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

Rigidity of Balanced Minimal Cycle Complexes

Ryoshun Oba Department of Mathematical Informatics, Graduate School of Information Science and Technology, University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, 113-8656, Tokyo Japan. Email: ryoshun_oba@mist.i.u-tokyo.ac.jp
Abstract

A (d1)(d-1)-dimensional simplicial complex Δ\Delta is balanced if its graph G(Δ)G(\Delta) is dd-colorable. Klee and Novik obtained the balanced lower bound theorem for balanced normal (d1)(d-1)-pseudomanifolds Δ\Delta with d3d\geq 3 by showing that the subgraph of G(Δ)G(\Delta) induced by the vertices colored in TT is rigid in 3\mathbb{R}^{3} for any 33 colors TT. We show that the same rigidity result, and thus the balanced lower bound theorem, holds for balanced minimal (d1)(d-1)-cycle complexes with d3d\geq 3. Motivated by the Stanley’s work on a colored system of parameters for the Stanley-Reisner ring of balanced simplicial complexes, we further investigate the infinitesimal rigidity of non-generic realization of balanced, and more broadly 𝒂\bm{a}-balanced, simplicial complexes. Among other results, we show that for d4d\geq 4, a balanced homology (d1)(d-1)-manifold can be realized as an infinitesimally rigid framework in d\mathbb{R}^{d} such that each vertex of color ii lies on the iith coordinate axis.

1 Introduction

Barnett’s lower bound theorem [2] asserts that the boundary complex Δ\Delta of a simplicial dd-polytope satisfies the inequality

f1(Δ)df0(Δ)(d+12),f_{1}(\Delta)\geq df_{0}(\Delta)-\binom{d+1}{2}, (1)

where fi(Δ)f_{i}(\Delta) denotes the number of ii-dimensional faces of Δ\Delta. Subsequently this was generalized to pseudomanifolds [7, 12]. Several variants of the lower bound theorem have been investigated (e.g. [9, 17]). In this paper we discuss the balanced lower bound theorem due to Goff, Klee, Novik [9] and Klee and Novik [13]. A (d1)(d-1)-dimensional simplicial complex Δ\Delta is balanced (or completely balanced) if its graph G(Δ)G(\Delta) is dd-colorable. The balanced lower bound theorem [13] asserts that, for d3d\geq 3, every balanced normal (d1)(d-1)-pseudomanifold Δ\Delta satisfies the inequality

2h2(Δ)(d1)h1(Δ),2h_{2}(\Delta)\geq(d-1)h_{1}(\Delta), (2)

where h1(Δ)=f0(Δ)dh_{1}(\Delta)=f_{0}(\Delta)-d, h2(Δ)=f1(Δ)(d1)f0(Δ)+(d2)h_{2}(\Delta)=f_{1}(\Delta)-(d-1)f_{0}(\Delta)+\binom{d}{2}.

A connection between the lower bound theorem and rigidity theory was pointed out by Kalai [12] (see also [10]). Kalai [12] noted that the inequality (1) follows immediately from the (generic) rigidity of the graph of Δ\Delta in d\mathbb{R}^{d}. Goff, Klee, Novik [9] pointed out that the inequality (2) holds if G(Δ)[κ1(T)]G(\Delta)[\kappa^{-1}(T)] is rigid in 3\mathbb{R}^{3} for every 33-set T[d]:={1,,d}T\subseteq[d]:=\{1,\ldots,d\}, where G[W]G[W] denotes the subgraph of GG induced by WW and κ\kappa is the proper dd-coloring of G(Δ)G(\Delta). Klee and Novik [13] showed this rigidity result, and thus the balanced lower bound theorem, for balanced normal (d1)(d-1)-pseudomanifolds for d3d\geq 3. This rigidity result for balanced normal pseudomanifolds [13, Lemma 3.5] is based on the rigidity theorem of pseudomanifolds by Fogelsanger [7]. Fogelsanger introduced a superclass of pseudomanifolds, called minimal cycle complexes, and show that minimal cycle complexes admit a decomposition into minimal cycle complexes in such a way that the decomposition behaves nicely with respect to edge contractions. He used this decomposition together with vertex splitting [20] and gluing to show that the graph of any minimal (d1)(d-1)-cycle complex is rigid in d\mathbb{R}^{d} for d3d\geq 3. The remarkable aspect of this proof of rigidity is that the inductive proof works within the same dimension, so it is applicable to a class of simplicial complexes which is not closed under taking links. Fogelsanger’s idea is recently used to show the global rigidity of pseudomanifolds [4] and 2\mathbb{Z}_{2}-symmetric rigidity of 2\mathbb{Z}_{2}-symmetric pseudomanifolds [5].

In this paper, we give a new application of Fogelsanger’s idea to prove rigidity results of balanced simplicial complexes. We generalize the rigidity result [13, Lemma 3.5] of balanced normal pseudomanifolds to balanced minimal cycle complexes (Theorem 4.1). Using an argument by Goff, Klee, Novik [9], this immediately implies the balanced lower bound theorem for minimal cycle complexes (Corollary 4.4).

Balanced simplicial complexes, and more broadly 𝒂\bm{a}-balanced simplicial complexes, have also been studied in the theory of Stanley-Reisner ring [16]. For 𝒂=(a1,,am)>0m\bm{a}=(a_{1},\ldots,a_{m})\in\mathbb{Z}_{>0}^{m} with i=1mai=d\sum_{i=1}^{m}a_{i}=d, a (d1)(d-1)-dimensional simplicial complex Δ\Delta on the vertex set V(Δ)V(\Delta) is 𝐚\bm{a}-balanced if there is a map κ:V(Δ)[m]\kappa:V(\Delta)\rightarrow[m] satisfying |Fκ1(i)|ai|F\cap\kappa^{-1}(i)|\leq a_{i} for any FΔF\in\Delta and i[m]i\in[m]. We call such a map κ\kappa an 𝐚\bm{a}-coloring of Δ\Delta. Given a simplicial complex Δ\Delta, the Stanley-Reisner ring of Δ\Delta is [Δ]:=[xv:vV(Δ)]/IΔ\mathbb{R}[\Delta]:=\mathbb{R}[x_{v}:v\in V(\Delta)]/I_{\Delta}, where IΔI_{\Delta} is the ideal generated by monomials vGxv\prod_{v\in G}x_{v} over all GΔG\not\in\Delta. For 𝒂>0m\bm{a}\in\mathbb{Z}_{>0}^{m} and an 𝒂\bm{a}-balanced simplicial complex Δ\Delta with an 𝒂\bm{a}-coloring κ\kappa, one can define an m\mathbb{N}^{m}-graded algebra structure of [Δ]\mathbb{R}[\Delta] by degxv=𝒆κ(v)\deg x_{v}=\bm{e}_{\kappa(v)}, where 𝒆im\bm{e}_{i}\in\mathbb{N}^{m} denotes the iith unit coordinate vector. Stanley [16] showed that, for an 𝒂\bm{a}-balanced (d1)(d-1)-dimensional simplicial complex Δ\Delta, [Δ]\mathbb{R}[\Delta] has a system of parameters θ1,,θd\theta_{1},\ldots,\theta_{d} such that exactly aia_{i} of θj\theta_{j}’s are of degree 𝒆i\bm{e}_{i} for i[m]i\in[m]. Such a system of parameters is called an 𝐚\bm{a}-colored s.o.p. for [Δ]\mathbb{R}[\Delta]. Cook et al. [11] showed that if Δ\Delta is a balanced simplicial 22-sphere, there is a (1,1,1)(1,1,1)-colored s.o.p. Θ=(θ1,θ2,θ3)\Theta=(\theta_{1},\theta_{2},\theta_{3}) for [Δ]\mathbb{R}[\Delta] and a linear form ω[Δ]1\omega\in\mathbb{R}[\Delta]_{1} such that the multiplication map (×ω):([Δ]/Θ[Δ])1([Δ]/Θ[Δ])2(\times\omega):(\mathbb{R}[\Delta]/\Theta\mathbb{R}[\Delta])_{1}\rightarrow(\mathbb{R}[\Delta]/\Theta\mathbb{R}[\Delta])_{2} is bijective. We conjecture that the statement of Cook et al. [11] holds for a larger class of simplicial complexes and any 𝒂>0m\bm{a}\in\mathbb{Z}_{>0}^{m} with a few exceptions as follows.

Conjecture 1.1.

For 𝒂>0m\bm{a}\in\mathbb{Z}_{>0}^{m} with d=i=1mai3d=\sum_{i=1}^{m}a_{i}\geq 3 and 𝒂(d1,1),(1,d1)\bm{a}\neq(d-1,1),(1,d-1), let Δ\Delta be an 𝒂\bm{a}-balanced minimal (d1)(d-1)-cycle complex. Then there is an 𝒂\bm{a}-colored s.o.p. Θ=(θ1,,θd)\Theta=(\theta_{1},\ldots,\theta_{d}) for [Δ]\mathbb{R}[\Delta] and a linear form ω[Δ]1\omega\in\mathbb{R}[\Delta]_{1} such that the multiplication map (×ω):([Δ]/Θ[Δ])1([Δ]/Θ[Δ])2(\times\omega):(\mathbb{R}[\Delta]/\Theta\mathbb{R}[\Delta])_{1}\rightarrow(\mathbb{R}[\Delta]/\Theta\mathbb{R}[\Delta])_{2} is injective.

A correspondence between Stanley-Reisner ring theory and (skeletal) rigidity theory has been noted in [14, 18]. In this correspondence, under the normalization ω=vV(Δ)xv\omega=\sum_{v\in V(\Delta)}x_{v}, a linear system of parameters (θ1,,θd)(\theta_{1},\ldots,\theta_{d}) for [Δ]\mathbb{R}[\Delta] is identified with a point configuration p:V(Δ)dp:V(\Delta)\rightarrow\mathbb{R}^{d} through θi=vV(Δ)p(v)ixv\theta_{i}=\sum_{v\in V(\Delta)}p(v)_{i}x_{v} for i[d]i\in[d] (see Corollary 5.3). Thus Conjecture 1.1 is equivalently formulated as the problem of finding an infinitesimally rigid realization of 𝒂\bm{a}-balanced minimal cycle complexes with non-generic point configurations. Let GG be a graph, κ:V(G)[m]\kappa:V(G)\rightarrow[m] a map, and 𝒂>0m\bm{a}\in\mathbb{Z}_{>0}^{m} an integer vector with i=1mai=d\sum_{i=1}^{m}a_{i}=d. We consider d\mathbb{R}^{d} as the direct product of ai\mathbb{R}^{a_{i}} over all i=1,,mi=1,\ldots,m. Then each xdx\in\mathbb{R}^{d} is denoted by x=(x1,,xm)x=(x_{1},\ldots,x_{m}) with xiaix_{i}\in\mathbb{R}^{a_{i}}. For i=1,,mi=1,\ldots,m, let Hi:={x=(x1,,xm):xj=0(ji)}dH_{i}:=\{x=(x_{1},\ldots,x_{m}):x_{j}=0~{}(j\neq i)\}\subseteq\mathbb{R}^{d}. We say that a point configuration p:V(G)dp:V(G)\rightarrow\mathbb{R}^{d} is (κ,𝐚)(\kappa,\bm{a})-sparse if p(v)Hκ(v)p(v)\in H_{\kappa(v)} for all vV(G)v\in V(G), and GG is (κ,𝐚)(\kappa,\bm{a})-sparse rigid if (G,p)(G,p) is infinitesimally rigid in d\mathbb{R}^{d} for some (κ,𝒂)(\kappa,\bm{a})-sparse point configuration pp. For example, if 𝒂=(1,,1)\bm{a}=(1,\ldots,1) and κ\kappa is a dd-coloring, then a (κ,𝒂)(\kappa,\bm{a})-sparse point configuration is the one such that each vertex of color ii is realized on the ii-th coordinate axis. Conjecture 1.1 can be restated as follows.

Conjecture 1.2.

For 𝒂>0m\bm{a}\in\mathbb{Z}_{>0}^{m} with d=i=1mai3d=\sum_{i=1}^{m}a_{i}\geq 3 and 𝒂(d1,1),(1,d1)\bm{a}\neq(d-1,1),(1,d-1), let Δ\Delta be an 𝒂\bm{a}-balanced minimal (d1)(d-1)-cycle complex and κ\kappa be an 𝒂\bm{a}-coloring of Δ\Delta. Then G(Δ)G(\Delta) is (κ,𝒂)(\kappa,\bm{a})-sparse rigid.

We will explain the equivalence of Conjecture 1.1 and Conjecture 1.2 in Section 5. The assumption on 𝒂\bm{a} in Conjecture 1.2 is necessary. Cook et al. [11] pointed out that there is a (2,1)(2,1)-balanced simplicial 22-sphere Δ\Delta and its (2,1)(2,1)-coloring κ\kappa such that G(Δ)G(\Delta) is not (κ,𝒂)(\kappa,\bm{a})-sparse rigid. As we will see in Example 7.6, the construction of Cook et al. [11] can be extended to higher dimension. In this paper we confirm Conjecture 1.2 for the following cases. In Theorem 6.1, we prove that Conjecture 1.2 holds if ai2a_{i}\geq 2 for all i[m]i\in[m] using Fogelsanger’s idea. In Theorem 7.2, we then apply the standard coning argument and show that Conjecture 1.2 holds if d4d\geq 4 and Δ\Delta is a homology (d1)(d-1)-manifold. The base cases in the coning argument are Gluck’s theorem [8], the result of Cook et al. [11, Theorem 1.1] and Theorem 6.1 for 𝒂=(2,2)\bm{a}=(2,2).

