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

Intersection properties of finite disk collections

Jesús F. Espinoza [email protected]  and  Cynthia G. Esquer-Pérez [email protected] Departamento de Matemáticas, Universidad de Sonora, México
Abstract.

In this work we study the intersection properties of a finite disk system in the euclidean space. We accomplish this by utilizing subsets of spheres with varying dimensions and analyze specific points within them, referred to as poles. Additionally, we introduce two applications: estimating the common scale factor for the radii that makes the re-scaled disks intersects in a single point, this is the Čech scale, and constructing the minimal Axis-Aligned Bounding Box (AABB) that encloses the intersection of all disks in the system.

1. Introduction

One of the new techniques developed for the analysis of large clusters of information, known as Big Data, is Topological Data Analysis (TDA). In TDA, simplicial complexes associated with the data are constructed. These structures include the Vietoris-Rips complex, the Čech complex, and the piecewise linear lower star complex, among others. Of special interest to us is the generalized Čech complex structure. Although the standard Čech complex is formed by intersecting a collection of disks with a fixed radius, the generalized version allows varying radii. This flexibility enables us to highlight specific data points by assigning or weighting them with larger and/or more rapidly expanding balls to them, while de-emphasizing others by using smaller and/or slower growing balls. This approach proves valuable for handling noisy data sets, offering an alternative to discarding data that may not meet a specific significance threshold [1].

Understanding the patterns of intersections and the timing of intersections among a set of disks in d\mathbb{R}^{d}, each with potentially different radii, is a fundamental problem. This leads to the exploration of the generalized Čech complex structure, which captures the intersection information of these disks, regardless of their radii. Rescaling the radii by the same factor, we obtain a filtered generalized Čech complex, where the associated simplicial complexes evolve as the scale parameter varies. In particular, in [4], algorithms are provided to calculate the generalized Čech complex in 2\mathbb{R}^{2}, and [3] presents an algorithm to determine the Čech scale for a collection of disks in the plane.

To establish the necessary foundation for our study, Section 2 introduces crucial concepts and notation that will be used throughout the article and we focus on analyzing the intersection of a disk system in d\mathbb{R}^{d}. We start by investigating the intersection of two disks in Subsection 2.1 and then expand our analysis to a system of mm disks in Subsection 2.2. By applying Helly’s Theorem, we prove that it is sufficient to examine the intersection of all subsystems consisting of d+1d+1 disks in order to determine if the system has a nonempty intersection.

In Section 3, we define Vietoris-Rips systems and Čech systems, together with presenting results regarding the Rips scale and the Čech scale, as well as their connections. In Subsection 3.1, we present an algorithm that can determine whether the intersection of the system is empty or non-empty. This is achieved by exclusively computing the poles of subsystems of disks (or spheres). Finally, in Subsection 3.2, we introduce the algorithm that computes an approximation to the Čech scale using the numerical bisection method.

Additionally, in Section 4 we incorporate the concept of an minimal axis-aligned bounding box (AABB) into our methodology. An AABB is a rectangular parallelepiped whose faces are perpendicular to the basis vectors. These bounding boxes frequently arise in spatial subdivision problems, such as ray tracing [5] and collision detection [2]. In this paper, we study AABBs to enclose the intersection of a finite collection of disks. This approach proves valuable for discerning whether the collection intersects at a singular point or not. In this section, we also provide an algorithm for constructing the AABB of a disk system.

2. Intersection properties of sphere systems

Throughout this work, we will refer to a dd-disk system MM, or simply a disk system, as a finite collection of closed disks in d\mathbb{R}^{d} with positive and not necessarily equal radii, i.e.,

M={Di(ci;ri)dcid,ri>0,1im<}.M=\{D_{i}(c_{i};r_{i})\subset\mathbb{R}^{d}\mid c_{i}\in\mathbb{R}^{d},r_{i}>0,1\leq i\leq m<\infty\}.

Moreover, in order to study the intersection properties of a disk system MM with the approach addressed in Sections 4 and 3 of this work, we will conduct a study in this section of the intersection properties of the spheres corresponding to the boundaries of each disk in MM, which we call a sphere system and denote by M\partial M,

M={DidDiM},\partial M=\{\partial D_{i}\subset\mathbb{R}^{d}\mid D_{i}\in M\},

where \partial denotes the topological boundary operator.

Following the notation in [6], we introduce the following generalization of the sphere.

Definition 1.

An ii-sphere in d\mathbb{R}^{d} is the intersection of a sphere with an affine subspace of dimension ii.

Of course, the notions of a sphere (as a (d1)(d-1)-dimensional surface) and a dd-sphere in d\mathbb{R}^{d} agree. However, an ii-sphere in d\mathbb{R}^{d} can also be viewed as the intersection of dd-spheres. For instance, the intersection of two spheres typically occurs in a hyperplane, forming a (d1)(d-1)-sphere in d\mathbb{R}^{d}. When another dd-sphere intersects this configuration, the result may be a (d1)(d-1)-sphere, a (d2)(d-2)-sphere, a 0-sphere (a single point), or it might even be empty, all within the same hyperplane. For a disk system M={Di(ci;ri)}M=\{D_{i}(c_{i};r_{i})\} composed of mm disks, where {c1,,cm}\{c_{1},\dots,c_{m}\} is a set in general position in d\mathbb{R}^{d}, the maximum dimension of the affine subspace associated with the ii-sphere, obtained from the intersection of all the spheres in M\partial M, is at most dm+1d-m+1, or equivalently, i=dm+1i=d-m+1. This conclusion is drawn from [6, Theorem 2.1] and the fact that the affine hull of {c1,,cm}\{c_{1},\dots,c_{m}\} is of dimension m1m-1. Consequently, the following result is proven.

Lemma 2.

Let M={D1(c1;r1),,Dm(cm;rm)}M=\{D_{1}(c_{1};r_{1}),\ldots,D_{m}(c_{m};r_{m})\} be disk system such that {c1,,cm}\{c_{1},\dots,c_{m}\} is a set in general position in d\mathbb{R}^{d}. Then, the possibilities for the set DiMDi\cap_{D_{i}\in M}\partial D_{i} are:

  1. (1)

    the empty set;

  2. (2)

    a single point;

  3. (3)

    a (dm+1)(d-m+1)-sphere.

Remarkable points in ii-spheres that will play a key role in the rest of the article are the poles. Let πi:d\pi_{i}:\mathbb{R}^{d}\longrightarrow\mathbb{R} be the canonical projection on the ii-th factor for i=1,,di=1,...,d, and let {e1,e2,,ed}\{e_{1},e_{2},...,e_{d}\} be the standard basis of d\mathbb{R}^{d}.

Definition 3.

Let eqe_{q} be the qq-th vector of the canonical base of d\mathbb{R}^{d}. An eqe_{q}-north (south) pole of an ii-sphere SS in d\mathbb{R}^{d} is a point on SS whose projection on the qq-th coordinate is maximum (minimum). In other words, xSx\in S is the eqe_{q}-north pole if πq(y)πq(x)\pi_{q}(y)\leq\pi_{q}(x) for all yS{x}y\in S-\{x\}, where πq\pi_{q} represents the projection onto the qq-th coordinate.

We denote the eqe_{q}-north pole of SS by sq+s_{q}^{+} and the eqe_{q}-south pole by sqs_{q}^{-}.

An ii-sphere can have a single eqe_{q}-pole (north or south) or an infinite number of them, which occurs when a normal vector to the affine space containing the ii-sphere is aligned with the vector eqe_{q}. We are interested in finding the eqe_{q}-poles of (dm+1)(d-m+1)-spheres originating from disk systems M={D1(c1;r1),,Dm(cm;rm)}M=\{D_{1}(c_{1};r_{1}),\ldots,D_{m}(c_{m};r_{m})\}, by taking the intersection j=1mDj\cap_{j=1}^{m}\partial D_{j}. Such (dm+1)(d-m+1)-spheres will be denoted by SM(c;r)S_{M}(c;r), to emphasize the disk system MM, as well as its center and radius.

Lemma 4.

Let M={D1,,Dm}M=\{D_{1},\ldots,D_{m}\} be a dd-disk system such that j=1mDj\bigcap_{j=1}^{m}D_{j}\neq\emptyset, and let pp be a point in j=1mDj\bigcap_{j=1}^{m}D_{j} such that πq(p)πq(x)\pi_{q}(p)\leq\pi_{q}(x) (resp. πq(p)πq(x)\pi_{q}(p)\geq\pi_{q}(x)) for every xx in j=1mDj\bigcap_{j=1}^{m}D_{j}. Then, there exists an ii-sphere S=Dj1DjiS=\partial D_{j_{1}}\cap\cdots\cap\partial D_{j_{i}} such that pp is in SS and pp is the eqe_{q}-south pole (resp. eqe_{q}-north pole) of SS.

Proof.

Since j=1mDj\cap_{j=1}^{m}D_{j}\neq\emptyset, then (j=1mDj)\partial(\cap_{j=1}^{m}D_{j})\neq\emptyset, (j=1mDj)j=1mDj\partial(\cap_{j=1}^{m}D_{j})\subset\cap_{j=1}^{m}D_{j} due to the closedness of the sets DjD_{j}, for j=1,,mj=1,\ldots,m, and p(j=1mDj)p\in\partial(\cap_{j=1}^{m}D_{j}).

