0% found this document useful (0 votes)
20 views20 pages

Intersection Properties of Finite Disk Collections

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
20 views20 pages

Intersection Properties of Finite Disk Collections

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 20

mathematics

Article
Intersection Properties of Finite Disk Collections
Jesús F. Espinoza *,† and Cynthia G. Esquer-Pérez †

Departamento de Matemáticas, Universidad de Sonora, Hermosillo C.P. 83000, Sonora, Mexico;


[email protected]
* Correspondence: [email protected]
† These authors contributed equally to this work.

Abstract: In this article, we study the intersection of a finite collection of disks in Euclidean space by
examining spheres of various dimensions and their poles (extreme values with respect to canonical
projections) contained within the intersection’s boundary. We derive explicit formulae for computing
these extreme values and present two applications. The first application involves computing the
smallest common rescaling factor for the radii of the disk system, which brings the system to a
single point of intersection. This calculation allows us to compute the generalized Čech filtration,
a crucial tool for the topological data analysis of weighted point clouds. The second application
focuses on determining the minimal Axis-Aligned Bounding Box (AABB) for the intersection of
a finite collection of disks in Euclidean space, addressing a significant problem in computational
geometry. We consider that this work aims to contribute to the fields of topological data analysis and
computational geometry by providing new tools for analyzing complex geometric structures.

Keywords: disk system; intersection of systems of disks; generalized Čech complex; minimal
axis-aligned bounding box

MSC: 68U05; 65D17; 05E45

1. Introduction
One of the new techniques developed for the analysis of large clusters of information,
Citation: Espinoza, J.F.; Esquer-Pérez, known as Big Data, is Topological Data Analysis (TDA). In TDA, simplicial complexes as-
C.G. Intersection Properties of Finite sociated with the data are constructed. These structures include the Vietoris–Rips complex,
Disk Collections. Mathematics 2024, 12, the Čech complex, and the piecewise linear lower star complex, among others. Of special
547. https://doi.org/10.3390/ interest to us is the generalized Čech complex structure ([1–4]). Although the standard Čech
math12040547 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
Academic Editor: Xiangmin Jiao
by assigning or weighting them with larger and/or more rapidly expanding balls while
Received: 11 January 2024 de-emphasizing others using smaller and/or slower-growing balls. This approach proves
Revised: 8 February 2024 valuable for handling noisy data sets, offering an alternative to discarding data that may
Accepted: 9 February 2024 not meet a specific significance threshold [5].
Published: 10 February 2024 Understanding the patterns of intersections and the timing of intersections among a
set of disks in Rd , each with potentially different radii, is a fundamental problem ([6–9]).
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
Copyright: © 2024 by the authors.
same factor, we obtain a filtered generalized Čech complex, where the associated simplicial
Licensee MDPI, Basel, Switzerland.
complexes evolve as the scale parameter varies. In particular, in [10], algorithms are
This article is an open access article
distributed under the terms and
provided to calculate the generalized Čech complex in R2 , and [11] presents an algorithm
conditions of the Creative Commons
to determine the Čech scale for a collection of disks in the plane.
Attribution (CC BY) license (https:// It is worth mentioning that our initial goal in writing this article was to compute the
creativecommons.org/licenses/by/ Čech scale of a disk system, generalizing [11], to provide a powerful tool for understanding
4.0/). the shape (i.e., the geometry) of complex data in higher dimensions. Facing the challenge

Mathematics 2024, 12, 547. https://doi.org/10.3390/math12040547 https://www.mdpi.com/journal/mathematics


Mathematics 2024, 12, 547 2 of 20

of generalizing the results in [11] beyond Euclidean spaces of dimension 2, we began with
computing the Axis-Aligned Bounding Box (AABB) for the intersection of a finite set of
disks, seeking a measure for such an intersection to incorporate it into a numerical method
with the objective of approximating the Čech scale. However, we soon realized that the
problem was more complex than anticipated. To compute the (minimal) AABB, it was
necessary to include the key points we refer later to as eq -poles. In retrospect, the key points
for solving the problem in dimension 2, as discussed in [11], can be considered to be the
poles of 0-dimensional spheres, a specific case of the i-spheres we study in this paper. Once
we characterized the intersection of a set of disks (with generally different radii) through
such eq -poles, we were able to solve both problems: computing (approximate) the Čech
scale and computing the minimal AABB for the intersection of a disk system.
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 Rd . We start by investigating the intersection of two
disks in Section 2.1 and then expand our analysis to a system of m disks in Section 2.2. By
applying Helly’s Theorem, we prove that it is sufficient to examine the intersection of all
subsystems consisting of d + 1 disks to determine if the system has a nonempty intersection.
In Section 3, we define Vietoris–Rips systems and Čech systems, together with pre-
senting results regarding the Rips scale and the Čech scale, as well as their connections. In
Section 3.1, we present an algorithm that can determine whether the intersection of the
system is empty or nonempty. This is achieved by exclusively computing the poles of
subsystems of disks (or spheres). Finally, in Section 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 a minimal Axis-Aligned
Bounding Box (AABB) into our methodology ([12–14]). An AABB is a rectangular par-
allelepiped whose faces are perpendicular to the basis vectors. These bounding boxes
frequently arise in spatial subdivision problems, such as ray tracing [15] and collision de-
tection [16]. 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 d-disk system M, or simply a disk system, as
a finite collection of closed disks in Rd with positive and not necessarily equal radii, i.e.,

M = { Di (ci ; ri ) ⊂ Rd | ci ∈ Rd , ri > 0, 1 ≤ i ≤ m < ∞}.

Moreover, in order to study the intersection properties of a disk system M with the
approach addressed in Sections 3 and 4 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 M, which we call a sphere system and denote by ∂M,

∂M = {∂Di ⊂ Rd | Di ∈ M },

where ∂ denotes the topological boundary operator.


Following the notation in [17], we introduce the following generalization of the sphere.

Definition 1. An i-sphere in Rd is the intersection of a sphere with an affine subspace of dimension i.

Of course, the notions of a sphere (as a (d − 1)-dimensional surface) and a d-sphere in


Rd agree. However, an i-sphere in Rd can also be viewed as the intersection of d-spheres.
For instance, the intersection of two spheres typically occurs in a hyperplane, forming a
(d − 1)-sphere in Rd . When another d-sphere intersects this configuration, the result may
be a (d − 1)-sphere, a (d − 2)-sphere, a 0-sphere (a single point), or it might even be empty,
Mathematics 2024, 12, 547 3 of 20

all within the same hyperplane. For a disk system M = { Di (ci ; ri )} composed of m disks,
where {c1 , . . . , cm } is a set in general position in Rd , the maximum dimension of the affine
subspace associated with the i-sphere, obtained from the intersection of all the spheres
in ∂M, is at most d − m + 1, or equivalently, i = d − m + 1. This conclusion is drawn
from ([17] (Theorem 2.1) and the fact that the affine hull of {c1 , . . . , cm } is of dimension
m − 1. Consequently, the following result is proven.