The paper is organized as follows. In Section 2, preliminaries on simplicial complexes and rigidity are given. In Section 3, we summarize the basic properties of Fogelsanger decomposition for minimal cycle complexes. In Section 4, we prove the rigidity of subgraph of G(Δ)G(\Delta) induced by colors and derive the balanced lower bound theorem for a balanced minimal cycle complex Δ\Delta. In Section 5, we discuss the equivalence of Conjecture 1.1 and Conjecture 1.2. In Section 6, we prove Conjecture 1.2 in the case when ai2a_{i}\geq 2 for all ii. In Section 7, we prove Conjecture 1.2 for homology manifolds. In Section 8, we give further observations related to Conjecture 1.2.

2 Preliminaries

2.1 Graphs and simplicial complexes

Throughout the paper, we only consider simple graphs and use the following basic notations. For a graph GG, the vertex set and the edge set is denoted by V(G)V(G) and E(G)E(G). For XV(G)X\subseteq V(G), G[X]=(X,E[X])G[X]=(X,E[X]) denotes the induced subgraph of GG by XX. For vV(G)v\in V(G), NG(v)N_{G}(v) denotes the set of vertices adjacent to vv in GG. Given a graph GG and an edge uvE(G)uv\in E(G), we denote G/uvG/uv to denote the simple graph obtained from GG by contracting vv onto uu. More precisely V(G/uv)=VvV(G/uv)=V-v, E(G/uv)=E[Vv]{uw:wNG(v)}E(G/uv)=E[V-v]\cup\{uw:w\in N_{G}(v)\}.

A simplicial complex Δ\Delta is a finite collection of finite sets such that if FΔF\in\Delta and GFG\subseteq F, GΔG\in\Delta. Elements of Δ\Delta are called faces of Δ\Delta. The dimension of a face FΔF\in\Delta is dimF:=|F|1\dim F:=|F|-1, and a face of dimension ii is called an ii-face. The dimension of Δ\Delta is dimΔ:=max{dimF:FΔ}\dim\Delta:=\max\{\dim F:F\in\Delta\}. A facet of Δ\Delta is a maximal face under inclusion and Δ\Delta is pure if all facets have the same dimension. For a finite collection 𝒮\mathcal{S} of finite sets, the simplicial complex spanned by 𝒮\mathcal{S} is 𝒮:={GF:F𝒮}\langle\mathcal{S}\rangle:=\{G\subseteq F:F\in\mathcal{S}\}. The vertex set V(Δ)V(\Delta) (resp. the edge set E(Δ)E(\Delta)) of Δ\Delta is the set of all 0-faces (resp. 11-faces). G(Δ)=(V(Δ),E(Δ))G(\Delta)=(V(\Delta),E(\Delta)) is called the graph (or 11-skeleton) of Δ\Delta. The link of a face FΔF\in\Delta is lkΔ(F)={GΔ:FG=,FGΔ}\operatorname{lk}_{\Delta}(F)=\{G\in\Delta:F\cap G=\emptyset,F\cup G\in\Delta\}. The (closed) star of a face FΔF\in\Delta is stΔ(F)={GΔ:FGΔ}\operatorname{st}_{\Delta}(F)=\{G\in\Delta:F\cup G\in\Delta\}. For WV(Δ)W\subseteq V(\Delta), define the restriction of Δ\Delta to WW to be Δ[W]:={FΔ:FW}\Delta[W]:=\{F\in\Delta:F\subseteq W\}.

Let Δ1,Δ2\Delta_{1},\Delta_{2} be pure simplicial complexes with V(Δ1)V(Δ2)=V(\Delta_{1})\cap V(\Delta_{2})=\emptyset and dimΔ1=dimΔ2\dim\Delta_{1}=\dim\Delta_{2}. Let F1F_{1} and F2F_{2} be a facet of Δ1\Delta_{1} and Δ2\Delta_{2} respectively, and let γ:F1F2\gamma:F_{1}\rightarrow F_{2} be a bijection. The connected sum of Δ1\Delta_{1} and Δ2\Delta_{2}, denoted as Δ1#γΔ2\Delta_{1}\#_{\gamma}\Delta_{2} (or simply Δ1#Δ2\Delta_{1}\#\Delta_{2}), is the simplicial complex obtained by identifying the vertices FF and GG and removing the facet corresponding to FF (which has been identified with GG). A (d1)(d-1)-sphere on n(d+1)n(\geq d+1) vertices written as a (nd)(n-d)-fold connected sum of the boundary complex of a dd-simplex is called a stacked (d1)(d-1)-sphere.

We say that a simplicial complex Δ\Delta is a simplicial (d1)(d-1)-sphere if its geometric realization is homeomorphic to 𝕊d1\mathbb{S}^{d-1}. A simplicial complex Δ\Delta is a homology (d1)(d-1)-manifold over 𝒌\bm{k} if H~(lkΔ(F);𝒌)H~(𝕊d|F|1;𝒌)\tilde{H}_{*}(\operatorname{lk}_{\Delta}(F);\bm{k})\cong\tilde{H}_{*}(\mathbb{S}^{d-|F|-1};\bm{k}) for every nonempty face FΔF\in\Delta. Here, H~(Δ;𝒌)\tilde{H}_{*}(\Delta;\bm{k}) denotes the reduced simplicial homology group of Δ\Delta with coefficients in 𝒌\bm{k}. A pure (d1)(d-1)-dimensional simplicial complex Δ\Delta is strongly connected if for every pair of facets FF and GG of Δ\Delta, there is a sequence of facets F=F0,F1,Fm=GF=F_{0},F_{1}\ldots,F_{m}=G such that |Fi1Fi|=d1|F_{i-1}\cap F_{i}|=d-1 for i[m]i\in[m]. A (d1)(d-1)-pseudomanifold is a strongly connected pure (d1)(d-1)-dimensional simplicial complex such that every (d2)(d-2)-face is contained in exactly two facets. A (d1)(d-1)-pseudomanifold is normal if the link of each face of dimension at most d3d-3 is connected. The class of normal pseudomanifolds is closed under taking links [1].

Let Δ\Delta be a (d1)(d-1)-dimensional simplicial complex. The ff-vector of Δ\Delta is an integer vector f(Δ):=(f1(Δ),,fd1(Δ))f(\Delta):=(f_{-1}(\Delta),\ldots,f_{d-1}(\Delta)), where fi(Δ)f_{i}(\Delta) is the number of ii-faces of Δ\Delta. The hh-vector of Δ\Delta is an integer vector h(Δ):=(h0(Δ),,hd(Δ))h(\Delta):=(h_{0}(\Delta),\ldots,h_{d}(\Delta)) whose entries are given by

hj(Δ):=i=0j(1)ji(didj)fi1(Δ).h_{j}(\Delta):=\sum_{i=0}^{j}(-1)^{j-i}\binom{d-i}{d-j}f_{i-1}(\Delta).

2.2 Rigidity

A framework in d\mathbb{R}^{d} is a pair (G,p)(G,p) of a graph GG and a map p:V(G)dp:V(G)\rightarrow\mathbb{R}^{d}. The map pp is called a point configuration. An infinitesimal motion of a framework (G,p)(G,p) in d\mathbb{R}^{d} is a map p˙:V(G)d\dot{p}:V(G)\rightarrow\mathbb{R}^{d} satisfying

(p(i)p(j))(p˙(i)p˙(j))=0ijE(G).(p(i)-p(j))\cdot(\dot{p}(i)-\dot{p}(j))=0\qquad ij\in E(G). (3)

An infinitesimal motion p˙\dot{p} defined by p˙(i)=Sp(i)+t\dot{p}(i)=Sp(i)+t (iV(G))(i\in V(G)) for a skew symmetric matrix SS and tdt\in\mathbb{R}^{d} is said to be trivial. A framework (G,p)(G,p) is infinitesimally rigid in d\mathbb{R}^{d} if every infinitesimal motion of (G,p)(G,p) is trivial. A graph GG is rigid in d\mathbb{R}^{d} if (G,p)(G,p) is infinitesimally rigid in d\mathbb{R}^{d} for some p:V(G)dp:V(G)\rightarrow\mathbb{R}^{d}.

The matrix R(G,p)|E(G)|×d|V(G)|R(G,p)\in\mathbb{R}^{|E(G)|\times d|V(G)|} representing (3) is called the rigidity matrix of (G,p)(G,p). More concretely, in R(G,p)R(G,p), dd consecutive columns are associated to each vertex and a row associated to ijE(G)ij\in E(G) is

[\@arstrutij\\0p(j)-p(i)0p(i)-p(j)0] ,\hbox{}\vbox{\kern 0.86108pt\hbox{$\kern 0.0pt\kern 2.5pt\kern-5.0pt\left[\kern 0.0pt\kern-2.5pt\kern-5.55557pt\vbox{\kern-0.86108pt\vbox{\vbox{ \halign{\kern\arraycolsep\hfil\@arstrut$\kbcolstyle#$\hfil\kern\arraycolsep& \kern\arraycolsep\hfil$\@kbrowstyle#$\ifkbalignright\relax\else\hfil\fi\kern\arraycolsep&& \kern\arraycolsep\hfil$\@kbrowstyle#$\ifkbalignright\relax\else\hfil\fi\kern\arraycolsep\cr 5.0pt\hfil\@arstrut$\scriptstyle$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle i$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle j$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle\\$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle\cdots 0\cdots$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle p(j)-p(i)$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle\cdots 0\cdots$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle p(i)-p(j)$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle\cdots 0\cdots$\hfil\kern 5.0pt\crcr}}}}\right]$}},

where p(i),p(j)p(i),p(j) are considered as a row vector. An element in the left kernel of R(G,p)R(G,p) is called an equilibrium stress (or an affine 22-stress) of (G,p)(G,p). We call the left kernel of R(G,p)R(G,p) the stress space of (G,p)(G,p). If p(V(G))p(V(G)) spans at least (d1)(d-1)-dimensional affine subspace, a framework (G,p)(G,p) is infinitesimally rigid if and only if rankR(G,p)=d|V(G)|(d+12)\rank R(G,p)=d|V(G)|-\binom{d+1}{2}.

To produce infinitesimally rigid frameworks from those with smaller size, gluing lemma and vertex splitting lemma are useful.

Lemma 2.1 (Gluing Lemma).

Let (G,p)(G,p) be a framework in d\mathbb{R}^{d}. Suppose that G1,G2G_{1},G_{2} are subgraphs of GG such that (Gi,p|V(Gi))(G_{i},p|_{V(G_{i})}) is infinitesimally rigid in d\mathbb{R}^{d} for i=1,2i=1,2 and p(V(G1)V(G2))p(V(G_{1})\cap V(G_{2})) spans at least (d1)(d-1)-dimensional affine subspace. Then (G,p)(G,p) is infinitesimally rigid in d\mathbb{R}^{d}.

Lemma 2.2 (Vertex Splitting Lemma [20]).

Let GG be a graph and uvE(G)uv\in E(G) be an edge satisfying |NG(u)NG(v)|d1|N_{G}(u)\cap N_{G}(v)|\geq d-1. Let CC be a size (d1)(d-1) subset of NG(u)NG(v)N_{G}(u)\cap N_{G}(v). Suppose that there is p:V(G/uv)dp:V(G/uv)\rightarrow\mathbb{R}^{d} such that (G/uv,p)(G/uv,p) is infinitesimally rigid in d\mathbb{R}^{d} and {p(w)p(u):wC}\{p(w)-p(u):w\in C\} is linearly independent. Let zdz\in\mathbb{R}^{d} be a vector not in the linear span of {p(w)p(u):wC}\{p(w)-p(u):w\in C\}.

Then there is tt\in\mathbb{R} such that for the extension p:V(G)dp^{\prime}:V(G)\rightarrow\mathbb{R}^{d} of pp defined by p(v)=p(u)+tzp^{\prime}(v)=p(u)+tz, (G,p)(G,p^{\prime}) is infinitesimally rigid in d\mathbb{R}^{d}.

Coning is a way to produce an infinitesimally rigid framework from that in one less dimension. The cone graph G{v}G*\{v\} of a graph GG is obtained from GG by adding a new vertex vv and edges from vv to all the original vertices in GG.

Lemma 2.3 (Cone Lemma [19]).