On the other hand, since (j=1mDj)i=1mDj\partial(\cap_{j=1}^{m}D_{j})\subset\cup_{i=1}^{m}\partial D_{j}, there exist indices j1,,jij_{1},\ldots,j_{i} such that pDjrp\in\partial D_{j_{r}} for any r=1,,ir=1,\ldots,i; let Λ(p)={j1,,ji}{1,,m}\Lambda(p)=\{j_{1},\ldots,j_{i}\}\subseteq\{1,\ldots,m\} be a maximal subset of indices such that pDjp\in D_{j} if and only if jΛ(p)j\in\Lambda(p). We claim that pr=1iDjrp\in\cap_{r=1}^{i}\partial D_{j_{r}} is the eqe_{q}-south pole of S:=r=1iDjrS:=\cap_{r=1}^{i}\partial D_{j_{r}}.

In effect, let VpdV_{p}\subset\mathbb{R}^{d} be an open neighborhood of pp sufficiently small such that:

  1. (1)

    Every xVp(j=1mDj)x\in V_{p}\cap\partial(\cap_{j=1}^{m}D_{j}) has as maximal set of indices a proper subset of Λ(p)\Lambda(p),

  2. (2)

    SVp(j=1mDj)S\cap V_{p}\subset\partial(\cap_{j=1}^{m}D_{j}).

The first condition can be guaranteed by the finiteness of the disk system MM, and the second condition is a consequence of the maximality of the set Λ(p)\Lambda(p). Therefore, πq(p)πq(x)\pi_{q}(p)\leq\pi_{q}(x) for every xSVpx\in S\cap V_{p}, which is equivalent to the fact that πq(p)πq(x)\pi_{q}(p)\leq\pi_{q}(x) for every xSx\in S, in the case of ii-spheres.

2.1. Sphere systems with two spheres

In the following two lemmas we provide the computations to determine the center, radius and poles for a (d1)(d-1)-sphere given by the intersection of two disks in d\mathbb{R}^{d}.

Lemma 5.

Let M={D1(c1;r1),D2(c2;r2)}M=\{D_{1}(c_{1};r_{1}),D_{2}(c_{2};r_{2})\} be a disk system with two dd-disks such that D1D2\partial D_{1}\cap\partial D_{2} is a (d1)(d-1)-sphere S=SM(c;r)S=S_{M}(c;r) with center cc and radius rr. Then,

c=12(1+r22r12c2c12)c1+12(1+r12r22c2c12)c2c=\frac{1}{2}\left(1+\frac{r_{2}^{2}-r_{1}^{2}}{\|c_{2}-c_{1}\|^{2}}\right)c_{1}+\frac{1}{2}\left(1+\frac{r_{1}^{2}-r_{2}^{2}}{\|c_{2}-c_{1}\|^{2}}\right)c_{2}

and

r=2s(sc2c1)(sr1)(sr2)c2c1r=\frac{2\sqrt{s(s-\|c_{2}-c_{1}\|)(s-r_{1})(s-r_{2})}}{\|c_{2}-c_{1}\|}

where s=12(c2c1+r1+r2)s=\frac{1}{2}(\|c_{2}-c_{1}\|+r_{1}+r_{2}).

Proof.

Let Π\Pi be the hyperplane containing the (d1)(d-1)-sphere SS, which is defined by the equation:

(1) i=1d(kihi)xi12i=1d(ki2hi2)=r12r222,\sum_{i=1}^{d}(k_{i}-h_{i})x_{i}-\frac{1}{2}\sum_{i=1}^{d}(k_{i}^{2}-h_{i}^{2})=\frac{r_{1}^{2}-r_{2}^{2}}{2},

where c1=(h1,,hd)c_{1}=(h_{1},\ldots,h_{d}) and c2=(k1,,kd)c_{2}=(k_{1},\ldots,k_{d}). Then the normal vector of the hyperplane Π\Pi is given by N:=c2c1=(k1h1,,kdhd)N:=c_{2}-c_{1}=(k_{1}-h_{1},\ldots,k_{d}-h_{d}), and the center cc of SS is determined by the intersection point of the hyperplane Π\Pi with the perpendicular line that passes through the center c1c_{1} of D1D_{1}. This line can be parameterized as γ:tc1+tN=(x1(t),,xd(t))\gamma:t\mapsto c_{1}+tN=(x_{1}(t),\ldots,x_{d}(t)), such that γ(0)=c1\gamma(0)=c_{1} and γ(1)=c2\gamma(1)=c_{2}. We can compute the intersection point c=γ(t)c=\gamma(t_{*}) of Π\Pi and γ([0,1])\gamma([0,1]), for any t[0,1]t_{*}\in[0,1], by substituting it in (1),

i=1d(kihi)(hi+t(kihi))=12(i=1d(ki2hi2)+(r12r22)).\sum_{i=1}^{d}(k_{i}-h_{i})(h_{i}+t_{*}(k_{i}-h_{i}))=\frac{1}{2}\left(\sum_{i=1}^{d}(k_{i}^{2}-h_{i}^{2})+(r_{1}^{2}-r_{2}^{2})\right).

And solving for tt_{*}, we obtain that t=12+r12r222c2c12t_{*}=\frac{1}{2}+\frac{r_{1}^{2}-r_{2}^{2}}{2\|c_{2}-c_{1}\|^{2}}. Hence, the center of SS is given by:

c=c1+tN=12(1+r22r12c2c12)c1+12(1+r12r22c2c12)c2\displaystyle c=c_{1}+t_{*}N=\frac{1}{2}\left(1+\frac{r_{2}^{2}-r_{1}^{2}}{\|c_{2}-c_{1}\|^{2}}\right)c_{1}+\frac{1}{2}\left(1+\frac{r_{1}^{2}-r_{2}^{2}}{\|c_{2}-c_{1}\|^{2}}\right)c_{2}

Next, we will compute the radius rr of SS. This radius can be determined as the height rr of the triangle with base c2c1\|c_{2}-c_{1}\| formed by the points c1c_{1}, c2c_{2}, and a point on SS. Thus, by the Heron’s formula we have that

r=2s(sc2c1)(sr1)(sr2)c2c1,r=\frac{2\sqrt{s(s-\|c_{2}-c_{1}\|)(s-r_{1})(s-r_{2})}}{\|c_{2}-c_{1}\|},

where s=12(c2c1+r1+r2)s=\frac{1}{2}(\|c_{2}-c_{1}\|+r_{1}+r_{2}) correspond to the semi-perimeter. ∎

We can proceed now to compute the poles of the (d1)(d-1)-sphere D1D2\partial D_{1}\cap\partial D_{2}.

Lemma 6.

Let D1(c1;r1)D_{1}(c_{1};r_{1}) and D2(c2;r2)D_{2}(c_{2};r_{2}) be two dd-disks such that D1D2\partial D_{1}\cap\partial D_{2} is a (d1)(d-1)-sphere S=S(c;r)S=S(c;r) with center cc and radius rr. Then, the eqe_{q}-poles of SS are sq±=c±idxieis_{q}^{\pm}=c\pm\sum_{i}^{d}x_{i}e_{i}, where

xi={r|πi(c2c1)πq(c2c1)|c2c1c2c12πq(c2c1)2,iqrc2c12πq(c2c1)2c2c1,i=q.x_{i}=\begin{cases}\dfrac{r|\pi_{i}(c_{2}-c_{1})\pi_{q}(c_{2}-c_{1})|}{\|c_{2}-c_{1}\|\sqrt{\|c_{2}-c_{1}\|^{2}-\pi_{q}(c_{2}-c_{1})^{2}}},&i\neq q\\ \\ \dfrac{r{\sqrt{\|c_{2}-c_{1}\|^{2}-\pi_{q}(c_{2}-c_{1})^{2}}}}{\|c_{2}-c_{1}\|},&i=q.\end{cases}
Proof.

For simplicity, we translate the hyperplane Π\Pi, which contains the (d1)(d-1)-sphere SS, as well as the sphere itself, to the origin; in such case, the corresponding equations are given by,

i=1d(kihi)xi\displaystyle\sum_{i=1}^{d}(k_{i}-h_{i})x_{i} =0,\displaystyle=0,
i=1dxi2\displaystyle\sum_{i=1}^{d}x_{i}^{2} =r2,\displaystyle=r^{2},

where hi:=πi(c1)h_{i}:=\pi_{i}(c_{1}) and ki:=πi(c2)k_{i}:=\pi_{i}(c_{2}) for i=1,,di=1,\ldots,d. In the case that kqhq=πq(c2c1)=0k_{q}-h_{q}=\pi_{q}(c_{2}-c_{1})=0, the normal vector N=c2c1N=c_{2}-c_{1} of the hyperplane Π\Pi is orthogonal to the basis vector eqe_{q}. Therefore, the eqe_{q}-poles of SS are c±reqc\pm re_{q}, which agree with the formulae of the lemma.

