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

Vietoris-Rips Complexes of Split-Decomposable Spaces

Mario Gómez
Abstract

Split-metric decompositions are an important tool in the theory of phylogenetics, particularly because of the link between the tight span and the class of totally decomposable spaces, a generalization of metric trees whose decomposition does not have a “prime” component. We use this connection to study the Vietoris-Rips complexes of totally decomposable spaces. In particular, we characterize the homotopy type of the Vietoris-Rips complex of circular decomposable spaces whose metric is monotone. We extend this result to compute the homology of certain non-monotone circular decomposable spaces. We also use the block decomposition of the tight span of a totally decomposable metric to induce a decomposition of the VR complex of a totally decomposable metric.

1 Introduction

Split-metric decompositions are an important tool in the theory of phylogenetics, particularly because of the link between the tight span and the class of totally decomposable spaces, a generalization of metric trees whose decomposition does not have a “prime” component. The connection with tight spans has been studied at least since the introduction of split-metric decompositions by Bandelt and Dress in [BD92], and culminated with the characterization of the polytopal structure of the tight span of a totally decomposable metric by Huber, Koolen, and Moulton in [HKM19]. A special type of totally decomposable spaces are the circular decomposable spaces introduced in [BD92]. Intuitively, a space XX is circular decomposable if there is a bijection f:VnXf:V_{n}\to X from the vertices of a regular nn-gon inscribed in the circle in such a way that the points of XX can be “separated” in the same way that VnV_{n} is separated by a line. See Definition 2.10 for a precise description.
We study the Vietoris-Rips (VR) complexes of totally decomposable spaces using two central results in Topological Data Analysis. One is the characterization of the VR complexes of subsets of the circle by Adamaszek and Adams [AA17]. Given the similarity between circular decomposable spaces and subsets of the circle, we investigate how similar their VR complexes are. In Section 3, we use the results of [AA17] to characterize the homotopy type of circular decomposable spaces whose metric is also monotone in the sense of [Far97] (see Theorem 3.14). We also give conditions on the isolation indices of a circular decomposable space that guarantee the metric to be monotone (Corollary 3.13). However, not every circular decomposable metric is also monotone. In Section 4, we compute the homology of circular decomposable spaces XX in terms of a strict subset YXY\subset X (Theorems 4.7 and 4.8). The results on the homotopy type of clique complexes of cyclic graphs [AA17] are instrumental in this section.
Lastly, in Section 5, we study how the block decomposition of the tight span of a totally decomposable metric XX affects the VR complex of XX. Our main result in this section is that the VR complex of XX decomposes as a disjoint union or wedge of the VR complexes of certain subsets XBXX_{B}\subset X, each corresponding to a block BB in the tight span (Theorem 5.12). We then compute the homology of XX as a direct sum of the homology groups of all XBX_{B}, with a small modification for H0H_{0} (Theorem 5.13).

Acknowledgments.

The author thanks Facundo Mémoli for introducing him to the paper [BD92] and for useful discussions and suggestions during the writing of this paper. The author also acknowledges funding from BSF Grant 2020124 received during the time he was finishing his PhD.

2 Preliminaries

In this section, we set the notation and collect the most important theorems and definitions from our sources.

2.1 Notation

Let (X,dX)(X,d_{X}) be a metric space. The diameter of σX\sigma\subset X is 𝐝𝐢𝐚𝐦(σ)=supx,yXdX(x,y)\mathbf{diam}(\sigma)=\sup_{x,y\in X}d_{X}(x,y). The (open) Vietoris-Rips complex (or VR complex for short) is the simplicial complex

VRr(X):={σ finite:𝐝𝐢𝐚𝐦(σ)<r}.\mathrm{VR}_{r}(X):=\{\sigma\text{ finite}:\mathbf{diam}(\sigma)<r\}.

We model the circle 𝕊1\mathbb{S}^{1} as the quotient /{\mathbb{R}}/{\mathbb{Z}} and equip it with the geodesic distance scaled so that 𝕊1\mathbb{S}^{1} has circumference 1. We say that a path γ:[0,1]𝕊1\gamma:[0,1]\to\mathbb{S}^{1} is clockwise if any lift γ~:[0,1]\widetilde{\gamma}:[0,1]\to{\mathbb{R}} is increasing – this defines the clockwise direction on 𝕊1\mathbb{S}^{1}.

Definition 2.1 (Cyclic order).

Given three points a,b,c𝕊1a,b,c\in\mathbb{S}^{1}, we define the cyclic order by setting abca\preceq b\preceq c if the clockwise path from aa to cc contains bb. We use aba\prec b to indicate that aba\neq b.

Definition 2.2 (Circular sum).

Fix n>0n>0 and consider a function f:{1,,n}f:\{1,\dots,n\}\to{\mathbb{R}}. Given a,b{1,,n}a,b\in\{1,\dots,n\}, we define the circular sum as

\modtwosumi=abf(i):=i=abf(i)\modtwosum_{i=a}^{b}f(i):=\sum_{i=a}^{b}f(i)

when aba\leq b and

\modtwosumi=abf(i):=i=anf(i)+i=1bf(i)\modtwosum_{i=a}^{b}f(i):=\sum_{i=a}^{n}f(i)+\sum_{i=1}^{b}f(i)

when a>ba>b. The value of nn will be clear from context.

Remark 2.3.

The expression i=abf(i)\sum_{i=a}^{b}f(i) when a>ba>b usually represents an empty sum and its value is 0 by convention, while the notation \modtwosumi=abf(i)\modtwosum_{i=a}^{b}f(i) is usually not 0. However, the degenerate cases in some results (like those in Section 3.2) could involve empty circular sums. Since we are deviating from the above convention, we never write empty circular sums, and always state degenerate cases separately if they would require an empty circular sum.

Two equivalent formulations of the circular sum are:

\modtwosumi=abf(i)=i=0baf(a+imod(n))=aibf(i)\modtwosum_{i=a}^{b}f(i)=\sum_{i=0}^{b-a}f(a+i\mod(n))=\sum_{a\preceq i\preceq b}f(i)

In fact, we think of \modtwosumi=abf(i)\modtwosum_{i=a}^{b}f(i) as a summation of ff over the vertices of a regular nn-gon inscribed on the circle. Indeed, if we label the vertices clockwise from 11 to nn, the sum \modtwosumi=abf(i)\modtwosum_{i=a}^{b}f(i) adds the value of ff at all the vertices in the clockwise path between the aa-th and the bb-th vertices (inclusive). If a>ba>b, the clockwise path starting at the aa-th vertex has to go through the nn-th vertex before getting to the bb-th vertex.

2.2 Background

Split systems.

Our main object of study is the split-metric decomposition of a finite metric. Given a finite set XX, a split is a partition S={A,B}S=\{A,B\} of XX. We denote splits by A|BA|B, B|AB|A, or A|A¯A|\overline{A} where A¯\overline{A} is the complement of AA in XX. Given xXx\in X, we denote by S(x)S(x) the element of SS that contains xx. A collection 𝒮={S}{\mathcal{S}}=\{S\} of splits is called a split system. A weighted split system is a pair (𝒮,α)({\mathcal{S}},\alpha) where α:𝒮>0\alpha:{\mathcal{S}}\to{\mathbb{R}}_{>0}. We write the value of α\alpha at S𝒮S\in{\mathcal{S}} as αS\alpha_{S}.

Weakly compatible split systems.

A split system is weakly compatible if there are no three splits S1,S2,S3S_{1},S_{2},S_{3} and four elements x0,x1,x2,x3Xx_{0},x_{1},x_{2},x_{3}\in X such that

Sj(xi)=Sj(x0) if and only if i=j.S_{j}(x_{i})=S_{j}(x_{0})\text{ if and only if }i=j.

Any finite metric space (X,d)(X,d) has an associated weakly compatible split system (which could be empty) with weights induced by the metric. For any a1,a2,b1,b2Xa_{1},a_{2},b_{1},b_{2}\in X (not necessarily distinct), define

β{a1,a2},{b1,b2}:=max{d(a1,b1)+d(a2,b2)d(a1,b2)+d(a2,b1)d(a1,a2)+d(b1,b2)}[d(a1,a2)+d(b1,b2)].\beta_{\{a_{1},a_{2}\},\{b_{1},b_{2}\}}:=\max\left\{\begin{array}[]{c}d(a_{1},b_{1})+d(a_{2},b_{2})\\ d(a_{1},b_{2})+d(a_{2},b_{1})\\ d(a_{1},a_{2})+d(b_{1},b_{2})\end{array}\right\}-\big{[}d(a_{1},a_{2})+d(b_{1},b_{2})\big{]}.

If A|BA|B is a split of XX, the isolation index of A|BA|B is the number

αA|B:=mina1,a2Ab1,b2Bβ{a1,a2},{b1,b2}.\alpha_{A|B}:=\min_{\begin{subarray}{c}a_{1},a_{2}\in A\\ b_{1},b_{2}\in B\end{subarray}}\beta_{\{a_{1},a_{2}\},\{b_{1},b_{2}\}}.

Note that β{a1,a2},{b1,b2}0\beta_{\{a_{1},a_{2}\},\{b_{1},b_{2}\}}\geq 0 and, as a consequence, αA|B0\alpha_{A|B}\geq 0. If αA|B0\alpha_{A|B}\neq 0, the split A|BA|B is called a dd-split. The set

𝒮(X,d):={A|B split of X such that αA|B0}{\mathcal{S}}(X,d):=\{A|B\text{ split of }X\text{ such that }\alpha_{A|B}\neq 0\}

is called the system of dd-splits of (X,d)(X,d) and it is weighted by the isolation indices αA|B\alpha_{A|B}. Bandelt and Dress proved that 𝒮(X,d){\mathcal{S}}(X,d) is always weakly compatible [BD92, Theorem 3]. However, not every metric space (X,d)(X,d) has dd-splits. We call such a space split-prime [BD92].

Example 2.4.

The complete bipartite graph K2,3K_{2,3} with the shortest path metric dd has no dd-splits. It is not hard to prove (see also the introduction of Section 2 of [BD92]) that any dd-split A|A¯A|\overline{A} must be dd-convex in the sense that x,yAx,y\in A and d(x,y)=d(x,z)+d(z,y)d(x,y)=d(x,z)+d(z,y) implies zAz\in A (likewise for A¯\overline{A}). It can be checked that K2,3K_{2,3} cannot be split into two dd-convex sets, so it is prime. The hypercube graph HnH_{n} with the shortest path metric is also split-prime for n3n\geq 3 (see the description after Proposition 3 in [BD92]).

Refer to caption
Figure 1: Left: The complete bipartite graph K2,3K_{2,3}. There are no splits of K2,3K_{2,3} into dd-convex sets, so K2,3K_{2,3} has no dd-splits. Right: The hypercube graph H3H_{3}. The distance between any two of the highlighted vertices is 2, so their β{a1,a2},{b1,b2}\beta_{\{a_{1},a_{2}\},\{b_{1},b_{2}\}} coefficient is 0. This prevents H3H_{3} from having any dd-split.

Split-metric decompositions.

Given a finite set XX and a split A|BA|B of XX, a split-metric is the pseudometric defined by

δA|B(x,y):={0x,yA or x,yB1otherwise.\delta_{A|B}(x,y):=\begin{cases}0&x,y\in A\text{ or }x,y\in B\\ 1&\text{otherwise}.\end{cases}

One of the main results of Bandelt and Dress is the decomposition of any finite metric as a linear combination of split-metrics and a split-prime residue.

Theorem 2.5 ([BD92, Theorem 2]).

Any metric d:X×Xd:X\times X\to{\mathbb{R}} on a finite set XX decomposes as

d=d0+A|B𝒮(X,d)αA|BδA|Bd=d_{0}+\sum_{A|B\in{\mathcal{S}}(X,d)}\alpha_{A|B}\cdot\delta_{A|B} (1)

where d0d_{0} is split-prime and the sum runs over all dd-splits A|BA|B. Moreover, the decomposition is unique in the following sense. Let 𝒮{\mathcal{S}}^{\prime} be a weakly compatible split system on XX with weights λS>0\lambda_{S}>0 for S𝒮S\in{\mathcal{S}}^{\prime}. If d=d0+S𝒮λSδSd=d_{0}+\sum_{S\in{\mathcal{S}}^{\prime}}\lambda_{S}\cdot\delta_{S}, then there is a bijection f:𝒮𝒮(X,d)f:{\mathcal{S}}^{\prime}\to{\mathcal{S}}(X,d) such that αf(S)=λS\alpha_{f(S)}=\lambda_{S}.

Equation (1) is called the split-metric decomposition of (X,d)(X,d), or split decomposition for brevity.

Remark 2.6.

Theorem 2 is stated in more generality in [BD92]. The version therein holds for any symmetric function d:X×Xd:X\times X\to{\mathbb{R}} with 0 diagonal if we allow negative coefficients in (1). We won’t need the full result, but note that if d0d_{0} is a (pseudo)metric, then dd is also a (pseudo)metric.

Definition 2.7.

Given a weighted weakly compatible split system (𝒮,α)({\mathcal{S}},\alpha) on XX, we define the pseudometric

d𝒮,α:=A|B𝒮(X,d)αA|BδA|Bd_{{\mathcal{S}},\alpha}:=\sum_{A|B\in{\mathcal{S}}(X,d)}\alpha_{A|B}\cdot\delta_{A|B}
Definition 2.8 (Totally decomposable spaces).

A metric space (X,d)(X,d) is called totally decomposable if d0=0d_{0}=0 in Equation (1). Equivalently, d=d𝒮,αd=d_{{\mathcal{S}},\alpha} where 𝒮=𝒮(X,d){\mathcal{S}}={\mathcal{S}}(X,d) and αS\alpha_{S} is the isolation index of S𝒮S\in{\mathcal{S}}.

Circular collections of splits.

Suppose XX is a set with nn points. Then a weakly compatible split system 𝒮{\mathcal{S}} on XX can have at most (n2)\binom{n}{2} splits. The reason is that the vector space of symmetric functions f:X×Xf:X\times X\to{\mathbb{R}} with 0 diagonal has dimension (n2)\binom{n}{2} and the uniqueness in Theorem 2.5 implies that the set of split-metrics {δS:S𝒮}\{\delta_{S}:S\in{\mathcal{S}}\} is linearly independent [BD92, Corollary 4].
There are examples of split systems that achieve this bound. Let VnV_{n} be the vertices of a regular nn-gon. Any line that passes through two distinct edges (vi1,vi)(v_{i-1},v_{i}) and (vj,vj+1)(v_{j},v_{j+1}) separates VnV_{n} into the sets {vi,,vj}\{v_{i},\dots,v_{j}\} and {vj+1,,vi1}\{v_{j+1},\dots,v_{i-1}\}. Since there are nn edges, this construction gives a split system 𝒮(Vn){\mathcal{S}}(V_{n}) with (n2)\binom{n}{2} different splits, and it can be verified that 𝒮(Vn){\mathcal{S}}(V_{n}) is weakly compatible. It turns out that, up to a bijection of the underlying sets, this is the only example.

Theorem 2.9 ([BD92, Theorem 5]).

The following conditions are equivalent for a weakly compatible split system 𝒮{\mathcal{S}} on a set XX with |X|=n|X|=n:

  1. 1.

    |𝒮|=(n2)|{\mathcal{S}}|=\binom{n}{2} ;

  2. 2.

    There exists a bijection f:VnXf:V_{n}\to X such that f(𝒮(Vn))=𝒮f({\mathcal{S}}(V_{n}))={\mathcal{S}};

  3. 3.