Let GG be a graph and G{v}G*\{v\} be its cone graph. Let p:V(G){v}d+1p:V(G)\cup\{v\}\rightarrow\mathbb{R}^{d+1} be a point configuration of cone graph such that p(v)p(u)p(v)\neq p(u) for all uV(G)u\in V(G). Let HdH\cong\mathbb{R}^{d} be a hyperplane in d+1\mathbb{R}^{d+1} not passing through p(v)p(v) and not parallel to the vectors {p(u)p(v):uV(G)}\{p(u)-p(v):u\in V(G)\}. Define a projection pH:V(G)dp_{H}:V(G)\rightarrow\mathbb{R}^{d} of pp by letting pH(u)p_{H}(u) be the intersection point of HH and the line passing through p(v)p(v) and p(u)p(u). Then (G{v},p)(G*\{v\},p) is infinitesimally rigid in d+1\mathbb{R}^{d+1} if and only if (G,pH)(G,p_{H}) is infinitesimally rigid in d\mathbb{R}^{d}.

3 Decomposition of minimal cycle complexes

Let Δ\Delta be a simplicial complex and uvE(Δ)uv\in E(\Delta). We define a simplicial complex Δ/uv\Delta/uv by

Δ/uv={FΔ:vF}{Fv+u:FstΔ(v)}.\Delta/uv=\{F\in\Delta:v\not\in F\}\cup\{F-v+u:F\in\operatorname{st}_{\Delta}(v)\}.

The operation ΔΔ/uv\Delta\rightarrow\Delta/uv is called an edge contraction of uvuv (onto uu). We remark that we do not allow multiple copies of the same faces, so for GlkΔ(u)lkΔ(v)G\in\operatorname{lk}_{\Delta}(u)\cap\operatorname{lk}_{\Delta}(v), faces G+uG+u and G+vG+v of Δ\Delta are identified under the edge contraction of uvuv.

Let Γ\Gamma be an abelian group and let Δ\Delta be a simplicial complex on the vertex set [n][n]. An ii-chain is a Γ\Gamma-coefficient formal sum of ii-faces. For an ii-chain c=FaFFc=\sum_{F}a_{F}F, define an (i1)(i-1)-chain c\partial c by c:=FaFF\partial c:=\sum_{F}a_{F}\partial F, where for F={x1,,xi+1}F=\{x_{1},\ldots,x_{i+1}\} (x1<<xi+1x_{1}<\cdots<x_{i+1}), F:=j=1i+1(1)jF{xj}\partial F:=\sum_{j=1}^{i+1}(-1)^{j}F\setminus\{x_{j}\}. An ii-cycle is an ii-chain cc satisfying c=0\partial c=0. An ii-chain c=FbFFc^{\prime}=\sum_{F}b_{F}F is a subchain of an ii-chain c=FaFFc=\sum_{F}a_{F}F if for each ii-face FF, bF=aFb_{F}=a_{F} or bF=0b_{F}=0 holds. An ii-cycle cc is a minimal ii-cycle if its only subchains which are ii-cycles are cc and 0. The support of an ii-chain c=FaFFc=\sum_{F}a_{F}F is suppc:={F:aF0}([n]i+1)\operatorname{supp}c:=\{F:a_{F}\neq 0\}\subseteq\binom{[n]}{i+1}, where ([n]i+1)\binom{[n]}{i+1} is the set of all (i+1)(i+1)-subsets of [n][n].

A (d1)(d-1)-dimensional simplicial complex Δ\Delta is called a minimal (d1)(d-1)-cycle complex (over Γ\Gamma) if (i) Δ={F}\Delta=\langle\{F\}\rangle for a dd-set FF or (ii) Δ=suppc\Delta=\langle\operatorname{supp}c\rangle for a nonzero minimal (d1)(d-1)-cycle cc. If the dimension is clear from the context, it is simply called a minimal cycle complex. We remark that although minimal cycle complexes are originally defined as simplicial complexes satisfying (ii), we include the case (i) to make the presentation simple. It is possible to unify (i) and (ii) by using multicomplex formulation as in [4]. A minimal cycle complex satisfying (i) (resp. (ii)) is said to be trivial (resp. nontrivial). A minimal cycle complex is always pure. Minimal cycle complexes may have singular point, so the class of minimal cycle complexes are not closed under taking links.

Pseudomanifolds are minimal cycle complexes over 2\mathbb{Z}_{2} [4]. 22-CM complexes over a field 𝒌\bm{k} are a minimal cycle complexes over a free 𝒌\bm{k}-module of finite rank [15]. Other examples of minimal cycle complexes arise in the context of simplicial matroid. A (d1)(d-1)-simplicial matroid on a subset \mathcal{E} of ([n]d)\binom{[n]}{d} over a field 𝒌\bm{k} is a matroid such that {F1,,Fk}\{F_{1},\ldots,F_{k}\}\subseteq\mathcal{E} is independent if and only if {F1,,Fk}\{\partial F_{1},\ldots,\partial F_{k}\} is linearly independent. A simplicial complex spanned by the circuit of (d1)(d-1)-simplicial matroid over 𝒌\bm{k} is a minimal (d1)(d-1)-cycle complex over 𝒌\bm{k}. A matroid on a ground set EE is connected if for any e,fEe,f\in E, there is a circuit containing both ee and ff. A pure (d1)(d-1)-dimensional simplicial complex whose facets define a connected (d1)(d-1)-simplicial matroid over 𝒌\bm{k} is a minimal (d1)(d-1)-cycle complex over a free 𝒌\bm{k}-module of finite rank. Thus a simplicial complex having a convex ear decomposition [3] is a minimal cycle complex.

We list basic properties of minimal cycle complexes.

Lemma 3.1.

For a nontrivial minimal (d1)(d-1)-cycle complex Δ\Delta, the followings hold:

  • (i)

    Δ\Delta is strongly connected.

  • (ii)

    Every (d2)(d-2)-face of Δ\Delta is contained in at least two facets of Δ\Delta. In particular, |V(Δ)|d+1|V(\Delta)|\geq d+1.

Proof.

Let cc be a minimal (d1)(d-1)-cycle satisfying Δ=suppc\Delta=\langle\operatorname{supp}c\rangle.

(i) Suppose that Δ\Delta is not strongly connected. Then there is a proper subset 𝒮\mathcal{S} of suppc\operatorname{supp}c such that no pair of F𝒮F\in\mathcal{S} and Gsuppc𝒮G\in\operatorname{supp}c\setminus\mathcal{S} shares the same (d2)(d-2)-face. Then, cc restricted to 𝒮\mathcal{S} is a (d1)(d-1)-cycle and is a proper subchain of cc, which is a contradiction.

(ii) Suppose that a (d2)(d-2)-face TT of Δ\Delta is contained in only one facet FF of Δ\Delta. Then (c)T=±cF0(\partial c)_{T}=\pm c_{F}\neq 0, which contradicts the fact that cc is a (d1)(d-1)-cycle. ∎

Fogelsanger [7] pointed out that a minimal (d1)(d-1)-cycle can be decomposed into minimal (d1)(d-1)-cycles in such a way that it behaves nicely with respect to edge contractions. We summarize the properties of this decomposition below, which is necessary in the later discussion, in terms of minimal cycle complexes. See [4, 7] for more details.

Lemma 3.2 (Fogelsanger [7]).

Suppose that Δ\Delta is a nontrivial minimal (d1)(d-1)-cycle complex and uvE(Δ)uv\in E(\Delta). Then there exists a sequence of nontrivial minimal (d1)(d-1)-cycle complexes Δ1+,,Δm+\Delta_{1}^{+},\ldots,\Delta_{m}^{+} satisfying the following properties:

  • (a)

    For each ii, uvE(Δi+)uv\in E(\Delta_{i}^{+}). Equivalently, there is a facet in Δi+\Delta_{i}^{+} containing both uu and vv.

  • (b)

    For each ii and each facet FF of Δi+\Delta_{i}^{+} with FΔF\not\in\Delta, uu and vv are contained in FF, and FuF-u and FvF-v are (d2)(d-2)-faces of Δi+Δ\Delta_{i}^{+}\cap\Delta.

  • (c)

    For each ii, Δi+/uv\Delta_{i}^{+}/uv is a minimal (d1)(d-1)-cycle complex.

  • (d)

    G(Δ)=i=1mG(Δi+)G(\Delta)=\bigcup_{i=1}^{m}G(\Delta_{i}^{+}).

  • (e)

    For each i2i\geq 2, (j<iΔj+)(\bigcup_{j<i}\Delta_{j}^{+}) and Δi+\Delta_{i}^{+} share at least one facet.

We remark that, as in Lemma 3.2 (b), for the minimal cycle complexes Δ1+,,Δm+\Delta_{1}^{+},\ldots,\Delta_{m}^{+} given in Lemma 3.2, Δi+\Delta_{i}^{+} may have a face not in Δ\Delta. We also remark that the decomposition given in Lemma 3.2 is not unique.

4 Rigidity of rank-selected subcomplexes

We use the notation [d]:={1,,d}[d]:=\{1,\ldots,d\}. A (d1)(d-1)-dimensional simplicial complex Δ\Delta is balanced (or completely balanced) if G(Δ)G(\Delta) is dd-colorable, or equivalently, there is a map κ:V(Δ)[d]\kappa:V(\Delta)\rightarrow[d] such that for any face FΔF\in\Delta, |Fκ1(i)|1|F\cap\kappa^{-1}(i)|\leq 1 for i[d]i\in[d]. Such a coloring is called a proper coloring of Δ\Delta. A proper coloring of a strongly connected simplicial complex is unique up to the permutation of colors and is called the proper coloring of Δ\Delta. The boundary complex of a dd-dimensional cross-polytope 𝒞d:=conv{±𝒆i:i[d]}\mathcal{C}_{d}^{*}:=\operatorname{conv}\{\pm\bm{e}_{i}:i\in[d]\}, where 𝒆i\bm{e}_{i} is the iith unit coordinate vector in d\mathbb{R}^{d}, is an example of a balanced (d1)(d-1)-simplicial complex.

Let Δ\Delta be a balanced, strongly connected, (d1)(d-1)-dimensional simplicial complex and κ\kappa be the proper coloring of Δ\Delta. For T[d]T\subseteq[d], the TT-rank selected subcomplex of Δ\Delta is ΔT:=Δ[κ1(T)]\Delta_{T}:=\Delta[\kappa^{-1}(T)]. Klee and Novik [13] showed that G(ΔT)G(\Delta_{T}) is rigid in |T|\mathbb{R}^{|T|} for any T[d]T\subseteq[d] with |T|3|T|\geq 3 if Δ\Delta is a balanced normal (d1)(d-1)-pseudomanifold with d3d\geq 3. We extend this result to the class of minimal (d1)(d-1)-cycle complexes as follows.

Theorem 4.1.

For d3d\geq 3, let Δ\Delta be a balanced minimal (d1)(d-1)-cycle complex. Then for T[d]T\subseteq[d] with |T|3|T|\geq 3, G(ΔT)G(\Delta_{T}) is rigid in |T|\mathbb{R}^{|T|}.

For a pure simplicial complex Δ\Delta, we say that a vertex subset UV(Δ)U\subseteq V(\Delta) is (k)(\geq k)-transversal if |UF|k|U\cap F|\geq k for every facet FΔF\in\Delta. Since, for a proper coloring κ\kappa of a pure simplicial complex Δ\Delta and T[d]T\subseteq[d], κ1(T)V(Δ)\kappa^{-1}(T)\subseteq V(\Delta) is (|T|)(\geq|T|)-transversal, Theorem 4.1 is a corollary of the following lemma.

Lemma 4.2.

For dk3d\geq k\geq 3, let Δ\Delta be a minimal (d1)(d-1)-cycle complex and UV(Δ)U\subseteq V(\Delta) be a (k)(\geq k)-transversal set of Δ\Delta. Then G(Δ[U])G(\Delta[U]) is rigid in k\mathbb{R}^{k}.

Proof.

We prove the statement by the induction on |V(Δ)||V(\Delta)|. If Δ\Delta is a trivial (d1)(d-1)-cycle complex, G(Δ)G(\Delta) is a complete graph. Hence, G(Δ[U])=G(Δ)[U]G(\Delta[U])=G(\Delta)[U] is a complete graph, which is rigid in k\mathbb{R}^{k}.

Suppose that Δ\Delta is a nontrivial minimal (d1)(d-1)-cycle complex. Pick u,vUu,v\in U with uvE(Δ)uv\in E(\Delta). Such u,vu,v always exist by k3k\geq 3. Let Δ1+,,Δm+\Delta_{1}^{+},\ldots,\Delta_{m}^{+} be the nontrivial minimal (d1)(d-1)-cycle complexes given in Lemma 3.2 with respect to Δ\Delta and uvuv. For each ii, let Ui:=UV(Δi+)U_{i}:=U\cap V(\Delta_{i}^{+}). For each facet FF of Δi+Δ\Delta_{i}^{+}\cap\Delta, we have |FUi|k|F\cap U_{i}|\geq k. For each facet FΔi+ΔF\in\Delta_{i}^{+}\setminus\Delta, as FF contains uu and FuF-u is a (d2)(d-2)-face of Δ\Delta by Lemma 3.2 (b), we have |FUi|=|(Fu)Ui|+1k|F\cap U_{i}|=|(F-u)\cap U_{i}|+1\geq k. Hence UiU_{i} is a (k)(\geq k)-transversal set of Δi+\Delta_{i}^{+}.