On the other hand, suppose that kqhq0k_{q}-h_{q}\neq 0. To find the eqe_{q}-poles of SS, we will use the Lagrange multiplier method. Consider the following function:

(2) xq=f(x1,x2,,xq^,,xd)=jqd(kjhj)xjkqhq,x_{q}=f(x_{1},x_{2},...,\widehat{x_{q}},...,x_{d})=\frac{-\sum_{j\neq q}^{d}(k_{j}-h_{j})x_{j}}{k_{q}-h_{q}},

subject to the restriction:

g(x1,x2,,xq^,,xd)=jqdxj2+(jqd(kjhj)xjkqhq)2r2=0g(x_{1},x_{2},...,\widehat{x_{q}},...,x_{d})=\sum_{j\neq q}^{d}x_{j}^{2}+\left(\frac{-\sum_{j\neq q}^{d}(k_{j}-h_{j})x_{j}}{k_{q}-h_{q}}\right)^{2}-r^{2}=0

Let λ\lambda be the Lagrange multiplier, we define

h(x1,,xq^,,xd,λ)=f(x1,,xq^,,xd)+λg(x1,,xq^,,xd)h(x_{1},...,\widehat{x_{q}},...,x_{d},\lambda)=f(x_{1},...,\widehat{x_{q}},...,x_{d})+\lambda g(x_{1},...,\widehat{x_{q}},...,x_{d})

For any iqi\neq q, consider the following system of equations:

hxi=kihikqhq+2λxi+2λ(jqd(kjhj)xjkqhq)(kihikqhq)=0.\frac{\partial h}{\partial x_{i}}=-\frac{k_{i}-h_{i}}{k_{q}-h_{q}}+2\lambda x_{i}+2\lambda\left(\frac{-\sum_{j\neq q}^{d}(k_{j}-h_{j})x_{j}}{k_{q}-h_{q}}\right)\left(-\frac{k_{i}-h_{i}}{k_{q}-h_{q}}\right)=0.

Then

kihikqhq+2λ(xi+kihi(kqhq)2jqd(kjhj)xj)\displaystyle-\frac{k_{i}-h_{i}}{k_{q}-h_{q}}+2\lambda\left(x_{i}+\frac{k_{i}-h_{i}}{(k_{q}-h_{q})^{2}}{\sum_{j\neq q}^{d}(k_{j}-h_{j})x_{j}}\right) =0\displaystyle=0
(kihi)(kqhq)+2λ((kqhq)2xi+(kihi)jqd(kjhj)xj)\displaystyle-(k_{i}-h_{i})(k_{q}-h_{q})+2\lambda\left((k_{q}-h_{q})^{2}x_{i}+(k_{i}-h_{i}){\sum_{j\neq q}^{d}(k_{j}-h_{j})x_{j}}\right) =0\displaystyle=0

Solving this system of equations for λ\lambda, we obtain that,

λ=(kihi)(kqhq)2((kqhq)2xi+(kihi)jqd(kjhj)xj)\lambda=\frac{(k_{i}-h_{i})(k_{q}-h_{q})}{2\left((k_{q}-h_{q})^{2}x_{i}+(k_{i}-h_{i}){\sum_{j\neq q}^{d}(k_{j}-h_{j})x_{j}}\right)}

Comparing the last expression for two indices ii~i\neq\tilde{i}, we have that,

xi2\displaystyle x_{i}^{2} =r2(kihi)2(ki~hi~)2+ji~,q(kjhj)2+1(kqhq)2(jq(kjhj)2)2\displaystyle=\frac{r^{2}(k_{i}-h_{i})^{2}}{(k_{\tilde{i}}-h_{\tilde{i}})^{2}+\sum_{j\neq\tilde{i},q}{(k_{j}-h_{j})}^{2}+\frac{1}{(k_{q}-h_{q})^{2}}\left(\sum_{j\neq q}(k_{j}-h_{j})^{2}\right)^{2}}
=r2(kihi)2jq(kjhj)2+1(kqhq)2(jq(kjhj)2)2\displaystyle=\frac{r^{2}(k_{i}-h_{i})^{2}}{\sum_{j\neq q}{(k_{j}-h_{j})}^{2}+\frac{1}{(k_{q}-h_{q})^{2}}\left(\sum_{j\neq q}(k_{j}-h_{j})^{2}\right)^{2}}
=r2(kihi)2(kqhq)2((kqhq)2+jq(kjhj)2)(jq(kjhj)2)\displaystyle=\frac{r^{2}(k_{i}-h_{i})^{2}(k_{q}-h_{q})^{2}}{\left((k_{q}-h_{q})^{2}+\sum_{j\neq q}(k_{j}-h_{j})^{2}\right)\left(\sum_{j\neq q}{(k_{j}-h_{j})}^{2}\right)}
=r2πi(c2c1)2πq(c2c1)2c2c12(c2c12πq(c2c1)2)\displaystyle=\frac{r^{2}\pi_{i}(c_{2}-c_{1})^{2}\pi_{q}(c_{2}-c_{1})^{2}}{\|c_{2}-c_{1}\|^{2}\left(\|c_{2}-c_{1}\|^{2}-\pi_{q}(c_{2}-c_{1})^{2}\right)}

Finally, for i=qi=q, we can use the last expression to substitute it in (2) and obtain the desired result. ∎

2.2. Sphere systems with more than two spheres

Now, let us proceed with the explicit calculation of the coefficients for the center cc of the (dm+1)(d-m+1)-sphere S=j=1mDjS=\cap_{j=1}^{m}\partial D_{j}. We can achieve this by considering the disk system translated to cmc_{m}, denoted as {Dj(cjcm;rj)}j=1m\{D_{j}(c_{j}-c_{m};r_{j})\}_{j=1}^{m}, and by defining the (dm+1)(d-m+1)-sphere S{cm}=j=1mDj(cjcm;rj)S-\{c_{m}\}=\cap_{j=1}^{m}\partial D_{j}(c_{j}-c_{m};r_{j}). This sphere is positioned at the intersection of hyperplanes (for more details, refer to [6]).

(3) (ckcm)Tx=12(rm2+ckcm2rk2)(c_{k}-c_{m})^{T}x=\frac{1}{2}(r_{m}^{2}+\|c_{k}-c_{m}\|^{2}-r_{k}^{2})

for all k=1,,m1k=1,...,m-1. Utilizing the information that the center of S{cm}S-\{c_{m}\} can be expressed as a combination of the centers ckcmc_{k}-c_{m} and substituting it into (3), we obtain a linear system of equations with dimensions (m1)×(m1)(m-1)\times(m-1):

j=1m1λj(ckcm)(cjcm)=12(rm2+ckcm2rk2)\sum_{j=1}^{m-1}\lambda_{j}(c_{k}-c_{m})\cdot(c_{j}-c_{m})=\frac{1}{2}(r_{m}^{2}+\|c_{k}-c_{m}\|^{2}-r_{k}^{2})

for k=1,,m1k=1,...,m-1. Solving the system of equations for λ=(λ1,,λm1)\lambda=(\lambda_{1},...,\lambda_{m-1}), we find the center of SS as follows:

c=λ1(c1cm)++λm1(cm1cm)+cmc=\lambda_{1}(c_{1}-c_{m})+...+\lambda_{m-1}(c_{m-1}-c_{m})+c_{m}

The radius of the sphere SS can be computed using the equation:

r2=rk2cck2r^{2}=r_{k}^{2}-\|c-c_{k}\|^{2}

for any k{1,,m1}k\in\{1,...,m-1\}.

Now that we have determined the center and radius of SS, as well as the affine space that contains it, we can proceed to compute its eqe_{q}-poles for each q{1,2,,d}q\in\{1,2,...,d\}. These poles reside in the affine space that contains SS and within a set that we define below.

Let SS be an ii-sphere in d\mathbb{R}^{d}, and let n1,,ndin_{1},...,n_{d-i} be orthogonal vectors to the affine space LL that contains SS. Consider the space MM generated by these vectors together with the vector eqe_{q} from the canonical basis of d\mathbb{R}^{d}. Let us denote nj=(nl(j))l=1dn_{j}=\left(n_{l}^{(j)}\right)_{l=1}^{d} for each j=1,,dij=1,...,d-i. Then, we can define L0L_{0}, the set LL translated to the origin, as follows:

L0\displaystyle L_{0} ={nj}j=1di\displaystyle=\left<\{n_{j}\}_{j=1}^{d-i}\right>^{\perp}
={xdxnj=0,j=1,,di}\displaystyle=\left\{x\in\mathbb{R}^{d}\mid x\cdot n_{j}=0,\hskip 5.69046pt\forall j=1,...,d-i\right\}

The set MM is defined as:

M\displaystyle M ={nj}j=1di{eq}\displaystyle=\left<\{n_{j}\}_{j=1}^{d-i}\cup\{e_{q}\}\right>
={x=j=1diλjnj+λdi+1eqdλj,j=1,,di+1}\displaystyle=\left\{x=\sum_{j=1}^{d-i}\lambda_{j}n_{j}+\lambda_{d-i+1}e_{q}\in\mathbb{R}^{d}\mid\lambda_{j}\in\mathbb{R},\hskip 5.69046pt\forall j=1,...,d-i+1\right\}
={(j=1diλjn1(j),,j=1diλjnq(j)+λdi+1,,j=1diλjnd(j))λj}\displaystyle=\left\{\left(\sum_{j=1}^{d-i}\lambda_{j}n_{1}^{(j)},...,\sum_{j=1}^{d-i}\lambda_{j}n_{q}^{(j)}+\lambda_{d-i+1},...,\sum_{j=1}^{d-i}\lambda_{j}n_{d}^{(j)}\right)\mid\lambda_{j}\in\mathbb{R}\right\}

Refer to Figure 1 for a visual representation of the subspaces LL and M+{c}M+\{c\}.