    The set {δS:S𝒮}\{\delta_{S}:S\in{\mathcal{S}}\} is a basis of the vector space of all symmetric functions d:X×Xd:X\times X\to{\mathbb{R}} with 0 diagonal.

Note that such an XX inherits the cyclic order from VnV_{n}. Explicitly, if we define xi:=f(vi)x_{i}:=f(v_{i}), then xixjxkx_{i}\preceq x_{j}\preceq x_{k} if and only if vivjvkv_{i}\preceq v_{j}\preceq v_{k}.

We later study not only 𝒮(Vn){\mathcal{S}}(V_{n}) but its subsets as well.

Definition 2.10.

Let 𝒮{\mathcal{S}} be a split system on a set XX. If there exists a bijection f:VnXf:V_{n}\to X such that 𝒮f(𝒮(Vn)){\mathcal{S}}\subset f({\mathcal{S}}(V_{n})), we say that 𝒮{\mathcal{S}} is a circular split system. Likewise, a finite metric space (X,dX)(X,d_{X}) is circular decomposable if 𝒮(X,dX){\mathcal{S}}(X,d_{X}) is circular. If in addition |𝒮|=(n2)|{\mathcal{S}}|=\binom{n}{2}, we say that 𝒮{\mathcal{S}} is a full circular system. We define full circular decomposable metrics analogously.

Clique complexes of cyclic graphs.

The main tool used in [AA17] to compute the homotopy type of VRr(𝕊1)\mathrm{VR}_{r}(\mathbb{S}^{1}) were the directed cyclic graphs. We give an undirected definition.

Definition 2.11.

Let GG be a graph with vertex set XX such that |X|=n|X|=n. We say that GG is cyclic if XX has a cyclic order x1x2xnx_{1}\prec x_{2}\prec\cdots\prec x_{n} and there exists a function σ:XX\sigma:X\to X such that the following two conditions hold:

  • For any acdbσ(a)a\preceq c\prec d\preceq b\preceq\sigma(a), if {a,b}\{a,b\} is an edge of GG, then {c,d}\{c,d\} is an edge of GG.

  • For any σ(a)bcda\sigma(a)\preceq b\preceq c\prec d\preceq a, if {a,b}\{a,b\} is an edge of GG, then {c,d}\{c,d\} is an edge of GG.

We also make the following definition for convenience.

Definition 2.12.

A clique complex Cl(G)\operatorname{Cl}(G) is called cyclic if its 1-skeleton GG is cyclic. Similarly, we define wf(Cl(G)):=wf(G)\operatorname{wf}(\operatorname{Cl}(G)):=\operatorname{wf}(G) when GG is cyclic.

Note that we can give an orientation to our cyclic graphs by orienting an edge {a,b}\{a,b\} from aa to bb if abσ(a)a\preceq b\preceq\sigma(a) or σ(b)ab\sigma(b)\preceq a\preceq b. Thanks to this, the results of [AA17] also hold for undirected cyclic graphs.

Theorem 2.13 ([AA17, Theorem 4.4]).

If GG is a cyclic graph, then

Cl(G){𝕊2l+1if l2l+1<wf(G)<l+12l+3 for some l=0,1,n2k1𝕊2lif wf(G)=l2l+1 and G dismantles to Cnk.\operatorname{Cl}(G)\simeq\begin{cases}\mathbb{S}^{2l+1}&\text{if }\frac{l}{2l+1}<\operatorname{wf}(G)<\frac{l+1}{2l+3}\text{ for some }l=0,1,\dots\\ \bigvee^{n-2k-1}\mathbb{S}^{2l}&\text{if }\operatorname{wf}(G)=\frac{l}{2l+1}\text{ and }G\text{ dismantles to }C_{n^{k}}.\end{cases}

We are mainly interested in studying metric spaces whose VR complexes are cyclic.

Definition 2.14.

A metric space (X,dX)(X,d_{X}) with |X|=n|X|=n is monotone111We adopted the term “monotone” from [Far97]. if it has a cyclic order x1x2xnx_{1}\prec x_{2}\prec\cdots\prec x_{n} and there exists a function σ:XX\sigma:X\to X such that the following two conditions hold:

  • For any acdbσ(a)a\preceq c\prec d\preceq b\preceq\sigma(a), we have dcddabd_{cd}\leq d_{ab}.

  • For any σ(a)bcda\sigma(a)\preceq b\prec c\preceq d\preceq a, dcddabd_{cd}\leq d_{ab}.

It is not hard to see that the VR complexes of a monotone metric space are cyclic for any r>0r>0.

Buneman complex.

There is a polytopal complex associated to any weakly compatible split system 𝒮{\mathcal{S}} on XX with weights αS>0\alpha_{S}>0. Let

U(𝒮):={AX:there exists S𝒮 with AS}.U({\mathcal{S}}):=\{A\subset X:\text{there exists }S\in{\mathcal{S}}\text{ with }A\in S\}.

Given ϕU(𝒮)\phi\in{\mathbb{R}}^{U({\mathcal{S}})}, let supp(ϕ):={AU(𝒮):ϕ(A)0}\mathrm{supp}(\phi):=\{A\in U({\mathcal{S}}):\phi(A)\neq 0\}. Define

H(𝒮,α):={ϕU(𝒮):ϕ(A)0 and ϕ(A)+ϕ(A¯)=12α(A|A¯) for all AU(𝒮)}.H({\mathcal{S}},\alpha):=\{\phi\in{\mathbb{R}}^{U({\mathcal{S}})}:\phi(A)\geq 0\text{ and }\phi(A)+\phi(\overline{A})=\tfrac{1}{2}\alpha(A|\overline{A})\text{ for all }A\in U({\mathcal{S}})\}.

Note that H(𝒮,α)H({\mathcal{S}},\alpha) is polytope isomorphic to a hypercube of dimension |𝒮||{\mathcal{S}}|. The Buneman complex of (𝒮,α)({\mathcal{S}},\alpha) is defined by

B(𝒮,α):={ϕH(𝒮,α):A1,A2supp(ϕ) and A1A2=XA1A2=}.B({\mathcal{S}},\alpha):=\{\phi\in H({\mathcal{S}},\alpha):A_{1},A_{2}\in\mathrm{supp}(\phi)\text{ and }A_{1}\cup A_{2}=X\Rightarrow A_{1}\cap A_{2}=\emptyset\}.

Both H(𝒮,α)H({\mathcal{S}},\alpha) and B(𝒮,α)B({\mathcal{S}},\alpha) are equipped with the L1L^{1} metric:

d1(ϕ,ϕ):=AU(𝒮)|ϕ(A)ϕ(A)|.d_{1}(\phi,\phi^{\prime}):=\sum_{A\in U({\mathcal{S}})}|\phi(A)-\phi^{\prime}(A)|.

The space (B(𝒮,α),d1)(B({\mathcal{S}},\alpha),d_{1}) admits an isometric embedding of (X,d𝒮,α)(X,d_{{\mathcal{S}},\alpha}) via the map x(ϕx:U(𝒮)0)x\mapsto(\phi_{x}:U({\mathcal{S}})\to{\mathbb{R}}_{\geq 0}) where

ϕx(A)={12αA|A¯if xA0otherwise.\phi_{x}(A)=\begin{cases}\frac{1}{2}\alpha_{A|\overline{A}}&\text{if }x\notin A\\ 0&\text{otherwise}.\end{cases}

Tight span.

Let (X,dX)(X,d_{X}) be a metric space. The tight span of XX is the set T(X,dX)XT(X,d_{X})\subset{\mathbb{R}}^{X} such that for all fT(X,dX)f\in T(X,d_{X}):

  1. 1.

    f(x)+f(y)dX(x,y)f(x)+f(y)\geq d_{X}(x,y) for all x,yXx,y\in X;

  2. 2.

    f(x)=supyX[dX(x,y)f(y)]f(x)=\sup_{y\in X}\big{[}d_{X}(x,y)-f(y)\big{]} for all xXx\in X.

These properties imply, in particular, that f(x)0f(x)\geq 0 for all fT(X,dX)f\in T(X,d_{X}) and xXx\in X.The tight span is equipped with the LL^{\infty} metric

d(f,g)=supxX|f(x)g(x)|,d_{\infty}(f,g)=\sup_{x\in X}|f(x)-g(x)|,

and there exists an isometric embedding XT(X,dX)X\hookrightarrow T(X,d_{X}) defined by

x(hx:ydX(x,y)).x\mapsto\big{(}h_{x}:y\mapsto d_{X}(x,y)\big{)}.

Tight spans are examples of injective metric spaces. A metric space EE is injective if for every 1-Lipschitz map f:XEf:X\to E and for every isometric embedding XX~X\hookrightarrow\widetilde{X}, there exists a 1-Lipschitz extension f~:X~E\widetilde{f}:\widetilde{X}\to E that makes the following diagram commute:

X{X}X~{\widetilde{X}}E{E}f\scriptstyle{f}f~\scriptstyle{\widetilde{f}}

In particular, T(X,dX)T(X,d_{X}) is the smallest injective space that contains XX, i.e. XT(X,dX)EX\hookrightarrow T(X,d_{X})\hookrightarrow E for any injective EE such that XEX\hookrightarrow E. Furthermore, injective spaces have nice topological and geometric structures.

Proposition 2.15.

Any injective metric space EE is contractible and geodesic.

See the comment after Proposition 2.1 of [LMO22].

Tight spans are an important tool in topological data analysis thanks to Theorem 2.16 below. We need one more definition before stating this result. Given a metric space (E,dE)(E,d_{E}), XEX\subset E and r>0r>0, the metric thickening of XX in EE is the set

Br(X;E):={eE| exists xX with dE(x,e)<r}.B_{r}(X;E):=\{e\in E|\text{ exists }x\in X\text{ with }d_{E}(x,e)<r\}.

If there is an isometric embedding ι:XE\iota:X\hookrightarrow E, we will write Br(X;E)B_{r}(X;E) instead of Br(ι(X);E)B_{r}(\iota(X);E).

Theorem 2.16 ([LMO22, Proposition 2.3]).

Let (X,dX)(X,d_{X}) be a metric space, and let EE be an injective space such that XEX\hookrightarrow E. Then the Vietoris-Rips complex VR2r(X,dX)\mathrm{VR}_{2r}(X,d_{X}) and the metric thickening Br(X;E)B_{r}(X;E) are homotopy equivalent for every r>0r>0.

Tight spans of totally decomposable metrics.

The tight span of totally decomposable metrics has been studied by several authors throughout the years [BD92, DH01, HKM06, HKM19]. We use the notation and results of [HKM19], one of the most recent papers. Given a weighted split system (𝒮,α)({\mathcal{S}},\alpha) on XX, define the map

κ:U(𝒮)\displaystyle\kappa:{\mathbb{R}}^{U({\mathcal{S}})} X\displaystyle\to{\mathbb{R}}^{X}
ϕ\displaystyle\phi (xd1(ϕ,ϕx)),\displaystyle\mapsto(x\mapsto d_{1}(\phi,\phi_{x})),

where d1d_{1} is the L1L^{1} metric on U(𝒮){\mathbb{R}}^{U({\mathcal{S}})}. The first salient property of κ\kappa is that it sends each element ϕx\phi_{x} to hxh_{x}. In other words, κ\kappa is “constant” on the embedded copy of XX inside the Buneman complex and the tight span. The map has much stronger properties when 𝒮{\mathcal{S}} is weakly compatible.

Theorem 2.17.

Let (𝒮,α)({\mathcal{S}},\alpha) be a weighted split system on XX. The map κ\kappa is 1-Lipschitz. Furthermore, κ(B(𝒮,α))=T(X,d𝒮,α)\kappa(B({\mathcal{S}},\alpha))=T(X,d_{{\mathcal{S}},\alpha}) if and only if 𝒮{\mathcal{S}} is weakly compatible.

See the introduction to Section 4 of [HKM06] to see why κ(ϕx)=hx\kappa(\phi_{x})=h_{x} and why κ\kappa is 1-Lipschitz. The equivalence between κ(B(𝒮,α))=T(X,d𝒮,α)\kappa(B({\mathcal{S}},\alpha))=T(X,d_{{\mathcal{S}},\alpha}) and weak compatibility is the main result of [DHM98].

Remark 2.18.

As natural as the choice of κ\kappa may seem, it is not an isometry in general. This is the main result of [DH01]. Note, however, that [DH01] uses the map Λd\Lambda_{d}. It can be checked that κ\kappa and Λd\Lambda_{d} differ by a constant using equation (2) in Section 4 of [HKM06].

The map κ\kappa also induces a strong relationship between the polytopal structures of the Buneman complex and the tight span. We need more definitions to state this result.
Let CC be a connected polytopal complex. A vertex vCv\in C is a cut-vertex if C{v}C-\{v\} is disconnected. A maximal subcomplex BCB\subset C that does not have cut-vertices is called a block. The set of blocks of CC is denoted as (C){\mathcal{B}}(C). The block structure of the Buneman complex B(𝒮,α)B({\mathcal{S}},\alpha) can be determined from the properties of 𝒮{\mathcal{S}} in a straightforward way.

Definition 2.19.

Let 𝒮{\mathcal{S}} be a split system on XX. Two splits S,S𝒮S,S^{\prime}\in{\mathcal{S}} are called compatible if the following equivalent conditions hold:

  • There exist ASA\in S, ASA^{\prime}\in S^{\prime} with AA=A\cap A^{\prime}=\emptyset;

  • There exist ASA\in S, ASA^{\prime}\in S^{\prime} with AA=XA\cup A^{\prime}=X;

  • There exist ASA\in S, ASA^{\prime}\in S^{\prime} with AAA\subset A^{\prime} or AAA^{\prime}\subset A.

If neither condition holds, we say that SS and SS^{\prime} are incompatible. We define the incompatibility graph I(𝒮)I({\mathcal{S}}) to be the graph with vertex set 𝒮{\mathcal{S}} and edge set consisting of all pairs {S,S}\{S,S^{\prime}\} where SS and SS^{\prime} are distinct incompatible splits.

Theorem 2.20.

Let (𝒮,α)({\mathcal{S}},\alpha) be a weighted split system on XX. There is a bijective correspondence between the connected components of the incompatibility graph and the blocks of B(𝒮,α)B({\mathcal{S}},\alpha). Moreover, for every 𝒮π0(I(𝒮)){\mathcal{S}}^{\prime}\in\pi_{0}(I({\mathcal{S}})), the block corresponding to 𝒮{\mathcal{S}}^{\prime} is isomorphic as a polytopal complex to B(𝒮,α|𝒮)B({\mathcal{S}}^{\prime},\alpha|_{{\mathcal{S}}^{\prime}}).

We denote the block of B(𝒮,α)B({\mathcal{S}},\alpha) corresponding to 𝒮π0(I(𝒮)){\mathcal{S}}^{\prime}\in\pi_{0}(I({\mathcal{S}})) as B𝒮(𝒮,α)B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha).

We need one more definition before stating the link between the polytopal structures of B(𝒮,α)B({\mathcal{S}},\alpha) and T(X,d𝒮,α)T(X,d_{{\mathcal{S}},\alpha}). A split system 𝒮{\mathcal{S}} is called octahedral if |𝒮|=4|{\mathcal{S}}|=4 and there exists a partition X=X1X6X=X_{1}\cup\cdots\cup X_{6} such that Si=(XiXi+1Xi+2)|(Xi+3Xi+4Xi+5)S_{i}=(X_{i}\cup X_{i+1}\cup X_{i+2})|(X_{i+3}\cup X_{i+4}\cup X_{i+5}) for 1i31\leq i\leq 3 (indices are taken modulo 6) and S4=(X1X3X5)|(X2X4X6)S_{4}=(X_{1}\cup X_{3}\cup X_{5})|(X_{2}\cup X_{4}\cup X_{6}).

