simplex. But this contradicts the above conclusion that tb~(K|{w}σ)=0\widetilde{tb}(K|_{\{w\}\cup\sigma})=0. So the claim is proved.

By the above claim, K=w(K\w)K=w*(K\backslash w) is a cone of ww with K\wK\backslash w. It follows that

dim(K\w)=d1.\dim(K\backslash w)=d-1.

So by (15), D~(K\w)=2md1=2m1dim(K\w)1\widetilde{D}(K\backslash w)=2^{m-d-1}=2^{m-1-\dim(K\backslash w)-1}, i.e. K\wK\backslash w is tight. Then by our induction, K\wK\backslash w is either a sphere-join or a simplex-sphere join, which implies that K=w(K\w)K=w*(K\backslash w) is a simplex-sphere join. The theorem is proved. ∎


4. Simplicial complexes with the maximum total bigraded Betti number

In this section, we give a complete answer to Question 4. First, we prove a lemma parallel to Lemma 2.1 on total bigraded Betti number.

Lemma 4.1.

For any two simplicial complexes K,LK,L with vertex set [m][m],

D~(K)+D~(L)D~(KL)+D~(KL).\widetilde{D}(K)+\widetilde{D}(L)\leqslant\widetilde{D}(K\cap L)+\widetilde{D}(K\cup L).
Proof.

From (8), it is easy to see that

𝒵KL=𝒵K𝒵L,𝒵KL=𝒵K𝒵L.\mathbb{R}\mathcal{Z}_{K\cup L}=\mathbb{R}\mathcal{Z}_{K}\cup\mathbb{R}\mathcal{Z}_{L},\quad\mathbb{R}\mathcal{Z}_{K\cap L}=\mathbb{R}\mathcal{Z}_{K}\cap\mathbb{R}\mathcal{Z}_{L}.

Then similarly to Lemma 2.1, the Mayer-Vietoris sequence for (𝒵K,𝒵L)(\mathbb{R}\mathcal{Z}_{K},\mathbb{R}\mathcal{Z}_{L}) gives

tb(𝒵K)+tb(𝒵L)tb(𝒵K𝒵L)+tb(𝒵K𝒵L),tb(\mathbb{R}\mathcal{Z}_{K})+tb(\mathbb{R}\mathcal{Z}_{L})\leqslant tb(\mathbb{R}\mathcal{Z}_{K}\cap\mathbb{R}\mathcal{Z}_{L})+tb(\mathbb{R}\mathcal{Z}_{K}\cup\mathbb{R}\mathcal{Z}_{L}),

which is equivalent to the statement of the lemma by (9). ∎

Theorem 4.2.

If KK is a simplicial complex with vertex set [m][m], then

D~(K;𝔽)g(m,[m13])+1\widetilde{D}(K;\mathbb{F})\leqslant g\left(m,\left[\frac{m-1}{3}\right]\right)+1

for any field 𝔽\mathbb{F}, where the equality holds if and only if K=Δ([m13]1)[m]K=\Delta_{\left(\left[\frac{m-1}{3}\right]-1\right)}^{[m]}.

Proof.

Let ΣD~(m)={KΣ(m)D~(K)=maxLΣ(m)D~(L)}\Sigma^{\widetilde{D}}(m)=\Big{\{}K\in\Sigma(m)\mid\widetilde{D}(K)=\underset{L\in\Sigma(m)}{\max}\widetilde{D}(L)\Big{\}}, which is a partially ordered set with respect to the inclusions of simplicial complexes.

Suppose KK is a minimal or a maximal element of ΣD~(m)\Sigma^{\widetilde{D}}(m). Then for every permutation ϕ\phi of the vertex set of KK, it follows from Lemma 4.1 that

D~(K)+D~(ϕK)D~(KϕK)+D~(KϕK).\widetilde{D}(K)+\widetilde{D}(\phi K)\leqslant\widetilde{D}(K\cap\phi K)+\widetilde{D}(K\cup\phi K).