Refer to caption
Figure 1. Visualization of the subspaces LL and M+{c}M+\{c\}.

As mentioned above, the eqe_{q}-poles of SS lie at the intersection of LL and M+{c}M+\{c\}, where cc is the center of SS. To simplify the calculations, we will utilize L0L_{0} and MM, and then translate them into cc. The intersection of MM and L0L_{0} can be expressed as follows:

ML0\displaystyle M\cap L_{0} ={xMxnk=0,k=1,,di}\displaystyle=\left\{x\in M\mid x\cdot n_{k}=0,\hskip 5.69046pt\forall k=1,...,d-i\right\}
={j=1diλjnj+λdi+1eqd|(j=1diλjnj+λdi+1eq)nk=0,λjk=1,,di}\displaystyle=\left\{\sum_{j=1}^{d-i}\lambda_{j}n_{j}+\lambda_{d-i+1}e_{q}\in\mathbb{R}^{d}\left|\left(\sum_{j=1}^{d-i}\lambda_{j}n_{j}+\lambda_{d-i+1}e_{q}\right)\cdot n_{k}=0,\hskip 5.69046pt\lambda_{j}\in\mathbb{R}\hskip 5.69046pt\forall k=1,...,d-i\right.\right\}
={j=1diλjnj+λdi+1eqd|j=1diλjnjnk+λdi+1nqk=0,λjk=1,,di}\displaystyle=\left\{\sum_{j=1}^{d-i}\lambda_{j}n_{j}+\lambda_{d-i+1}e_{q}\in\mathbb{R}^{d}\left|\sum_{j=1}^{d-i}\lambda_{j}n_{j}\cdot n_{k}+\lambda_{d-i+1}n_{q}^{k}=0,\hskip 5.69046pt\lambda_{j}\in\mathbb{R}\hskip 5.69046pt\forall k=1,...,d-i\right.\right\}

Let us consider a disk system in d\mathbb{R}^{d}, denoted {Dj(cj;rj)}j=1m\{D_{j}(c_{j};r_{j})\}_{j=1}^{m}, where m<dm<d. The intersection of their boundaries forms a (dm+1)(d-m+1)-sphere SS. In this case, the subspace MM has dimension mm, or dim(M)=m1dim(M)=m-1 if eqMe_{q}\in M. We choose the normal vectors for the affine space containing SS as nj=cjcmn_{j}=c_{j}-c_{m}, where j=1,,m1j=1,...,m-1. Then

M={(j=1m1λj(c1(j)c1(m)),,j=1m1λj(cq(j)cq(m))+λm,,j=1m1λj(cd(j)cd(m)))|λj}M=\left\{\left(\sum_{j=1}^{m-1}\lambda_{j}\left(c_{1}^{(j)}-c_{1}^{(m)}\right),...,\sum_{j=1}^{m-1}\lambda_{j}\left(c_{q}^{(j)}-c_{q}^{(m)}\right)+\lambda_{m},...,\sum_{j=1}^{m-1}\lambda_{j}\left(c_{d}^{(j)}-c_{d}^{(m)}\right)\right)\left|\lambda_{j}\in\mathbb{R}\right.\right\}

By rewriting, we have

ML0\displaystyle M\cap L_{0} ={j=1m1λjnj+λmeqd|j=1m1λjnjnk+λmnqk=0,λj,k=1,,m1}\displaystyle=\left\{\sum_{j=1}^{m-1}\lambda_{j}n_{j}+\lambda_{m}e_{q}\in\mathbb{R}^{d}\left|\sum_{j=1}^{m-1}\lambda_{j}n_{j}\cdot n_{k}+\lambda_{m}n_{q}^{k}=0,\hskip 5.69046pt\lambda_{j}\in\mathbb{R},\hskip 5.69046pt\forall k=1,...,m-1\right.\right\}

If S(c;r)=j=1mDjS(c;r)=\cap_{j=1}^{m}\partial D_{j} is the (dm+1)(d-m+1)-sphere with center in cc and radius rr, then the eqe_{q}-poles of SS are the eqe_{q}-poles of S{cm}S-\{c_{m}\} but translated by cc. The poles of S{cm}S-\{c_{m}\} are located in ML0M\cap L_{0}. If pp is an eqe_{q}-pole of S{cm}S-\{c_{m}\}, then it can be expressed as

p=j=1m1λjnj+eqλmp=\sum_{j=1}^{m-1}\lambda_{j}n_{j}+e_{q}\lambda_{m}

for some λj\lambda_{j}\in\mathbb{R}, j=1,,mj=1,...,m and the following conditions holds:

j=1m1λjnjnk+λmnqk=0\sum_{j=1}^{m-1}\lambda_{j}n_{j}\cdot n_{k}+\lambda_{m}n_{q}^{k}=0

for each k=1,,m1k=1,...,m-1 and

p2=r2\|p\|^{2}=r^{2}

Thus, if p=j=1m1λjnj+eqλmp=\sum_{j=1}^{m-1}\lambda_{j}n_{j}+e_{q}\lambda_{m} is an eqe_{q}-pole of S{cm}S-\{c_{m}\}, the following equations are satisfied for λ1,λ2,,λm\lambda_{1},\lambda_{2},...,\lambda_{m}\in\mathbb{R}:

(4) j=1m1λjnjnk+λmnqk=0\sum_{j=1}^{m-1}\lambda_{j}n_{j}\cdot n_{k}+\lambda_{m}n_{q}^{k}=0
(5) i=1d(j=1m1λjnij)2+2λmj=1m1λjnqj+λm2=r2\sum_{i=1}^{d}\left(\sum_{j=1}^{m-1}\lambda_{j}n_{i}^{j}\right)^{2}+2\lambda_{m}\sum_{j=1}^{m-1}\lambda_{j}n_{q}^{j}+\lambda_{m}^{2}=r^{2}

for all k=1,,m1k=1,...,m-1, with rr the radius of the (dm+1)(d-m+1)-sphere SS. From (4) we have the system

(n1n1n1n2.n1nm1n2n1n2n2.n2nm1nm1n1nm1n2.nm1nm1)(λ1λ2λm1)+λm(nq1nq2nqm1)=0.\begin{pmatrix}n_{1}\cdot n_{1}&n_{1}\cdot n_{2}&....&n_{1}\cdot n_{m-1}\\ n_{2}\cdot n_{1}&n_{2}\cdot n_{2}&....&n_{2}\cdot n_{m-1}\\ ...\\ n_{m-1}\cdot n_{1}&n_{m-1}\cdot n_{2}&....&n_{m-1}\cdot n_{m-1}\\ \end{pmatrix}\begin{pmatrix}\lambda_{1}\\ \lambda_{2}\\ ...\\ \lambda_{m-1}\end{pmatrix}+\lambda_{m}\begin{pmatrix}n_{q}^{1}\\ n_{q}^{2}\\ ...\\ n_{q}^{m-1}\end{pmatrix}=0.

Let us denote AA as the matrix (ninj)i,j(n_{i}\cdot n_{j})_{i,j} and BB as the vector (nqj)j=1m1(-n_{q}^{j})_{j=1}^{m-1}. Then, we have Aλ=λmBA\lambda=\lambda_{m}B, where λ=(λ1,,λm1)\lambda=(\lambda_{1},...,\lambda_{m-1}). Solving for λ\lambda, we obtain λj=λm(A1B)[j]\lambda_{j}=\lambda_{m}(A^{-1}B)[j] for each j=1,,m1j=1,...,m-1 (where (A1B)[j](A^{-1}B)[j] denotes the entry jj of the (m1)×1(m-1)\times 1 vector (A1B)(A^{-1}B)). By substituting the value of λj\lambda_{j} into (5), we obtain the quadratic equation:

λm2[i=1d(j=1m1(A1B)[j]nij)2+2j=1m1(A1B)[j]nqj+1]r2=0\lambda_{m}^{2}\left[\sum_{i=1}^{d}\left(\sum_{j=1}^{m-1}(A^{-1}B)[j]n_{i}^{j}\right)^{2}+2\sum_{j=1}^{m-1}(A^{-1}B)[j]n_{q}^{j}+1\right]-r^{2}=0

Let us define

Γi=j=1m1(A1B)[j]nij\Gamma_{i}=\sum_{j=1}^{m-1}(A^{-1}B)[j]n_{i}^{j}

for all i=1,,di=1,...,d. Solving this equation, we find:

λm=±ri=1dΓi2+2Γq+1\lambda_{m}=\frac{\pm r}{\sqrt{\sum_{i=1}^{d}\Gamma_{i}^{2}+2\Gamma_{q}+1}}
λj=±r(A1B)[j]i=1dΓi2+2Γq+1\lambda_{j}=\frac{\pm r(A^{-1}B)[j]}{\sqrt{\sum_{i=1}^{d}\Gamma_{i}^{2}+2\Gamma_{q}+1}}

for j=1,,m1j=1,...,m-1. Therefore, the eqe_{q}-poles of SS, for q=1,,dq=1,...,d, are:

p=j=1m1λjnj+eqλm+c=±ri=1dΓi2+2Γq+1(j=1m1(A1B)[j]nj+eq)+cp=\sum_{j=1}^{m-1}\lambda_{j}n_{j}+e_{q}\lambda_{m}+c=\frac{\pm r}{\sqrt{\sum_{i=1}^{d}\Gamma_{i}^{2}+2\Gamma_{q}+1}}\left(\sum_{j=1}^{m-1}(A^{-1}B)[j]n_{j}+e_{q}\right)+c

3. Vietoris-Rips and Čech systems

Our goal in this section is to provide a comprehensive understanding of the disk system, the Vietoris-Rips system, and the Čech system. Additionally, we introduce some results that establish a certain connection between both disk systems. Investigating the features and qualities of data and spaces can provide us with useful knowledge about their geometric and topological characteristics.