Theorem 2.21.

Let (𝒮,α)({\mathcal{S}},\alpha) be a weighted weakly compatible split system on XX. Then κ\kappa induces a bijection between (B(𝒮,α)){\mathcal{B}}(B({\mathcal{S}},\alpha)) and (T(X,d𝒮,α)){\mathcal{B}}(T(X,d_{{\mathcal{S}},\alpha})) such that:

  • If 𝒮π0(I(𝒮)){\mathcal{S}}^{\prime}\in\pi_{0}(I({\mathcal{S}})) is not octahedral, then the block B𝒮(𝒮,α)B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha) is isomorphic to the block κ(BcS(𝒮,α))T(X,d𝒮,α)\kappa(B_{cS^{\prime}}({\mathcal{S}},\alpha))\subset T(X,d_{{\mathcal{S}},\alpha}) as polytopal complexes.

  • If 𝒮π0(I(𝒮)){\mathcal{S}}^{\prime}\in\pi_{0}(I({\mathcal{S}})) is octahedral, then κ(B𝒮(𝒮,α))\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)) is a block isomorphic to a rhombic dodecahedron.

3 Properties of circular decomposable metrics

According to Theorem 2.9, a split system 𝒮{\mathcal{S}} on XX with maximal cardinality is necessarily circular. In this section, we investigate the properties of a circular decomposable metric space (X,dX)(X,d_{X}).
We fix n:=|X|n:=|X| throughout this section. We assume that X={1,,n}X=\{1,\dots,n\} and that the bijection f:VnXf:V_{n}\to X from Theorem 2.9 satisfies f(vi)=if(v_{i})=i. If 𝒮(X,dX){\mathcal{S}}(X,d_{X}) is a full circular system, the splits have the form Sij:=Aij|A¯ijS_{ij}:=A_{ij}|\overline{A}_{ij} where Aij:={i,,j}A_{ij}:=\{i,\dots,j\} and 1ij<n1\leq i\leq j<n. Note that there are (n2)\binom{n}{2} splits of the form SijS_{ij} and that they are all distinct. Then if αij:=αSij\alpha_{ij}:=\alpha_{S_{ij}}, we have

dX=1ij<nαijδSijd_{X}=\sum_{1\leq i\leq j<n}\alpha_{ij}\delta_{S_{ij}} (2)

by Theorem 2.5. If 𝒮(X,dX){\mathcal{S}}(X,d_{X}) is circular but not full, then 𝒮(X,dX)f(𝒮(Vn)){\mathcal{S}}(X,d_{X})\subsetneq f({\mathcal{S}}(V_{n})) (see Definition 2.10). In that case, we set αij=0\alpha_{ij}=0 for any split Sij𝒮(X,dX)S_{ij}\notin{\mathcal{S}}(X,d_{X}) and Equation (2) still holds. We write dab:=dX(a,b)d_{ab}:=d_{X}(a,b) for any 1a,bn1\leq a,b\leq n. Recall that XX inherits the cyclic order from VnV_{n} (see Definition 2.1) so that ijki\preceq j\preceq k if the clockwise path from viv_{i} to vkv_{k} contains vjv_{j}.

3.1 An expression for dXd_{X} in terms of αij\alpha_{ij}

Our objective in this section is to find an expression for dabd_{ab} that does not contain any δSij\delta_{S_{ij}} terms. For that, it will be convenient to extend the definition of SijS_{ij} and αij\alpha_{ij} to any 1i,jn1\leq i,j\leq n. In fact, the current expression for AijA_{ij} when 1ij<n1\leq i\leq j<n is equivalent to{k:ikj}\{k:i\preceq k\preceq j\}, so we extend this definition to any 1i,jn1\leq i,j\leq n. Notice that:

  • If j=nj=n, Aij={i,,n}=A¯1,i1A_{ij}=\{i,\dots,n\}=\overline{A}_{1,i-1} for 1<i<n1<i<n;

  • If 1j<in1\leq j<i\leq n, Aij={i,,n}{1,,j}=A¯j+1,i1A_{ij}=\{i,\dots,n\}\cup\{1,\dots,j\}=\overline{A}_{j+1,i-1} unless i=j+1i=j+1;

  • If i=1i=1 and j=nj=n or i=j+1i=j+1, then Aij=XA_{ij}=X.

Hence, we define:

  • For 1j<in1\leq j<i\leq n and 1i<j=n1\leq i<j=n, Sij:=Sj+1,i1S_{ij}:=S_{j+1,i-1} and αij:=αj+1,i1\alpha_{ij}:=\alpha_{j+1,i-1} where indices are taken modulo nn (unless i=1i=1 and j=nj=n or i=j+1i=j+1);

  • If i=j+1 mod ni=j+1\text{ mod }n, we set αij:=0\alpha_{ij}:=0 and leave SijS_{ij} undefined.

The reason to leave SijS_{ij} undefined if i=j+1 mod ni=j+1\text{ mod }n (which includes the case of i=1i=1 and j=nj=n) is because Aij=XA_{ij}=X forces A¯ij=\overline{A}_{ij}=\emptyset. A pair {A,A¯}\{A,\overline{A}\} is only a valid split if both AA and A¯\overline{A} are non-empty. However, defining the coefficients αij\alpha_{ij} as 0 will be convenient for later calculations.

Remark 3.1.

We summarize the above discussion for future reference. For any 1i,jn1\leq i,j\leq n, the set AijA_{ij} is defined as {k:ikj}\{k:i\preceq k\preceq j\}, and the isolation indices αij\alpha_{ij} satisfy the relations

  • αij=αj+1,i1\alpha_{ij}=\alpha_{j+1,i-1}, and

  • αj+1,j=0\alpha_{j+1,j}=0.

If ij+1 mod ni\neq j+1\text{ mod }n, we define the split Sij:=Aij|A¯ijS_{ij}:=A_{ij}|\overline{A}_{ij}, which satisfies the relation Sj+1,i1=SijS_{j+1,i-1}=S_{ij}.

Lemma 3.2.

Let a,b,c,d{1,,n}a,b,c,d\in\{1,\dots,n\}. Then i=a+1bj=cd1αij=i=c+1dj=ab1αij\displaystyle\sum_{i=a+1}^{b}\sum_{j=c}^{d-1}\alpha_{ij}=\sum_{i=c+1}^{d}\sum_{j=a}^{b-1}\alpha_{ij} (indices are treated modulo nn). The analogous equation holds if we replace \sum with \modtwosum\modtwosum.

Proof.

By Remark 3.1,

i=a+1bj=cd1αij=j=cd1i=a+1bαj+1,i1=j=c+1di=ab1αji.\sum_{i=a+1}^{b}\sum_{j=c}^{d-1}\alpha_{ij}=\sum_{j=c}^{d-1}\sum_{i=a+1}^{b}\alpha_{j+1,i-1}=\sum_{j=c+1}^{d}\sum_{i=a}^{b-1}\alpha_{ji}.

The lemma follows by changing the roles of ii and jj in the last sum. ∎

Lemma 3.3.

For any 1a,bn1\leq a,b\leq n with aba\neq b, dab=\modtwosumi=a+1b\modtwosumj=ba1αij\displaystyle d_{ab}=\modtwosum_{i=a+1}^{b}\modtwosum_{j=b}^{a-1}\alpha_{ij}.

Proof.

Suppose a<ba<b. If 1ij<n1\leq i\leq j<n, then aAija\in A_{ij} and bAijb\notin A_{ij} if and only if iaj<bi\leq a\leq j<b. Similarly, aAija\notin A_{ij} and bAijb\in A_{ij} if and only if a<ibja<i\leq b\leq j. Let χ\chi be an indicator function. Then by equation (2) and Lemma 3.2,

dab\displaystyle d_{ab} =1ij<nαijδSij(a,b)\displaystyle=\sum_{1\leq i\leq j<n}\alpha_{ij}\delta_{S_{ij}}(a,b)
=1ij<nαijχ(iaj<b)+1ij<nαijχ(a<ibj)\displaystyle=\sum_{1\leq i\leq j<n}\alpha_{ij}\cdot\chi(i\leq a\leq j<b)+\sum_{1\leq i\leq j<n}\alpha_{ij}\cdot\chi(a<i\leq b\leq j)
=i=1aj=ab1αij+i=a+1bj=bn1αij=i=1aj=ab1αij+i=b+1nj=ab1αij\displaystyle=\sum_{i=1}^{a}\sum_{j=a}^{b-1}\alpha_{ij}+\sum_{i=a+1}^{b}\sum_{j=b}^{n-1}\alpha_{ij}=\sum_{i=1}^{a}\sum_{j=a}^{b-1}\alpha_{ij}+\sum_{i=b+1}^{n}\sum_{j=a}^{b-1}\alpha_{ij}
=\modtwosumi=b+1a\modtwosumj=ab1αij.\displaystyle=\modtwosum_{i=b+1}^{a}\modtwosum_{j=a}^{b-1}\alpha_{ij}.

Using Lemma 3.2 one more time gives the result. If a>ba>b, we obtain the desired formula for dab=dbad_{ab}=d_{ba} by swapping the roles of aa and bb in the equation above. ∎

3.2 Monotone circular decomposable metrics

Given that circular split systems have such a close relationship with finite subsets of 𝕊1\mathbb{S}^{1}, we want to explore how similar their VR complexes are. A crucial feature of subsets of 𝕊1\mathbb{S}^{1} is that their VR complexes are cyclic. However, this is not the case for every circular decomposable metric.

Example 3.4.

Consider the wedge of 𝕊1\mathbb{S}^{1} and an interval as shown in Figure 2. Let Y={x1yx3x4x5}Y=\{x_{1}\prec y\prec x_{3}\prec x_{4}\prec x_{5}\} and Z={x1,,x5}Z=\{x_{1},\dots,x_{5}\}. YY has a circular split system and all its VR complexes are cyclic because Y𝕊1Y\subset\mathbb{S}^{1}. Since dZ(x2,xi)=dY(y,xi)+(15+ϵ)d_{Z}(x_{2},x_{i})=d_{Y}(y,x_{i})+(\tfrac{1}{5}+\epsilon) for i2i\neq 2, it follows that α{x2},Z{x2}=α{y},Y{y}+(15+ϵ)\alpha_{\{x_{2}\},Z\setminus\{x_{2}\}}=\alpha_{\{y\},Y\setminus\{y\}}+(\frac{1}{5}+\epsilon). Hence, ZZ also has a circular split system and it inherits the cycic order from YY so that x1x2x3x4x5x_{1}\prec x_{2}\prec x_{3}\prec x_{4}\prec x_{5}. However, dZ(x1,x2),dZ(x2,x3)>dZ(x1,x3)d_{Z}(x_{1},x_{2}),d_{Z}(x_{2},x_{3})>d_{Z}(x_{1},x_{3}), so VRr(Z)\mathrm{VR}_{r}(Z) is not cyclic for 25r<25+ϵ\frac{2}{5}\leq r<\frac{2}{5}+\epsilon. Intuitively, the edge [x1,x3][x_{1},x_{3}] appears in VRr(Z)\mathrm{VR}_{r}(Z) before either [x1,x2][x_{1},x_{2}] or [x2,x3][x_{2},x_{3}] even though x2x_{2} is “between” x1x_{1} and x3x_{3}.

Refer to caption
Figure 2: The space Y:={x1,y,x3,x4,x5}Y:=\{x_{1},y,x_{3},x_{4},x_{5}\} consists of the vertices of a regular pentagon inscribed on the circle. We attach an edge ee of length 15+ϵ\frac{1}{5}+\epsilon to the circle at the point yy and define x2x_{2} as the boundary of ee different from yy. The circular decomposition of YY induces a circular decomposition of Z:={x1,x2,x3,x4,x5}Z:=\{x_{1},x_{2},x_{3},x_{4},x_{5}\} that satisfies dZ(x1,x2),dZ(x2,x3)>dZ(x1,x3)d_{Z}(x_{1},x_{2}),d_{Z}(x_{2},x_{3})>d_{Z}(x_{1},x_{3}). See Example 3.4.

Given that not every circular decomposable metric space XX has cyclic VR complexes, we look for conditions on the isolation indices that force the VR complexes to be cyclic. We begin with the following Lemma.

Lemma 3.5.

If abca\prec b\prec c, we have

  • dabdacd_{ab}\leq d_{ac} if and only if \modtwosumi=b+1c\modtwosumj=ab1αij\modtwosumi=b+1c\modtwosumj=ca1αij\modtwosum_{i=b+1}^{c}\modtwosum_{j=a}^{b-1}\alpha_{ij}\leq\modtwosum_{i=b+1}^{c}\modtwosum_{j=c}^{a-1}\alpha_{ij}.

  • dbcdacd_{bc}\leq d_{ac} if and only if \modtwosumi=a+1b\modtwosumj=bc1αij\modtwosumi=a+1b\modtwosumj=ca1αij\modtwosum_{i=a+1}^{b}\modtwosum_{j=b}^{c-1}\alpha_{ij}\leq\modtwosum_{i=a+1}^{b}\modtwosum_{j=c}^{a-1}\alpha_{ij}.

Proof.

The lemma follows from the observations that

dab=\modtwosumi=a+1b\modtwosumj=ba1αij\displaystyle d_{ab}=\modtwosum_{i=a+1}^{b}\modtwosum_{j=b}^{a-1}\alpha_{ij} =\modtwosumi=a+1b\modtwosumj=ca1αij+\modtwosumi=a+1b\modtwosumj=bc1αij\displaystyle=\modtwosum_{i=a+1}^{b}\modtwosum_{j=c}^{a-1}\alpha_{ij}+\modtwosum_{i=a+1}^{b}\modtwosum_{j=b}^{c-1}\alpha_{ij}
dac=\modtwosumi=a+1c\modtwosumj=ca1αij\displaystyle d_{ac}=\modtwosum_{i=a+1}^{c}\modtwosum_{j=c}^{a-1}\alpha_{ij} =\modtwosumi=a+1b\modtwosumj=ca1αij+\modtwosumi=b+1c\modtwosumj=ca1αij\displaystyle=\modtwosum_{i=a+1}^{b}\modtwosum_{j=c}^{a-1}\alpha_{ij}+\modtwosum_{i=b+1}^{c}\modtwosum_{j=c}^{a-1}\alpha_{ij}
=\modtwosumi=b+1c\modtwosumj=ca1αij+\modtwosumi=a+1b\modtwosumj=ca1αij,\displaystyle=\modtwosum_{i=b+1}^{c}\modtwosum_{j=c}^{a-1}\alpha_{ij}+\modtwosum_{i=a+1}^{b}\modtwosum_{j=c}^{a-1}\alpha_{ij},
dbc=\modtwosumi=b+1c\modtwosumj=cb1αij\displaystyle d_{bc}=\modtwosum_{i=b+1}^{c}\modtwosum_{j=c}^{b-1}\alpha_{ij} =\modtwosumi=b+1c\modtwosumj=ca1αij+\modtwosumi=b+1c\modtwosumj=ab1αij\displaystyle=\modtwosum_{i=b+1}^{c}\modtwosum_{j=c}^{a-1}\alpha_{ij}+\modtwosum_{i=b+1}^{c}\modtwosum_{j=a}^{b-1}\alpha_{ij}

and Lemma 3.2, which gives the identities \modtwosumi=a+1b\modtwosumj=bc1αij=\modtwosumi=b+1c\modtwosumj=ab1αij\displaystyle\modtwosum_{i=a+1}^{b}\modtwosum_{j=b}^{c-1}\alpha_{ij}=\modtwosum_{i=b+1}^{c}\modtwosum_{j=a}^{b-1}\alpha_{ij} and \modtwosumi=b+1c\modtwosumj=ab1αij=\modtwosumi=a+1b\modtwosumj=bc1αij\displaystyle\modtwosum_{i=b+1}^{c}\modtwosum_{j=a}^{b-1}\alpha_{ij}=\modtwosum_{i=a+1}^{b}\modtwosum_{j=b}^{c-1}\alpha_{ij}. ∎

In order for the VR complex of XX to be cyclic, we need [a,a+1][a,a+1] to be an edge of VRr(X)\mathrm{VR}_{r}(X) whenever [a,a+2][a,a+2] is. Similarly, [a,a+2][a,a+2] must be an edge whenever [a,a+3][a,a+3] is, and so on. In the other direction, [a1,a][a-1,a] must be an edge whenever [a2,a][a-2,a] is. Rephrasing these conditions in terms of the metric shows that, at a minimum, we need

da,a+1da,a+2da,k and dk,ada2,ada1,ad_{a,a+1}\leq d_{a,a+2}\leq\cdots\leq d_{a,k}\text{ and }d_{k,a}\geq\cdots\geq d_{a-2,a}\geq d_{a-1,a} (3)

for some kk. We give a name to this condition.