Claim 4.3.

|NG(Δi+)(u)NG(Δi+)(v)Ui|k1|N_{G(\Delta_{i}^{+})}(u)\cap N_{G(\Delta_{i}^{+})}(v)\cap U_{i}|\geq k-1 holds for each ii.

Proof of claim.

Let FF be a facet of Δi+\Delta_{i}^{+} containing both uu and vv. Such a facet always exists by Lemma 3.2(a). As UiU_{i} is a (k)(\geq k)-transversal set of Δi+\Delta_{i}^{+}, we have |FUi|k|F\cap U_{i}|\geq k. Thus we have |(Fuv)Ui|k2|(F-u-v)\cap U_{i}|\geq k-2. If |(Fuv)Ui|k1|(F-u-v)\cap U_{i}|\geq k-1, the claim follows as (Fuv)Ui(F-u-v)\cap U_{i} is included in NG(Δi+)(u)NG(Δi+)(v)N_{G(\Delta_{i}^{+})}(u)\cap N_{G(\Delta_{i}^{+})}(v). If |(Fuv)Ui|=k2|(F-u-v)\cap U_{i}|=k-2, by k3k\geq 3, we can pick w(Fuv)Uiw\in(F-u-v)\cap U_{i}. By Lemma 3.1(ii), there is another facet GΔi+G\in\Delta_{i}^{+} different from FF which includes FwF-w. Since |GUi|k|G\cap U_{i}|\geq k, the unique element xGFx\in G\setminus F must be in UiU_{i}. Now (Fuv+x)Ui(F-u-v+x)\cap U_{i} is the desired subset. ∎

For each ii, since u,vUiu,v\in U_{i}, UivU_{i}-v is a (k)(\geq k)-transversal set of Δi+/uv\Delta_{i}^{+}/uv and G(Δi+[Ui])/uv=G((Δi+/uv)[Uiv])G(\Delta_{i}^{+}[U_{i}])/uv=G((\Delta_{i}^{+}/uv)[U_{i}-v]). Also Δi+/uv\Delta_{i}^{+}/uv is a minimal (d1)(d-1)-cycle complex by Lemma 3.2 (c). Hence, by the induction hypothesis, G((Δi+/uv)[Uiv])G((\Delta_{i}^{+}/uv)[U_{i}-v]) is rigid in k\mathbb{R}^{k}. By Lemma 2.2 and Claim 4.3, G(Δi+[Ui])G(\Delta_{i}^{+}[U_{i}]) is rigid in k\mathbb{R}^{k}. Now the rigidity of G(Δ[U])=i=1mG(Δi+[Ui])G(\Delta[U])=\bigcup_{i=1}^{m}G(\Delta_{i}^{+}[U_{i}]) follows from Lemma 2.1 and Lemma 3.2 (d), (e). ∎

The balanced analogue of the lower bound theorem [2] has been investigated in [9, 13]. For positive integers n,dn,d with nn divisible by dd, a stacked cross-polytopal (d1)(d-1)-sphere on nn vertices is the connected sum of nd1\frac{n}{d}-1 copies of the boundary complex of the cross-polytope 𝒞d\mathcal{C}_{d}^{*}. A stacked cross-polytopal (d1)(d-1)-sphere Δ\Delta satisfies 2h2(Δ)=(d1)h1(Δ)2h_{2}(\Delta)=(d-1)h_{1}(\Delta), and the balanced lower bound theorem [13] asserts that 2h2(Δ)(d1)h1(Δ)2h_{2}(\Delta)\geq(d-1)h_{1}(\Delta) holds for any balanced normal (d1)(d-1)-pseudomanifolds with d3d\geq 3. Goff, Klee, Novik [9] showed that for a balanced pure simplicial complex Δ\Delta, this inequality follows from the rigidity of G(ΔT)G(\Delta_{T}) in 3\mathbb{R}^{3} for every 33-set T[d]T\subseteq[d]. Hence Theorem 4.1 implies the following generalization of balanced lower bound theorem to balanced minimal cycle complexes.

Corollary 4.4.

Let Δ\Delta be a balanced minimal (d1)(d-1)-cycle complex with d3d\geq 3. Then 2h2(Δ)(d1)h1(Δ)2h_{2}(\Delta)\geq(d-1)h_{1}(\Delta).

Remark 4.5.

Characterizing simplicial complexes achieving the tight equality in the lower bound theorem is also a well-studied problem. Klee and Novik [13] showed that for a balanced normal (d1)(d-1)-pseudomanifold Δ\Delta with d4d\geq 4, 2h2(Δ)=(d1)h1(Δ)2h_{2}(\Delta)=(d-1)h_{1}(\Delta) holds if and only if Δ\Delta is a stacked cross-polytopal (d1)(d-1)-sphere. It is interesting to know when the equality 2h2(Δ)=(d1)h1(Δ)2h_{2}(\Delta)=(d-1)h_{1}(\Delta) occurs for a balanced minimal (d1)(d-1)-cycle complex Δ\Delta.

5 Stanley-Reisner ring and rigidity

For an \mathbb{N}-graded algebra AA, the iith homogeneous component is denoted as AiA_{i}. An element of A1A_{1} is called a linear form. For a sequence of linear forms Θ=(θ1,,θd)\Theta=(\theta_{1},\ldots,\theta_{d}) of [Δ]\mathbb{R}[\Delta], the ideal of [Δ]\mathbb{R}[\Delta] generated by θ1,,θd\theta_{1},\ldots,\theta_{d} is denoted as Θ[Δ]\Theta\mathbb{R}[\Delta]. For a (d1)(d-1)-dimensional simplicial complex Δ\Delta, a sequence of dd linear forms θ1,,θd[Δ]1\theta_{1},\ldots,\theta_{d}\in\mathbb{R}[\Delta]_{1} is called a linear system of parameters (l.s.o.p. for short) for [Δ]\mathbb{R}[\Delta] if dim[Δ]/Θ[Δ]<\dim_{\mathbb{R}}\mathbb{R}[\Delta]/\Theta\mathbb{R}[\Delta]<\infty. For Stanley-Reisner ring, the following criterion of an l.s.o.p. is known (see [16]).

Lemma 5.1.

Let Δ\Delta be a (d1)(d-1)-dimensional simplicial complex and θ1,,θd[Δ]1\theta_{1},\ldots,\theta_{d}\in\mathbb{R}[\Delta]_{1} be linear forms. Define p:V(Δ)dp:V(\Delta)\rightarrow\mathbb{R}^{d} by the relation θi=vV(Δ)p(v)ixv\theta_{i}=\sum_{v\in V(\Delta)}p(v)_{i}x_{v} for i[d]i\in[d]. Then θ1,,θd\theta_{1},\ldots,\theta_{d} is an l.s.o.p. for [Δ]\mathbb{R}[\Delta] if and only if {p(v):vF}\{p(v):v\in F\} is linearly independent for every FΔF\in\Delta.

The connection between Stanley-Reisner ring theory and rigidity theory was pointed out by Lee [14]. Lee [14] defined the notion of linear and affine rr-stresses of a simplicial complex Δ\Delta and proved that, for a sequence of linear forms Θ=(θ1,,θd)\Theta=(\theta_{1},\ldots,\theta_{d}) and a linear form ω=vV(Δ)xv\omega=\sum_{v\in V(\Delta)}x_{v}, ([Δ]/Θ[Δ])r\left(\mathbb{R}[\Delta]/\Theta\mathbb{R}[\Delta]\right)_{r} (resp. ([Δ]/(Θ,ω)[Δ])r\left(\mathbb{R}[\Delta]/(\Theta,\omega)\mathbb{R}[\Delta]\right)_{r}) is linearly isomorphic to the space of linear (resp. affine) rr-stresses of Δ\Delta. For the remaining argument, the case of r=1,2r=1,2 is related, which is summarized as below.

Lemma 5.2 (Lee [14]).

Let Δ\Delta be a (d1)(d-1)-dimensional simplicial complex. Let Θ=(θ1,,θd)\Theta=(\theta_{1},\ldots,\theta_{d}) be a sequence of linear forms of [Δ]\mathbb{R}[\Delta]. Let ω=vV(Δ)xv\omega=\sum_{v\in V(\Delta)}x_{v} and define p:V(Δ)dp:V(\Delta)\rightarrow\mathbb{R}^{d} by the relation θi=vV(Δ)p(v)ixv\theta_{i}=\sum_{v\in V(\Delta)}p(v)_{i}x_{v} for i[d]i\in[d]. The followings hold:

  • (i)

    ([Δ]/Θ[Δ])1(\mathbb{R}[\Delta]/\Theta\mathbb{R}[\Delta])_{1} is linearly isomorphic to {tV(Δ):vV(Δ)tvp(v)=0}\{t\in\mathbb{R}^{V(\Delta)}:\sum_{v\in V(\Delta)}t_{v}p(v)=0\} .

  • (ii)

    ([Δ]/(Θ,ω)[Δ])2\left(\mathbb{R}[\Delta]/(\Theta,\omega)\mathbb{R}[\Delta]\right)_{2} is linearly isomorphic to kerR(G(Δ),p)\ker R(G(\Delta),p)^{\top}.

  • (iii)

    ([Δ]/Θ[Δ])2(\mathbb{R}[\Delta]/\Theta\mathbb{R}[\Delta])_{2} is linearly isomorphic to kerR(G(Δ){u},p)\ker R(G(\Delta)*\{u\},p^{\prime})^{\top}, where G(Δ){u}G(\Delta)*\{u\} is the cone graph of G(Δ)G(\Delta) and p:V(Δ){u}dp^{\prime}:V(\Delta)\cup\{u\}\rightarrow\mathbb{R}^{d} is the extension of pp defined by p(u)=0p^{\prime}(u)=0.

We have the following corollary. Although the result is already known, we include the proof for completeness.

Corollary 5.3.

Let Δ\Delta be a strongly connected (d1)(d-1)-dimensional simplicial complex. Let Θ=(θ1,,θd)\Theta=(\theta_{1},\ldots,\theta_{d}) be an l.s.o.p. for [Δ]\mathbb{R}[\Delta]. Let ω=vV(Δ)xv\omega=\sum_{v\in V(\Delta)}x_{v} and define p:V(Δ)dp:V(\Delta)\rightarrow\mathbb{R}^{d} by the relation θi=vV(Δ)p(v)ixv\theta_{i}=\sum_{v\in V(\Delta)}p(v)_{i}x_{v} for i[d]i\in[d]. Then the multiplication map (×ω):([Δ]/Θ[Δ])1([Δ]/Θ[Δ])2(\times\omega):\left(\mathbb{R}[\Delta]/\Theta\mathbb{R}[\Delta]\right)_{1}\rightarrow\left(\mathbb{R}[\Delta]/\Theta\mathbb{R}[\Delta]\right)_{2} is injective if and only if (G(Δ),p)(G(\Delta),p) is infinitesimally rigid in d\mathbb{R}^{d}.

Proof.

Since Θ\Theta is an l.s.o.p. for [Δ]\mathbb{R}[\Delta], p(V(Δ))p(V(\Delta)) linearly spans d\mathbb{R}^{d} by Lemma 5.1. Hence by Lemma 5.2 (i), dim([Δ]/Θ[Δ])1=f0(Δ)d=h1(Δ)\dim\left(\mathbb{R}[\Delta]/\Theta\mathbb{R}[\Delta]\right)_{1}=f_{0}(\Delta)-d=h_{1}(\Delta). By Lemma 5.2 (ii), Coker(×ω)([Δ]/(Θ,ω)[Δ])2\operatorname{Coker}(\times\omega)\cong\left(\mathbb{R}[\Delta]/(\Theta,\omega)\mathbb{R}[\Delta]\right)_{2} is linearly isomorphic to kerR(G(Δ),p)\ker R(G(\Delta),p)^{\top}. Hence dimCoker(×ω)=f1(Δ)df0(Δ)+(d+12)(=h2(Δ)h1(Δ))\dim\operatorname{Coker}(\times\omega)=f_{1}(\Delta)-df_{0}(\Delta)+\binom{d+1}{2}(=h_{2}(\Delta)-h_{1}(\Delta)) if and only if (G(Δ),p)(G(\Delta),p) is infinitesimally rigid in d\mathbb{R}^{d}. Thus, the statement follows if dim([Δ]/Θ[Δ])2=h2(Δ)\dim(\mathbb{R}[\Delta]/\Theta\mathbb{R}[\Delta])_{2}=h_{2}(\Delta) always holds.