Before we look into the definitions of Vietoris-Rips and Čech systems, let us give a brief overview. These systems are essential in the field of topological data analysis for recognizing and comprehending the geometric structure of point cloud data. Vietoris-Rips complex and Čech complex share the goal of capturing the topology of the underlying metric space, both provide different ways of recognizing connections and associations among data points. The Vietoris-Rips complex tends to be more efficient and scalable for large datasets, while the Čech complex can be more accurate but computationally more expensive. The choice between the two depends on the nature of the dataset and the specific goals of the topological analysis. Now, let us move on to defining these fundamental concepts.

Definition 7.

Let M={D1,D2,,Dm}M=\{D_{1},D_{2},\ldots,D_{m}\} be a dd-disk system. We say MM is a Vietoris-Rips system if DiDjD_{i}\cap D_{j}\neq\emptyset for each pair i,j{1,2,,m}i,j\in\{1,2,\ldots,m\}. Furthermore, if the dd-disk system MM has the nonempty intersection property DiMDi\bigcap_{D_{i}\in M}D_{i}\neq\emptyset, then MM is called a Čech system.

For each λ0\lambda\geq 0, we define a collection of dd-disks MλM_{\lambda} with the same centers as those in the dd-disk system MM, but with radii rescaled by λ\lambda. When λ>0\lambda>0, MλM_{\lambda} is a dd-disk system again. M1M_{1} is equal to MM, and M0M_{0} is the set of the centers of the dd-disks in MM.

In the field of topological data analysis, the Rips scale and the Čech scale are essential parameters for determining the closeness and connectivity between data points. These two scales offer different perspectives on how we measure and comprehend geometric relationships within point-cloud data. To understand their importance in capturing the underlying topological structure, let us look at their definitions. The Vietoris-Rips scale νM\nu_{M}, of a dd-disk system MM is the smallest λ\lambda\in\mathbb{R} such that MλM_{\lambda} is a Vietoris-Rips system. Similarly, the Čech scale μM\mu_{M}, of MM is the smallest λ\lambda\in\mathbb{R} such that MλM_{\lambda} is a Čech system. This is

νM\displaystyle\nu_{M} =inf{λMλ is a Vietoris-Rips system}\displaystyle=\inf\{\lambda\in\mathbb{R}\mid M_{\lambda}\mbox{ is a Vietoris-Rips system}\}
μM\displaystyle\mu_{M} =inf{λMλ is a Čech system}\displaystyle=\inf\{\lambda\in\mathbb{R}\mid M_{\lambda}\mbox{ is a \v{C}ech system}\}

Next, we present some easily observable properties for both scales. It can be easily seen that MM is a Vietoris-Rips system if and only if νM1\nu_{M}\leq 1 (in particular νMνM=1\nu_{M_{\nu_{M}}}=1); similarly, MM is a Čech system if and only if μM1\mu_{M}\leq 1.

Note that for a given dd-disk system M={D1,D2,,Dm}M=\{D_{1},D_{2},\ldots,D_{m}\} the Vietoris-Rips scale is

νM=maxi<j{cicj/(ri+rj)}\nu_{M}=\max_{i<j}\{\|c_{i}-c_{j}\|/(r_{i}+r_{j})\}

where cic_{i} and rir_{i} are the center and radii of DiD_{i}. An additional observation is that, in cases where the disk system consists of either one or two disks, the Vietoris-Rips scale coincides with the Čech scale. It is evident that every Čech system is also a Vietoris-Rips system; however, the reverse assertion, in general, is not true.

Conversely, if the dd-disk system contains at least three disks, determining the Čech scale becomes more complex. In the context of Čech scale, the following remark is important and play a key role in implementation (see [3] for details).

Remark 8.

If μM\mu_{M} is the Čech scale for MM, then the μM\mu_{M}-rescaled system MμMM_{\mu_{M}}, has only one point in the intersection DiMDi(ci;μMri)\bigcap_{D_{i}\in M}D_{i}(c_{i};\mu_{M}r_{i}).

As we have mentioned, a Čech system is also a Vietoris-Rips system, but the converse is not true. What we can affirm is that if a system is a Vietoris-Rips system, then the system rescaled by the factor 2d/(d+1)\sqrt{2d/(d+1)} is also a Čech system. This is established by the following lemma, the proof of which can be found in [3].

Lemma 9.

Let M={Di(ci;ri)}M=\{D_{i}(c_{i};r_{i})\} be a dd-disk system in euclidean space d\mathbb{R}^{d}. If Di(ci;ri)Dj(cj;rj)D_{i}(c_{i};r_{i})\cap D_{j}(c_{j};r_{j})\neq\emptyset for every pair of disks in MM, then

DiMDi(ci;2d/(d+1)ri).\bigcap_{D_{i}\in M}D_{i}(c_{i};\sqrt{2d/(d+1)}\,r_{i})\neq\emptyset.

One of the implications of the previous result is that, for any given disk system MM, we can bound the Čech scale using the Vietoris-Rips scale νM\nu_{M}. This is stated by the following corollary.

Corollary 10.

If MM is an arbitrary dd-disk system and νM\nu_{M} is its Vietoris-Rips scale, then its Čech scale satisfies νMμM2d/(d+1)νM\nu_{M}\leq\mu_{M}\leq\sqrt{2d/(d+1)}\,\nu_{M}. Therefore, for every dd-disk system MM, the rescaled disk system M2d/(d+1)νMM_{\sqrt{2d/(d+1)}\,\nu_{M}} is always a Čech system. In particular, if 2d/(d+1)νM1\sqrt{2d/(d+1)}\,\nu_{M}\leq 1 then MνMM_{\nu_{M}} is a Čech system.

3.1. Algorithm for determining Čech system.

In the previous section, we have determined the eqe_{q}-poles for the intersection of any number of disks in d\mathbb{R}^{d}. If any of these poles is in all disks of the disk system, it indicates that the system conforms to the criteria of a Čech system. It is important to recognize that this result streamlines our calculation process, focusing on specific points to establish whether the system exhibits a non-empty intersection.

Given a system of mm disks in d\mathbb{R}^{d} where m>dm>d, it is enough to verify if every subsystem of d+1d+1 disks qualifies as a Čech system to conclude that the entire system of disks has a non-empty intersection. This assertion is supported by the Helly’s Theorem.

Now, we introduce an algorithm that determines whether a disk system qualifies as a Čech system. In simpler terms, if the disk system exhibits a non-empty intersection, the algorithm outputs ”TRUE”; otherwise, it outputs ”FALSE”. The algorithm operates by seeking poles within the intersections of the disk boundaries, which, as we have observed, correspond to ii-spheres. It initiates the search for poles within individual disks and then progresses to the pairwise intersections of the disk boundaries ((d2)sphere(d-2)-sphere), continuing the process iteratively. If a pole is found within the remaining disks, the system is classified as a Čech system.

1
Input : A dd-disk system M={Dj}j=1mM=\{D_{j}\}_{j=1}^{m}
Output : A logical TRUE/FALSE to indicate if MM is a Čech system
2 Initialize: Is_Cech_SystemFALSE\texttt{Is\_Cech\_System}\leftarrow\texttt{FALSE}
3 for k1k\leftarrow 1 to mm do
4       Let 𝒮\mathcal{S} be the set of (dk+1)(d-k+1)-spheres of M\partial M
5       for SS in 𝒮\mathcal{S} do
6             for q1q\leftarrow 1 to dd do
7                   Compute the set {sq±}\{s_{q}^{\pm}\} of eqe_{q}-poles of SS
8                   for s{sq±}s\leftarrow\{s_{q}^{\pm}\} do
9                         if sj=1dDjs\in\cap_{j=1}^{d}D_{j} then
10                               Is_Cech_SystemTRUE\texttt{Is\_Cech\_System}\leftarrow\texttt{TRUE}
11                               Go to line 16
12                         end if
13                        
14                   end for
15                  
16             end for
17            
18       end for
19      
20 end for
return (Is_Cech_System)
Algorithm 1 Cech.system
Theorem 11.

Let M={D1,,Dm}M=\{D_{1},\ldots,D_{m}\} be a dd-disk system. Then, MM is a Čech system if and only if Cech.system(M)=(M)= TRUE.

Proof.

If Cech.system(M)=(M)= TRUE, then the Cech.system algorithm (Algorithm 1) found a pole contained in the intersection j=1mDj\cap_{j=1}^{m}D_{j}, therefore j=1mDj\cap_{j=1}^{m}D_{j}\neq\emptyset and it follows that MM is a Čech system.