Definition 3.6.

Let (X,dX)(X,d_{X}) be a metric space with a cyclic order \preceq. We say that XX is unimodal at aa if there exists kk such that:

  • dabdacd_{ab}\leq d_{ac} for all abcka\preceq b\preceq c\preceq k, and

  • dabdacd_{ab}\geq d_{ac} for all kbcak\preceq b\preceq c\preceq a.

We denote the minimal such kk as M(a)M(a).

Unimodality by itself is not enough to guarantee monotonicity of XX, but it is pretty close.

Lemma 3.7.

Let (XdX)(Xd_{X}) be a finite metric space with a cyclic order \preceq. If XX is unimodal at every point and abM(a)a\preceq b\preceq M(a) if and only if M(a)M(b)aM(a)\preceq M(b)\preceq a, then XX is monotone.

Proof.

Suppose acdbM(a)a\preceq c\prec d\preceq b\preceq M(a). By hypothesis, acM(a)a\preceq c\prec M(a) implies M(a)M(c)aM(a)\preceq M(c)\preceq a. Hence, cdbM(a)M(c)c\prec d\preceq b\preceq M(a)\preceq M(c). By unimodality at cc, dcddcbd_{cd}\leq d_{cb}. Similarly, abM(a)a\prec b\preceq M(a) implies M(a)M(b)acbM(a)\preceq M(b)\preceq a\preceq c\preceq b, so unimodality at bb implies dcbdabd_{cb}\leq d_{ab}. The second condition in Definition 2.14 follows by exchanging \preceq for \succeq in the argument above. ∎

For the remainder of the section, we find conditions on the isolation indices so that the circular decomposable metric (X,dX)(X,d_{X}) is monotone. It turns out that we can prove many more inequalities using the properties of circular decomposable metrics.

Lemma 3.8.

Let (X,dX)(X,d_{X}) be a circular decomposable metric space. Then for any bab\neq a,

da,b1dab\modtwosumj=ab1αbj\modtwosumj=ba1αbj.d_{a,b-1}\leq d_{ab}\Leftrightarrow\modtwosum_{j=a}^{b-1}\alpha_{bj}\leq\modtwosum_{j=b}^{a-1}\alpha_{bj}.

As a consequence, if acba\prec c\prec b and da,b1dabd_{a,b-1}\leq d_{ab}, then dc,b1dcbd_{c,b-1}\leq d_{cb}.

Proof.

By Lemma 3.5, da,b1dabd_{a,b-1}\leq d_{ab} if and only if \modtwosumj=ab2αbj\modtwosumj=ba1αbj\modtwosum_{j=a}^{b-2}\alpha_{bj}\leq\modtwosum_{j=b}^{a-1}\alpha_{bj}. The result follows because αb,b1=0\alpha_{b,b-1}=0 (see Remark 3.1). If acba\prec c\prec b,

da,b1dab\displaystyle d_{a,b-1}\leq d_{ab} \modtwosumj=ab1αbj\modtwosumj=ba1αbj\displaystyle\Leftrightarrow\modtwosum_{j=a}^{b-1}\alpha_{bj}\leq\modtwosum_{j=b}^{a-1}\alpha_{bj}
\modtwosumj=cb1αbj\modtwosumj=bc1αbjdc,b1dcb.\displaystyle\Rightarrow\modtwosum_{j=c}^{b-1}\alpha_{bj}\leq\modtwosum_{j=b}^{c-1}\alpha_{bj}\Leftrightarrow d_{c,b-1}\leq d_{cb}.

The proof of Lemma 3.8 uses the fact that the sum \modtwosumj=bc1αbj\modtwosum_{j=b}^{c-1}\alpha_{bj} is at its smallest when c=b+1c=b+1 and increases as cc cycles through b+1,b+2,,b1b+1,b+2,\dots,b-1. Likewise, the sum \modtwosumj=cb1αbj\modtwosum_{j=c}^{b-1}\alpha_{bj} achieves its maximum when c=bc=b and steadily decreases in the same range. It turns out that the point when \modtwosumj=bc1αbj\modtwosum_{j=b}^{c-1}\alpha_{bj} becomes larger than \modtwosumj=cb1αbj\modtwosum_{j=c}^{b-1}\alpha_{bj} determines whether XX is monotone or not, and we give this point a name.

Definition 3.9.

Let (X,dX)(X,d_{X}) be a circular decomposable metric and fix bXb\in X. We define b~\widetilde{b} to be the first element of the sequence {b+1,b+2,,b1}\{b+1,b+2,\dots,b-1\} such that:

  • \modtwosumj=bc1αbj\modtwosumj=cb1αbj\modtwosum_{j=b}^{c-1}\alpha_{bj}\leq\modtwosum_{j=c}^{b-1}\alpha_{bj} when bcb~b\prec c\prec\widetilde{b}; and

  • \modtwosumj=bc1αbj\modtwosumj=cb1αbj\modtwosum_{j=b}^{c-1}\alpha_{bj}\geq\modtwosum_{j=c}^{b-1}\alpha_{bj} when b~cb\widetilde{b}\preceq c\prec b.

Let’s prove that a~\widetilde{a} determines M(a)M(a).

Lemma 3.10.

Let (X,dX)(X,d_{X}) be a circular decomposable metric. Suppose there exists a value MaM_{a} such that abMaa\preceq b\preceq M_{a} if and only if b~ab\widetilde{b}\preceq a\preceq b. Then XX is unimodal at aa and Ma=M(a)M_{a}=M(a).

Proof.

Suppose abMaa\prec b\preceq M_{a}. By definition of b~\widetilde{b}, b~ab\widetilde{b}\preceq a\preceq b implies \modtwosumj=ba1αbj\modtwosumj=ab1αbj\modtwosum_{j=b}^{a-1}\alpha_{bj}\geq\modtwosum_{j=a}^{b-1}\alpha_{bj} which, by Lemma 3.8, is equivalent to da,b1dabd_{a,b-1}\leq d_{ab}. Repeating this argument for all a+1bMaa+1\prec b\prec M_{a} gives da,a+1da,a+2da,Mad_{a,a+1}\leq d_{a,a+2}\leq\cdots\leq d_{a,M_{a}}. Conversely, if MabaM_{a}\prec b\prec a, then we have bab~b\prec a\prec\widetilde{b}. The definition of b~\widetilde{b} and Lemma 3.8 now give da,b1dabd_{a,b-1}\geq d_{ab}. Hence, da,Mada,Ma+1da,a1d_{a,M_{a}}\geq d_{a,M_{a}+1}\geq\cdots d_{a,a-1}. The conclusion follows. ∎

Heuristically, if all isolation indices have roughly the same value, the distances daa~d_{a\widetilde{a}} and dbb~d_{b\widetilde{b}} should be similar. In consequence, we could expect the following property to hold:

abca~b~c~.a\prec b\prec c\Rightarrow\widetilde{a}\prec\widetilde{b}\prec\widetilde{c}. (\star)

Throughout the following lemmas, we show that a circular decomposable metric that satisfies this assumption must be monotone. We start by proving the existence of an MaM_{a} that satisfies the conditions of Lemma 3.10. This will give a way to find M(a)M(a) using a~\widetilde{a}.

Lemma 3.11.

Let (X,dX)(X,d_{X}) be a circular decomposable metric that satisfies Property (\star3.2). Let MaM_{a} be the last bb in the sequence a+1,a+2,,a1a+1,a+2,\dots,a-1 such that a~b~a\widetilde{a}\preceq\widetilde{b}\preceq a. Then MaM_{a} satisfies abMaa\preceq b\preceq M_{a} if and only if a~b~a\widetilde{a}\preceq\widetilde{b}\preceq a.

Proof.

By definition of MaM_{a}, ab~a~a\prec\widetilde{b}\prec\widetilde{a} holds for any MabaM_{a}\prec b\prec a. Thus, we only need to verify the lemma for abMaa\preceq b\preceq M_{a}. By Property (\star3.2) and the definition of MaM_{a}, a~b~Ma~a\widetilde{a}\preceq\widetilde{b}\preceq\widetilde{M_{a}}\preceq a. Hence, a~b~a\widetilde{a}\preceq\widetilde{b}\preceq a. ∎

In the next step, we verify that Property (\star3.2) implies the conditions of Lemma 3.7.

Lemma 3.12.

Let (X,dX)(X,d_{X}) be a circular decomposable metric that satisfies Property (\star3.2). Then abM(a)a\preceq b\preceq M(a) if and only if M(a)M(b)aM(a)\preceq M(b)\preceq a.

Proof.

We first show that abM(a)a\preceq b\preceq M(a) implies M(a)M(b)aM(a)\preceq M(b)\preceq a. Let cc be an element of the sequence b+1,b+2,,b1b+1,b+2,\dots,b-1. By Lemmas 3.10 and 3.11, M(b)M(b) is the last value cc that satisfies b~c~b\widetilde{b}\preceq\widetilde{c}\preceq b. Our objective is to determine the range of values of cc in which M(b)M(b) can appear.
First, we show that M(b)M(b) occurs after M(a)M(a). If abcM(a)a\preceq b\prec c\preceq M(a), Lemma 3.11 implies that a~c~ab\widetilde{a}\preceq\widetilde{c}\preceq a\preceq b. Property (\star3.2) gives a~b~c~\widetilde{a}\preceq\widetilde{b}\prec\widetilde{c}, which combines with the previous inequality to give b~cb\widetilde{b}\prec c\prec b. In particular, c=M(a)c=M(a) satisfies that condition, so M(b)M(b) occurs at the same time or after M(a)M(a).
Now we show that M(b)M(b) cannot occur after aa. If acbM(a)a\prec c\prec b\prec M(a), the characterization of M(a)M(a) says that a~b~a\widetilde{a}\preceq\widetilde{b}\preceq a. By Property (\star3.2), a~c~b~\widetilde{a}\prec\widetilde{c}\prec\widetilde{b}. Hence, a~c~b~ab\widetilde{a}\prec\widetilde{c}\prec\widetilde{b}\preceq a\prec b. Thus, b~c~b\widetilde{b}\prec\widetilde{c}\prec b does not hold, so M(b)M(b) cannot occur after aa. This finishes the proof that abM(a)M(a)M(b)aa\preceq b\preceq M(a)\Rightarrow M(a)\preceq M(b)\preceq a, and the proof of M(a)baaM(b)M(a)M(a)\preceq b\preceq a\Rightarrow a\preceq M(b)\preceq M(a) is analogous. ∎

Corollary 3.13.

Any circular decomposable metric that satisfies Property (\star3.2) is monotone.

Proof.

Lemmas 3.11 and 3.10 show that XX is unimodal at every point, and Lemma 3.12 shows that M(a)M(a) satisfies the hypothesis of Lemma 3.7. Thus, (X,dX)(X,d_{X}) is monotone. ∎

With the previous corollary, we can write a formula for the homotopy type of the VR complex of any circular decomposable metric that satisfies Property (\star3.2).

Theorem 3.14.

Let (X,dX)(X,d_{X}) be a circular decomposable metric that satisfies Property (\star3.2) and fix r>0r>0. Let GG be the 1-skeleton of VRr(X)\mathrm{VR}_{r}(X). Then

VRr(X){𝕊2l+1if l2l+1<wf(G)<l+12l+3 for some l=0,1,n2k1𝕊2lif wf(G)=l2l+1 and G dismantles to Cnk.\mathrm{VR}_{r}(X)\simeq\begin{cases}\mathbb{S}^{2l+1}&\text{if }\frac{l}{2l+1}<\operatorname{wf}(G)<\frac{l+1}{2l+3}\text{ for some }l=0,1,\dots\\ \bigvee^{n-2k-1}\mathbb{S}^{2l}&\text{if }\operatorname{wf}(G)=\frac{l}{2l+1}\text{ and }G\text{ dismantles to }C_{n^{k}}.\end{cases}
Proof.

By Corollary 3.13, (X,dX)(X,d_{X}) is monotone. Hence, GG is cyclic for any r>0r>0 and Theorem 2.13 applies. ∎

4 Non-monotone circular decomposable metrics

In this section, we describe how to compute the homology groups of circular decomposable metrics that are not monotone. However, circular decomposable metrics are not too far from being monotone, so we can use the properties of monotone metrics and cyclic graphs to compute the homology of more general circular decomposable metrics. Concretely, our strategy to compute H(VRr(X))H_{*}(\mathrm{VR}_{r}(X)) will be to split VRr(X)\mathrm{VR}_{r}(X) into the cyclic piece and the non-cyclic piece, compute the homology of the cyclic piece with 2.13, and use the Mayer-Vietoris sequence to find the homology of VRr(X)\mathrm{VR}_{r}(X).

4.1 Cyclic component of VRr(X)\mathrm{VR}_{r}(X) for a circular decomposable non-monotone XX.

Fix r>0r>0 and suppose (X,dX)(X,d_{X}) is a circular decomposable metric. For ease of notation, we denote VRr(X)\mathrm{VR}_{r}(X) with VXV_{X} and its 1-skeleton with GXG_{X}. We use the following function to test if XX is monotone.

Definition 4.1.

Let (X,dX)(X,d_{X}) be a circular decomposable metric and fix r>0r>0. Given aXa\in X, let σX(a)\sigma_{X}(a) be the last zz in the sequence a+1,a+2,,a1a+1,a+2,\dots,a-1 such that dX(c,d)rd_{X}(c,d)\leq r for all acdza\preceq c\preceq d\preceq z.