Lemma 1. Let M = { D1 (c1 ; r1 ), . . . , Dm (cm ; rm )} be disk system such that {c1 , . . . , cm } is a set
in general position in Rd . Then, the possibilities for the set ∩ Di ∈ M ∂Di are:
1. the empty set;
2. a single point;
3. a (d − m + 1)-sphere.

Remarkable points in i-spheres that will play a key role in the rest of the article are the
poles. Let πi : Rd −→ R be the canonical projection on the i-th factor for i = 1, ..., d, and let
{e1 , e2 , ..., ed } be the standard basis of Rd .

Definition 2. Let eq be the q-th vector of the canonical base of Rd . An eq -north (south) pole of an
i-sphere S in Rd is a point on S whose projection on the q-th coordinate is maximum (minimum). In
other words, x ∈ S is the eq -north pole if πq (y) ≤ πq ( x ) for all y ∈ S − { x }, where πq represents
the projection onto the q-th coordinate.
We denote the eq -north pole of S by s+ −
q and the eq -south pole by sq .

An i-sphere can have a single eq -pole (north or south) or an infinite number of them,
which occurs when a normal vector to the affine space containing the i-sphere is aligned
with the vector eq . We are interested in finding the eq -poles of (d − m + 1)-spheres orig-
inating from disk systems M = { D1 (c1 ; r1 ), . . . , Dm (cm ; rm )}, by taking the intersection
∩mj=1 ∂D j . Such ( d − m + 1)-spheres will be denoted by S M ( c; r ) to emphasize the disk
system M, as well as its center and radius.

Lemma 2. Let M = { D1 , . . . , Dm } be a d-disk system such that m j=1 D j ̸ = ∅, and let p be a


T

point in j=1 D j such that πq ( p) ≤ πq ( x ) (resp. πq ( p) ≥ πq ( x )) for every x in m


Tm T
j=1 D j . Then,
there exists an i-sphere S = ∂D j1 ∩ · · · ∩ ∂D ji such that p is in S and p is the eq -south pole (resp.
eq -north pole) of S.

Proof. Since ∩m j=1 D j ̸ = ∅, then ∂ (∩ j=1 D j ) ̸ = ∅, ∂ (∩ j=1 D j ) ⊂ ∩ j=1 D j due to the closedness
m m m

of the sets D j , for j = 1, . . . , m, and p ∈ ∂(∩m j =1 D j ).


On the other hand, since ∂(∩m D
j =1 j ) ⊂ ∪im=1 ∂D j , there exist indices j1 , . . . , ji such that
p ∈ ∂D jr for any r = 1, . . . , i; let Λ( p) = { j1 , . . . , ji } ⊆ {1, . . . , m} be a maximal subset of
indices such that p ∈ D j if and only if j ∈ Λ( p). We claim that p ∈ ∩ri =1 ∂D jr is the eq -south
pole of S := ∩ri =1 ∂D jr .
In effect, let Vp ⊂ Rd be an open neighborhood of p sufficiently small such that:
1. j=1 D j ) has as maximal set of indices a proper subset of Λ ( p ),
Every x ∈ Vp ∩ ∂(∩m
2. S ∩ Vp ⊂ ∂(∩m
j =1 D j ).
The first condition can be guaranteed by the finiteness of the disk system M, and the
second condition is a consequence of the maximality of the set Λ( p). Therefore, πq ( p) ≤
πq ( x ) for every x ∈ S ∩ Vp , which is equivalent to the fact that πq ( p) ≤ πq ( x ) for every
x ∈ S, in the case of i-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 (d − 1)-sphere given by the intersection of two disks in Rd .
Mathematics 2024, 12, 547 4 of 20

Lemma 3. Let M = { D1 (c1 ; r1 ), D2 (c2 ; r2 )} be a disk system with two d-disks such that ∂D1 ∩
∂D2 is a (d − 1)-sphere S = S M (c; r ) with center c and radius r. Then,
! !
1 r22 − r12 1 r12 − r22
c= 1+ c + 1+ c2
2 ∥ c2 − c1 ∥2 1 2 ∥ c2 − c1 ∥2

and p
2 s(s − ∥c2 − c1 ∥)(s − r1 )(s − r2 )
r= ,
∥ c2 − c1 ∥
where s = 21 (∥c2 − c1 ∥ + r1 + r2 ).

Proof. Let Π be the hyperplane containing the (d − 1)-sphere S, which is defined by the
equation:
d
1 d r2 − r2
∑ (ki − hi )xi − 2 ∑ (k2i − h2i ) = 1 2 2 , (1)
i =1 i =1

where c1 = (h1 , . . . , hd ) and c2 = (k1 , . . . , k d ). Then the normal vector of the hyperplane Π
is given by N := c2 − c1 = (k1 − h1 , . . . , k d − hd ), and the center c of S is determined by
the intersection point of the hyperplane Π with the perpendicular line that passes through
the center c1 of D1 . This line can be parameterized as γ : t 7→ c1 + tN = ( x1 (t), . . . , xd (t)),
such that γ(0) = c1 and γ(1) = c2 . We can compute the intersection point c = γ(t∗ ) of Π
and γ([0, 1]), for any t∗ ∈ [0, 1], by substituting it in (1),
!
d
1 d 2
∑ (ki − hi )(hi + t∗ (ki − hi )) = 2 ∑ (ki − hi ) + (r1 − r2 ) .
2 2 2
i =1 i =1

1 r12 −r22
And solving for t∗ , we obtain that t∗ = 2 + 2∥ c2 − c1 ∥2
. Hence, the center of S is given
by:
! !
1 r22 − r12 1 r12 − r22
c = c1 + t ∗ N = 1+ c + 1+ c2 .
2 ∥ c2 − c1 ∥2 1 2 ∥ c2 − c1 ∥2

Next, we will compute the radius r of S. This radius can be determined as the height r
of the triangle with base ∥c2 − c1 ∥ formed by the points c1 , c2 , and a point on S. Thus, by
Heron’s formula, we have
p
2 s(s − ∥c2 − c1 ∥)(s − r1 )(s − r2 )
r= ,
∥ c2 − c1 ∥

where s = 21 (∥c2 − c1 ∥ + r1 + r2 ) corresponds to the semi-perimeter.