On the other hand, if j=1mDj\bigcap_{j=1}^{m}D_{j}\neq\emptyset, let pp be a point in j=1mDj\bigcap_{j=1}^{m}D_{j} satisfying π1(p)π1(x)\pi_{1}(p)\leq\pi_{1}(x) for every xx in j=1mDj\bigcap_{j=1}^{m}D_{j}. By Lemma 4, it follows that pp belongs to an ii-sphere and must be an e1e_{1}-south pole. Therefore, by the exhaustive search of Algorithm 1 across all poles, its output is Cech.system(M)=(M)= TRUE. ∎

3.2. Algorithm to compute the Čech scale.

Finding the minimum parameter for which the rescaled system of disks has a non-empty intersection is significant because it helps identify a critical threshold at which the disks come into contact. This parameter, known as the Čech scale, provides valuable information about the proximity or overlap of the disks, which can be crucial in various applications such as collision detection in computer graphics, spatial packing problems, and modeling physical phenomena. In this section, we introduce an algorithm to compute an approximation of the Čech scale for a system of mm disks in d\mathbb{R}^{d}.

1
Input : A dd-disk system MM in d\mathbb{R}^{d} and precision parameter η>0\eta>0
Output : μM\mu_{M}, a Čech scale approximation
2 Compute the Vietoris-Rips scale of MM: νM\nu_{M}
3 if Cech.system(MνM)=(M_{\nu_{M}})= TRUE then
4       μMνM\mu_{M}\leftarrow\nu_{M}
5       Go to line 16
6      
7else
8       Initialize: μMνM\mu_{M}^{*}\leftarrow\nu_{M}, μM2d/(d+1)νM\mu_{M}\leftarrow\sqrt{2d/(d+1)}\nu_{M}
9       while μMμM>η\mu_{M}-\mu_{M}^{*}>\eta do
10             Compute: λμM+μM2\lambda\leftarrow\dfrac{\mu_{M}^{*}+\mu_{M}}{2}
11             if Cech.system(Mλ)=(M_{\lambda})= TRUE then
12                   Update: μMλ\mu_{M}\leftarrow\lambda
13            else
14                   Update: μMλ\mu_{M}^{*}\leftarrow\lambda
15             end if
16            
17       end while
18      
19 end if
return (μM\mu_{M})
Algorithm 2 Cech.scale

The given code presents an algorithm to compute an approximation of the Čech scale of a disk system in Euclidean space using Algorithm 1 and a precision parameter η>0\eta>0. It initializes the scale factor λ\lambda to the Rips scale νM\nu_{M}. If this scale satisfies Cech.system(MνM)=(M_{\nu_{M}})= TRUE, it indicates that the Čech scale has been found. Otherwise, we initiate a cycle in which we compute Cech.system of the system rescaled by a factor λ\lambda. The Čech scale is known to fall between the Rips scale and the value 2d/(d+1)νM\sqrt{{2d}/{(d+1)}}\nu_{M} (Generalized Vietoris-Rips, Corollary 10). To approximate the Čech scale, we employ the bisection method as long as the interval enclosing the Čech scale has a length greater than η\eta. Finally, the algorithm returns an approximation of the Čech scale.

Utilizing the previously described algorithm, we can construct the filtered generalized Čech complex for a disk system MM. Let 𝒞(M)\mathscr{C}(M) denote the set of Čech subsystems, and 𝒞λ(M)\mathscr{C}_{\lambda}(M) the set of Čech subsystems for the rescaled disk system MλM_{\lambda}. The Čech filtration of the MM system forms a maximal chain of Čech complexes

𝒞(M):𝒞0(M)𝒞λ1(M)𝒞λ2(M)𝒞μM(M),\mathscr{C}_{*}(M):\mathscr{C}_{0}(M)\subsetneq\mathscr{C}_{\lambda_{1}}(M)\subsetneq\mathscr{C}_{\lambda_{2}}(M)\subsetneq...\subsetneq\mathscr{C}_{\mu_{M}}(M),

where each λi\lambda_{i} represents the Čech scale of the system MλiM_{\lambda_{i}}. Since the Čech scale of a disk system indicates the factor by which we must rescale the system to make it Čech, defining a level of the filtration, 𝒞λ(M)\mathscr{C}_{\lambda}(M), simply requires determining the Čech scale of the system MλM_{\lambda}.

4. Minimal Axis-Aligned Bounding Box.

In this section, we introduce the concept of the minimal axis-aligned bounding box (AABB) for the intersection of dd-disks and present methods for its computation. The AABB provides a simplified representation of the disk intersection, making it easier to obtain valuable information about the disks. This information could be useful for computing the Čech scale of a disk system.

Definition 12.

Let MM be a disk system in d\mathbb{R}^{d}. The minimal axis-aligned bounding box of MM, denoted as AABB(M)AABB(M), is defined as the smallest axis-aligned bounding box that contains the intersection D=DiMDiD=\cap_{D_{i}\in M}D_{i} , given by

AABB(M):=DB~B~AABB(M):=\bigcap_{D\subset\tilde{B}}\tilde{B}

where B~\tilde{B} ranges over all axis-aligned bounding boxes that contain DD.

Note that the AABB can be expressed as

AABB(M)=k=1d[infπk(D),supπk(D)]AABB(M)=\prod_{k=1}^{d}[\inf\pi_{k}(\partial D),\sup\pi_{k}(\partial D)]

where πk:d\pi_{k}:\mathbb{R}^{d}\longrightarrow\mathbb{R} is the canonical projection onto the kk-th factor, and D\partial D denotes the boundary of DD. In other words, the AABB of a disk system MM is given by the Cartesian product of intervals, where each interval is determined by the minimum and maximum values of the corresponding projection of the disk boundaries.

4.1. Minimal axis-aligned bounding box for two disks

Let’s consider the situation when the disk system MM is composed of two disks D1D_{1} and D2D_{2} in d\mathbb{R}^{d}. If D1D2D_{1}\cap D_{2}\neq\emptyset and D1D2D_{1}\neq D_{2}, the subset D1D2\partial D_{1}\cap\partial D_{2} can take one of three forms: an empty set, a single common point (when the disks are tangent), or a (d2)(d-2)-dimensional sphere. In the last case, we denote the (d2)(d-2)-dimensional sphere or the (d1)(d-1)-sphere D1D2\partial D_{1}\cap\partial D_{2} by S1,2S_{1,2}. To calculate infπi((D1D2))\inf\pi_{i}(\partial(D_{1}\cap D_{2})) and supπi((D1D2))\sup\pi_{i}(\partial(D_{1}\cap D_{2})) for each i{1,2,,d}i\in\{1,2,...,d\}, we can use:

(6) infπi((D1D2))={πi(c1r1ei)if c1r1eiD2,πi(c2r2ei)if c2r2eiD1,inf(πi(D1D2))otherwise,\displaystyle\inf\pi_{i}(\partial(D_{1}\cap D_{2}))=\begin{cases}\pi_{i}(c_{1}-r_{1}e_{i})&\text{if }c_{1}-r_{1}e_{i}\in D_{2},\\ \pi_{i}(c_{2}-r_{2}e_{i})&\text{if }c_{2}-r_{2}e_{i}\in D_{1},\\ \inf(\pi_{i}(\partial D_{1}\cap\partial D_{2}))&\text{otherwise},\end{cases}
supπi((D1D2))={πi(c1+r1ei)if c1+r1eiD2,πi(c2+r2ei)if c2+r2eiD1,sup(πi(D1D2))otherwise.\displaystyle\sup\pi_{i}(\partial(D_{1}\cap D_{2}))=\begin{cases}\pi_{i}(c_{1}+r_{1}e_{i})&\text{if }c_{1}+r_{1}e_{i}\in D_{2},\\ \pi_{i}(c_{2}+r_{2}e_{i})&\text{if }c_{2}+r_{2}e_{i}\in D_{1},\\ \sup(\pi_{i}(\partial D_{1}\cap\partial D_{2}))&\text{otherwise}.\end{cases}

Indeed, by Lemma 4, the extremes of the AABB are the projections of certain poles, either from the (d1)(d-1)-sphere or some dd-sphere. The dd-spheres represent the boundaries of each disk, with poles given by cj±rjeic_{j}\pm r_{j}e_{i}, and the (d1)(d-1)-sphere is S1,2=D1D2S_{1,2}=\partial D_{1}\cap\partial D_{2}, whose poles are computed using Lemma 6. It is worth noting that there are no further options for (dm+1)(d-m+1)-spheres in the case of a two-disk system.

To simplify the notation, we will use Bi,jB_{i,j} to denote the AABB of the intersection of disks DiD_{i} and DjD_{j}. Figure 2 illustrates the AABB of the intersection of two disks in the plane.

Refer to caption
Figure 2. AABB of two disks.

According to (6) and Lemma 5, we have a method to calculate the axis-aligned boundary box (AABB) for systems of two disks.

Knowing how to compute the AABB of two disks is not sufficient to determine the AABB for a disk system with more than two disks in d\mathbb{R}^{d}. In the following examples, we demonstrate that the AABB of a disk system is not simply the intersection of all AABBs of two disks in 3\mathbb{R}^{3}.