Notice that for any acdbσX(a)a\preceq c\preceq d\preceq b\preceq\sigma_{X}(a), both dcdd_{cd} and dabd_{ab} are smaller than rr, so {a,b}VX{c,d}VX\{a,b\}\in V_{X}\Rightarrow\{c,d\}\in V_{X} holds. However, we don’t know if the same implication holds when σ(a)bcda\sigma(a)\preceq b\preceq c\preceq d\preceq a. This implication holds if XX is monotone and, in that case, we can use the results from the previous section to find H(VX)H_{*}(V_{X}). If XX is not monotone, there exists aσX(a)ba\prec\sigma_{X}(a)\prec b and bcdab\preceq c\prec d\preceq a such that dabr<dcdd_{ab}\leq r<d_{cd}. Intuitively, VRr(X)\mathrm{VR}_{r}(X) is not cyclic because the edge {a,b}\{a,b\} appeared at a lower rr than it was supposed to. These are the edges we want to remove, so we set

EX:={{a,b}X:aσ(a)b and dabr<dcd for some bcda}.E_{X}:=\left\{\{a,b\}\subset X:a\prec\sigma(a)\prec b\text{ and }d_{ab}\leq r<d_{cd}\text{ for some }b\preceq c\prec d\preceq a\right\}.
Definition 4.2.

Let r>0r>0 and suppose (X,dX)(X,d_{X}) is a circular decomposable metric. We form a graph GXcG_{X}^{c} by removing the edges in EXE_{X} from GXG_{X} and define VXc:=Cl(GXc)V_{X}^{c}:=\operatorname{Cl}(G_{X}^{c}). We call VXcV_{X}^{c} the cyclic component of VRr(X)\mathrm{VR}_{r}(X).

By definition, VXcV_{X}^{c} is a cyclic complex because we removed the edges that made VXV_{X} non-cyclic. Plus, if VXV_{X} was cyclic to begin with, then VXc=VXV_{X}^{c}=V_{X}. On the other hand, let YY be the vertex set of St(EX)\operatorname{St}(E_{X}), the closed star of EXE_{X} in VXV_{X}. Notice that the subcomplex of VRr(X)\mathrm{VR}_{r}(X) induced on YY equals VY=VRr(Y)V_{Y}=\mathrm{VR}_{r}(Y) and St(EX)VY\operatorname{St}(E_{X})\subset V_{Y}.
Intuitively, we have separated VXV_{X} into its cyclic part VXcV_{X}^{c} and the subcomplex VYV_{Y} induced by edges that prevent VXV_{X} from being cyclic. We want to use these complexes and the Mayer-Vietoris sequence to compute the homology of VXV_{X}, and there are a couple of facts to verify. First, let VY:=VXcVYV_{Y}^{\prime}:=V_{X}^{c}\cap V_{Y}.

Lemma 4.3.

VYV_{Y}^{\prime} is a cyclic clique complex.

Proof.

Firstly, note that YY inherits the cyclic order from XX so that Y={y1y2ym}Y=\{y_{1}\prec y_{2}\prec\cdots\prec y_{m}\}. Also, VYV_{Y}^{\prime} is a clique complex because both VXcV_{X}^{c} and VYV_{Y} are. To verify that VYV_{Y}^{\prime} is cyclic, define σY:YY\sigma_{Y}:Y\to Y by setting σY(yi)\sigma_{Y}(y_{i}) to be the last zz from the sequence yi+1,yi+2,,yi1y_{i+1},y_{i+2},\dots,y_{i-1} such that yizσX(yi)y_{i}\preceq z\preceq\sigma_{X}(y_{i}).
Let a,b,c,dYa,b,c,d\in Y such that acdbσY(a)a\preceq c\prec d\preceq b\preceq\sigma_{Y}(a). To show that VYV_{Y}^{\prime} is cyclic, we have to prove that {a,b}VY\{a,b\}\in V_{Y}^{\prime} implies {c,d}VY\{c,d\}\in V_{Y}^{\prime}. If {a,b}VYVXc\{a,b\}\in V_{Y}^{\prime}\subset V_{X}^{c}, the fact that VXcV_{X}^{c} is cyclic and abσY(a)σX(a)a\prec b\preceq\sigma_{Y}(a)\preceq\sigma_{X}(a) means that {c,d}VXcVX\{c,d\}\in V_{X}^{c}\subset V_{X}. This also implies that {c,d}VY\{c,d\}\in V_{Y} because VYV_{Y} is the induced subcomplex of VXV_{X} on YY. Hence, {c,d}VXcVY=VY\{c,d\}\in V_{X}^{c}\cap V_{Y}=V_{Y}^{\prime}. The second condition in Definition 2.11 follows analogously. ∎

Now we make sure that we didn’t lose any simplices when splitting VXV_{X} into VXcV_{X}^{c} and VYV_{Y}.

Lemma 4.4.

VX=VXcVYV_{X}=V_{X}^{c}\cup V_{Y}.

Proof.

Let σ\sigma be a simplex of VXV_{X}. If σ\sigma does not contain an edge from EXE_{X}, then σVXc\sigma\in V_{X}^{c} by definition of VXcV_{X}^{c}. If σ\sigma does contain an edge from EXE_{X}, then σSt(EX)VY\sigma\in\operatorname{St}(E_{X})\subset V_{Y}. ∎

Lastly, we record the following for future use.

Proposition 4.5.

Let ιX:VYVXc\iota_{X}:V_{Y}^{\prime}\hookrightarrow V_{X}^{c}, ιY:VYVY\iota_{Y}:V_{Y}^{\prime}\hookrightarrow V_{Y}, τX:VXcVX\tau_{X}:V_{X}^{c}\hookrightarrow V_{X} and τ:VYVX\tau:V_{Y}\hookrightarrow V_{X} be the natural inclusions. The homology groups of VXV_{X}, VXcV_{X}^{c}, VYV_{Y}, and VYV_{Y}^{\prime} satisfy the following Mayer-Vietoris sequence:

Hk(VY)(ιX,ιY)Hk(VXc)Hk(VY)τXτYHk(VX)Hk1(VY).\cdots\to H_{k}(V_{Y}^{\prime})\xrightarrow{(\iota_{X},\iota_{Y})}H_{k}(V_{X}^{c})\oplus H_{k}(V_{Y})\xrightarrow{\tau_{X}-\tau_{Y}}H_{k}(V_{X})\to H_{k-1}(V_{Y}^{\prime})\to\cdots. (4)

4.2 Recursive computation of H(VRr(X))H_{*}(\mathrm{VR}_{r}(X)).

There is an obstacle to using Proposition 4.5. If X=YX=Y, VYV_{Y} equals VXV_{X} because we defined VYV_{Y} as an induced subcomplex. We would also have VY=VXcV_{Y}^{\prime}=V_{X}^{c}, which means that the Mayer-Vietoris sequence would give the trivial statements H(VX)=H(VY)H_{*}(V_{X})=H_{*}(V_{Y}) and H(VXc)=H(VY)H_{*}(V_{X}^{c})=H_{*}(V_{Y}^{\prime}). Hence, we assume XYX\neq Y in this section.
First we focus on finding H(VX)H_{*}(V_{X}) when VYV_{Y}^{\prime} is in the non-critical regime, i.e. l2l+1<wf(VY)<l+12l+3\frac{l}{2l+1}<\operatorname{wf}(V_{Y}^{\prime})<\frac{l+1}{2l+3} for some ll\in{\mathbb{Z}}. We begin with a technical property.

Lemma 4.6.

Let f:ABf:A\to B, g:ACg:A\to C, h:BDh:B\to D, k:CDk:C\to D be group homomorphisms and suppose

A(f,g)BChkD0A\xrightarrow{(f,g)}B\oplus C\xrightarrow{h-k}D\to 0

is exact. If gg is an injection/surjection/isomorphism, then so is hh.

Proof.

Suppose gg is injective. If h(b)=0h(b)=0, then h(b)k(0)=0h(b)-k(0)=0. By exactness, there exists aAa\in A such that f(a)=bf(a)=b and g(a)=0g(a)=0. Since gg is injective, a=0a=0, so b=f(a)=0b=f(a)=0.
Suppose gg is surjective. Given dDd\in D, there exist bBb\in B and cCc\in C such that h(b)k(c)=dh(b)-k(c)=d by exactness. Also, there exists aAa\in A such that g(a)=cg(a)=c by assumption. Exactness implies that (hfkg)(a)=0(h\circ f-k\circ g)(a)=0. Thus, d=h(b)k(c)=h(b)k(g(a))=h(bf(a))d=h(b)-k(c)=h(b)-k(g(a))=h(b-f(a)). ∎

Theorem 4.7.

Suppose l2l+1<wf(VY)<l+12l+3\frac{l}{2l+1}<\operatorname{wf}(V_{Y}^{\prime})<\frac{l+1}{2l+3}.

  1. 1.

    If ιY:H2l+1(VY)H2l+1(VY)\iota_{Y}:H_{2l+1}(V_{Y}^{\prime})\to H_{2l+1}(V_{Y}) is injective and coker(ιY)\operatorname{coker}(\iota_{Y}) is torsion-free, then

    H2l+1(VX)\displaystyle H_{2l+1}(V_{X}) H2l+1(VXc)(H2l+1(VY)/H2l+1(VY))\displaystyle\cong H_{2l+1}(V_{X}^{c})\oplus\left(H_{2l+1}(V_{Y})/H_{2l+1}(V_{Y}^{\prime})\right)
    H~k(VX)\displaystyle\widetilde{H}_{k}(V_{X}) H~k(VXc)H~k(VY) for k2l+1.\displaystyle\cong\widetilde{H}_{k}(V_{X}^{c})\oplus\widetilde{H}_{k}(V_{Y})\text{ for }k\neq 2l+1.
  2. 2.

    If H2l+1(VY)=0H_{2l+1}(V_{Y})=0, then

    H2l+1(VX)\displaystyle H_{2l+1}(V_{X}) 0\displaystyle\cong 0
    H2l+2(VX)\displaystyle H_{2l+2}(V_{X}) H2l+2(VXc)H2l+2(VY)(H2l+1(VY)/H2l+1(VXc))\displaystyle\cong H_{2l+2}(V_{X}^{c})\oplus H_{2l+2}(V_{Y})\oplus\left(H_{2l+1}(V_{Y}^{\prime})/H_{2l+1}(V_{X}^{c})\right)
    H~k(VX)\displaystyle\widetilde{H}_{k}(V_{X}) H~k(VXc)H~k(VY) for k2l+1,2l+2.\displaystyle\cong\widetilde{H}_{k}(V_{X}^{c})\oplus\widetilde{H}_{k}(V_{Y})\text{ for }k\neq 2l+1,2l+2.
Proof.

Recall Proposition 4.5:

Hk(VY)Hk(VXc)Hk(VY)Hk(VX)Hk1(VY).\cdots\to H_{k}(V_{Y}^{\prime})\to H_{k}(V_{X}^{c})\oplus H_{k}(V_{Y})\to H_{k}(V_{X})\to H_{k-1}(V_{Y}^{\prime})\to\cdots.

By Theorem 2.13, H~k(VY)\widetilde{H}_{k}(V_{Y}^{\prime}) equals {\mathbb{Z}} if k=2l+1k=2l+1 and 0 otherwise, so H~k(VX)H~k(VXc)H~k(VY)\widetilde{H}_{k}(V_{X})\cong\widetilde{H}_{k}(V_{X}^{c})\oplus\widetilde{H}_{k}(V_{Y}) for k2l+1,2l+2k\neq 2l+1,2l+2. The remaining part of the sequence is as follows:

0\displaystyle 0 H2l+2(VXc)H2l+2(VY)H2l+2(VX)\displaystyle\to H_{2l+2}(V_{X}^{c})\oplus H_{2l+2}(V_{Y})\to H_{2l+2}(V_{X}) (5)
H2l+1(VY)𝜄H2l+1(VXc)H2l+1(VY)H2l+1(VX)0.\displaystyle\xrightarrow{\partial_{*}}H_{2l+1}(V_{Y}^{\prime})\xrightarrow{\iota}H_{2l+1}(V_{X}^{c})\oplus H_{2l+1}(V_{Y})\to H_{2l+1}(V_{X})\to 0.

If ιY:H2l+1(VY)H2l+1(VY)\iota_{Y}:H_{2l+1}(V_{Y}^{\prime})\to H_{2l+1}(V_{Y}) is injective, then so is ι\iota and the bottom row becomes a short exact sequence. This forces =0\partial_{*}=0, so H2l+2(VX)H2l+2(VXc)H2l+2(VY)H_{2l+2}(V_{X})\cong H_{2l+2}(V_{X}^{c})\oplus H_{2l+2}(V_{Y}).
Finding H2l+1(VX)H_{2l+1}(V_{X}) requires two cases depending on whether H2l+1(VXc)H_{2l+1}(V_{X}^{c}) is trivial or not. If H2l+1(VXc)=0H_{2l+1}(V_{X}^{c})=0, the fact that the bottom row of (5) is a short exact sequence yields H2l+1(VX)H2l+1(VY)/H2l+1(VY)H_{2l+1}(V_{X})\cong H_{2l+1}(V_{Y})/H_{2l+1}(V_{Y}^{\prime}). If H2l+1(VXc)0H_{2l+1}(V_{X}^{c})\neq 0, we must have l2l+1<wf(VXc)<l+12l+3\frac{l}{2l+1}<\operatorname{wf}(V_{X}^{c})<\frac{l+1}{2l+3} by Theorem 2.13. The inclusion of cyclic complexes VYVXcV_{Y}^{\prime}\hookrightarrow V_{X}^{c} is induced by the inclusion of their 1-skeleta which, in turn, is a cyclic homomorphism of cyclic graphs by [AA17, Lemma 3.6]. Then, by [AA17, Proposition 4.9], VYVXcV_{Y}^{\prime}\simeq V_{X}^{c}. Hence, ιX\iota_{X} is an isomorphism, so Lemma 4.6 gives that τY:H2l+1(VY)H2l+1(VX)\tau_{Y}:H_{2l+1}(V_{Y})\to H_{2l+1}(V_{X}) is an isomorphism. Moreover, since coker(ιY)\operatorname{coker}(\iota_{Y}) is torsion-free by hypothesis, the short exact sequence

0H2l+1(VY)H2l+1(VY)coker(ιY)00\to H_{2l+1}(V_{Y}^{\prime})\to H_{2l+1}(V_{Y})\to\operatorname{coker}(\iota_{Y})\to 0

splits. Thus,

H2l+1(VX)\displaystyle H_{2l+1}(V_{X}) H2l+1(VY)H2l+1(VY)coker(ιY)\displaystyle\cong H_{2l+1}(V_{Y})\cong H_{2l+1}(V_{Y}^{\prime})\oplus\operatorname{coker}(\iota_{Y})
H2l+1(VXc)(H2l+1(VY)/H2l+1(VY)).\displaystyle\cong H_{2l+1}(V_{X}^{c})\oplus\left(H_{2l+1}(V_{Y})/H_{2l+1}(V_{Y}^{\prime})\right).

In our second claim we have H2l+1(VY)=0H_{2l+1}(V_{Y})=0. In particular, ιY\iota_{Y} is surjective. The reasoning in the paragraph above shows that H2l+1(VXc)H_{2l+1}(V_{X}^{c}) is either 0 or isomorphic to H2l+1(VY)H_{2l+1}(V_{Y}^{\prime}), and in either case, ιX\iota_{X} is surjective. Hence, ι\iota is surjective, so H2l+1(VX)=0H_{2l+1}(V_{X})=0 by exactness of (5). This observation also implies that (5) reduces to

0H2l+2(VXc)H2l+2(VY)H2l+2(VX)H2l+1(VY)ιYH2l+1(VXc)0.0\to H_{2l+2}(V_{X}^{c})\oplus H_{2l+2}(V_{Y})\to H_{2l+2}(V_{X})\xrightarrow{\partial_{*}}H_{2l+1}(V_{Y}^{\prime})\xrightarrow{\iota_{Y}}H_{2l+1}(V_{X}^{c})\to 0.