To see this, let G(Δ){u}G(\Delta)*\{u\} be the cone graph of G(Δ)G(\Delta) and let p:V(Δ){u}dp^{\prime}:V(\Delta)\cup\{u\}\rightarrow\mathbb{R}^{d} be the point configuration as in Lemma 5.2 (iii). By Lemma 5.1, for each facet FF of Δ\Delta, F+uF+u is a clique in G(Δ){u}G(\Delta)*\{u\} and p(F+u)p^{\prime}(F+u) affinely spans d\mathbb{R}^{d}, so (G(Δ)[F+u],p|F+u)(G(\Delta)[F+u],p^{\prime}|_{F+u}) is infinitesimally rigid in d\mathbb{R}^{d}. Again by Lemma 5.1, for facets FF and GG of Δ\Delta with |FG|=d1|F\cap G|=d-1, p((FG)+u)p^{\prime}((F\cap G)+u) affinely spans a (d1)(d-1)-dimensional subspace. As Δ\Delta is strongly connected, one can order the facets of Δ\Delta as F1,,FmF_{1},\ldots,F_{m} in such a way that, for each i2i\geq 2, there is j<ij<i satisfying |FiFj|=d1|F_{i}\cap F_{j}|=d-1. Hence by the repeated application of Lemma 2.1, (G(Δ){u},p)(G(\Delta)*\{u\},p^{\prime}) is infinitesimally rigid in d\mathbb{R}^{d}. Therefore by Lemma 5.2(iii), we get dim([Δ]/Θ[Δ])2=dimkerR(G(Δ){u},p)=(f1(Δ)+f0(Δ))d(f0(Δ)+1)+(d+12)=h2(Δ)\dim(\mathbb{R}[\Delta]/\Theta\mathbb{R}[\Delta])_{2}=\dim\ker R(G(\Delta)*\{u\},p^{\prime})^{\top}=(f_{1}(\Delta)+f_{0}(\Delta))-d(f_{0}(\Delta)+1)+\binom{d+1}{2}=h_{2}(\Delta) as desired. ∎

Let us recall the definition of 𝒂\bm{a}-balancedness. For 𝒂>0m\bm{a}\in\mathbb{Z}_{>0}^{m} with i=1mai=d\sum_{i=1}^{m}a_{i}=d, a (d1)(d-1)-dimensional simplicial complex Δ\Delta on the vertex set V(Δ)V(\Delta) is 𝐚\bm{a}-balanced if there is a map κ:V(Δ)[m]\kappa:V(\Delta)\rightarrow[m] satisfying |Fκ1(i)|ai|F\cap\kappa^{-1}(i)|\leq a_{i} for any FΔF\in\Delta. We call such a map κ\kappa an 𝐚\bm{a}-coloring of Δ\Delta. If Δ\Delta is pure, this condition is equivalent to |Fκ1(i)|=ai|F\cap\kappa^{-1}(i)|=a_{i} for any facet FF of Δ\Delta. Stanley [16] showed that an 𝒂\bm{a}-balanced simplicial complex Δ\Delta admits a special type of l.s.o.p. for [Δ]\mathbb{R}[\Delta]. We remark that we always use the term “linear forms” to mean degree one forms in the usual \mathbb{N}-grading.

Proposition 5.4 (Stanley [16]).

Let Δ\Delta be an 𝒂\bm{a}-balanced (d1)(d-1)-dimensional simplicial complex for 𝒂>0m\bm{a}\in\mathbb{Z}_{>0}^{m} and κ:V(Δ)[m]\kappa:V(\Delta)\rightarrow[m] be an 𝒂\bm{a}-coloring of Δ\Delta. Make [Δ]\mathbb{R}[\Delta] into an m\mathbb{N}^{m}-graded algebra by defining degxv=𝒆κ(v)m\deg x_{v}=\bm{e}_{\kappa(v)}\in\mathbb{N}^{m}. Then [Δ]\mathbb{R}[\Delta] has an l.s.o.p. Θ=(θ1,,θd)\Theta=(\theta_{1},\ldots,\theta_{d}) such that each θi\theta_{i} is homogeneous in m\mathbb{N}^{m}-grading and exactly aia_{i} elements among Θ\Theta are of degree 𝒆i\bm{e}_{i} for each i[m]i\in[m].

An l.s.o.p. satisfying the property in Proposition 5.4 is called an 𝐚\bm{a}-colored s.o.p. We now prove the equivalence of Conjecture 1.1 and Conjecture 1.2. We recall the definition of (κ,𝒂)(\kappa,\bm{a})-sparse rigidity. Let GG be a graph, κ:V(G)[m]\kappa:V(G)\rightarrow[m] a map, and 𝒂>0m\bm{a}\in\mathbb{Z}_{>0}^{m} an integer vector with i=1mai=d\sum_{i=1}^{m}a_{i}=d. Let Hi:=0××ai××0d=i=1maiH_{i}:=0\times\cdots\times\mathbb{R}^{a_{i}}\times\cdots\times 0\subseteq\mathbb{R}^{d}=\prod_{i=1}^{m}\mathbb{R}^{a_{i}} for i[m]i\in[m]. We say that a point configuration p:V(G)dp:V(G)\rightarrow\mathbb{R}^{d} is (κ,𝐚)(\kappa,\bm{a})-sparse if p(v)Hκ(v)p(v)\in H_{\kappa(v)} for all vV(G)v\in V(G), and GG is (κ,𝐚)(\kappa,\bm{a})-sparse rigid if (G,p)(G,p) is infinitesimally rigid in d\mathbb{R}^{d} for some (κ,𝒂)(\kappa,\bm{a})-sparse point configuration pp. Note that by Lemma 5.1, under the identification θi=vV(Δ)p(v)ix(v)\theta_{i}=\sum_{v\in V(\Delta)}p(v)_{i}x(v) for i[d]i\in[d], θ1,,θd\theta_{1},\ldots,\theta_{d} is an 𝒂\bm{a}-colored s.o.p. for [Δ]\mathbb{R}[\Delta] if and only if p:V(Δ)dp:V(\Delta)\rightarrow\mathbb{R}^{d} is (κ,𝒂)(\kappa,\bm{a})-sparse and {p(v):vF}\{p(v):v\in F\} is linearly independent for every FΔF\in\Delta.

Proof of equivalence of Conjecture 1.1 and Conjecture 1.2.

If Conjecture 1.1 holds, for any generic choice of ω\omega, (×ω):([Δ]/Θ[Δ])1([Δ]/Θ[Δ])2(\times\omega):\left(\mathbb{R}[\Delta]/\Theta\mathbb{R}[\Delta]\right)_{1}\rightarrow\left(\mathbb{R}[\Delta]/\Theta\mathbb{R}[\Delta]\right)_{2} is injective. So, in Conjecture 1.1, we may add extra constraints that ω=vV(Δ)avxv\omega=\sum_{v\in V(\Delta)}a_{v}x_{v} with av0a_{v}\neq 0 for all vV(Δ)v\in V(\Delta). Moreover, by setting avxva_{v}x_{v} to xvx_{v}, we may further suppose ω=vV(Δ)xv\omega=\sum_{v\in V(\Delta)}x_{v} in the statement of Conjecture 1.1. Now, the equivalence of Conjecture 1.1 and Conjecture 1.2 follows from Corollary 5.3. ∎

6 (κ,𝒂)(\kappa,\bm{a})-sparse rigidity of minimal cycle complexes for 𝒂2m\bm{a}\in\mathbb{Z}_{\geq 2}^{m}

In this section we shall verify Conjecture 1.2 in the case when ai2a_{i}\geq 2 for all ii as follows.

Theorem 6.1.

Let 𝒂=(a1,,am)>0m\bm{a}=(a_{1},\ldots,a_{m})\in\mathbb{Z}_{>0}^{m} be a positive integer vector with ai2a_{i}\geq 2 for all i[m]i\in[m] and d:=i=1mai3d:=\sum_{i=1}^{m}a_{i}\geq 3. Let Δ\Delta be an 𝒂\bm{a}-balanced minimal (d1)(d-1)-cycle complex with an 𝒂\bm{a}-coloring κ\kappa. Then G(Δ)G(\Delta) is (κ,𝒂)(\kappa,\bm{a})-sparse rigid.

The proof of Theorem 6.1 consists of several lemmas. To state lemmas, we introduce LL-sparse rigidity as a generalization of (κ,𝒂)(\kappa,\bm{a})-sparse rigidity. For x=(x1,,xd)dx=(x_{1},\ldots,x_{d})\in\mathbb{R}^{d}, we denote suppx:={i[d]:xi0}\operatorname{supp}x:=\{i\in[d]:x_{i}\neq 0\}. Let GG be a graph and L:V(G)2[d]L:V(G)\rightarrow 2^{[d]} be a map. A point configuration p:V(G)dp:V(G)\rightarrow\mathbb{R}^{d} is LL-sparse if suppp(v)L(v)\operatorname{supp}p(v)\subseteq L(v) for every vV(G)v\in V(G). We say that an LL-sparse point configuration p:V(G)dp:V(G)\rightarrow\mathbb{R}^{d} is generic over \mathbb{Q} if {p(v)j:vV(G),jL(v)}\{p(v)_{j}:v\in V(G),j\in L(v)\} is algebraically independent over \mathbb{Q}. A graph GG is LL-sparse rigid if (G,p)(G,p) is infinitesimally rigid in d\mathbb{R}^{d} for some LL-sparse point configuration pp. Note that for κ:V(G)[m]\kappa:V(G)\rightarrow[m] and 𝒂>0m\bm{a}\in\mathbb{Z}_{>0}^{m} with d=i=1maid=\sum_{i=1}^{m}a_{i}, (κ,𝒂)(\kappa,\bm{a})-sparse rigidity coincides with Lκ,𝒂L_{\kappa,\bm{a}}-sparse rigidity, where Lκ,𝒂:V(G)[d]L_{\kappa,\bm{a}}:V(G)\rightarrow\mathbb{R}^{[d]} is defined by partitioning [d][d] into disjoint sets I1,,ImI_{1},\ldots,I_{m} with |Ii|=ai|I_{i}|=a_{i} and letting Lκ,𝒂(v)=Iκ(v)L_{\kappa,\bm{a}}(v)=I_{\kappa(v)}. A (κ,𝒂)(\kappa,\bm{a})-sparse point configuration pp is said to be generic if pp is generic as an Lκ,𝒂L_{\kappa,\bm{a}}-sparse point configuration. Eftekhari et al. [6] addressed the special setting of LL-sparse rigidity in which, given XV(G)X\subseteq V(G), L(v)=[d1]L(v)=[d-1] for vXv\in X and L(v)=[d]L(v)=[d] for vXv\not\in X, and they gave a combinatorial characterization when d=2d=2. Cook et al. [11] gave a combinatorial characterization for (κ,𝒂)(\kappa,\bm{a})-sparse rigidity of maximal planar graphs GG with 𝒂=(2,1)\bm{a}=(2,1) and κ:V(G){1,2}\kappa:V(G)\rightarrow\{1,2\} in which κ1(2)\kappa^{-1}(2) is a stable set in GG.

A set of points XdX\subseteq\mathbb{R}^{d} is affinely independent if the dimension of the affine span of XX is |X|1|X|-1.

Lemma 6.2.

Let GG be a graph and UV(G)U\subseteq V(G) be a subset of vertices with size at most d+1d+1. Let L:V(G)2[d]L:V(G)\rightarrow 2^{[d]} be a map, and p:V(G)dp:V(G)\rightarrow\mathbb{R}^{d} be a generic LL-sparse point configuration. Then p(U)p(U) is affinely independent if and only if |vWL(v)|+1|W||\bigcup_{v\in W}L(v)|+1\geq|W| for every subset WUW\subseteq U.

Proof.

Let k=|U|k=|U| and U={v1,,vk}U=\{v_{1},\ldots,v_{k}\}. Let A(d+1)×kA\in\mathbb{R}^{(d+1)\times k} be the matrix defined by

A=[p(v1)p(vk)11].\displaystyle A=\begin{bmatrix}p(v_{1})&\cdots&p(v_{k})\\ 1&\cdots&1\\ \end{bmatrix}.

Then p(U)p(U) is affinely independent if and only if rankA=k\rank A=k. Consider the bipartite graph G(A)G(A) between UU and [d+1][d+1] such that there is an edge between vUv\in U and i[d+1]i\in[d+1] if and only if Ai,v0A_{i,v}\neq 0.

For R[d+1]R\subseteq[d+1] with |R|=k|R|=k, let PRP_{R} be the determinant of submatrices of AA indexed by RR. Then

PR=σsσj=1kAvj,σ(vj),P_{R}=\sum_{\sigma}s_{\sigma}\prod_{j=1}^{k}A_{v_{j},\sigma(v_{j})}, (4)

where the sum is taken over all perfect matchings σ\sigma of G(A)[UR]G(A)[U\sqcup R] and sσ=±1s_{\sigma}=\pm 1. When PRP_{R} is considered as a polynomial of {p(v)i:vU,iL(v)}\{p(v)_{i}:v\in U,i\in L(v)\}, each monomial appears at most once in the summand of the right hand side of (4). Since pp is a generic LL-sparse point configuration, PR0P_{R}\neq 0 if and only if there is a perfect matching in G(A)[UR]G(A)[U\sqcup R].

Hence rankA=k\rank A=k if and only if there is a size |U||U| matching in the bipartite graph G(A)G(A), which is equivalent to the condition given in the statement by Hall’s theorem. ∎

Let GG be a graph and L:V(G)2[d]L:V(G)\rightarrow 2^{[d]} be a map. For UV(G)U\subseteq V(G), consider the following “Hall condition” (H):

  • (H)

    |vWL(v)|+1|W||\bigcup_{v\in W}L(v)|+1\geq|W| holds for every subset WUW\subseteq U.

Lemma 6.3.