This implies that K=ϕKK=\phi K. So KK must be Δ(d)[m]\Delta^{[m]}_{(d)} where d=dim(K)d=\dim(K). Observe that any nonempty full subcomplex of Δ(d)[m]\Delta^{[m]}_{(d)} on kk vertices with kd+1k\leqslant d+1 is a simplex. So to compute D~(Δ(d)[m])\widetilde{D}(\Delta^{[m]}_{(d)}) via the formula (2), we only need to consider the full subcomplexes of Δ(d)[m]\Delta^{[m]}_{(d)} with more than d+1d+1 vertices. Moreover, it is easy to see that the reduced homology group of any nonempty full subcomplex of Δ(d)[m]\Delta^{[m]}_{(d)} always concentrates at degree dd. Then an easy calculation shows that

D~(Δ(d)[m])=i=0md2(mmi)(mi1d+1)+1=(3)g(m,d+1)+1.\widetilde{D}(\Delta^{[m]}_{(d)})=\sum_{i=0}^{m-d-2}\binom{m}{m-i}\binom{m-i-1}{d+1}+1\overset{\eqref{Equ:g-def}}{=}g(m,d+1)+1.

So the theorem follows from Lemma 5.1 in the Appendix. ∎

Remark 4.3.

For a simplicial complex KK with vertex set [m][m], the following combinatorial invariants of KK were studied by Codenotti, Spreer and Santos [8]:

τi(K)=1m+1J[m]β~i(K|J)(m|J|),i1.\tau_{i}(K)=\frac{1}{m+1}\sum\limits_{J\subseteq[m]}\frac{\widetilde{\beta}_{i}(K|_{J})}{\binom{m}{|J|}},\ i\geqslant-1.

Formally, τi(K)\tau_{i}(K) is some sort of weighted average of the ii-th Betti number of all the full subcomplexes of KK, which has a similar flavor as our D~(K)\widetilde{D}(K).


5. Appendix

Lemma 5.1.

For 0d<m0\leqslant d<m, g(m,d)=j=d+1m(mj)(j1d)g(m,d)=\sum_{j=d+1}^{m}\binom{m}{j}\binom{j-1}{d} reaches the maximum when and only when d=[m13]d=\left[\frac{m-1}{3}\right].

Proof.

The cases m4m\leqslant 4 can be checked by hand, so we assume m5m\geqslant 5 in the rest of the proof. It is easy to verify that

(16) g(m,d)+g(m,d1)=2md(md).g(m,d)+g(m,d-1)=2^{m-d}\binom{m}{d}.

So g(m,d)g(m,d2)=2md(md)2md+1(md1)g(m,d)-g(m,d-2)=2^{m-d}\binom{m}{d}-2^{m-d+1}\binom{m}{d-1}, which implies that

g(m,d)g(m,d2)>0d<m+13.g(m,d)-g(m,d-2)>0\,\Longleftrightarrow\,d<\frac{m+1}{3}.
g(m,d)=g(m,d2)d=m+13.g(m,d)=g(m,d-2)\,\Longleftrightarrow\,d=\frac{m+1}{3}.

Then we can deduce that

  • if m=3nm=3n, then g(m,d)g(m,d) is maximal only when d=n1d=n-1 or nn;

  • if m=3n+1m=3n+1, then g(m,d)g(m,d) is maximal only when d=n1d=n-1 or nn;

  • if m=3n+2m=3n+2, then g(m,d)g(m,d) is maximal only when d=n1,nd=n-1,n or n+1n+1 (since g(n1)=g(n+1)g(n-1)=g(n+1) in this case).

So to prove the Lemma, we only need to prove: for any n1n\geqslant 1,

  • (a)

    g(3n,n1)>g(3n,n)g(3n,n-1)>g(3n,n);

  • (b)

    g(3n+1,n1)<g(3n+1,n)g(3n+1,n-1)<g(3n+1,n);

  • (c)

    g(3n+2,n1)<g(3n+2,n)g(3n+2,n-1)<g(3n+2,n).