We can proceed now to compute the poles of the (d − 1)-sphere ∂D1 ∩ ∂D2 .

Lemma 4. Let D1 (c1 ; r1 ) and D2 (c2 ; r2 ) be two d-disks such that ∂D1 ∩ ∂D2 is a (d − 1)-sphere
S = S(c; r ) with center c and radius r. Then, the eq -poles of S are s± d
q = c ± ∑i xi ei , where

r |πi (c2 − c1 )πq (c2 − c1 )|



, i ̸= q


 q
2 − π ( c − c )2




 ∥ c 2 − c 1 ∥ ∥ c 2 − c 1 ∥ q 2 1
xi =

 q


 r ∥ c2 − c1 ∥2 − π q ( c2 − c1 )2
, i = q.


∥ c2 − c1 ∥

Mathematics 2024, 12, 547 5 of 20

Proof. For simplicity, we translate the hyperplane Π, which contains the (d − 1)-sphere
S, as well as the sphere itself, to the origin; in such case, the corresponding equations are
given by

d
∑ (ki − hi )xi = 0,
i =1
d
∑ xi2 = r2 ,
i =1

where hi := πi (c1 ) and k i := πi (c2 ) for i = 1, . . . , d. In the case that k q − hq = πq (c2 − c1 ) =


0, the normal vector N = c2 − c1 of the hyperplane Π is orthogonal to the basis vector eq .
Therefore, the eq -poles of S are c ± req , which agree with the formulae of the lemma.
On the other hand, suppose that k q − hq ̸= 0. To find the eq -poles of S, we will use the
Lagrange multiplier method. Consider the following function:

− ∑dj̸=q (k j − h j ) x j
xq = f ( x1 , x2 , ..., xbq , ..., xd ) = , (2)
k q − hq

subject to the restriction:


!2
d − ∑dj̸=q (k j − h j ) x j
g( x1 , x2 , ..., xbq , ..., xd ) = ∑ x2j +
k q − hq
− r2 = 0.
j̸=q

Let λ be the Lagrange multiplier, we define

h( x1 , ..., xbq , ..., xd , λ) = f ( x1 , ..., xbq , ..., xd ) + λg( x1 , ..., xbq , ..., xd ).

For any i ̸= q, consider the following system of equations:

− ∑dj̸=q (k j − h j ) x j
!
k i − hi k − hi

∂h
=− + 2λxi + 2λ − i = 0.
∂xi k q − hq k q − hq k q − hq

Then
!
k − hi k i − hi d
(k q − hq )2 j∑
− i + 2λ xi + (k j − h j ) x j =0
k q − hq ̸=q
!
d
−(k i − hi )(k q − hq ) + 2λ (k q − hq )2 xi + (k i − hi ) ∑ (k j − h j ) x j = 0.
j̸=q

Solving this system of equations for λ, we obtain

(k i − hi )(k q − hq )
λ=  .
2 (k q − hq )2 xi + (k i − hi )∑dj̸=q (k j − h j ) x j

Comparing the last expression for two indices i ̸= ĩ, we have


Mathematics 2024, 12, 547 6 of 20

r 2 ( k i − h i )2
xi2 =  2
(k ĩ − hĩ )2 + ∑ j̸=ĩ,q (k j − h j )2 + 1
( k q − h q )2 ∑ j ̸ = q ( k j − h j )2
r 2 ( k i − h i )2
=  2
2
∑ j̸=q (k j − h j ) + (kq −1hq )2 ∑ j̸=q (k j − h j )2
r 2 ( k i − h i )2 ( k q − h q )2
=   
( k q − h q )2 + ∑ j ̸ = q ( k j − h j )2 ∑ j ̸ = q ( k j − h j )2
r 2 π i ( c2 − c1 )2 π q ( c2 − c1 )2
= .
∥ c2 − c1 ∥2 ∥ c2 − c1 ∥2 − π q ( c2 − c1 )2

Finally, for i = 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
c of the (d − m + 1)-sphere S = ∩m j=1 ∂D j . We can achieve this by considering the disk
system translated to cm , denoted as { D j (c j − cm ; r j )}m
j=1 , and by defining the ( d − m + 1)-
m
sphere S − {cm } = ∩ j=1 ∂D j (c j − cm ; r j ). This sphere is positioned at the intersection of
hyperplanes (for more details, refer to [17]).

1 2
(ck − cm ) T x = (r + ∥ck − cm ∥2 − rk2 ) (3)
2 m
for all k = 1, . . . , m − 1. Utilizing the information that the center of S − {cm } can be
expressed as a combination of the centers ck − cm and substituting it into (3), we obtain a
linear system of equations with dimensions (m − 1) × (m − 1):

m −1
1 2
∑ λ j (ck − cm ) · (c j − cm ) = (r + ∥ck − cm ∥2 − rk2 ),
2 m
j =1

for k = 1, ..., m − 1. Solving the system of equations for λ = (λ1 , . . . , λm−1 ), we find the
center of S as follows:

c = λ 1 ( c 1 − c m ) + · · · + λ m −1 ( c m −1 − c m ) + c m .

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

r2 = rk2 − ∥c − ck ∥2

for any k ∈ {1, . . . , m − 1}.


Now that we have determined the center and radius of S, as well as the affine space
that contains it, we can proceed to compute its eq -poles for each q ∈ {1, 2, . . . , d}. These
poles reside in the affine space that contains S and within a set that we define below.
Let S be an i-sphere in Rd , and let n1 , . . . , nd−i be orthogonal vectors to the affine space
L that contains S. Consider the space M generated by these vectors together with the vector
( j) d
 
eq from the canonical basis of Rd . Let us denote n j = nl for each j = 1, . . . , d − i.
l =1
Then, we can define L0 , the set L translated to the origin, as follows:
D E⊥
L0 = {n j }dj=−1i
n o
= x ∈ Rd | x · n j = 0, ∀ j = 1, . . . , d − i .
Mathematics 2024, 12, 547 7 of 20

The set M is defined as:


D E
M = {n j }dj=−1i ∪ {eq }
( )
d −i
= x= ∑ λ j n j + λ d − i +1 e q ∈ R d
| λ j ∈ R, ∀ j = 1, ..., d − i + 1
j =1
( ! )
d −i d −i d −i
∑ ∑ + λd−i+1 , ..., ∑
( j) ( j) ( j)
= λ j n1 , ..., λ j nq λ j nd | λj ∈ R .
j =1 j =1 j =1

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

Figure 1. Visualization of the subspaces L and M + {c}.