Let GG be a graph and uvE(G)uv\in E(G) be an edge satisfying |NG(u)NG(v)|d1|N_{G}(u)\cap N_{G}(v)|\geq d-1. Let L:V(G)2[d]L:V(G)\rightarrow 2^{[d]} be a map satisfying L(u)=L(v)L(u)=L(v), and define LL^{\prime} as the restriction of LL to V(G/uv)V(G/uv). Suppose that there is a (d1)(d-1)-set CNG(u)NG(v)C\subseteq N_{G}(u)\cap N_{G}(v) such that C+u+vC+u+v satisfies the condition (H) with respect to GG and LL. Then GG is LL-sparse rigid if G/uvG/uv is LL^{\prime}-sparse rigid.

Proof.

Suppose that G/uvG/uv is LL^{\prime}-sparse rigid. Let p:V(G/uv)dp:V(G/uv)\rightarrow\mathbb{R}^{d} be a generic LL^{\prime}-sparse point configuration. Then (G/uv,p)(G/uv,p) is infinitesimally rigid in d\mathbb{R}^{d}. By Lemma 6.2, the condition (H) of C+u+vC+u+v implies that the set of points {p(w):wC{u}}\{p(w):w\in C\cup\{u\}\} is affinely independent. Hence {p(w)p(u):wC}\{p(w)-p(u):w\in C\} is linearly independent.

Among vectors zdz\in\mathbb{R}^{d} with supp(z)=L(v)\operatorname{supp}(z)=L(v), pick a generic one zz. Then, by Lemma 6.2, the condition (H) of C+u+vC+u+v implies that {p(w):wC{u}}{p(u)+z}\{p(w):w\in C\cup\{u\}\}\cup\{p(u)+z\} is affinely independent. Hence zz is not in the linear span of {p(w)p(u):wC}\{p(w)-p(u):w\in C\}. Thus by Lemma 2.2, there is an extension pp^{\prime} of pp such that p(v)=p(u)+tzp^{\prime}(v)=p(u)+tz for some tt\in\mathbb{R} such that (G,p)(G,p^{\prime}) is infinitesimally rigid. As pp^{\prime} is an LL-sparse point configuration, GG is LL-rigid. ∎

As a special case, we have the following corollary for (κ,𝒂)(\kappa,\bm{a})-sparse rigidity.

Corollary 6.4.

Let GG be a graph and uvE(G)uv\in E(G) be an edge satisfying |NG(u)NG(v)|d1|N_{G}(u)\cap N_{G}(v)|\geq d-1. Let κ:V(G)[m]\kappa:V(G)\rightarrow[m] be a map satisfying κ(u)=κ(v)\kappa(u)=\kappa(v), and let κ\kappa^{\prime} be the restriction of κ\kappa to V(G/uv)V(G/uv). For UV(G)U\subseteq V(G), define tκ(U):=(|Uκ1(i)|)i0mt_{\kappa}(U):=(|U\cap\kappa^{-1}(i)|)_{i}\in\mathbb{Z}_{\geq 0}^{m}. For 𝒂>0m\bm{a}\in\mathbb{Z}_{>0}^{m}, suppose that there is a (d1)(d-1)-set CNG(u)NG(v)C\subseteq N_{G}(u)\cap N_{G}(v) such that tκ(C+u+v)=𝒂+𝒆jt_{\kappa}(C+u+v)=\bm{a}+\bm{e}_{j} for some j[m]j\in[m], where 𝒆j\bm{e}_{j} denotes the jjth unit coordinate vector. Then GG is (κ,𝒂)(\kappa,\bm{a})-sparse rigid if G/uvG/uv is (κ,𝒂)(\kappa^{\prime},\bm{a})-sparse rigid.

Proof.

One can easily check that in the setting of (κ,𝒂)(\kappa,\bm{a})-sparse rigidity, (d+1)(d+1)-set UU satisfies the condition (H) if and only if tκ(U)=𝒂+𝒆jt_{\kappa}(U)=\bm{a}+\bm{e}_{j} for some j[m]j\in[m]. Now the statement follows from Lemma 6.3. ∎

Proof of Theorem 6.1.

We prove the statement by the induction on |V(Δ)||V(\Delta)|. When Δ\Delta is trivial, G(Δ)=KdG(\Delta)=K_{d}, a complete graph on dd vertices. By Lemma 6.2, p(V(G))p(V(G)) spans (d1)(d-1)-dimensional affine space for any generic (κ,𝒂)(\kappa,\bm{a})-sparse pp. Hence G(Δ)G(\Delta) is (κ,𝒂)(\kappa,\bm{a})-sparse rigid.

Suppose that Δ\Delta is nontrivial. By a12a_{1}\geq 2, we can pick uvE(G)uv\in E(G) with κ(u)=κ(v)=1\kappa(u)=\kappa(v)=1. Let Δ1+,,Δt+\Delta_{1}^{+},\ldots,\Delta_{t}^{+} be the minimal (d1)(d-1)-cycle complexes given in Lemma 3.2 with respect to Δ\Delta and uvuv. For UV(G)U\subseteq V(G), define the type of UV(G)U\subseteq V(G) by tκ(U):=(|Uκ1(i)|)i0mt_{\kappa}(U):=(|U\cap\kappa^{-1}(i)|)_{i}\in\mathbb{Z}_{\geq 0}^{m}.

Claim 6.5.

For each ii, there is a (d1)(d-1)-set CNG(Δi+)(u)NG(Δi+)(v)C\subseteq N_{G(\Delta_{i}^{+})}(u)\cap N_{G(\Delta_{i}^{+})}(v) such that tκ(C+u+v)=𝒂+𝒆kt_{\kappa}(C+u+v)=\bm{a}+\bm{e}_{k} for some k[m]k\in[m].

Proof of claim.

Let FF^{*} be a facet of Δi+\Delta_{i}^{+} containing uu and vv, which exists by Lemma 3.2 (a). By Lemma 3.1 (ii), for every wFuvw\in F^{*}-u-v, there exists x(w)x(\neq w) such that Fw+xF^{*}-w+x is a facet of Δi+\Delta_{i}^{+}. Then C:=Fuv+xC:=F^{*}-u-v+x is included in the common neighborhood of uu and vv in G(Δi+)G(\Delta_{i}^{+}). We show that for an appropriate choice of ww, tκ(C+u+v)=𝒂+𝒆kt_{\kappa}(C+u+v)=\bm{a}+\bm{e}_{k} holds for some k[m]k\in[m].

By Lemma 3.2 (b), for every facet FF of Δi+\Delta_{i}^{+} containing uu and vv, FvF-v is a (d2)(d-2)-face of Δ\Delta, and thus tκ(Fv)=𝒂𝒆jt_{\kappa}(F-v)=\bm{a}-\bm{e}_{j} for some j[m]j\in[m]. Hence, we have the following property (\star):

(\star) for every facet FF of Δi+\Delta_{i}^{+} containing uu and vv, tκ(F)=𝒂𝒆j+𝒆1t_{\kappa}(F)=\bm{a}-\bm{e}_{j}+\bm{e}_{1} for some j[m]j\in[m].

By property (\star) of FF^{*}, tκ(F)=𝒂𝒆j+𝒆1t_{\kappa}(F^{*})=\bm{a}-\bm{e}_{j}+\bm{e}_{1} for some j[m]j\in[m]. If j=1j=1, for any choice of ww, we have tκ(C+u+v)=𝒂+𝒆κ(x)t_{\kappa}(C+u+v)=\bm{a}+\bm{e}_{\kappa(x)} as desired. If j1j\neq 1, pick ww from Fκ1(j)F^{*}\cap\kappa^{-1}(j), which is not empty by aj2a_{j}\geq 2. Since Fw+xF^{*}-w+x also satisfies (\star), it follows that κ(x)=j\kappa(x)=j. Hence we have tκ(C+u+v)=𝒂+𝒆1t_{\kappa}(C+u+v)=\bm{a}+\bm{e}_{1} as desired. ∎

Let κi\kappa_{i} (resp. κi\kappa_{i}^{\prime}) be the restriction of κ\kappa to V(Δi+)V(\Delta_{i}^{+}) (resp. V(Δi+/uv)V(\Delta_{i}^{+}/uv)). Then Δi+/uv\Delta_{i}^{+}/uv is 𝒂\bm{a}-balanced and κi\kappa_{i}^{\prime} is an 𝒂\bm{a}-coloring of Δi+/uv\Delta_{i}^{+}/uv. Δi+/uv\Delta_{i}^{+}/uv is also a minimal (d1)(d-1)-cycle complex by Lemma 3.2 (c). Thus by induction hypothesis, G(Δi+/uv)G(\Delta_{i}^{+}/uv) is (κi,𝒂)(\kappa_{i}^{\prime},\bm{a})-sparse rigid. By Corollary 6.4 and Claim 6.5, G(Δi+)G(\Delta_{i}^{+}) is (κi,𝒂)(\kappa_{i},\bm{a})-sparse rigid for each ii.

By Lemma 6.2 and the property (\star), for a facet FF of Δi+\Delta_{i}^{+} and a generic (κ,𝒂)(\kappa,\bm{a})-sparse point configuration pp, p(F)p(F) spans a (d1)(d-1)-dimensional affine subspace. Hence by Lemma 3.2 (d) and (e), we can deduce that G(Δ)=i=1tG(Δi+)G(\Delta)=\bigcup_{i=1}^{t}G(\Delta_{i}^{+}) is (κ,𝒂)(\kappa,\bm{a})-sparse rigid by the repeated application of Lemma 2.1. ∎

7 (κ,𝒂)(\kappa,\bm{a})-sparse rigidity of homology manifolds

In this section, we prove that Conjecture 1.2 holds for any 𝒂(d1),(1,d1)\bm{a}\neq(d-1),(1,d-1) if d4d\geq 4 and Δ\Delta is a homology (d1)(d-1)-manifolds.

Conjecture 1.2 was verified for balanced simplicial 22-spheres by Cook et al. [11].

Theorem 7.1 (Cook et al. [11]).

For a balanced simplicial 22-sphere Δ\Delta with a proper 33-coloring κ\kappa, G(Δ)G(\Delta) is (κ,𝒂)(\kappa,\bm{a})-sparse rigid.

Kalai [12] defined a class 𝒞d\mathcal{C}_{d} of (d1)(d-1)-dimensional pseudomanifolds for d3d\geq 3 as follows: 𝒞3\mathcal{C}_{3} is the class of simplicial 22-sphere, and for d4d\geq 4, a (d1)(d-1)-dimensional pseudomanifold Δ\Delta belongs to 𝒞d\mathcal{C}_{d} if lkΔ(v)𝒞d1\operatorname{lk}_{\Delta}(v)\in\mathcal{C}_{d-1} for every vV(Δ)v\in V(\Delta). For d4d\geq 4, 𝒞d\mathcal{C}_{d} includes all homology (d1)(d-1)-manifolds (over any field). Kalai [12] showed that if Δ𝒞d\Delta\in\mathcal{C}_{d} (d3d\geq 3), G(Δ)G(\Delta) is rigid in d\mathbb{R}^{d}. In the case of (κ,𝒂)(\kappa,\bm{a})-sparse rigidity, we have the following theorem.

Theorem 7.2.

For d3d\geq 3, let 𝒂=(a1,,am)>0m\bm{a}=(a_{1},\ldots,a_{m})\in\mathbb{Z}_{>0}^{m} be a positive integer vector with i=1mai=d\sum_{i=1}^{m}a_{i}=d. Let Δ\Delta be an 𝒂\bm{a}-balanced pseudomanifold satisfying Δ𝒞d\Delta\in\mathcal{C}_{d}, and let κ\kappa be an 𝒂\bm{a}-coloring of Δ\Delta. If 𝒂(d1,1),(1,d1)\bm{a}\neq(d-1,1),(1,d-1), G(Δ)G(\Delta) is (κ,𝒂)(\kappa,\bm{a})-sparse rigid.

In the proof of Theorem 7.2, we use cone lemma for (κ,𝒂)(\kappa,\bm{a})-sparse rigidity, which we first prove in the generality of LL-sparse rigidity.

Lemma 7.3.

Let GG be a graph and G{v}G*\{v\} be its cone graph. Let L:V(G){v}2[d+1]L:V(G)\cup\{v\}\rightarrow 2^{[d+1]} be a map with L(v)L(v)\neq\emptyset. Suppose that there is iL(v)i\in L(v) such that, for each uV(G)u\in V(G), either iL(u)i\not\in L(u) or |L(v)L(u)|1|L(v)\setminus L(u)|\leq 1 holds. Define L:V(G)2[d+1]{i}L^{\prime}:V(G)\rightarrow 2^{[d+1]\setminus\{i\}} by L(u):=L(u)L(v){i}L^{\prime}(u):=L(u)\cup L(v)\setminus\{i\} if iL(u)i\in L(u) and L(u):=L(u)L^{\prime}(u):=L(u) otherwise, and identify 2[d+1]{i}2^{[d+1]\setminus\{i\}} with 2[d]2^{[d]}. Then GG is LL^{\prime}-sparse rigid if and only if G{v}G*\{v\} is LL-sparse rigid.

Proof.