Note that by (16),

g(m,d)>g(m,d1)g(m,d1)<2md1(md)g(m,d1)(md)<2md1.g(m,d)>g(m,d-1)\,\Longleftrightarrow\,g(m,d-1)<2^{m-d-1}\binom{m}{d}\,\Longleftrightarrow\,\frac{g(m,d-1)}{\binom{m}{d}}<2^{m-d-1}.

We directly compute

g(m,d1)(md)=\displaystyle\frac{g(m,d-1)}{\binom{m}{d}}= 1(md)j=dm(mj)(j1d1)=jmj1(md)j=0md(mmj)(mj1d1)\displaystyle\ \frac{1}{\binom{m}{d}}\sum_{j=d}^{m}\binom{m}{j}\binom{j-1}{d-1}\stackrel{{\scriptstyle j\to m-j}}{{=}}\frac{1}{\binom{m}{d}}\sum_{j=0}^{m-d}\binom{m}{m-j}\binom{m-j-1}{d-1}
=\displaystyle= j=0mddmj(mdj)=01𝑑xd1j=0md(mdj)xmdjdx\displaystyle\ \sum_{j=0}^{m-d}\frac{d}{m-j}\binom{m-d}{j}=\int_{0}^{1}dx^{d-1}\sum_{j=0}^{m-d}\binom{m-d}{j}x^{m-d-j}\mathrm{d}x
(17) =\displaystyle= 01𝑑xd1(1+x)mddx.\displaystyle\ \int_{0}^{1}dx^{d-1}(1+x)^{m-d}\mathrm{d}x.

So we obtain

g(m,d)>g(m,d1)\displaystyle g(m,d)>g(m,d-1) 01𝑑xd1(1+x)mddx<2md1.\displaystyle\Longleftrightarrow\,\int_{0}^{1}dx^{d-1}(1+x)^{m-d}\mathrm{d}x<2^{m-d-1}.

Moreover, by the Cauchy-Schwarz inequality:

01𝑑xd1(1+x)mddx\displaystyle\int_{0}^{1}dx^{d-1}(1+x)^{m-d}\mathrm{d}x d01(xd1)2dx01[(1+x)md]2dx\displaystyle\leqslant d\,\sqrt{\int_{0}^{1}(x^{d-1})^{2}\,\mathrm{d}x}\,\sqrt{\int_{0}^{1}[(1+x)^{m-d}]^{2}\,\mathrm{d}x}
2md18d2(2d1)(2m2d+1).\displaystyle\leqslant 2^{m-d-1}\sqrt{\frac{8d^{2}}{(2d-1)(2m-2d+1)}}.

So to prove (b) and (c), we only need to show 8d2<(2d1)(2m2d+1)8d^{2}<(2d-1)(2m-2d+1) for (m,d)=(3n+1,n)(m,d)=(3n+1,n) and (m,d)=(3n+2,n)(m,d)=(3n+2,n), which is easy to check. Similarly,

g(m,d)<g(m,d1)\displaystyle g(m,d)<g(m,d-1) g(m,d)<2md1(md)g(m,d)(md)<2md1\displaystyle\,\Longleftrightarrow\,g(m,d)<2^{m-d-1}\binom{m}{d}\,\Longleftrightarrow\,\frac{g(m,d)}{\binom{m}{d}}<2^{m-d-1}
01(md)xd(1+x)md1dx<2md1,\displaystyle\Longleftrightarrow\,\int_{0}^{1}(m-d)x^{d}(1+x)^{m-d-1}\mathrm{d}x<2^{m-d-1},

where we use the result of (5) with dd substituted by d+1d+1 and the relation (md+1)=mdd+1(md)\binom{m}{d+1}=\frac{m-d}{d+1}\binom{m}{d}. Then again by the Cauchy-Schwarz inequality, we obtain

01(md)xd(1+x)md1dx2md12(md)2(2d+1)(2m2d1).\int_{0}^{1}(m-d)x^{d}(1+x)^{m-d-1}\mathrm{d}x\leqslant 2^{m-d-1}\sqrt{\frac{2(m-d)^{2}}{(2d+1)(2m-2d-1)}}.