As mentioned above, the eq -poles of S lie at the intersection of L and M + {c}, where c
is the center of S. To simplify the calculations, we will utilize L0 and M, and then translate
them into c. The intersection of M and L0 can be expressed as follows:

M ∩ L0 = { x ∈ M | x · nk = 0, ∀k = 1, ..., d − i }
   
 d −i d −i 
= ∑ λ j n j + λ d − i +1 e q ∈ R d 
∑ λ j n j + λd−i+1 eq  · nk = 0, λ j ∈ R ∀k = 1, ..., d − i
 j =1 j =1

 
 d −i d −i 
= ∑ λ j n j + λd−i+1 eq ∈ Rd ∑ λ j n j · nk + λd−i+1 nq = 0, λ j ∈ R ∀k = 1, ..., d − i .
(k)
 j =1 j =1

Mathematics 2024, 12, 547 8 of 20

Let us consider a disk system in Rd , denoted { D j (c j ; r j )}m j=1 , where m < d. The
intersection of their boundaries forms a (d − m + 1)-sphere S. In this case, the subspace
M has dimension m, or dim( M ) = m − 1 if eq ∈ M. We choose the normal vectors for the
affine space containing S as n j = c j − cm , where j = 1, ..., m − 1. Then
  
 m −1   m −1   m −1   
M =  ∑ λ j c1 − c1
( j )
, ..., ∑ λ j cq − cq
( j )
+ λm , ..., ∑ λ j cd − cd
( m ) ( m ) ( j ) ( m )  λj ∈ R .
 j =1 j =1 j =1

By rewriting, we have
 
 m −1 m −1 
 j∑ ∑
(k)
M ∩ L0 = λ j n j + λ m e q ∈ Rd λ j n j · nk + λm nq = 0, λ j ∈ R, ∀k = 1, ..., m − 1 .
=1 j =1

If S(c; r ) = ∩m
j=1 ∂D j is the ( d − m + 1)-sphere with center in c and radius r, then the
eq -poles of S are the eq -poles of S − {cm } but translated by c. The poles of S − {cm } are in
M ∩ L0 . If p is an eq -pole of S − {cm }, then it can be expressed as

m −1
p= ∑ λ j n j + eq λ m ,
j =1

for some λ j ∈ R, j = 1, ..., m and the following conditions holds:

m −1

(k)
λ j n j · nk + λm nq = 0,
j =1

for each k = 1, ..., m − 1 and


∥ p ∥2 = r 2 .
−1
Thus, if p = ∑m j=1 λ j n j + eq λm is an eq -pole of S − { cm }, the following equations are
satisfied for λ1 , λ2 , ..., λm ∈ R:
m −1

(k)
λ j n j · nk + λm nq = 0 (4)
j =1

!2
d m −1 m −1
∑ ∑ ∑
( j) ( j)
λ j ni + 2λm λ j nq + λ2m = r2 , (5)
i =1 j =1 j =1

for all k = 1, ..., m − 1, with r the radius of the (d − m + 1)-sphere S. From (4) we have the
system
 (1) 
n1 · n1 n1 · n2 n 1 · n m −1 nq
  
.... λ1
 n2 · n1  (2) 
n2 · n2 .... n 2 · n m −1   λ2   nq 
 ...  + λm   = 0.
  
 ...  ... 
n m −1 · n 1 n m −1 · n 2 .... n m −1 · n m −1 λ m −1 n
( m −1)
q

−1 ( j)
Let us denote A as the matrix (ni · n j )i,j and B as the vector (−nq )m
j=1 . Then, we have
Aλ = λm B, where λ = (λ1 , ..., λm−1 ). Solving for λ, we obtain λ j = λm ( A−1 B)[ j] for each
j = 1, ..., m − 1 (where ( A−1 B)[ j] denotes the entry j of the (m − 1) × 1 vector ( A−1 B)). By
substituting the value of λ j into (5), we obtain the quadratic equation:
 !2 
d m −1 m −1
λ2m  ∑ ∑ ( A−1 B)[ j]ni ∑ ( A−1 B)[ j]nq
( j) ( j)
+2 + 1 − r2 = 0.
i =1 j =1 j =1
Mathematics 2024, 12, 547 9 of 20

Let us define
m −1
∑ ( A−1 B)[ j]ni
( j)
Γi = ,
j =1

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

±r
λm = q
∑id=1 Γ2i + 2Γq + 1

±r ( A−1 B)[ j]
λj = q ,
∑id=1 Γ2i + 2Γq + 1
for j = 1, ..., m − 1. Therefore, the eq -poles of S, for q = 1, ..., d, are:
!
m −1 m −1
±r
p= ∑ λ j n j + eq λm + c = q ∑ (A −1
B)[ j]n j + eq + c.
j =1 ∑id=1 Γ2i + 2Γq + 1 j =1

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 investigate 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 recog-
nizing and comprehending the geometric structure of point-cloud data. The 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 spe-
cific goals of the topological analysis. Now, let us move on to defining these fundamental
concepts.

Definition 3. Let M = { D1 , D2 , . . . , Dm } be a d-disk system. We say M is a Vietoris–Rips


system if Di ∩ D j ̸= ∅ for each pair i, j ∈ {1, 2, . . . , m}. Furthermore, if the d-disk system M has
the nonempty intersection property Di ∈ M Di ̸= ∅, then M is called a Čech system.
T

For each λ ≥ 0, we define a collection of d-disks Mλ with the same centers as those
in the d-disk system M, but with radii rescaled by λ. When λ > 0, Mλ is a d-disk system
again. M1 is equal to M, and M0 is the set of the centers of the d-disks in M.
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 , of a d-disk system M is the smallest λ ∈ R such that Mλ is a Vietoris–Rips system.
Similarly, the Čech scale µ M , of M is the smallest λ ∈ R such that Mλ is a Čech system.
This is

νM = inf{λ ∈ R | Mλ is a Vietoris–Rips system}


µ M = inf{λ ∈ R | Mλ is a Čech system}.
Mathematics 2024, 12, 547 10 of 20

Next, we present some easily observable properties for both scales. It can be easily seen
that M is a Vietoris–Rips system if and only if νM ≤ 1 (in particular νMν = 1); similarly,
M
M is a Čech system if and only if µ M ≤ 1.
Please note that for a given d-disk system M = { D1 , D2 , . . . , Dm } the Vietoris–Rips
scale is
νM = max{∥ci − c j ∥/(ri + r j )},
i< j

where ci and ri are the center and radii of Di . 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 d-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 plays a key role in implementation (see [11] for details).