Lastly, the fact that H2l+1(VXc)=0H_{2l+1}(V_{X}^{c})=0 or H2l+1(VXc)H2l+1(VY)H_{2l+1}(V_{X}^{c})\cong H_{2l+1}(V_{Y}^{\prime}) implies that this sequence becomes a split short exact sequence upon replacing the terms H2l+1(VY)H2l+1(VXc)0H_{2l+1}(V_{Y}^{\prime})\to H_{2l+1}(V_{X}^{c})\to 0 with H2l+1(VY)/H2l+1(VXc)0H_{2l+1}(V_{Y}^{\prime})/H_{2l+1}(V_{X}^{c})\to 0. This finishes the proof. ∎

Now assume wf(VY)=l2l+1\operatorname{wf}(V_{Y}^{\prime})=\frac{l}{2l+1} for some lZl\in Z. Unlike the non-critical regime, H2l(VY)H_{2l}(V_{Y}^{\prime}) usually has more than one generator.

Theorem 4.8.

Suppose wf(VY)=l2l+1\operatorname{wf}(V_{Y}^{\prime})=\frac{l}{2l+1} for some ll\in{\mathbb{Z}}. Let KK and RR be the kernel and coimage, respectively, of the map H2l(VY)H2l(VXc)H2l(Vy)H_{2l}(V_{Y}^{\prime})\to H_{2l}(V_{X}^{c})\oplus H_{2l}(V_{y}). If both KK and RR are free, then

H2l(VX)\displaystyle H_{2l}(V_{X}) (H2l(VXc)H2l(VY))/ι(R);\displaystyle\cong\left(H_{2l}(V_{X}^{c})\oplus H_{2l}(V_{Y})\right)/\iota(R);
H2l+1(VX)\displaystyle H_{2l+1}(V_{X}) H2l+1(VXc)H2l+1(VY)K;\displaystyle\cong H_{2l+1}(V_{X}^{c})\oplus H_{2l+1}(V_{Y})\oplus K;
H~k(VX)\displaystyle\widetilde{H}_{k}(V_{X}) H~k(VXc)H~k(VY) for k2l,2l+1.\displaystyle\cong\widetilde{H}_{k}(V_{X}^{c})\oplus\widetilde{H}_{k}(V_{Y})\text{ for }k\neq 2l,2l+1.
Proof.

By Theorem 2.13, H~k(VY)\widetilde{H}_{k}(V_{Y}^{\prime}) is non-zero only when k=2lk=2l. Then Proposition 4.5 gives H~k(VX)H~k(VXc)H~k(VY)\widetilde{H}_{k}(V_{X})\cong\widetilde{H}_{k}(V_{X}^{c})\oplus\widetilde{H}_{k}(V_{Y}) for k2l,2l+1k\neq 2l,2l+1 and

0\displaystyle 0 H2l+1(VXc)H2l+1(VY)H2l+1(VX)\displaystyle\to H_{2l+1}(V_{X}^{c})\oplus H_{2l+1}(V_{Y})\to H_{2l+1}(V_{X})
H2l(VY)𝜄H2l(VXc)H2l(VY)H2l(VX)0.\displaystyle\xrightarrow{\partial_{*}}H_{2l}(V_{Y}^{\prime})\xrightarrow{\iota}H_{2l}(V_{X}^{c})\oplus H_{2l}(V_{Y})\to H_{2l}(V_{X})\to 0.

Since RR is free, the short exact sequence 0KH2l(VY)R00\to K\to H_{2l}(V_{Y}^{\prime})\to R\to 0 splits, so H2l(VY)KRH_{2l}(V_{Y}^{\prime})\cong K\oplus R. This allows us to separate the sequence above into

0\displaystyle 0 H2l+1(VXc)H2l+1(VY)H2l+1(VX)K0, and\displaystyle\to H_{2l+1}(V_{X}^{c})\oplus H_{2l+1}(V_{Y})\to H_{2l+1}(V_{X})\to K\to 0,\text{ and}
0\displaystyle 0 R𝜄H2l(VXc)H2l(VY)H2l(VX)0.\displaystyle\to R\xrightarrow{\iota}H_{2l}(V_{X}^{c})\oplus H_{2l}(V_{Y})\to H_{2l}(V_{X})\to 0.

The first sequence splits because KK is free, and the second row yields H2l(VX)coker(ι|R)H_{2l}(V_{X})\cong\operatorname{coker}(\iota|_{R}). The conclusion follows. ∎

Remark 4.9.

If we use homology with field coefficients, Theorems 4.7 and 4.8 can be used to recursively compute H(VX)H_{*}(V_{X}). The kernel, cokernel, and coimage of a linear map between finite dimensional vector spaces are themselves vector spaces and, with field coefficients, homology groups are vector spaces.

5 Wedge decomposition of VRr(X)\mathrm{VR}_{r}(X).

In the last section, we use the block decomposition of the Buneman complex and the tight span to show that VRr(X)\mathrm{VR}_{r}(X) decomposes as a wedge sum of smaller VR complexes. During this section, we assume that (X,dX)(X,d_{X}) is a totally decomposable space that is not necessarily circular decomposable.

Let’s find the split decomposition of a subset of XX that determines a block.

Definition 5.1.

Let (𝒮,α)({\mathcal{S}},\alpha) be a weighted weakly compatible split system on XX. Let 𝒮π0(I(𝒮)){\mathcal{S}}^{\prime}\in\pi_{0}(I({\mathcal{S}})) and define Y={yX:ϕyB𝒮(𝒮,α)}Y=\{y\in X:\phi_{y}\in B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)\}. We define a weighted split system (𝒮Y,αY)({\mathcal{S}}_{Y},\alpha_{Y}) on YY as follows. For each S=A|A¯𝒮S=A|\overline{A}\in{\mathcal{S}}^{\prime}, define

SY:=(AY)|(A¯Y).S_{Y}:=(A\cap Y)|(\overline{A}\cap Y).

Set 𝒮Y={SY:S𝒮}{\mathcal{S}}_{Y}=\{S_{Y}:S\in{\mathcal{S}}^{\prime}\}, and define αY:𝒮Y\alpha_{Y}:{\mathcal{S}}_{Y}\to{\mathbb{R}} by (αY)SY=αS(\alpha_{Y})_{S_{Y}}=\sum\alpha_{S} where the sum runs over all S𝒮S\in{\mathcal{S}}^{\prime} that extend SYS_{Y}. We denote YY by X𝒮X_{{\mathcal{S}}^{\prime}}.

Lemma 5.2.

If Y=X𝒮Y=X_{{\mathcal{S}}^{\prime}} for some 𝒮π0(I(𝒮)){\mathcal{S}}^{\prime}\in\pi_{0}(I({\mathcal{S}})), (d𝒮,α)|Y×Y=d𝒮Y,αY(d_{{\mathcal{S}},\alpha})|_{Y\times Y}=d_{{\mathcal{S}}_{Y},\alpha_{Y}}. As a consequence, 𝒮Y{\mathcal{S}}_{Y} is weakly compatible.

Proof.

Let S𝒮𝒮S\in{\mathcal{S}}\setminus{\mathcal{S}}^{\prime}. Since 𝒮{\mathcal{S}}^{\prime} is a block of B(𝒮,α)B({\mathcal{S}},\alpha), the connected component Cπ0(B(𝒮,α))C\in\pi_{0}(B({\mathcal{S}},\alpha)) that contains SS is different from 𝒮{\mathcal{S}}^{\prime}. In other words, B𝒮(𝒮,α)B(𝒮,α)CB_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)\subset B({\mathcal{S}},\alpha)\setminus C, so by [HKM19, Theorem 7], ϕy(A)=ϕy(A)\phi_{y}(A)=\phi_{y^{\prime}}(A) for any yyYy\neq y^{\prime}\in Y and ASA\in S. Then

dX(y,y)=d1(ϕy,ϕy)=AU(𝒮)|ϕy(A)ϕy(A)|=AU(𝒮)|ϕy(A)ϕy(A)|.d_{X}(y,y^{\prime})=d_{1}(\phi_{y},\phi_{y^{\prime}})=\sum_{A\in U({\mathcal{S}})}|\phi_{y}(A)-\phi_{y^{\prime}}(A)|=\sum_{A\in U({\mathcal{S}}^{\prime})}|\phi_{y}(A)-\phi_{y^{\prime}}(A)|.

Let A|A¯𝒮A|\overline{A}\in{\mathcal{S}}^{\prime}. If y,yAy,y^{\prime}\in A, then ϕy(A)=ϕy(A)\phi_{y}(A)=\phi_{y^{\prime}}(A) and ϕy(A¯)=ϕy(A¯)\phi_{y}(\overline{A})=\phi_{y^{\prime}}(\overline{A}). The same holds if y,yA¯y,y^{\prime}\in\overline{A}. On the other hand, if yAy\in A and yA¯y^{\prime}\in\overline{A}, then ϕy(A)=0\phi_{y}(A)=0 and ϕy(A)=αA|A¯/2\phi_{y^{\prime}}(A)=\alpha_{A|\overline{A}}/2 while ϕy(A)=αA|A¯/2\phi_{y}(A)=\alpha_{A|\overline{A}}/2 and ϕy(A)=0\phi_{y^{\prime}}(A)=0. In other words,

|ϕy(A)ϕy(A)|+|ϕy(A¯)ϕy(A¯)|=αA|A¯δA|A¯(y,y).|\phi_{y}(A)-\phi_{y^{\prime}}(A)|+|\phi_{y}(\overline{A})-\phi_{y^{\prime}}(\overline{A})|=\alpha_{A|\overline{A}}\cdot\delta_{A|\overline{A}}(y,y^{\prime}).

The same holds if yA¯y\in\overline{A} and yAy^{\prime}\in A, so

dX(y,y)=AU(𝒮)|ϕy(A)ϕy(A)|\displaystyle d_{X}(y,y^{\prime})=\sum_{A\in U({\mathcal{S}}^{\prime})}|\phi_{y}(A)-\phi_{y^{\prime}}(A)| =A|A¯𝒮|ϕy(A)ϕy(A)|+|ϕy(A¯)ϕy(A¯)|\displaystyle=\sum_{A|\overline{A}\in{\mathcal{S}}^{\prime}}|\phi_{y}(A)-\phi_{y^{\prime}}(A)|+|\phi_{y}(\overline{A})-\phi_{y^{\prime}}(\overline{A})|
=S𝒮αSδS(y,y)\displaystyle=\sum_{S\in{\mathcal{S}}^{\prime}}\alpha_{S}\cdot\delta_{S}(y,y^{\prime})
=S0𝒮Y(S𝒮extends S0αSY)δS0(y,y).\displaystyle=\sum_{S_{0}\in{\mathcal{S}}_{Y}}\left(\sum_{\begin{subarray}{c}S\in{\mathcal{S}}^{\prime}\\ \text{extends }S_{0}\end{subarray}}\alpha_{S_{Y}}\right)\cdot\delta_{S_{0}}(y,y^{\prime}).

The last equation holds because a split S𝒮S\in{\mathcal{S}}^{\prime} can only extend SYS_{Y}. By Theorem 2.5, (d𝒮,α)|Y×Y=(d𝒮,α|S)|Y×Y(d_{{\mathcal{S}},\alpha})|_{Y\times Y}=(d_{{\mathcal{S}}^{\prime},\alpha|_{S^{\prime}}})|_{Y\times Y}. The lemma follows because 𝒮{\mathcal{S}}^{\prime} restricts to 𝒮Y{\mathcal{S}}_{Y} in YY. ∎

Recall that the map κ\kappa between the Buneman complex and the Tight span is not an isometry in general. However, it does preserve the distance from an element in the Buneman complex to every point in the embedding of XX.

Proposition 5.3.

Let (𝒮,α)({\mathcal{S}},\alpha) be a weighted weakly compatible split system on XX. For any ϕB(𝒮,α)\phi\in B({\mathcal{S}},\alpha) and xXx\in X,

d1(ϕ,ϕx)=d(κ(ϕ),hx).d_{1}(\phi,\phi_{x})=d_{\infty}(\kappa(\phi),h_{x}).

As a consequence, the map κ:B(𝒮,α)T(X,d𝒮,α)\kappa:B({\mathcal{S}},\alpha)\to T(X,d_{{\mathcal{S}},\alpha}) sends Br(X;B(𝒮,α))B_{r}(X;B({\mathcal{S}},\alpha)) onto Br(X,T(X;d𝒮,α))B_{r}(X,T(X;d_{{\mathcal{S}},\alpha})).

Proof.

Let ϕB(𝒮,α)\phi\in B({\mathcal{S}},\alpha). Observe that

d(κ(ϕ),hx)=supyX|κ(ϕ)(y)hx(y)|=supyX|d1(ϕ,ϕy)dX(x,y)|.d_{\infty}(\kappa(\phi),h_{x})=\sup_{y\in X}|\kappa(\phi)(y)-h_{x}(y)|=\sup_{y\in X}|d_{1}(\phi,\phi_{y})-d_{X}(x,y)|.

Since dX(x,y)=d1(ϕx,ϕy)d_{X}(x,y)=d_{1}(\phi_{x},\phi_{y}), the reverse triangle inequality yields |d1(ϕ,ϕy)d1(x,y)|d1(ϕ,ϕx)|d_{1}(\phi,\phi_{y})-d_{1}(x,y)|\leq d_{1}(\phi,\phi_{x}). This upper bound is actually realized when we set y=xy=x, so d(κ(ϕ),hx)=d1(ϕ,ϕx)d_{\infty}(\kappa(\phi),h_{x})=d_{1}(\phi,\phi_{x}). Lastly, observe that κ\kappa maps Br(X;B(𝒮,α))B_{r}(X;B({\mathcal{S}},\alpha)) onto Br(X;T(X,d𝒮,α))B_{r}(X;T(X,d_{{\mathcal{S}},\alpha})) because

ϕBr(X;B(𝒮,α))\displaystyle\phi\in B_{r}(X;B({\mathcal{S}},\alpha)) xX such that d1(ϕx,ϕ)<r\displaystyle\Leftrightarrow\exists x\in X\text{ such that }d_{1}(\phi_{x},\phi)<r
xX such that d(hx,κ(ϕ))<r\displaystyle\Leftrightarrow\exists x\in X\text{ such that }d_{\infty}(h_{x},\kappa(\phi))<r
ϕBr(X;T(X,d𝒮,α)).\displaystyle\Leftrightarrow\phi\in B_{r}(X;T(X,d_{{\mathcal{S}},\alpha})).

Now we use the block structure of the Buneman complex to induce a decomposition of the VR complex of XX in terms of the VR complex of each X𝒮X_{{\mathcal{S}}^{\prime}} for each 𝒮π0(I(𝒮)){\mathcal{S}}^{\prime}\in\pi_{0}(I({\mathcal{S}})).

Definition 5.4.

Let (𝒮,α)({\mathcal{S}},\alpha) be a weighted weakly compatible split system on XX, and let C𝒮,αC_{{\mathcal{S}},\alpha} be the set of cut-vertices of B(𝒮,α)B({\mathcal{S}},\alpha). Let G𝒮,αG_{{\mathcal{S}},\alpha} be the graph that has vertex set κ(C𝒮,α)(T(X,d𝒮,α)\kappa(C_{{\mathcal{S}},\alpha})\cup{\mathcal{B}}(T(X,d_{{\mathcal{S}},\alpha}). We draw an edge between any cκ(C𝒮,α)c\in\kappa(C_{{\mathcal{S}},\alpha}) and every block BB such that κ(c)B\kappa(c)\in B.