So to prove (a), we only need to show 2(md)2<(2d+1)(2m2d1)2(m-d)^{2}<(2d+1)(2m-2d-1) for (m,d)=(3n,n)(m,d)=(3n,n), which is also easy to check. This finishes the proof. ∎


References

  • [1] A. Abbondandolo, Morse theory for Hamiltonian systems, Chapman &\& Hall/CRC Res. Notes Math., 425, 2001.
  • [2] C. Allday and V. Puppe, Cohomological methods in transformation groups, Cambridge Studies in Advanced Mathematics 32, Cambridge University Press, 1993.
  • [3] I. Anderson, Combinatorics of finite sets, Dover Publications, Inc., Mineola, NY, 2002.
  • [4] A. Björner and G. Kalai, An extended Euler-Poincaré theorem, Acta Math. 161 (1988), no. 3-4, 279–303.
  • [5] V. M. Buchstaber and T. E. Panov, Toric Topology. Mathematical Surveys and Monographs, Vol. 204, American Mathematical Society, Providence, RI, 2015.
  • [6] X. Cao and Z. Lü, Möbius transform, moment-angle complexes and Halperin-Carlsson conjecture, J. Algebraic Combin.35 (2012), no. 1, 121–140.
  • [7] G. Carlsson, Free (/2)k(\mathbb{Z}/2)^{k} actions and a problem in commutative algebra, Lecture Notes in Math., 1217, Springer, Berlin, 1986, 79–83.
  • [8] G. Codenotti, J. Spreer and F. Santos, Average Betti numbers of induced subcomplexes in triangulations of manifolds, Electron. J. Combin. 27 (2020), no.3, Paper No. 3.40.
  • [9] M. W. Davis and T. Januszkiewicz, Convex polytopes, Coxeter orbifolds and torus actions. Duke Math. J. 62 (1991), no. 2, 417–451.
  • [10] P. Erdös, C. Ko and R. Rado, Intersection theorems for systems of finite sets, Quart. J. Math. Oxford Ser. (2) 12 (1961), 313–320.
  • [11] J. Griggs, Collections of subsets with the Sperner property, Trans. Amer. Math. Soc. 269 (1982), no. 2, 575–591.
  • [12] S. Halperin, Rational homotopy and torus actions, In: Aspects of Topology. London Math. Soc. Lecture Note Series, 93. Cambridge Univ. Press, Cambridge (1985), 293–306.
  • [13] A. Hatcher, Algebraic topology, Cambridge University Press, Cambridge, 2002.
  • [14] M. Hochster, Cohen-Macaulay rings, combinatorics, and simplicial complexes, Ring Theory II (Proc. Second Oklahoma Conference), B. R. McDonald and R. Morris, eds. Dekker, New York (1977), 171–223.
  • [15] G. Kalai, Characterization of ff-vectors of families of convex sets in RdR^{d}, I. Necessity of Eckhoff’s conditions, Israel J. Math. 48 (1984), no. 2-3, 175–195.
  • [16] G. Kalai, ff-vectors of acyclic complexes, Discrete Math. 55 (1985), no. 1, 97–99.
  • [17] K. W. Lih, Sperner families over a subset, J. Combin. Theory Ser. A 29 (1980), no. 2, 182–185.
  • [18] E. Sperner, Ein Satz über Untermenge einer endlichen Menge, Math. Z., 27 (1928), 544–548.
  • [19] R. Stanley, Combinatorics and commutative algebra, 2nd edition. Birkhäuser Boston, 2007.
  • [20] Y. M. Ustinovsky, Toral rank conjecture for moment-angle complexes, Mathematical Notes 90 (2011), no. 2, 279–283.
  • [21] L. Yu and M. Masuda, On descriptions of products of simplices, Chinese Ann. Math. Ser. B 42 (2021), no. 5, 777–790.