Remark 1. If µ M is the Čech scale for M, then the µ M -rescaled system Mµ M , has only one
T
point in the intersection Di ∈ M Di (ci ; µ M ri ).

As we have mentioned, a Čech system is also a Vietoris–Rips system, but the converse
is not true. What we can affirm
p is that if a system is a Vietoris–Rips system, then the
system rescaled by the factor 2d/(d + 1) is also a Čech system. This is established by the
following lemma, the proof of which can be found in [11].

Lemma 5. Let M = { Di (ci ; ri )} be a d-disk system in Euclidean space Rd . If Di (ci ; ri ) ∩


D j (c j ; r j ) ̸= ∅ for every pair of disks in M, then
q
2d/(d + 1) ri ) ̸= ∅.
\
Di ( c i ;
Di ∈ M

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

Corollary 1. If M is an arbitrary
p d-disk system and νM is its Vietoris–Rips scale, then its Čech
scale satisfies νM ≤ µ M ≤ 2d/(d + 1) νM . Therefore, for every d-disk
p system M, the rescaled
disk system M√2d/(d+1) ν is always a Čech system. In particular, if 2d/(d + 1) νM ≤ 1 then
M
MνM is a Čech system.

3.1. Algorithm for Determining Čech System


In the previous section, we have determined the eq -poles for the intersection of any
number of disks in Rd . 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 nonempty intersection.
Given a system of m disks in Rd where m > d, it is enough to verify if every subsystem
of d + 1 disks qualifies as a Čech system to conclude that the entire system of disks has a
nonempty 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 nonempty 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 i-spheres. It initiates the search for poles within individual disks and then
progresses to the pairwise intersections of the disk boundaries ((d − 2) − sphere), continuing
Mathematics 2024, 12, 547 11 of 20

the process iteratively. If a pole is found within the remaining disks, the system is classified
as a Čech system.

Theorem 1. Let M = { D1 , . . . , Dm } be a d-disk system. Then, M is a Čech system if and only if


Cech.system( M) = TRUE.

Proof. If Cech.system( M ) = TRUE, then the Cech.system algorithm (Algorithm 1) found


j=1 D j ; therefore ∩ j=1 D j ̸ = ∅ and it follows that M is
a pole contained in the intersection ∩m m

a Čech system.
On the other hand, if m
Tm
j=1 D j ̸ = ∅, let p be a point in j=1 D j satisfying π1 ( p ) ≤ π1 ( x )
T
Tm
for every x in j=1 D j . By Lemma 2, it follows that p belongs to an i-sphere and must be
an e1 -south pole. Therefore, by the exhaustive search of Algorithm 1 across all poles, its
output is Cech.system( M ) = TRUE.

Algorithm 1: Cech.system
Input : A d-disk system M = { D j }m
j =1
Output : A logical TRUE/FALSE to indicate if M is a Čech system
1 Initialize: Is_Cech_System ← FALSE
2 for k ← 1 to m do
3 Let S be the set of (d − k + 1)-spheres of ∂M
4 for S in S do
5 for q ← 1 to d do
6 Compute the set {s± q } of eq -poles of S
7 for s ← {s± q } do
8 if s ∈ ∩dj=1 D j then
9 Is_Cech_System ← TRUE
10 Go to line 16
11 end
12 end
13 end
14 end
15 end
16 return (Is_Cech_System)

3.2. Algorithm to Compute the Čech Scale


Finding the minimum parameter for which the rescaled system of disks has a nonempty
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 (Algorithm 2) to compute
an approximation of the Čech scale for a system of m disks in Rd .
The given algorithm presents an algorithm to compute an approximation of the
Čech scale of a disk system in Euclidean space using Algorithm 1 and a precision pa-
rameter η > 0. It initializes the scale factor λ to the Rips scale νM . If this scale satisfies
Cech.system( MνM ) = 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 p by a factor
λ. The Čech scale is known to fall between the Rips scale and the value 2d/(d + 1)νM
(Generalized Vietoris–Rips, Corollary 1). 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. Finally, the algorithm returns an approximation of the Čech scale.
Mathematics 2024, 12, 547 12 of 20

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

C∗ ( M ) : C0 ( M) ⊊ Cλ1 ( M ) ⊊ Cλ2 ( M ) ⊊ ... ⊊ Cµ M ( M),

where each λi represents the Čech scale of the system Mλi . 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), requires determining the Čech scale of the system Mλ .

Algorithm 2: Cech.scale
Input : A d-disk system M in Rd and precision parameter η > 0
Output : µ M , a Čech scale approximation
1 Compute the Vietoris–Rips scale of M: νM
2 if Cech.system( MνM ) = TRUE then
3 µ M ← νM
4 Go to line 16
5 else
Initialize: µ∗M ← νM , µ M ← 2d/(d + 1)νM
p
6
7 while µ M − µ∗M > η do
µ∗ + µ M
8 Compute: λ ← M
2
9 if Cech.system( Mλ ) = TRUE then
10 Update: µ M ← λ
11 else
12 Update: µ∗M ← λ
13 end
14 end
15 end
16 return (µ M )

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 d-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 4. Let M be a disk system in Rd . The minimal Axis-Aligned Bounding Box of M,


denoted as AABB( M ), is defined as the smallest Axis-Aligned Bounding Box that contains the
intersection D = ∩ Di ∈ M Di , given by
\
AABB( M ) := B̃,
D ⊂ B̃

where B̃ ranges over all axis-aligned bounding boxes that contain D.

Please note that the AABB can be expressed as

d
AABB( M ) = ∏ [inf πk (∂D), sup πk (∂D)],
k =1
Mathematics 2024, 12, 547 13 of 20

where πk : Rd −→ R is the canonical projection onto the k-th factor, and ∂D denotes the
boundary of D. In other words, the AABB of a disk system M 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 us consider the situation when the disk system M is composed of two disks D1
and D2 in Rd . If D1 ∩ D2 ̸= ∅ and D1 ̸= D2 , the subset ∂D1 ∩ ∂D2 can take one of three
forms: an empty set, a single common point (when the disks are tangent), or a (d − 2)-
dimensional sphere. In the last case, we denote the (d − 2)-dimensional sphere or the
(d − 1)-sphere ∂D1 ∩ ∂D2 by S1,2 . To calculate inf πi (∂( D1 ∩ D2 )) and sup πi (∂( D1 ∩ D2 ))
for each i ∈ {1, 2, ..., d}, we can use:

 π i ( c1 − r1 ei )
 if c1 − r1 ei ∈ D2 ,
inf πi (∂( D1 ∩ D2 )) = πi (c2 − r2 ei ) if c2 − r2 ei ∈ D1 , (6)

inf(πi (∂D1 ∩ ∂D2 )) otherwise,


 π i ( c1 + r1 ei )
 if c1 + r1 ei ∈ D2 ,
sup πi (∂( D1 ∩ D2 )) = πi (c2 + r2 ei ) if c2 + r2 ei ∈ D1 ,

sup(πi (∂D1 ∩ ∂D2 )) otherwise.