Identify H:={xd+1:xi=0}H:=\{x\in\mathbb{R}^{d+1}:x_{i}=0\} with d\mathbb{R}^{d}. Suppose that G{v}G*\{v\} is LL-sparse rigid. Let p:V(G){v}d+1p:V(G)\cup\{v\}\rightarrow\mathbb{R}^{d+1} be a generic LL-sparse configuration. Then (G{v},p)(G*\{v\},p) is infinitesimally rigid in d+1\mathbb{R}^{d+1}. As iL(v)i\in L(v), we have p(v)Hp(v)\not\in H, and p(u)p(v)p(u)-p(v) is not parallel to HH for any uV(G)u\in V(G). For each uV(G)u\in V(G), let pH(u)p_{H}(u) be the intersection of HH and the line passing through p(u)p(u) and p(v)p(v). By Lemma 2.3, (G,pH)(G,p_{H}) is infinitesimally rigid in d\mathbb{R}^{d}. By the definition of LL^{\prime}, pHp_{H} is LL^{\prime}-sparse. Thus GG is LL^{\prime}-sparse rigid.

To see the other direction, suppose that GG is LL^{\prime}-sparse rigid. There is q:V(G)dHq:V(G)\rightarrow\mathbb{R}^{d}\cong H such that (G,q)(G,q) is infinitesimally rigid in d\mathbb{R}^{d} and qq is LL^{\prime}-sparse and suppq(u)=L(u)\operatorname{supp}q(u)=L^{\prime}(u) for all uV(G)u\in V(G). We define p:V{v}d+1p:V\cup\{v\}\rightarrow\mathbb{R}^{d+1} as follows. Let p(v)d+1p(v)\in\mathbb{R}^{d+1} be a point such that supp(p(v))=L(v)\operatorname{supp}(p(v))=L(v) and p(v)jq(u)jp(v)_{j}\neq q(u)_{j} for any jL(v)j\in L(v) and uV(G)u\in V(G). For uV(G)u\in V(G), if iL(u)i\not\in L(u) or L(u)L(v)L(u)\subseteq L(v), let p(u):=q(u)p(u):=q(u). For uV(G)u\in V(G), if iL(u)i\in L(u) and L(u)L(v)L(u)\not\subseteq L(v), by the assumption on LL, we must have L(v)L(u)={j}L(v)\setminus L(u)=\{j\} for some j(i)j(\neq i). For such uu, as q(u)j0,p(v)jq(u)_{j}\neq 0,p(v)_{j}, there is a unique real number t(0,1)t(\neq 0,1)\in\mathbb{R} such that (tq(u)+(1t)p(v))j=0(tq(u)+(1-t)p(v))_{j}=0, so let p(u):=tq(u)+(1t)p(v)p(u):=tq(u)+(1-t)p(v). One can see that pp is LL-sparse. By Lemma 2.3,(G{v},p),(G*\{v\},p) is infinitesimally rigid in d+1\mathbb{R}^{d+1}. Hence G{v}G*\{v\} is LL-sparse rigid. ∎

We obtain the following corollary of Lemma 7.3 for (κ,𝒂)(\kappa,\bm{a})-sparse rigidity.

Corollary 7.4.

Let GG be a graph, κ:V(G)[m]\kappa:V(G)\rightarrow[m] a map, and 𝒂>0m\bm{a}\in\mathbb{Z}_{>0}^{m} a positive integer vector. Let G{v}G*\{v\} be the cone graph of GG. Suppose that GG is (κ,𝒂)(\kappa,\bm{a})-sparse rigid. We have the followings:

  • (i)

    For the extension κ:V(G){v}[m+1]\kappa^{\prime}:V(G)\cup\{v\}\rightarrow[m+1] of κ\kappa defined by κ(v)=m+1\kappa^{\prime}(v)=m+1, G{v}G*\{v\} is (κ,(𝒂,1))(\kappa^{\prime},(\bm{a},1))-sparse rigid.

  • (ii)

    For the extension κ:V(G){v}[m]\kappa^{\prime}:V(G)\cup\{v\}\rightarrow[m] of κ\kappa defined by κ(v)=i\kappa^{\prime}(v)=i for some i[m]i\in[m], G{v}G*\{v\} is (κ,𝒂+𝒆i)(\kappa^{\prime},\bm{a}+\bm{e}_{i})-sparse rigid.

Proof of Theorem 7.2.

If m=1m=1, the statement is the result of Kalai [12]. If m=2m=2, the statement follows from Theorem 6.1. We prove the statement for m3m\geq 3 by the induction on dd. If d=3d=3, we have 𝒂=(1,1,1)\bm{a}=(1,1,1), and the statement follows from Theorem 7.1. Suppose that d4d\geq 4 and Δ𝒞d\Delta\in\mathcal{C}_{d} is 𝒂\bm{a}-balanced. Let κ\kappa be an 𝒂\bm{a}-coloring of Δ\Delta. Define UV(G)U\subseteq V(G) as follows: if aj=1a_{j}=1 for all jj, let U:=κ1({1,2})U:=\kappa^{-1}(\{1,2\}), and if aj2a_{j}\geq 2 for some j[m]j\in[m], let U:=κ1(j)U:=\kappa^{-1}(j).

Claim 7.5.

For each vUv\in U, G(stΔ(v))G(\operatorname{st}_{\Delta}(v)) is (κ|V(stΔ(v)),𝒂)(\kappa|_{V(\operatorname{st}_{\Delta}(v))},\bm{a})-sparse rigid.

Proof of claim.

First consider the case in which aj=1a_{j}=1 for all jj. In this case m=d4m=d\geq 4. Let κ:V(lkΔ(v))[m1]\kappa^{\prime}:V(\operatorname{lk}_{\Delta}(v))\rightarrow[m-1] be the restriction of κ\kappa to V(lkΔ(v))V(\operatorname{lk}_{\Delta}(v)), where [m1][m-1] is identified with [m]{κ(v)}[m]\setminus\{\kappa(v)\}. Then κ\kappa^{\prime} is a 𝒃\bm{b}-coloring of lkΔ(v)\operatorname{lk}_{\Delta}(v), where 𝒃=(1,,1)m1\bm{b}=(1,\ldots,1)\in\mathbb{Z}^{m-1}. As lkΔ(v)𝒞d1\operatorname{lk}_{\Delta}(v)\in\mathcal{C}_{d-1}, by the induction hypothesis, G(lkΔ(v))G(\operatorname{lk}_{\Delta}(v)) is (κ,𝒃)(\kappa^{\prime},\bm{b})-sparse rigid. Thus by Corollary 7.4 (i), the claim follows.

Now consider the case in which aj2a_{j}\geq 2 for some jj. Then κ|V(lkΔ(v))\kappa|_{V(\operatorname{lk}_{\Delta}(v))} is an (𝒂𝒆j)(\bm{a}-\bm{e}_{j})-coloring of lkΔ(v)\operatorname{lk}_{\Delta}(v), so G(lkΔ(v))G(\operatorname{lk}_{\Delta}(v)) is (κ|V(lkΔ(v)),𝒃)(\kappa|_{V(\operatorname{lk}_{\Delta}(v))},\bm{b})-sparse rigid by the induction hypothesis. Thus by Corollary 7.4 (ii), the claim follows. ∎

By definition of UU, |UF|2|U\cap F|\geq 2 for every facet FF of Δ\Delta. As Δ\Delta is a pseudomanifold, Δ\Delta is strongly connected. Hence G(Δ)[U]G(\Delta)[U] is connected. Thus, the vertices of UU can be ordered as v1,,vtv_{1},\ldots,v_{t} in such a way that j<istΔ(vj)\bigcup_{j<i}\operatorname{st}_{\Delta}(v_{j}) and stΔ(vi)\operatorname{st}_{\Delta}(v_{i}) share at least one facet for each i2i\geq 2. By Lemma 6.2, for any generic (κ,𝒂)(\kappa,\bm{a})-sparse configuration p:V(Δ)dp:V(\Delta)\rightarrow\mathbb{R}^{d} and any facet FF of Δ\Delta, p(F)p(F) spans a (d1)(d-1)-dimensional affine subspace. Hence, by Lemma 2.1 and Claim 7.5, G(Δ)=vUG(stΔ(v))G(\Delta)=\bigcup_{v\in U}G(\operatorname{st}_{\Delta}(v)) is (κ,𝒂)(\kappa,\bm{a})-sparse rigid. ∎

The assumption 𝒂(d1),(1,d1)\bm{a}\neq(d-1),(1,d-1) in Theorem 7.2 is necessary. Cook et al. [11] showed that there is a (2,1)(2,1)-balanced simplicial 22-sphere Δ\Delta and its (2,1)(2,1)-coloring κ\kappa such that G(Δ)G(\Delta) is not (κ,𝒂)(\kappa,\bm{a})-sparse rigid. Based on the construction given in [11], for each d3d\geq 3, we give a (d1,1)(d-1,1)-balanced simplicial (d1)(d-1)-sphere Δ\Delta and a (d1,1)(d-1,1)-coloring κ\kappa of Δ\Delta such that G(Δ)G(\Delta) is not (κ,(d1,1))(\kappa,(d-1,1))-sparse rigid. Moreover, dimkerR(G(Δ),p)(d+12)\dim\ker R(G(\Delta),p)-\binom{d+1}{2} can be arbitrarily large for any (κ,(d1,1))(\kappa,(d-1,1))-sparse point configuration pp.

Example 7.6.

Let Γ\Gamma be a stacked (d1)(d-1)-simplicial sphere. Let Δ\Delta be the simplicial (d1)(d-1)-sphere obtained from Γ\Gamma by subdividing all facets of Γ\Gamma. Then Δ\Delta is a stacked (d1)(d-1)-sphere with f0(Δ)+fd1(Δ)f_{0}(\Delta)+f_{d-1}(\Delta) vertices. Define κ:V(Δ){1,2}\kappa:V(\Delta)\rightarrow\{1,2\} by κ(v)=1\kappa(v)=1 if vV(Γ)v\in V(\Gamma) and κ(v)=2\kappa(v)=2 otherwise. Then κ\kappa is a (d1,1)(d-1,1)-coloring of Δ\Delta. Let p:V(Δ)dp:V(\Delta)\rightarrow\mathbb{R}^{d} be a (κ,(d1,1))(\kappa,(d-1,1))-sparse point configuration of G(Δ)G(\Delta). As Δ\Delta is a stacked sphere, we have dimkerR(G(Δ),p)=dimkerR(G(Δ),p)(d+12)\dim\ker R(G(\Delta),p)^{\top}=\dim\ker R(G(\Delta),p)-\binom{d+1}{2}. The subframework (G(Γ),p|V(Γ))(G(\Gamma),p|_{V(\Gamma)}) of (G(Δ),p)(G(\Delta),p) is in a (d1)(d-1)-dimensional affine subspace and has df0(Γ)(d+12)df_{0}(\Gamma)-\binom{d+1}{2} edges. Thus the dimension of the linear space of the equilibrium stresses of (G(Γ),p|V(Γ))(G(\Gamma),p|_{V(\Gamma)}) is at least df0(Γ)(d+12)((d1)f0(Γ)(d2))=f0(Γ)ddf_{0}(\Gamma)-\binom{d+1}{2}-((d-1)f_{0}(\Gamma)-\binom{d}{2})=f_{0}(\Gamma)-d. Hence dimkerR(G(Δ),p)(d+12)\dim\ker R(G(\Delta),p)-\binom{d+1}{2} can be arbitrarily large by setting f0(Γ)f_{0}(\Gamma) large enough.

8 Further observations on LL-sparse rigidity

We can prove further results similar to Theorem 6.1. The following proposition treats the case in which Δ\Delta satisfies the weaker condition than 𝒂\bm{a}-balancedness.

Proposition 8.1.

For m1m\geq 1, let 𝒂=(a1,,am)>0m\bm{a}=(a_{1},\ldots,a_{m})\in\mathbb{Z}_{>0}^{m} be a positive integer vector with a14a_{1}\geq 4 and ai2a_{i}\geq 2 for all i[m]i\in[m]. Let Δ\Delta be a minimal (d1)(d-1)-cycle complex, where d=i=1maid=\sum_{i=1}^{m}a_{i}. Let κ:V(Δ)[m]\kappa:V(\Delta)\rightarrow[m] be a map such that, for any facet FF of Δ\Delta, (|Fκ1(i)|)i=𝒂+𝒆j𝒆1>0m\left(|F\cap\kappa^{-1}(i)|\right)_{i}=\bm{a}+\bm{e}_{j}-\bm{e}_{1}\in\mathbb{Z}_{>0}^{m} holds for some j[m]j\in[m]. Then, G(Δ)G(\Delta) is (κ,𝒂)(\kappa,\bm{a})-sparse rigid.

Proof.