Example 13.

Let

M={D1((4,1,0);2),D2((4,1,0);2),D3((0,0,0);3)}3M=\{D_{1}((4,1,0);\sqrt{2}),D_{2}((4,-1,0);\sqrt{2}),D_{3}((0,0,0);3)\}\subset\mathbb{R}^{3}

be a Vietoris-Rips system in 3\mathbb{R}^{3} with the following projection onto the xyxy-plane:

Refer to caption
Figure 3. Disk system MM projected onto the xyxy-plane.

By computing the boxes Bi,jB_{i,j} for all i<ji<j, we obtain:

B1,2\displaystyle B_{1,2} =[3,5]×[2,2]×[1,1]\displaystyle=[3,5]\times[-\sqrt{2},\sqrt{2}]\times[-1,1]
B1,3\displaystyle B_{1,3} =[42,3]×[0,1.41]×[0.72,0.72]\displaystyle=[4-\sqrt{2},3]\times[0,1.41]\times[-0.72,0.72]
B2,3\displaystyle B_{2,3} =[42,3]×[1.4,0]×[0.72,0.72]\displaystyle=[4-\sqrt{2},3]\times[-1.4,0]\times[-0.72,0.72]

Therefore, i<jBi,j=[3,3]×[0,0]×[0.72,0.72]\cap_{i<j}B_{i,j}=[3,3]\times[0,0]\times[-0.72,0.72]. However, the disk system intersects at the point P=(3,0,0)P=(3,0,0), which means AABB(M)={P}AABB(M)=\{P\}. In other words, AABB(M)AABB(M) is not equal to i<jBi,j\cap_{i<j}B_{i,j}.

We know that if DD\neq\emptyset, then i<jBi,j\cap_{i<j}B_{i,j}\neq\emptyset. However, the converse is not always true. The following example illustrates this fact.

Example 14.

Let N={D1,D2,D3((0,1,0);10),D4((3,0,1);0.9)}3N=\{D_{1},D_{2},D_{3}((0,1,0);\sqrt{10}),D_{4}((3,0,1);0.9)\}\subset\mathbb{R}^{3} be a Vietoris-Rips system. The projection of disks D1,D2,D_{1},D_{2}, and D3D_{3} onto the xyxy-plane is illustrated in Figure 4.

Refer to caption
Figure 4. Disk system NN projected onto the xyxy-plane.

We will now compute the intersections Bi,jB_{i,j} for different pairs of disks:

B1,2\displaystyle B_{1,2} =[3,5]×[2,2]×[1,1]\displaystyle=[3,5]\times[-\sqrt{2},\sqrt{2}]\times[-1,1]
B1,3\displaystyle B_{1,3} =[42,10]×[0,2]×[1,1]\displaystyle=[4-\sqrt{2},\sqrt{10}]\times[0,2]\times[-1,1]
B1,4\displaystyle B_{1,4} =[2.6,3.9]×[0.4,0.9]×[0.100007,1.3]\displaystyle=[2.6,3.9]\times[-0.4,0.9]\times[0.100007,1.3]
B2,3\displaystyle B_{2,3} =[2.6,3]×[0.8,0]×[0.44,4.47]\displaystyle=[2.6,3]\times[-0.8,0]\times[-0.44,4.47]
B2,4\displaystyle B_{2,4} =[2.6,3.9]×[0.9,0.4]×[0.100007,1.3]\displaystyle=[2.6,3.9]\times[-0.9,0.4]\times[0.100007,1.3]
B3,4\displaystyle B_{3,4} =[2.1,3.11]×[0.73,0.9]×[0.1,1.7]\displaystyle=[2.1,3.11]\times[-0.73,0.9]\times[0.1,1.7]

The intersection of all pairwise intersections, i<jBi,j\cap_{i<j}B_{i,j}, is given by [3,3]×[0,0]×[0.1,1.7][3,3]\times[0,0]\times[0.1,1.7]. However, the intersection of disks D1,D2,D_{1},D_{2}, and D3D_{3} is a single point PP, which is not contained in D4D_{4} (by construction). Therefore, DD is an empty set, but i<jBi,j\cap_{i<j}B_{i,j} is not.

These examples clearly illustrate that when dealing with the AABB of three disks or more, knowing the AABB for pairs of disks is insufficient. In Example 13, we observe that the intersection of AABB({Di,Dj})AABB(\{D_{i},D_{j}\}) is not equal to the AABB of the disk system MM. Similarly, in Example 14, we find that the intersection of three disks is empty, yet the intersection of AABB({Di,Dj})AABB(\{D_{i},D_{j}\}) contains points.

Therefore, the next crucial step is to determine how to calculate the AABB of a disk system consisting of more than two disks in d\mathbb{R}^{d}.

4.2. Minimal axis-aligned bounding box for more than two disks

Given a system MM consisting of mm disks in d\mathbb{R}^{d}, we can compute the eqe_{q}-poles for any subcollection of disks. Using these eqe_{q}-poles, we can determine the axis-aligned bounding box (AABB) of MM.

If mdm\leq d, we calculate the eqe_{q}-poles of the (dm+1)(d-m+1)-sphere Di\cap\partial D_{i}, and with these poles, we define the AABB of MM by taking infπqD=πq(p)\inf\pi_{q}\partial D=\pi_{q}(p), where pp is the eqe_{q}-south pole of the (dm+1)(d-m+1)-sphere (similarly for supπqD\sup\pi_{q}\partial D).

In the case where m=d+1m=d+1, we consider AABB(M(1)),,AABB(M(m))AABB(M(1)),...,AABB(M(m)) as a collection of minimal axis-aligned boxes for the disk system M(i)=M{Di}M(i)=M-\{D_{i}\} in d\mathbb{R}^{d}. If the intersection of any d+1d+1 of these sets is nonempty, then the intersection of the entire collection gives us the minimal axis-aligned box of the disk system MM. This can be expressed as AABB(M)=AABB(M(i))AABB(M)=\cap AABB(M(i)). The next theorem confirms this finding.

Theorem 15 (Helly’s theorem for minimal axis-aligned boxes).

Let MM be a Rips system with d+1d+1 disks in d\mathbb{R}^{d}. Then, the minimal axis-aligned bounding box for the intersection set D=j=1d+1DjD=\cap_{j=1}^{d+1}D_{j} satisfies:

AABB(M)=j=1d+1AABB(M(j))AABB(M)=\bigcap_{j=1}^{d+1}AABB(M(j))

where M(j)=M{Dj}M(j)=M-\{D_{j}\}.

Proof.

We know that AABB(M)j=1d+1AABB(M(j))AABB(M)\subseteq\bigcap_{j=1}^{d+1}AABB(M(j)). Now, our objective is to establish the reverse inclusion, that is, j=1d+1AABB(M(j))AABB(M)\bigcap_{j=1}^{d+1}AABB(M(j))\subseteq AABB(M). In order to derive a contradiction, suppose that the reverse inclusion is not true. By definition, we have:

AABB(M)=i=1d[infπi(D),supπi(D)]AABB(M)=\prod_{i=1}^{d}\left[\inf\pi_{i}(\partial D),\sup\pi_{i}(\partial D)\right]

and

j=1d+1AABB(M(j))=i=1d[maxk{infπi(jkDj)},mink{supπi(jkDj)}]\bigcap_{j=1}^{d+1}AABB(M(j))=\prod_{i=1}^{d}\left[\max_{k}\{\inf\pi_{i}(\partial\cap_{j\neq k}D_{j})\},\min_{k}\{\sup\pi_{i}(\partial\cap_{j\neq k}D_{j})\}\right]

Without loss of generality, let’s assume that:

π1(p)>maxk{infπ1(jkDj)}k=1d+1\pi_{1}(p)>\max_{k}\left\{\inf\pi_{1}(\cap_{j\neq k}\partial D_{j})\right\}_{k=1}^{d+1}

where pDp\in D satisfies π1(p)=infπ1(D)\pi_{1}(p)=\inf\pi_{1}(\partial D).

Let qjq_{j} be the point in kjDk\cap_{k\neq j}D_{k} such that π1(qj)=infπ1(kjDk)\pi_{1}(q_{j})=\inf\pi_{1}(\cap_{k\neq j}D_{k}) for each j=1,2,,d+1j=1,2,...,d+1. Note that qjDjq_{j}\notin D_{j} because qjq_{j} is not in DD. Now, let γj\gamma_{j} be the line segment that connects qjq_{j} and pp.