Indeed, by Lemma 2, the extremes of the AABB are the projections of certain poles,
either from the (d − 1)-sphere or some d-sphere. The d-spheres represent the boundaries of
each disk, with poles given by c j ± r j ei , and the (d − 1)-sphere is S1,2 = ∂D1 ∩ ∂D2 , whose
poles are computed using Lemma 4. It is worth noting that there are no further options for
(d − m + 1)-spheres in the case of a two-disk system.
To simplify the notation, we will use Bi,j to denote the AABB of the intersection of
disks Di and D j . Figure 2 illustrates the AABB of the intersection of two disks in the plane.

Figure 2. AABB of two disks.


Mathematics 2024, 12, 547 14 of 20

According to (6) and Lemma 3, we have a method to calculate the axis-aligned bound-
ary 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 Rd . 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 R3 .
Example 1. Let
√ √
M = { D1 ((4, 1, 0); 2), D2 ((4, −1, 0); 2), D3 ((0, 0, 0); 3)} ⊂ R3

be a Vietoris–Rips system in R3 with projection onto the xy-plane shown in Figure 3.

Figure 3. Disk system M projected onto the xy-plane.

By computing the boxes Bi,j for all i < j, we obtain:


√ √
B1,2 = [3, 5] × [− 2, 2] × [−1, 1]

B1,3 = [4 − 2, 3] × [0, 1.41] × [−0.72, 0.72]

B2,3 = [4 − 2, 3] × [−1.4, 0] × [−0.72, 0.72]

Therefore, ∩i< j Bi,j = [3, 3] × [0, 0] × [−0.72, 0.72]. However, the disk system intersects
at the point P = (3, 0, 0), which means AABB( M ) = { P}. In other words, AABB( M ) is
not equal to ∩i< j Bi,j .
We know that if D ̸= ∅, then ∩i< j Bi,j ̸= ∅. However, the converse is not always true.
The following example illustrates this fact.

Example 2. Let N = { D1 , D2 , D3 ((0, 1, 0); 10), D4 ((3, 0, 1); 0.9)} ⊂ R3 be a Vietoris–Rips
system. The projection of disks D1 , D2 , and D3 onto the xy-plane is illustrated in Figure 4.
We will now compute the intersections Bi,j for different pairs of disks:
Mathematics 2024, 12, 547 15 of 20

√ √
B1,2 = [3, 5] × [− 2, 2] × [−1, 1]
√ √
B1,3 = [4 − 2, 10] × [0, 2] × [−1, 1]
B1,4 = [2.6, 3.9] × [−0.4, 0.9] × [0.100007, 1.3]
B2,3 = [2.6, 3] × [−0.8, 0] × [−0.44, 4.47]
B2,4 = [2.6, 3.9] × [−0.9, 0.4] × [0.100007, 1.3]
B3,4 = [2.1, 3.11] × [−0.73, 0.9] × [0.1, 1.7]

The intersection of all pairwise intersections, ∩i< j Bi,j , is given by [3, 3] × [0, 0] ×
[0.1, 1.7]. However, the intersection of disks D1 , D2 , and D3 is a single point P, which is not
contained in D4 (by construction). Therefore, D is an empty set, but ∩i< j Bi,j is not.

Figure 4. Disk system N projected onto the xy-plane.

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 1, we observe that the
intersection of AABB({ Di , D j }) is not equal to the AABB of the disk system M. Similarly,
in Example 2, we find that the intersection of three disks is empty, yet the intersection of
AABB({ Di , 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 Rd .
Mathematics 2024, 12, 547 16 of 20

4.2. Minimal Axis-Aligned Bounding Box for More Than Two Disks
Given a system M consisting of m disks in Rd , we can compute the eq -poles for any
subcollection of disks. Using these eq -poles, we can determine the Axis-Aligned Bounding
Box (AABB) of M.
If m ≤ d, we calculate the eq -poles of the (d − m + 1)-sphere ∩∂Di , and with these
poles, we define the AABB of M by taking inf πq ∂D = πq ( p), where p is the eq -south pole
of the (d − m + 1)-sphere (similarly for sup πq ∂D).
In the case where m = d + 1, we consider AABB( M (1)), ..., AABB( M(m)) as a col-
lection of minimal axis-aligned boxes for the disk system M (i ) = M − { Di } in Rd . If the
intersection of any d + 1 of these sets is nonempty, then the intersection of the entire collec-
tion gives us the minimal axis-aligned box of the disk system M. This can be expressed as
AABB( M) = ∩ AABB( M (i )). The next theorem confirms this finding.

Theorem 2 (Helly-type theorem for minimal axis-aligned bounding boxes). Let M be a Rips
system with d + 1 disks in Rd . Then, the minimal Axis-Aligned Bounding Box for the intersection
set D = ∩dj=+11 D j satisfies:
+1
d\
AABB( M ) = AABB( M ( j))
j =1

where M( j) = M − { D j }.
T d +1
Proof. We know that AABB( M ) ⊆ j =1 AABB( M ( j)). Now, our objective is to establish
T d +1
the reverse inclusion, i.e., j=1 AABB( M ( j)) ⊆ AABB( M). To derive a contradiction,
suppose that the reverse inclusion is not true. By definition, we have:

d
AABB( M ) = ∏[inf πi (∂D), sup πi (∂D)]
i =1

and
+1
d\ d  
AABB( M ( j)) = ∏ max{inf πi (∂ ∩ j̸=k D j )}, min{sup πi (∂ ∩ j̸=k D j )} .
k k
j =1 i =1

Without loss of generality, let us assume that:


n o d +1
π1 ( p) > max inf π1 (∩ j̸=k ∂D j ) ,
k k =1

where p ∈ D satisfies π1 ( p) = inf π1 (∂D ).