We think of G𝒮,αG_{{\mathcal{S}},\alpha} as the skeleton of T(X,d𝒮,α)T(X,d_{{\mathcal{S}},\alpha}). We make each block in (T(X,d𝒮,α)){\mathcal{B}}(T(X,d_{{\mathcal{S}},\alpha})) into a vertex and join two intersecting blocks B1,B2B_{1},B_{2} through a path B1,c,B2B_{1},c,B_{2} where κ(c)\kappa(c) is the cut-vertex in the intersection of B1B_{1} and B2B_{2}. Moreover, G𝒮,αG_{{\mathcal{S}},\alpha} has a simple structure because the tight span is contractible.

Lemma 5.5.

The graph G𝒮,αG_{{\mathcal{S}},\alpha} is a tree.

Proof.

Recall that the tight span of any metric space is contractible (Proposition 2.15). In particular, each block B(T(X,d𝒮,α))B\in{\mathcal{B}}(T(X,d_{{\mathcal{S}},\alpha})) is contractible and deformation retracts onto a space homeomorphic to the star of BB as a vertex of G𝒮,αG_{{\mathcal{S}},\alpha} through a homotopy that fixes the cut-vertices in Bκ(C𝒮,α)B\cap\kappa(C_{{\mathcal{S}},\alpha}). Thus, T(X,d𝒮,α)T(X,d_{{\mathcal{S}},\alpha}) deformation retracts onto G𝒮,αG_{{\mathcal{S}},\alpha} through a homotopy relative to κ(C𝒮,α)\kappa(C_{{\mathcal{S}},\alpha}). G𝒮,αG_{{\mathcal{S}},\alpha} must then be a contractible graph, hence a tree. ∎

On the other hand, the fact that G𝒮,αG_{{\mathcal{S}},\alpha} is a tree has implications for the tight span.

Lemma 5.6.

For every 𝒮π0(I(𝒮)){\mathcal{S}}^{\prime}\in\pi_{0}(I({\mathcal{S}})), the block κ(B𝒮(𝒮,α))T(X,d𝒮,α)\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha))\subset T(X,d_{{\mathcal{S}},\alpha}) is a hyperconvex space, hence injective.

Proof.

Let piκ(B𝒮(𝒮,α))p_{i}\in\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)) and ri>0r_{i}>0 such that d(pi,pj)ri+rjd_{\infty}(p_{i},p_{j})\leq r_{i}+r_{j}. To prove that κ(B𝒮(𝒮,α))\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)) is hyperconvex, we need to find qκ(B𝒮(𝒮,α))q\in\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)) such that d(pi,q)rid_{\infty}(p_{i},q)\leq r_{i} for all ii. We know that T(X,d𝒮,α)T(X,d_{{\mathcal{S}},\alpha}) is injective, so there exists a point qT(X,d𝒮,α)q\in T(X,d_{{\mathcal{S}},\alpha}) that satisfies d(pi,q)rid_{\infty}(p_{i},q)\leq r_{i}. If qκ(B𝒮(𝒮,α))q\in\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)), we are done. If not, we claim there exists a unique cut-vertex cC𝒮,αB𝒮(𝒮,α)c\in C_{{\mathcal{S}},\alpha}\cap B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha) such that qq and κ(B𝒮(𝒮,α))\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)) are in different connected components of T(X,d𝒮,α){κ(c)}T(X,d_{{\mathcal{S}},\alpha})\setminus\{\kappa(c)\}. First, recall that κ(B𝒮(𝒮,α))\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)) is a block of T(X,d𝒮,α)T(X,d_{{\mathcal{S}},\alpha}) by Theorem 2.21, so there exists a cut-vertex κ(c)C𝒮,αB𝒮(𝒮,α)\kappa(c)\in C_{{\mathcal{S}},\alpha}\cap B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha) that separates qq from κ(B𝒮(𝒮,α))\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)). If there were two ccC𝒮,αB𝒮(𝒮,α)c\neq c^{\prime}\in C_{{\mathcal{S}},\alpha}\cap B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha) such that both κ(c)\kappa(c) and κ(c)\kappa(c^{\prime}) separated qq from κ(B𝒮(𝒮,α))\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)), then there would be a sequence of blocks B1B2BB_{1}\neq B_{2}\neq\dots\neq B_{\ell} and cut-vertices κ(c)=c0,c1,,c=κ(c)\kappa(c)=c_{0},c_{1},\dots,c_{\ell}=\kappa(c^{\prime}) of T(X,d𝒮,α)T(X,d_{{\mathcal{S}},\alpha}) such that

κ(B𝒮(𝒮,α))κ(c0)B1c1B2Bc=κ(c)κ(B𝒮(𝒮,α)).\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha))\ni\kappa(c_{0})\in B_{1}\ni c_{1}\in B_{2}\cdots B_{\ell}\ni c_{\ell}=\kappa(c^{\prime})\in\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)).

This induces a non-trivial cycle in G𝒮,αG_{{\mathcal{S}},\alpha}, which cannot exist by Lemma 5.5. This verifies the uniqueness of cC𝒮,αB𝒮(𝒮,α)c\in C_{{\mathcal{S}},\alpha}\cap B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha). Since T(X,d𝒮,α)T(X,d_{{\mathcal{S}},\alpha}) is geodesic, there exist geodesics γi\gamma_{i} from pip_{i} to qq of length less than rir_{i}. Since κ(c)\kappa(c) separates qq from κ(B𝒮(𝒮,α))\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)), these geodesics must contain κ(c)\kappa(c). Hence, d(pi,κ(c))rid_{\infty}(p_{i},\kappa(c))\leq r_{i} and κ(c)κ(B𝒮(𝒮,α))\kappa(c)\in\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)). Hence, κ(B𝒮(𝒮,α))\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)) is hyperconvex. ∎

Notice that the block structure of B(𝒮,α)B({\mathcal{S}},\alpha) induces a decomposition of VRr(X)\mathrm{VR}_{r}(X). By Theorem 2.16, VR2r(X)\mathrm{VR}_{2r}(X) is homotopy equivalent to Br(X,T(X,d𝒮,α))B_{r}(X,T(X,d_{{\mathcal{S}},\alpha})), and this decomposes into the union of the intersection of B2r(X,T(X,d𝒮,α))B_{2r}(X,T(X,d_{{\mathcal{S}},\alpha})) with each block of T(X,d𝒮,α)T(X,d_{{\mathcal{S}},\alpha}). However, thanks to Lemma 5.6, we will prove that this intersection is the VR complex of a subset of XX. That subset is given by the following definition.

Definition 5.7.

Let 𝒮π0(I(𝒮)){\mathcal{S}}^{\prime}\in\pi_{0}(I({\mathcal{S}})). For each edge (c,B𝒮(𝒮,α))G𝒮,α(c,B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha))\in G_{{\mathcal{S}},\alpha}, consider all the points xXx\in X such that ϕx\phi_{x} and B=κ(B𝒮(𝒮,α))B=\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)) are in different connected components of B(𝒮,α){c}B({\mathcal{S}},\alpha)\setminus\{c\}. Among those, choose xcx_{c} to be any point that minimizes d1(c,ϕxc)d_{1}(c,\phi_{x_{c}}) and define

X𝒮r:=X𝒮{xc:(c,B𝒮(𝒮,α))G𝒮,α}.X_{{\mathcal{S}}^{\prime}}^{r}:=X_{{\mathcal{S}}}\cup\{x_{c}:(c,B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha))\in G_{{\mathcal{S}},\alpha}\}.

We verify that the above definition does not depend on the choices of xcx_{c}.

Lemma 5.8.

The isometry type of X𝒮rX_{{\mathcal{S}}^{\prime}}^{r} is independent of the choice of xcx_{c}.

Proof.

Our strategy will be to show that for any choice of xcx_{c} and xcx_{c^{\prime}} for c,cC𝒮,ακ(B𝒮(𝒮,α))c,c\in C_{{\mathcal{S}},\alpha}\cap\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)), the distances between xcx_{c} and every point in X𝒮{xc}X_{{\mathcal{S}}^{\prime}}\cup\{x_{c^{\prime}}\} depends only on cc and and cc^{\prime}. Thus, given cC𝒮,ακ(B𝒮(𝒮,α))c\in C_{{\mathcal{S}},\alpha}\cap\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)), let xx be any point such that ϕx\phi_{x} is in the connected component of T(X,d𝒮,α){c}T(X,d_{{\mathcal{S}},\alpha})\setminus\{c\} that does not intersect κ(B𝒮(𝒮,α))\kappa(B_{\mathcal{S}}^{\prime}({\mathcal{S}},\alpha)). By Theorem 2.21, κ(c)\kappa(c) is a cut-vertex of T(X,d𝒮,α)T(X,d_{{\mathcal{S}},\alpha}). This, together with the fact that the tight span is geodesic (Proposition 2.15), implies that

d(hx,ϕ)=d(hx,c)+d(c,ϕ)d_{\infty}(h_{x},\phi)=d_{\infty}(h_{x},c)+d_{\infty}(c,\phi) (6)

for any ϕκ(B𝒮(𝒮,α))\phi\in\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)). In particular, this holds for ϕ=hy\phi=h_{y} for any yX𝒮y\in X_{{\mathcal{S}}^{\prime}}. Hence, dX(x,y)=d(hx,hy)d_{X}(x,y)=d_{\infty}(h_{x},h_{y}) only depends on d(hx,κ(c))d_{\infty}(h_{x},\kappa(c)). Since xcx_{c} minimizes d(hx,c)d_{\infty}(h_{x},c), the distance dX(xc,y)d_{X}(x_{c},y) does not depend on the choice of xcx_{c}. If cc^{\prime} is another cut-vertex in C𝒮,ακ(B𝒮(𝒮,α))C_{{\mathcal{S}},\alpha}\cap\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)), we set ϕ=c\phi=c^{\prime} in the above equation and get

d(hxc,c)=d(hxc,c)+d(c,c).d_{\infty}(h_{x_{c}},c^{\prime})=d_{\infty}(h_{x_{c}},c)+d_{\infty}(c,c^{\prime}).

Moreover, any geodesic from hxch_{x_{c}} to cc^{\prime} goes through cc by definition of xcx_{c}. Swapping the roles of cc and cc^{\prime} implies that a geodesic from hxch_{x_{c}} to hxch_{x_{c^{\prime}}} passes through both cc and cc^{\prime}, so

dX(xc,xc)=d(hxc,hxc)=d(hxc,c)+d(c,c)+d(c,hxc),d_{X}(x_{c},x_{c^{\prime}})=d_{\infty}(h_{x_{c}},h_{x_{c^{\prime}}})=d_{\infty}(h_{x_{c}},c)+d_{\infty}(c,c^{\prime})+d_{\infty}(c^{\prime},h_{x_{c^{\prime}}}),

which, again, does not depend on the choice of xcx_{c} and xcx_{c^{\prime}}. ∎

Lemma 5.9.

Let 𝒮π0(I(𝒮)){\mathcal{S}}^{\prime}\in\pi_{0}(I({\mathcal{S}})). Let TT be the space formed by attaching an edge of length d(xc,c)d_{\infty}(x_{c},c) to κ(B𝒮(𝒮,α))T(X,dX)\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha))\subset T(X,d_{X}) at cc for every cut-vertex cC𝒮,ακ(B𝒮(𝒮,α))c\in C_{{\mathcal{S}},\alpha}\cap\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)). Then TT is injective and X𝒮rTX_{{\mathcal{S}}^{\prime}}^{r}\hookrightarrow T.

Proof.

By Lemma 5.6, the block κ(B𝒮(𝒮,α))T(X,dX)\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha))\subset T(X,d_{X}) is an injective space. Let EcE_{c} be a line segment of length d(xc,c)d_{\infty}(x_{c},c) for every cut-vertex cC𝒮,ακ(B𝒮(𝒮,α))c\in C_{{\mathcal{S}},\alpha}\cap\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)). A line segment with the usual metric is hyperconvex hence injective, and the wedge of two injective spaces is injective by [LMO22, Lemma 6.2]. Hence, TT is injective. Now consider the map i:X𝒮rTi:X_{{\mathcal{S}}^{\prime}}^{r}\hookrightarrow T that sends each xX𝒮x\in X_{{\mathcal{S}}} to hxκ(B𝒮(𝒮,α))h_{x}\in\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)) and each xcX𝒮rX𝒮x_{c}\in X_{{\mathcal{S}}^{\prime}}^{r}\setminus X_{{\mathcal{S}}} to the boundary point of EcE_{c} that is different from cc. By equation (6), ii is an isometric embedding. ∎

Below we realize the VR complex of X𝒮rX_{{\mathcal{S}}^{\prime}}^{r} as a subset of a block of T(X,d𝒮,α)T(X,d_{{\mathcal{S}},\alpha}).

Lemma 5.10.

If 𝒮π0(I(𝒮)){\mathcal{S}}^{\prime}\in\pi_{0}(I({\mathcal{S}})),

VR2r(X𝒮r)(Br(X𝒮r;T(X,d𝒮,α))κ(B𝒮(𝒮,α))){xcX𝒮r:d(ϕxc,c)<r}.\mathrm{VR}_{2r}(X_{{\mathcal{S}}^{\prime}}^{r})\simeq\left(B_{r}(X_{{\mathcal{S}}^{\prime}}^{r};T(X,d_{{\mathcal{S}},\alpha}))\cap\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha))\right)\sqcup\{x_{c}\in X_{{\mathcal{S}}^{\prime}}^{r}:d_{\infty}(\phi_{x_{c}},c)<r\}.
Proof.

Let B1,,BB_{1},\dots,B_{\ell} be the connected components of Br(X𝒮r;T(X,d𝒮,α))κ(B𝒮(𝒮,α))B_{r}(X_{{\mathcal{S}}^{\prime}}^{r};T(X,d_{{\mathcal{S}},\alpha}))\cap\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)). If d1(xc,c)rd_{1}(x_{c},c)\leq r, equation (6) implies that Br(hxc,T(X,d𝒮,α))κ(B𝒮(𝒮,α))B_{r}(h_{x_{c}},T(X,d_{{\mathcal{S}},\alpha}))\cap\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)) is a ball centered at κ(c)\kappa(c) of radius rd(hx,κ(c))r-d_{\infty}(h_{x},\kappa(c)). If d1(xc,c)>rd_{1}(x_{c},c)>r, then Br(hxc,T(X,d𝒮,α))κ(B𝒮(𝒮,α))=B_{r}(h_{x_{c}},T(X,d_{{\mathcal{S}},\alpha}))\cap\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha))=\emptyset. Hence, Br(X𝒮r;T(X,d𝒮,α))κ(B𝒮(𝒮,α))B_{r}(X_{{\mathcal{S}}^{\prime}}^{r};T(X,d_{{\mathcal{S}},\alpha}))\cap\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)) equals Br(X𝒮;T(X,d𝒮,α))κ(B𝒮(𝒮,α))B_{r}(X_{{\mathcal{S}}^{\prime}};T(X,d_{{\mathcal{S}},\alpha}))\cap\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)) union with the balls centered at κ(c)\kappa(c) of radius rd(hx,κ(c))r-d_{\infty}(h_{x},\kappa(c)) for those cc for which d1(ϕxc,c)rd_{1}(\phi_{x_{c}},c)\leq r.
On the other hand, let TT be the injective space from Lemma 5.9. Let EcE_{c} be the edge attached at κ(c)\kappa(c) for cC𝒮,ακ(B𝒮(𝒮,α))c\in C_{{\mathcal{S}},\alpha}\cap\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)). A homotopy equivalence that contracts the edges of TT of length less than rr to their gluing points in κ(B𝒮(𝒮,α))\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)) induces a homotopy equivalence Br(X𝒮r;T)Br(X𝒮r;T)κ(B𝒮(𝒮,α))B_{r}(X_{{\mathcal{S}}^{\prime}}^{r};T)\simeq B_{r}(X_{{\mathcal{S}}^{\prime}}^{r};T)\cap\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)). We can then apply a homotopy equivalence to contract every edge EcE_{c} to xcx_{c} whenever d1(ϕxc,c)>rd_{1}(\phi_{x_{c}},c)>r. Thus, by Theorem 2.16,

VR2r(X𝒮r)\displaystyle\mathrm{VR}_{2r}(X_{{\mathcal{S}}^{\prime}}^{r}) Br(X𝒮r;T)Br(X𝒮r;T)κ(B𝒮(𝒮,α)){xcX𝒮r:d1(ϕxc,c)<r}\displaystyle\simeq B_{r}(X_{{\mathcal{S}}^{\prime}}^{r};T)\simeq B_{r}(X_{{\mathcal{S}}^{\prime}}^{r};T)\cap\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha))\cup\{x_{c}\in X_{{\mathcal{S}}^{\prime}}^{r}:d_{1}(\phi_{x_{c}},c)<r\}
Br(X𝒮r;T(X,d𝒮,α))κ(B𝒮(𝒮,α)){xcX𝒮r:d1(ϕxc,c)<r}.\displaystyle\cong B_{r}(X_{{\mathcal{S}}^{\prime}}^{r};T(X,d_{{\mathcal{S}},\alpha}))\cap\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha))\cup\{x_{c}\in X_{{\mathcal{S}}^{\prime}}^{r}:d_{1}(\phi_{x_{c}},c)<r\}.