The proof is similar to that of Theorem 6.1. Define a type tκ(U)t_{\kappa}(U) of UV(Δi+)U\subseteq V(\Delta_{i}^{+}) as (|Uκ1(i)|)i0m(|U\cap\kappa^{-1}(i)|)_{i}\in\mathbb{Z}_{\geq 0}^{m}. Note that by Lemma 6.2, for a generic (κ,𝒂)(\kappa,\bm{a})-sparse point configuration p:V(Δ)dp:V(\Delta)\rightarrow\mathbb{R}^{d} and UV(Δ)U\subseteq V(\Delta) with |U|=d+1|U|=d+1, p(U)p(U) is affinely independent if and only if tκ(U)=𝒂+𝒆jt_{\kappa}(U)=\bm{a}+\bm{e}_{j} for some j[m]j\in[m].

The proof is done by the induction on |V(Δ)||V(\Delta)|. As a base case, if Δ\Delta is a trivial minimal (d1)(d-1)-cycle complex, G(Δ)=KdG(\Delta)=K_{d} and by Lemma 6.2, for a generic (κ,𝒂)(\kappa,\bm{a})-sparse point configuration p:V(Δ)dp:V(\Delta)\rightarrow\mathbb{R}^{d}, p(V(Δ))p(V(\Delta)) spans a (d1)(d-1)-dimensional affine space. Hence G(Δ)G(\Delta) is (κ,𝒂)(\kappa,\bm{a})-sparse rigid.

Suppose that Δ\Delta is a nontrivial minimal (d1)(d-1)-cycle complex. Pick u,vκ1(1)u,v\in\kappa^{-1}(1) with uvE(G)uv\in E(G), which exist by a12a_{1}\geq 2. Let Δ1+,,Δt+\Delta_{1}^{+},\ldots,\Delta_{t}^{+} be the nontrivial minimal (d1)(d-1)-cycle complexes given in Lemma 3.2 with respect to Δ\Delta and uvuv.

Claim 8.2.

For each ii, there is a (d1)(d-1)-set CNG(Δi+)(u)NG(Δi+)(v)C\subseteq N_{G(\Delta_{i}^{+})}(u)\cap N_{G(\Delta_{i}^{+})}(v) such that tκ(C+u+v)=𝒂+𝒆kt_{\kappa}(C+u+v)=\bm{a}+\bm{e}_{k} for some k[m]k\in[m].

Proof of claim.

Pick a facet FΔi+F^{*}\in\Delta_{i}^{+} containing uu and vv. For every wFuvw\in F^{*}-u-v, there is x(w)x(\neq w) such that Fw+xF^{*}-w+x is a facet of Δi+\Delta_{i}^{+}. We show that for an appropriate choice of wFuvw\in F^{*}-u-v, C:=F+xuvC:=F^{*}+x-u-v satisfies tκ(C+u+v)=𝒂+𝒆jt_{\kappa}(C+u+v)=\bm{a}+\bm{e}_{j} for some j[m]j\in[m].

By Lemma 3.2 (b), for every facet FF of Δi+\Delta_{i}^{+} containing vv, tκ(Fv)=𝒃𝒆jt_{\kappa}(F-v)=\bm{b}-\bm{e}_{j} for some j[m]j\in[m] and some 𝒃{𝒂+𝒆k𝒆1:k[m]}\bm{b}\in\{\bm{a}+\bm{e}_{k}-\bm{e}_{1}:k\in[m]\}. Hence, we have the following:

(\star) for every facet FF of Δi+\Delta_{i}^{+} containing vv, tκ(F)=𝒂+𝒆k𝒆jt_{\kappa}(F)=\bm{a}+\bm{e}_{k}-\bm{e}_{j} for some k,j[m]k,j\in[m].

By the property (\star) of FF^{*}, tκ(F)=𝒂+𝒆k𝒆jt_{\kappa}(F^{*})=\bm{a}+\bm{e}_{k}-\bm{e}_{j} for some k,j[m]k,j\in[m]. If k=jk=j, for any choice of ww, we have tκ(C+u+v)=𝒂+𝒆κ(x)t_{\kappa}(C+u+v)=\bm{a}+\bm{e}_{\kappa(x)} as desired. Thus assume that kjk\neq j. If j=1j=1, pick wFuvw\in F^{*}-u-v with κ(w)=1\kappa(w)=1. Note that such ww exists by a14a_{1}\geq 4. Then for Fw+xF^{*}-w+x to satisfy ()(\star), κ(x)\kappa(x) must be 11, and we get tκ(C+u+v)=𝒂+𝒆kt_{\kappa}(C+u+v)=\bm{a}+\bm{e}_{k} as desired. If j1j\neq 1, by aj2a_{j}\geq 2, we can pick wFuvw\in F-u-v with κ(w)=j\kappa(w)=j. Then from the property ()(\star) of Fw+xF^{*}-w+x, we have κ(x)=j\kappa(x)=j. Thus we get tκ(C+u+v)=𝒂+𝒆kt_{\kappa}(C+u+v)=\bm{a}+\bm{e}_{k} as desired. ∎

The restriction of κ\kappa to V(Δi+)V(\Delta_{i}^{+}) (resp. V(Δi+/uv)V(\Delta_{i}^{+}/uv)) satisfies the assumption for Δi+\Delta_{i}^{+} (resp. Δi+/uv\Delta_{i}^{+}/uv). Hence by the same argument as the proof of Theorem 6.1, from Claim 8.2, we can deduce that G(Δ)G(\Delta) is (κ,𝒂)(\kappa,\bm{a})-sparse rigid by Lemma 2.1, Lemma 2.2, Lemma 3.2 (d) and (e). ∎

The following proposition asserts that the assumption ai2a_{i}\geq 2 for all ii in Theorem 6.1 can be weakened to a12a_{1}\geq 2 (and no assumption on aia_{i} for i1i\neq 1) if vertices in κ1(1)\kappa^{-1}(1) are allowed to have full support.

Proposition 8.3.

Let d>a2d>a\geq 2 be integers and let 𝒃>0m\bm{b}\in\mathbb{Z}_{>0}^{m} be an integer vector with da=i=1mbid-a=\sum_{i=1}^{m}b_{i}. Let Δ\Delta be a minimal (d1)(d-1)-cycle complex and XV(Δ)X\subseteq V(\Delta) be a (a)(\geq a)-transversal set of Δ\Delta. Let κ:V(Δ)X[m]\kappa:V(\Delta)\setminus X\rightarrow[m] be a map satisfying |Fκ1(i)|bi|F\cap\kappa^{-1}(i)|\leq b_{i} for every FΔF\in\Delta and i[m]i\in[m].

Let I1,,ImI_{1},\ldots,I_{m} be disjoint subsets of [d][d] with |Ii|=bi|I_{i}|=b_{i} for i[m]i\in[m] and define a map L:V(Δ)2[d]L^{*}:V(\Delta)\rightarrow 2^{[d]} by L(v)=[d]L^{*}(v)=[d] if vXv\in X and L(v)=Iκ(v)L^{*}(v)=I_{\kappa(v)} if vV(Δ)Xv\in V(\Delta)\setminus X. Then G(Δ)G(\Delta) is LL^{*}-sparse rigid.

Proof.

For UV(Δ)U\subseteq V(\Delta), let t(U):=(|UX|,|Uκ1(1)|,,|Uκ1(m)|)0m+1t(U):=(|U\cap X|,|U\cap\kappa^{-1}(1)|,\ldots,|U\cap\kappa^{-1}(m)|)\in\mathbb{Z}_{\geq 0}^{m+1}. The assumption on XX and κ\kappa is that, for each facet FF of Δ\Delta, t(F)=(a+i=1mdi,𝒃𝒅)t(F)=(a+\sum_{i=1}^{m}d_{i},\bm{b}-\bm{d}) for some 𝒅0m\bm{d}\in\mathbb{Z}_{\geq 0}^{m} with 𝒅𝒃\bm{d}\leq\bm{b}. One can check that a (d+1)(d+1)-set UV(Δ)U\subseteq V(\Delta) satisfies the condition (H) for LL^{*} if t(U)=(a+i=1mdi,𝒃𝒅)+𝒆jt(U)=(a+\sum_{i=1}^{m}d_{i},\bm{b}-\bm{d})+\bm{e}_{j} for some j[m+1]j\in[m+1] and some 𝒅0m\bm{d}\in\mathbb{Z}_{\geq 0}^{m} with 𝒅𝒃\bm{d}\leq\bm{b}.

The proof is done by the induction on |V(Δ)||V(\Delta)|. When Δ\Delta is trivial, the statement easily follows. Suppose that Δ\Delta is a nontrivial minimal (d1)(d-1)-cycle complex. Pick u,vXu,v\in X with uvE(Δ)uv\in E(\Delta), which exist by a2a\geq 2. Let Δ1+,,Δt+\Delta_{1}^{+},\ldots,\Delta_{t}^{+} be the nontrivial minimal (d1)(d-1)-cycle complexes given in Lemma 3.2 with respect to Δ\Delta and uvuv.

Claim 8.4.

For each ii, there exists CNG(Δi+)(u)NG(Δi+)(v)C\in N_{G(\Delta_{i}^{+})}(u)\cap N_{G(\Delta_{i}^{+})}(v) such that C+u+vC+u+v satisfies the condition (H) for LL^{*}.

Proof of claim.

For each facet FF of Δi+\Delta_{i}^{+} containing vv, we have t(F)=(a+i=1mdi,𝒃𝒅)t(F)=(a+\sum_{i=1}^{m}d_{i},\bm{b}-\bm{d}) for some 𝒅0m\bm{d}\in\mathbb{Z}_{\geq 0}^{m} with 𝒅𝒃\bm{d}\leq\bm{b}. Pick FΔi+F^{*}\in\Delta_{i}^{+} with u,vFu,v\in F^{*}, and then pick wFuvw\in F^{*}-u-v. Then there is x(w)x(\neq w) such that F+xwF^{*}+x-w is a facet of Δi+\Delta_{i}^{+}. Now C:=F+xuvC:=F^{*}+x-u-v is the desired set. ∎

By the similar argument to the proof of Theorem 6.1, the LL^{*}-sparse rigidity of G(Δ)G(\Delta) is deduced. ∎

Acknowledgement

I would like to thank my advisor Shin-ichi Tanigawa for helpful discussions and helpful comments on the presentation.

References

  • [1] B. Bagchi and B. Datta. Lower bound theorem for normal pseudomanifolds. Expositiones Mathematicae, 26(4):327–351, 2008.
  • [2] D. Barnette. A proof of the lower bound conjecture for convex polytopes. Pacific Journal of Mathematics, 46(2):349–354, 1973.
  • [3] M. Chari. Two decompositions in topological combinatorics with applications to matroid complexes. Transactions of the American Mathematical Society, 349(10):3925–3943, 1997.
  • [4] J. Cruickshank, B. Jackson, and S. Tanigawa. Global rigidity of triangulated manifolds. arXiv:2204.02503, 2022.
  • [5] J. Cruickshank, B. Jackson, and S. Tanigawa. Rigidity of symmetric simplicial complexes and the lower bound theorem. arXiv preprint arXiv:2304.04693, 2023.
  • [6] Y. Eftekhari, B. Jackson, A. Nixon, B. Schulze, S. Tanigawa, and W. Whiteley. Point-hyperplane frameworks, slider joints, and rigidity preserving transformations. Journal of Combinatorial Theory, Series B, 135:44–74, 2019.
  • [7] A. L. Fogelsanger. The generic rigidity of minimal cycles. PhD thesis, Cornell University Ithaca, 1988.
  • [8] H. Gluck. Almost all simply connected closed surfaces are rigid. Geometric topology. Lecture Notes in Math., pages 225–239, 1975.
  • [9] M. Goff, S. Klee, and I. Novik. Balanced complexes and complexes without large missing faces. Arkiv för matematik, 49(2):335–350, 2011.
  • [10] M. Gromov. Partial Differential Relations, volume 9. Springer Science & Business Media, 1986.
  • [11] D. Cook II, M. Juhnke-Kubitzke, S. Murai, and E. Nevo. Lefschetz properties of balanced 3-polytopes. Rocky Mountain J. Math, 48:769–790, 2018.
  • [12] G. Kalai. Rigidity and the lower bound theorem 1. Inventiones Mathematicae, 88(1):125–151, 1987.
  • [13] S. Klee and I. Novik. Lower bound theorems and a generalized lower bound conjecture for balanced simplicial complexes. Mathematika, 62(2):441–477, 2016.
  • [14] C. Lee. PL-spheres, convex polytopes, and stress. Discrete & Computational Geometry, 15(4):389–421, 1996.
  • [15] E. Nevo. Rigidity and the lower bound theorem for doubly Cohen–Macaulay complexes. Discrete & Computational Geometry, 39:411–418, 2008.
  • [16] R. Stanley. Balanced Cohen-Macaulay complexes. Transactions of the American Mathematical Society, 249(1):139–157, 1979.
  • [17] R. Stanley. On the number of faces of centrally-symmetric simplicial polytopes. Graphs and Combinatorics, 3(1):55–66, 1987.
  • [18] T.-S. Tay, N. White, and W. Whiteley. Skeletal rigidity of simplicial complexes, I,II. European Journal of Combinatorics, 16(4):381–403, 1995.
  • [19] W. Whiteley. Cones, infinity and one-story buildings. Structural Topology 1983 núm 8, 1983.
  • [20] W. Whiteley. Vertex splitting in isostatic frameworks. Structutal Topology, 1990, núm. 16, 1990.