Let q j be the point in ∩k̸= j Dk such that π1 (q j ) = inf π1 (∩k̸= j Dk ) for each j = 1, 2, . . . ,
d + 1. Please note that q j ∈ / D j because q j is not in D. Now, let γ j be the line segment that
connects q j and p.
Choose ϵ > 0 small enough such that the hyperplane P : x1 = π1 ( p) − ϵ does not
contain any q j , and π1 (q j ) < π1 ( p) − ϵ for all j = 1, 2, . . . , d + 1 (such hyperplane P exists
because π1 ( p) > π1 (q j ) for all j = 1, . . . , d + 1). Since q j is in ∩k̸= j Dk , the hyperplane P
intersects every disk in M ( j) for each j, and γ j ⊂ ∩k̸= j Dk intersects P at a point that is in
∩k̸= j Dk (see Figure 5). Furthermore, Dk ∩ P is a (d − 1)-dimensional disk. Therefore, we
have a collection D = { D j ∩ P| j = 1, 2, . . . , d + 1} of d + 1 disks, each of dimension d − 1,
such that every subset A of D consisting of d disks has the nonempty intersection property.
By Helly’s Theorem, the intersection of all (d − 1)-disks in D is not empty. Therefore, there
exists a point q ∈ D with π1 (q) < π1 ( p). However, this contradicts the fact that p ∈ D is
such that π1 ( p) = inf π1 (∂D ).
Therefore, we conclude that dj=+11 AABB( M( j)) = AABB( M ).
T
Mathematics 2024, 12, 547 17 of 20

Figure 5. The hyperplane P.

Lemma 6. Let M be a Vietoris–Rips system of d + 1 disks in Rd . If D = ∩id=+11 Di = ∅ and


AABB( M( j)) ̸= ∅ for each j = 1, . . . , d + 1, then ∩id=+11 AABB( M (i )) consists only of intervals
of the form [ a, b] with a > b (inverted intervals).

Proof. Suppose that ∩id=+11 AABB( M(i )) contains an interval. Without loss of generality, let
us assume it is the interval

[inf π1 (∂D (k)), sup π1 (∂D (l ))],

for some k, l ∈ {1, 2, . . . , d + 1}, where D (k) = ∩i̸=k Di . By definition of ∩id=+11 AABB( M (i )),
we have inf π1 (∂D (k)) ≥ inf π1 (∂D (i )) for all i ̸= k and sup π1 (∂D (l )) ≤ sup π1 (∂D ( j))
for all j ̸= l. Now, consider a hyperplane P : x1 = p with inf π1 (∂D (k)) ≤ p ≤
sup π1 (∂D (k)). The hyperplane P intersects D ( j) for every j = 1, . . . , d + 1 (since P cuts
Mathematics 2024, 12, 547 18 of 20

through all the boxes AABB( M ( j))). We also know that P ∩ Di is a d − 1-dimensional disk
for each i.
Therefore, we have a collection of d disks in a d − 1-dimensional space, and by Helly’s
Theorem, this collection must have a nonempty intersection. This implies that D ̸= ∅,
which contradicts our assumption that D is empty. Hence, all intervals in ∩id=+11 AABB( M(i ))
must be inverted intervals of the form [ a, b] with a > b.

As an example, we provide the lower bounds of the AABB for a system of three
disks in Rd , i.e., we compute inf πi (∂D ) for each i = 1, . . . , d (analogous sup πi ∂D) with
D = ∩3j=1 D j . Let M = { D1 , D2 , D3 } be a Rips system in Rd , then, for each i = 1, . . . , d, the
computation of the AABB for the disk system M is given by:

 π i ( c k − r k ei )
 if ck − rk ei ∈ D
inf πi (∂D ) = max j<k {inf πi (∂D j ∩ ∂Dk )} if (*)
inf πi (∩3j=1 ∂D j )

otherwise

where (*) denotes the case where q ∈ D with q ∈ ∂D j ∩ ∂Dk such that

πi (q) = max{inf πi (∂D j ∩ ∂Dk )}.


j<k

The preceding calculations are a result of Lemma 2. It is known that the extremes of
the AABB lie in the projections of specific poles, either from a d-sphere, (d − 1)-sphere, or
the (d − 2)-sphere.
Given a disk system M in Rd , we can compute the minimal Axis-Aligned Bounding
Box (AABB) for the intersection of all disks in M. 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 M is not a point, we can rescale M by a scale factor λ such that the AABB of
Mλ becomes a point. The value of λ is referred to as the Čech scale of the system M.

4.3. Minimal Axis-Aligned Bounding Box Algorithm


Now, we present an algorithm (Algorithm 3: AABB.minimal) to calculate the minimal
Axis-Aligned Bounding Box (AABB) of a system of m disks in Rd . As previously explained,
the AABB’s extremes are defined by projecting the poles of specific i-spheres. Thus, in
computing the AABB, we will determine the poles for each i-sphere within the disk system.
The Algorithm 3 starts by initializing the set of poles, P, 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 M (for this, we require the center and
radius, which are computed in Section 2.2). Next, we identify the poles of each sphere, and
if any of them are present in all the disks of M, we add them to the set P. If P 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 P is empty,
it indicates that there is no intersection in the disk system.
Mathematics 2024, 12, 547 19 of 20

Algorithm 3: AABB.minimal
Input : A d-disk system M = { D j }m j =1
Output : The minimal AABB for the system M
1 Initialize: P = ∅
2 for k ← 1 to m do
3 Let S be the set of (d − k + 1)-spheres of ∂M
4 for S in S do
5 for q ← 1 to d do
6 Compute the set {s± q } of eq -poles of S
7 for s ← {s± q } do
8 if s ∈ ∩dj=1 D j then
9 Add: P ← s
10 end
11 end
12 end
13 end
14 end
15 if P ̸ = ∅ then
16 for q ← 1 to d do
17 aq = min P {πq (s− q )}
18 bq = max P {πq (s+ q )}
19 end
20 return (Πdq=1 [ aq , bq ])
21 else
22 return (The disk system does not intersect)
23 end

5. Conclusions
Addressing the problem of computing the Čech filtration for a finite disk system,
as well as the computation of the minimal Axis-Aligned Bounding Box (AABB) for the
intersection of such a disk system, in this paper, we have designed three algorithms:
Cech.system (a TRUE/FALSE test to indicate whether a disk system has a nonempty in-
tersection), Cech.scale (a numerical method approach to approximate the Čech scale of
a disk system with any given precision), and AABB.minimal (an exhaustive search algo-
rithm for eq -poles whose output is the minimal AABB for the intersection of the disks
in any disk system). The algorithms we have described are fundamentally based on
Lemmas 1 and 2 in Section 2. The latter is a key result for this work, as it directs our atten-
tion to a finite collection of points: the north/south poles of spheres of different dimensions,
which are determined by the intersection of other spheres. These points allow us to suffi-
ciently characterize the intersection of a finite collection of disks so that it can be bounded
by axis-aligned boxes and to determine if such an intersection is nonempty. Furthermore, to
make this result computable (as required in applications, such as topological data analysis),
we have developed explicit formulae for computing such poles in Sections 2.1 and 2.2.