Definition 5.11.

Given r0r\geq 0 and a choice of X𝒮rX_{{\mathcal{S}}^{\prime}}^{r} for each 𝒮π0(I(𝒮)){\mathcal{S}}^{\prime}\in\pi_{0}(I({\mathcal{S}})), we define G𝒮,αrG_{{\mathcal{S}},\alpha}^{r} as follows. For each block 𝒮π0(I(𝒮)){\mathcal{S}}^{\prime}\in\pi_{0}(I({\mathcal{S}})),

  • Replace the vertex κ(B𝒮(𝒮,α))\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)) with one vertex for each connected component B1,,BLB_{1},\dots,B_{L} of Br(X𝒮r,T(X,d𝒮,α))κ(B𝒮(𝒮,α))B_{r}(X_{{\mathcal{S}}^{\prime}}^{r},T(X,d_{{\mathcal{S}},\alpha}))\cap\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)).

  • An edge of the form (c,κ(B𝒮(𝒮,α)))\left(c,\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha))\right) is replaced with the edge (c,Bi)(c,B_{i}) if cBic\in B_{i} and d1(ϕxc,c)rd_{1}(\phi_{x_{c}},c)\leq r (if any).

Note that, similarly to Lemma 5.8, G𝒮,αrG_{{\mathcal{S}},\alpha}^{r} does not depend on the specific choice of xcx_{c}.

Now, think of G𝒮,αrG_{{\mathcal{S}},\alpha}^{r} as an indexing category where an edge {c,B}\{c,B\} with cC𝒮,αc\in C_{{\mathcal{S}},\alpha} becomes an arrow cBc\to B. Define a functor F𝒮,αr:G𝒮,αrTopF_{{\mathcal{S}},\alpha}^{r}:G_{{\mathcal{S}},\alpha}^{r}\to\mathrm{Top} by

cC𝒮,α\displaystyle c\in C_{{\mathcal{S}},\alpha} cT(X,d𝒮,α)\displaystyle\mapsto c\in T(X,d_{{\mathcal{S}},\alpha})
Bπ0(Br(X𝒮r;T(X,d𝒮,α))κ(B𝒮(𝒮,α)))\displaystyle B\in\pi_{0}\left(B_{r}(X_{{\mathcal{S}}^{\prime}}^{r};T(X,d_{{\mathcal{S}},\alpha}))\cap\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha))\right) BT(X,d𝒮,α).\displaystyle\mapsto B\subset T(X,d_{{\mathcal{S}},\alpha}).

F𝒮,αrF_{{\mathcal{S}},\alpha}^{r} sends each arrow (c,B)(c,B) to the inclusion map F𝒮,αr(c)F𝒮,αr(B)F_{{\mathcal{S}},\alpha}^{r}(c)\hookrightarrow F_{{\mathcal{S}},\alpha}^{r}(B).

This functor F𝒮,αrF_{{\mathcal{S}},\alpha}^{r} encodes the block decomposition of VR2r(X)\mathrm{VR}_{2r}(X).

Theorem 5.12.

For any r0r\geq 0,

colimF𝒮,αrVR2r(X){cC𝒮,α:c is an isolated vertex of G𝒮,αr}.\operatorname{colim}F_{{\mathcal{S}},\alpha}^{r}\simeq\mathrm{VR}_{2r}(X)\sqcup\{c\in C_{{\mathcal{S}},\alpha}:c\text{ is an isolated vertex of }G_{{\mathcal{S}},\alpha}^{r}\}.
Proof.

Given 𝒮π0(I(𝒮)){\mathcal{S}}^{\prime}\in\pi_{0}(I({\mathcal{S}})), let E𝒮:=Br(X𝒮;T(X,d𝒮,α))κ(B𝒮(𝒮,α))E_{{\mathcal{S}}^{\prime}}:=B_{r}(X_{{\mathcal{S}}^{\prime}};T(X,d_{{\mathcal{S}},\alpha}))\cap\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha)). Notice that any two B1π0(E𝒮1)B_{1}\in\pi_{0}(E_{{\mathcal{S}}_{1}}), B2π0(E𝒮2)B_{2}\in\pi_{0}(E_{{\mathcal{S}}_{2}}) intersect in a common cut-vertex cc if and only if G𝒮,αrG_{{\mathcal{S}},\alpha}^{r} has edges (c,B1),(c,B2)(c,B_{1}),(c,B_{2}). Let \sim denote the identification of the point cB1c\in B_{1} and the point cB2c\in B_{2} whenever (c,B1),(c,B2)(c,B_{1}),(c,B_{2}) are edges of G𝒮,αrG_{{\mathcal{S}},\alpha}^{r}. Then

colimF𝒮,αr\displaystyle\operatorname{colim}F_{{\mathcal{S}},\alpha}^{r} (v vertex of G𝒮,αrF𝒮,αr(v))/\displaystyle\simeq\left(\bigsqcup_{v\text{ vertex of }G_{{\mathcal{S}},\alpha}^{r}}F_{{\mathcal{S}},\alpha}^{r}(v)\right)/\sim
[𝒮π0(I(𝒮))(Bπ0(E𝒮)B)]{cC𝒮,α:c is an isolated vertex of G𝒮,αr}\displaystyle\simeq\left[\bigcup_{{\mathcal{S}}^{\prime}\in\pi_{0}(I({\mathcal{S}}))}\left(\bigcup_{B\in\pi_{0}(E_{{\mathcal{S}}^{\prime}})}B\right)\right]\sqcup\{c\in C_{{\mathcal{S}},\alpha}:c\text{ is an isolated vertex of }G_{{\mathcal{S}},\alpha}^{r}\}
=(𝒮π0(I(𝒮))E𝒮){cC𝒮,α:c is an isolated vertex of G𝒮,αr}.\displaystyle=\left(\bigcup_{{\mathcal{S}}^{\prime}\in\pi_{0}(I({\mathcal{S}}))}E_{{\mathcal{S}}^{\prime}}\right)\sqcup\{c\in C_{{\mathcal{S}},\alpha}:c\text{ is an isolated vertex of }G_{{\mathcal{S}},\alpha}^{r}\}.

Then by Theorem 2.16,

VR2r(X)\displaystyle\mathrm{VR}_{2r}(X) Br(X;T(X,d𝒮,α))\displaystyle\simeq B_{r}(X;T(X,d_{{\mathcal{S}},\alpha}))
=𝒮π0(I(𝒮))Br(X;T(X,d𝒮,α))κ(B𝒮(𝒮,α))\displaystyle=\bigcup_{{\mathcal{S}}^{\prime}\in\pi_{0}(I({\mathcal{S}}))}B_{r}(X;T(X,d_{{\mathcal{S}},\alpha}))\cap\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha))
=𝒮π0(I(𝒮))Br(X𝒮r;T(X,d𝒮,α))κ(B𝒮(𝒮,α))\displaystyle=\bigcup_{{\mathcal{S}}^{\prime}\in\pi_{0}(I({\mathcal{S}}))}B_{r}(X_{{\mathcal{S}}^{\prime}}^{r};T(X,d_{{\mathcal{S}},\alpha}))\cap\kappa(B_{{\mathcal{S}}^{\prime}}({\mathcal{S}},\alpha))
=𝒮π0(I(𝒮))E𝒮.\displaystyle=\bigcup_{{\mathcal{S}}^{\prime}\in\pi_{0}(I({\mathcal{S}}))}E_{{\mathcal{S}}^{\prime}}.

The work of this section pays off below. We prove that the homology of the VR complex of XX decomposes in terms of the homology of VRr(X𝒮r)\mathrm{VR}_{r}(X_{{\mathcal{S}}^{\prime}}^{r}).

Theorem 5.13.

Let (X,dX)(X,d_{X}) be a totally decomposable metric with split system (𝒮,α)({\mathcal{S}},\alpha). For any k1k\geq 1,

Hk(VRr(X))𝒮π0(I(𝒮))Hk(VRr(X𝒮r)).H_{k}(\mathrm{VR}_{r}(X))\cong\bigoplus_{{\mathcal{S}}^{\prime}\in\pi_{0}(I({\mathcal{S}}))}H_{k}(\mathrm{VR}_{r}(X_{{\mathcal{S}}^{\prime}}^{r})).

For k=0k=0, the equation holds after identifying the class of xx with the class of any xcX𝒮rx_{c}\in X_{{\mathcal{S}}^{\prime}}^{r} for which x=xcx=x_{c}.

Proof.

Let 𝒮1,,𝒮L{\mathcal{S}}_{1},\dots,{\mathcal{S}}_{L} be the connected components of I(𝒮)I({\mathcal{S}}) and let B1,,BLB_{1},\dots,B_{L} be their corresponding blocks in T(X,d𝒮,α)T(X,d_{{\mathcal{S}},\alpha}). Let Xi:=X𝒮ir/2X_{i}:=X_{{\mathcal{S}}_{i}}^{r/2} and Vi:=Br/2(Xi;T(X,d𝒮,α))BiV_{i}:=B_{r/2}(X_{i};T(X,d_{{\mathcal{S}},\alpha}))\cap B_{i}. Lastly, let V¯:=V1V\overline{V}_{\ell}:=V_{1}\cup\cdots\cup V_{\ell}. We claim that

Hk(V¯)i=1Hk(Vi).H_{k}(\overline{V}_{\ell})\cong\bigoplus_{i=1}^{\ell}H_{k}(V_{i}). (7)

The claim is clear for =1\ell=1. For >1\ell>1, we have the Mayer-Vietoris sequence

Hk(V¯1V)Hk(V¯1)Hk(V)Hk(V¯)Hk1(V¯1V).\cdots\to H_{k}(\overline{V}_{\ell-1}\cap V_{\ell})\to H_{k}(\overline{V}_{\ell-1})\oplus H_{k}(V_{\ell})\to H_{k}(\overline{V}_{\ell})\to H_{k-1}(\overline{V}_{\ell-1}\cap V_{\ell})\to\cdots.

Notice that VViV_{\ell}\cap V_{i} is either empty or a cut-vertex cic_{i}, so V¯1V\overline{V}_{\ell-1}\cap V_{\ell} is a set of isolated points. Hence, (7) holds for k2k\geq 2. In order for the claim to hold for k=1k=1, we need the boundary map H1(V¯)H0(V¯1V)H_{1}(\overline{V}_{\ell})\to H_{0}(\overline{V}_{\ell-1}\cap V_{\ell}) to be 0. This happens if and only if the map

H0(V¯1V)H0(V¯1)H0(V)H_{0}(\overline{V}_{\ell-1}\cap V_{\ell})\to H_{0}(\overline{V}_{\ell-1})\oplus H_{0}(V_{\ell})

is injective. In other words, we need to show that if any pair of non-empty intersections {ci}=ViV\{c_{i}\}=V_{i}\cap V_{\ell} and {cj}=VjV\{c_{j}\}=V_{j}\cap V_{\ell} has cicjc_{i}\neq c_{j}, then cic_{i} and cjc_{j} belong to different connected components of V¯1\overline{V}_{\ell-1} or VV_{\ell}. Indeed, this must happen by Lemma 5.5 because if both cic_{i} and cjc_{j} belonged to the same connected components of V¯1\overline{V}_{\ell-1} and VV_{\ell}, then G𝒮,αG_{{\mathcal{S}},\alpha} would have a cycle. Hence, (7) holds for all =1,,m\ell=1,\dots,m. The theorem for k1k\geq 1 now follows from Lemma 5.10.
Lastly, let k=0k=0. Every point xXx\in X belongs to the Xi0X_{i_{0}} for which ϕxBi0\phi_{x}\in B_{i_{0}}. However, xx could also belong to XiX_{i} if x=xcx=x_{c} for some cC𝒮,αBic\in C_{{\mathcal{S}},\alpha}\cap B_{i}, ii0i\neq i_{0}. This introduces duplicate connected components in 𝒮π0(I(𝒮))Hk(VRr(X𝒮r))\displaystyle\bigoplus_{{\mathcal{S}}^{\prime}\in\pi_{0}(I({\mathcal{S}}))}H_{k}(\mathrm{VR}_{r}(X_{{\mathcal{S}}^{\prime}}^{r})). We get the correct H0(VRr(X))H_{0}(\mathrm{VR}_{r}(X)) after identifying these classes. ∎

References

  • [AA17] Michał Adamaszek and Henry Adams. The Vietoris-Rips complexes of a circle. Pacific Journal of Mathematics, 290(1):1–40, 2017.
  • [BD92] Hans-Jürgen Bandelt and Andreas Dress. A canonical decomposition theory for metrics on a finite set. Advances in Mathematics, 92(1):47–105, 1992.
  • [DH01] Andreas Dress and Katharina T. Huber. Totally split-decomposable metrics of combinatorial dimension two. Annals of Combinatorics, 5(1):99–112, Jun 2001.
  • [DHM98] Andreas Dress, Katharina T. Huber, and Vincent Moulton. A comparison between two distinct continuous models in projective cluster theory: The median and the tight-span construction. Annals of Combinatorics, 2(4):299–311, Dec 1998.
  • [Far97] Martin Farach. Recognizing circular decomposable metrics. Journal of Computational Biology, 4(2):157–162, 1997. PMID: 9228614.
  • [HKM06] Katharina T. Huber, Jack H. Koolen, and Vincent Moulton. On the structure of the tight-span of a totally split-decomposable metric. European Journal of Combinatorics, 27(3):461–479, 2006.
  • [HKM19] Katharina T. Huber, Jack H. Koolen, and Vincent Moulton. The polytopal structure of the tight-span of a totally split-decomposable metric. Discrete Mathematics, 342(3):868–878, 2019.
  • [LMO22] Sunhyuk Lim, Facundo Memoli, and Osman Berat Okutan. Vietoris-Rips persistent homology, injective metric spaces, and the filling radius. Algebraic and Geometric Topology, 2022.