Choose ϵ>0\epsilon>0 small enough such that the hyperplane P:x1=π1(p)ϵP:x_{1}=\pi_{1}(p)-\epsilon does not contain any qjq_{j}, and π1(qj)<π1(p)ϵ\pi_{1}(q_{j})<\pi_{1}(p)-\epsilon for all j=1,2,,d+1j=1,2,...,d+1 (such hyperplane PP exists because π1(p)>π1(qj)\pi_{1}(p)>\pi_{1}(q_{j}) for all j=1,,d+1j=1,...,d+1). Since qjq_{j} is in kjDk\cap_{k\neq j}D_{k}, the hyperplane PP intersects every disk in M(j)M(j) for each jj, and γjkjDk\gamma_{j}\subset\cap_{k\neq j}D_{k} intersects PP at a point that is in kjDk\cap_{k\neq j}D_{k} (see Figure 5). Furthermore, DkPD_{k}\cap P is a (d1)(d-1)-dimensional disk. Therefore, we have a collection 𝒟={DjP|j=1,2,,d+1}\mathcal{D}=\{D_{j}\cap P|j=1,2,...,d+1\} of d+1d+1 disks, each of dimension d1d-1, such that every subset AA of 𝒟\mathcal{D} consisting of dd disks has the non-empty intersection property. By Helly’s Theorem, the intersection of all (d1)(d-1)-disks in 𝒟\mathcal{D} is not empty. Therefore, there exists a point qDq\in D with π1(q)<π1(p)\pi_{1}(q)<\pi_{1}(p). However, this contradicts the fact that pDp\in D is such that π1(p)=infπ1(D)\pi_{1}(p)=\inf\pi_{1}(\partial D).

Therefore, we conclude that j=1d+1AABB(M(j))=AABB(M)\bigcap_{j=1}^{d+1}AABB(M(j))=AABB(M).

Refer to caption
Figure 5.

Lemma 16.

Let MM be a Vietoris-Rips system of d+1d+1 disks in d\mathbb{R}^{d}. If D=i=1d+1Di=D=\cap_{i=1}^{d+1}D_{i}=\emptyset and AABB(M(j))AABB(M(j))\neq\emptyset for each j=1,,d+1j=1,...,d+1, then i=1d+1AABB(M(i))\cap_{i=1}^{d+1}AABB(M(i)) consists only of intervals of the form [a,b][a,b] with a>ba>b (inverted intervals).

Proof.

Suppose that i=1d+1AABB(M(i))\cap_{i=1}^{d+1}AABB(M(i)) contains an interval. Without loss of generality, let’s assume it is the interval

[infπ1(D(k)),supπ1(D(l))][\inf\pi_{1}(\partial D(k)),\sup\pi_{1}(\partial D(l))]

for some k,l{1,2,,d+1}k,l\in\{1,2,...,d+1\} where D(k)=ikDiD(k)=\cap_{i\neq k}D_{i}. By definition of i=1d+1AABB(M(i))\cap_{i=1}^{d+1}AABB(M(i)), we have infπ1(D(k))infπ1(D(i))\inf\pi_{1}(\partial D(k))\geq\inf\pi_{1}(\partial D(i)) for all iki\neq k and supπ1(D(l))supπ1(D(j))\sup\pi_{1}(\partial D(l))\leq\sup\pi_{1}(\partial D(j)) for all jlj\neq l. Now, consider a hyperplane P:x1=pP:x_{1}=p with infπ1(D(k))psupπ1(D(k))\inf\pi_{1}(\partial D(k))\leq p\leq\sup\pi_{1}(\partial D(k)). The hyperplane PP intersects D(j)D(j) for every j=1,,d+1j=1,...,d+1 (since PP cuts through all the boxes AABB(M(j))AABB(M(j))). We also know that PDiP\cap D_{i} is a d1d-1-dimensional disk for each ii.

Therefore, we have a collection of dd disks in a d1d-1 dimensional space, and by Helly’s Theorem, this collection must have a non-empty intersection. This implies that DD\neq\emptyset, which contradicts our assumption that DD is empty. Hence, all intervals in i=1d+1AABB(M(i))\cap_{i=1}^{d+1}AABB(M(i)) must be inverted intervals of the form [a,b][a,b] with a>ba>b. ∎

As an example, we provide the lower bounds of the AABB for a system of three disks in d\mathbb{R}^{d}, this is, we compute infπi(D)\inf\pi_{i}(\partial D) for each i=1,,di=1,...,d (analogous supπiD\sup\pi_{i}\partial D) with D=j=13DjD=\cap_{j=1}^{3}D_{j}. Let M={D1,D2,D3}M=\{D_{1},D_{2},D_{3}\} be a Rips system in d\mathbb{R}^{d}, then, for each i=1,,di=1,...,d, the computation of the AABB for the disk system MM is given by:

infπi(D)={πi(ckrkei)if ckrkeiDmaxj<k{infπi(DjDk)}if (*)infπi(j=13Dj)otherwise\displaystyle\inf\pi_{i}(\partial D)=\begin{cases}\pi_{i}(c_{k}-r_{k}e_{i})&\text{if $c_{k}-r_{k}e_{i}\in D$}\\ \max_{j<k}\{\inf\pi_{i}(\partial D_{j}\cap\partial D_{k})\}&\text{if (*)}\\ \inf\pi_{i}(\cap_{j=1}^{3}\partial D_{j})&\text{otherwise}\end{cases}

where (*) denotes the case where qDq\in D with qDjDkq\in\partial D_{j}\cap\partial D_{k} such that

πi(q)=maxj<k{infπi(DjDk)}.\pi_{i}(q)=\max_{j<k}\{\inf\pi_{i}(\partial D_{j}\cap\partial D_{k})\}.

The preceding calculations are a result of Lemma 4. It is known that the extremes of the AABB lie in the projections of specific poles, either from a dd-sphere, (d1)(d-1)-sphere, or the (d2)(d-2)-sphere.

Given a disk system MM in d\mathbb{R}^{d}, we can compute the minimal axis-aligned bounding box (AABB) for the intersection of all disks in MM. If the AABB is a point, then it represents the intersection of the disks. This property allows us to identify when the AABB is a point. If the AABB of MM is not a point, we can rescale MM by a scale factor λ\lambda such that the AABB of MλM_{\lambda} becomes a point. The value of λ\lambda is referred to as the Čech scale of the system MM.

4.3. Minimal axis-aligned bounding box Algorithm.

Now, we present an algorithm to calculate the minimal axis-aligned bounding box (AABB) of a system of mm disks in d\mathbb{R}^{d}. As previously explained, the AABB’s extremes are defined by projecting the poles of specific ii-spheres. Thus, in computing the AABB, we will determine the poles for each ii-sphere within the disk system.

1
Input : A dd-disk system M={Dj}j=1mM=\{D_{j}\}_{j=1}^{m}
Output : The minimal AABB for the system MM
2 Initialize: P=P=\emptyset
3 for k1k\leftarrow 1 to mm do
4       Let 𝒮\mathcal{S} be the set of (dk+1)(d-k+1)-spheres of M\partial M
5       for SS in 𝒮\mathcal{S} do
6             for q1q\leftarrow 1 to dd do
7                   Compute the set {sq±}\{s_{q}^{\pm}\} of eqe_{q}-poles of SS
8                   for s{sq±}s\leftarrow\{s_{q}^{\pm}\} do
9                         if sj=1dDjs\in\cap_{j=1}^{d}D_{j} then
10                               Add: PsP\leftarrow s
11                              
12                         end if
13                        
14                   end for
15                  
16             end for
17            
18       end for
19      
20 end for
21if PP\neq\emptyset then
22       for q1q\leftarrow 1 to dd do
23             aq=minP{πq(sq)}a_{q}=min_{P}\{\pi_{q}(s_{q}^{-})\}
24             bq=maxP{πq(sq+)}b_{q}=max_{P}\{\pi_{q}(s_{q}^{+})\}
25            
26       end for
27      return (Πq=1d[aq,bq]\Pi_{q=1}^{d}[a_{q},b_{q}])
28else
29       return (The disk system does not intersect)
30 end if
Algorithm 3 AABB.minimal

The algorithm starts by initializing the set of poles, PP, as an empty set. In each iteration of the first loop, we determine the spheres formed by the intersection of the boundaries of the disk subcollections in the system MM (for this, we require the center and radius, which are computed in Subsection 2.2). Next, we identify the poles of each sphere and if any of them are present in all the disks of MM, we add them to the set PP. If PP is not empty, we proceed to calculate the extremes for each dimension of the AABB using the set of north poles for the upper bounds and the set of south poles for the lower bounds. In this case, the output is the product of the intervals defined by the computed extremes. If PP is empty, it indicates that there is no intersection in the disk system.

5. Acknowledgements

C.G.E.P acknowledges CONACYT for the financial support provided through a National Fellowship (CVU-638165).

References

  • [1] G. Bell, A. Lawson, J. Martin, J. Rudzinski, and C. Smyth, Weighted persistent homology, Involve, a Journal of Mathematics, 12 (2019), pp. 823–837.
  • [2] P. Cai, Y. Cai, I. Chandrasekaran, and J. Zheng, Collision detection using axis aligned bounding boxes, Simulation, Serious Games and Their Applications, (2013), pp. 1–14.
  • [3] J. F. Espinoza, R. Hernández-Amador, H. A. Hernández-Hernández, and B. Ramonetti-Valencia, A numerical approach for the filtered generalized čech complex, Algorithms, 13 (2019), p. 11.
  • [4] N. K. Le, P. Martins, L. Decreusefond, and A. Vergne, Construction of the generalized cech complex, 2015.
  • [5] J. Mahovsky and B. Wyvill, Fast ray-axis aligned bounding box overlap tests with plucker coordinates, Journal of Graphics Tools, 9 (2004), pp. 35–46.
  • [6] D. S. Maioli, C. Lavor, and D. S. Gonçalves, A note on computing the intersection of spheres in n\mathbb{R}^{n}, The ANZIAM Journal, 59 (2017), p. 271–279.