Author Contributions: Conceptualization, J.F.E. and C.G.E.-P.; methodology, J.F.E. and C.G.E.-P.; investi-
gation, J.F.E. and C.G.E.-P.; writing—original draft preparation, J.F.E. and C.G.E.-P.; writing—review and
editing, J.F.E. and C.G.E.-P.; visualization, C.G.E.-P.; supervision, J.F.E.; All authors have read and
agreed to the published version of the manuscript.
Funding: C.G.E.P acknowledges CONAHCYT for the financial support provided through a National
Fellowship (CVU-638165).
Data Availability Statement: No new data were created or analyzed in this study. Data sharing is
not applicable to this article.
Mathematics 2024, 12, 547 20 of 20

Conflicts of Interest: The authors declare no conflicts of interest.

References
1. Kerber, M.; Sharathkumar, R. Approximate Čech Complex in Low and High Dimensions. In Algorithms and Computation; Cai, L.,
Cheng, S.W., Lam, T.W., Eds.; Springer: Berlin/Heidelberg, Germany, 2013; pp. 666–676.
2. Le, N.K.; Vergne, A.; Martins, P.; Decreusefond, L. Optimize wireless networks for energy saving by distributed computation of
Čech complex. In Proceedings of the 2017 IEEE 13th International Conference on Wireless and Mobile Computing, Networking
and Communications (WiMob), Rome, Italy, 9–11 October 2017; pp. 1–8. https://doi.org/10.1109/WiMOB.2017.8115760.
3. Chazal, F.; Michel, B. An Introduction to Topological Data Analysis: Fundamental and Practical Aspects for Data Scientists. Front.
Artif. Intell. 2021, 4, 108. https://doi.org/10.3389/frai.2021.667963.
4. Meng, Z.; Anand, D.V.; Lu, Y.; Wu, J.; Xia, K. Weighted persistent homology for biomolecular data analysis. Sci. Rep. 2020,
10, 2079. https://doi.org/10.1038/s41598-019-55660-3.
5. Bell, G.; Lawson, A.; Martin, J.; Rudzinski, J.; Smyth, C. Weighted persistent homology. Involv. J. Math. 2019, 12, 823–837.
https://doi.org/10.2140/involve.2019.12.823.
6. Haines, E.; Günther, J.; Akenine-Möller, T. Precision Improvements for Ray/Sphere Intersection. In Ray Tracing Gems: High-Quality
and Real-Time Rendering with DXR and Other APIs; Haines, E., Akenine-Möller, T., Eds.; Apress: Berkeley, CA, USA, 2019; pp. 87–94.
https://doi.org/10.1007/978-1-4842-4427-2_7.
7. Petitjean, M., Spheres Unions and Intersections and Some of their Applications in Molecular Modeling. In Distance Geometry:
Theory, Methods, and Applications; Mucherino, A., Lavor, C., Liberti, L., Maculan, N., Eds.; Springer: New York, NY, USA, 2013;
pp. 61–83. https://doi.org/doi.org/10.1007/978-1-4614-5128-0_4.
8. Ratschek, H.; Rokne, J. Box-sphere intersection tests. Comput.-Aided Des. 1994, 26, 579–584. https://doi.org/10.1016/0010-448
5(94)90089-2.
9. Ramos, E.A. Intersection of unit-balls and diameter of a point set in R3 . Comput. Geom. 1997, 8, 57–65. https://doi.org/10.1016/
S0925-7721(96)00010-7.
10. Le, N.K.; Martins, P.; Decreusefond, L.; Vergne, A. Construction of the Generalized Czech Complex. In Proceedings of the 2015
IEEE 81st Vehicular Technology Conference (VTC Spring), Glasgow, UK, 11–14 May 2015; pp. 1–5. https://doi.org/10.1109/
VTCSpring.2015.7145759.
11. Espinoza, J.F.; Hernández-Amador, R.; Hernández-Hernández, H.A.; Ramonetti-Valencia, B. A Numerical Approach for the
Filtered Generalized Čech Complex. Algorithms 2019, 13, 11. https://doi.org/10.3390/a13010011.
12. Elfizar, S.; Sastria, G. Analysis of Axis Aligned Bounding Box in Distributed Virtual Environment. Int. J. Comput. Appl. 2014,
105, 29–33. https://doi.org/10.5120/18351-9475.
13. Wang, C.C.; Wang, M.; Sun, J.; Mojtahedi, M. A Safety Warning Algorithm Based on Axis Aligned Bounding Box Method to
Prevent Onsite Accidents of Mobile Construction Machineries. Sensors 2021, 21, 7075. https://doi.org/10.3390/s21217075.
14. Sulaiman, H.A.; Othman, M.A.; Ismail, M.M.; Said, M.A.M.; Ramlee, A.; Misran, M.H.; Bade, A.; Abdullah, M.H. Distance
computation using axis aligned bounding box (AABB) parallel distribution of dynamic origin point. In Proceedings of the
2013 Annual International Conference on Emerging Research Areas and 2013 International Conference on Microelectronics,
Communications and Renewable Energy, Kanjirapally, India, 4–6 June 2013; pp. 1–6.
15. Mahovsky, J.; Wyvill, B. Fast Ray-Axis Aligned Bounding Box Overlap Tests with Plucker Coordinates. J. Graph. Tools 2004,
9, 35–46. https://doi.org/10.1080/10867651.2004.10487597.
16. Cai, P.; Indhumathi, C.; Cai, Y.; Zheng, J.; Gong, Y.; Lim, T.S.; Wong, P., Collision Detection Using Axis Aligned Bounding
Boxes. In Simulations, Serious Games and Their Applications; Cai, Y., Goei, S.L., Eds.; Springer: Singapore, 2014; pp. 1–14.
https://doi.org/10.1007/978-981-4560-32-0_1.
17. Maioli, D.S.; Lavor, C.; Gonçalves, D.S. A note on computing the intersection of spheres in Rn . Anziam J. 2017, 59, 271–279.
https://doi.org/10.1017/S1446181117000372.

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.

You might also like