Processing math: 100%
Research article Special Issues

On the hyperbolicity of Delaunay triangulations

  • If X is a geodesic metric space and x1,x2,x3X, a geodesic triangle T={x1,x2,x3} is the union of the three geodesics [x1x2], [x2x3] and [x3x1] in X. The space X is hyperbolic if there exists a constant δ0 such that any side of any geodesic triangle in X is contained in the δ-neighborhood of the union of the two other sides. In this paper, we study the hyperbolicity of an important kind of Euclidean graphs called Delaunay triangulations. Furthermore, we characterize the Delaunay triangulations contained in the Euclidean plane that are hyperbolic.

    Citation: Walter Carballosa, José M. Rodríguez, José M. Sigarreta. On the hyperbolicity of Delaunay triangulations[J]. AIMS Mathematics, 2023, 8(12): 28780-28790. doi: 10.3934/math.20231474

    Related Papers:

    [1] Belén Ariño-Morera, Angélica Benito, Álvaro Nolla, Tomás Recio, Emilio Seoane . Looking at Okuda's artwork through GeoGebra: A Citizen Science experience. AIMS Mathematics, 2023, 8(8): 17433-17447. doi: 10.3934/math.2023890
    [2] José Luis Díaz Palencia, Abraham Otero . Instability analysis and geometric perturbation theory to a mutual beneficial interaction between species with a higher order operator. AIMS Mathematics, 2022, 7(9): 17210-17224. doi: 10.3934/math.2022947
    [3] Yutong Zhou . Idempotent completion of right suspended categories. AIMS Mathematics, 2022, 7(7): 13442-13453. doi: 10.3934/math.2022743
    [4] Naveed Iqbal, Meshari Alesemi . Soliton dynamics in the $ (2+1) $-dimensional Nizhnik-Novikov-Veselov system via the Riccati modified extended simple equation method. AIMS Mathematics, 2025, 10(2): 3306-3333. doi: 10.3934/math.2025154
    [5] Muhammad Imran Asjad, Naeem Ullah, Asma Taskeen, Fahd Jarad . Study of power law non-linearity in solitonic solutions using extended hyperbolic function method. AIMS Mathematics, 2022, 7(10): 18603-18615. doi: 10.3934/math.20221023
    [6] Mahmoud A. E. Abdelrahman, Wael W. Mohammed, Meshari Alesemi, Sahar Albosaily . The effect of multiplicative noise on the exact solutions of nonlinear Schrödinger equation. AIMS Mathematics, 2021, 6(3): 2970-2980. doi: 10.3934/math.2021180
    [7] Chun Huang, Zhao Li . Soliton solutions of conformable time-fractional perturbed Radhakrishnan-Kundu-Lakshmanan equation. AIMS Mathematics, 2022, 7(8): 14460-14473. doi: 10.3934/math.2022797
    [8] Shahbaz Ali, Muhammad Khalid Mahmmod, Raúl M. Falcón . A paradigmatic approach to investigate restricted hyper totient graphs. AIMS Mathematics, 2021, 6(4): 3761-3771. doi: 10.3934/math.2021223
    [9] Imrana Kousar, Saima Nazeer, Abid Mahboob, Sana Shahid, Yu-Pei Lv . Numerous graph energies of regular subdivision graph and complete graph. AIMS Mathematics, 2021, 6(8): 8466-8476. doi: 10.3934/math.2021491
    [10] Muhammad Zeeshan Hanif, Naveed Yaqoob, Muhammad Riaz, Muhammad Aslam . Linear Diophantine fuzzy graphs with new decision-making approach. AIMS Mathematics, 2022, 7(8): 14532-14556. doi: 10.3934/math.2022801
  • If X is a geodesic metric space and x1,x2,x3X, a geodesic triangle T={x1,x2,x3} is the union of the three geodesics [x1x2], [x2x3] and [x3x1] in X. The space X is hyperbolic if there exists a constant δ0 such that any side of any geodesic triangle in X is contained in the δ-neighborhood of the union of the two other sides. In this paper, we study the hyperbolicity of an important kind of Euclidean graphs called Delaunay triangulations. Furthermore, we characterize the Delaunay triangulations contained in the Euclidean plane that are hyperbolic.



    The concept of Gromov hyperbolicity grasps the essence of negatively curved spaces like the classical hyperbolic space, Riemannian manifolds of negative sectional curvature, and of discrete spaces like the Cayley graphs of many finitely generated groups and trees [1,2,3]. Initially, Gromov spaces were applied to the study of automatic groups in computer science [4,5].

    The hyperbolicity of a geodesic metric space is equivalent to the hyperbolicity of a graph related to it [6]. This conceptualization has multiple practical applications such as networks and algorithms [7], random graphs [8,9,10], or real networks [11,12]. For example, it has been shown in [13,14] that the internet topology and the topology of many complex networks embed with better accuracy into a hyperbolic space than into an Euclidean space of comparable dimension. Another important application of Gromov hyperbolic spaces is the study of the spread of viruses through on the internet [15,16] and secure transmission of information on the network [17].

    By a tessellation graph we mean the 1-skeleton (i.e., the set of 1-cells) of a CW 2-complex contained in the Euclidean plane R2 such that every point in R2 is contained in some face (2-cell) of the complex. The edges of a tessellation graph are straight segments and have its Euclidean length. Note that this class of graphs contains as particular cases many planar graphs.

    Delaunay triangulations are an important kind of Euclidean graphs. It is important to note that Delaunay triangulations are the dual of Voronoi graphs (the 1-skeleton of the Voronoi diagram) in the Euclidean plane [18,19,20]. Euclidean graphs have been used as a modeling tool; for example, in circuit layout [21]. They also arise indirectly in the solution to geometric problems such as motion planning [22]. Delaunay triangulations are good candidates for approximating graphs, since if S has N points, then the triangulation contains only a linear number of edges and can be computed from S in O(NlogN) time [23].

    Delaunay triangulations tend to have desirable geometric properties, such as maximizing the minimum angle of the triangles, which leads to more regular and well-conditioned triangles. This quality is crucial in applications where accurate and robust geometric computations are required. Delaunay triangulations are widely used for generating triangular meshes. They provide a way to divide a complex domain into a set of non-overlapping triangles, which is useful in finite element analysis, computational physics, and computer-aided design (CAD) applications. Also, Delaunay triangulations can be employed for interpolating values or reconstructing surfaces from scattered data points. They provide a natural and optimal way to connect data points, allowing for the creation of smooth surfaces or accurate approximations based on the input data. Besides, Delaunay triangulations serve as a fundamental building block in many computational geometry algorithms. They are used as a data structure to efficiently solve problems such as nearest neighbor search, point location, convex hull computation, and geometric intersection problems.

    Since the Euclidean plane is non-hyperbolic, it is natural to conjecture that "every tessellation of the Euclidean plane with convex tiles induces a non-hyperbolic graph". In [24] it was shown that the conjecture is false. However, we prove in this work that it is true for every Delaunay triangulation which is a triangulation of the Euclidean plane, see Corollary 3.14. Furthermore, Theorem 3.13 characterizes the Delaunay triangulations contained in the Euclidean plane that are hyperbolic in terms of only two geometric parameters associated with the triangulation (they may or may not be a tessellation of the plane).

    If X is a metric space, the curve γ:[a,b]X is a geodesic if L(γ|[t,s])=d(γ(t),γ(s))=st for every s,t[a,b] with t<s. The metric space X is said geodesic if for every couple of points in X there exists a geodesic joining them; we denote by [xy] any geodesic joining x and y. Hence, any geodesic metric space is connected. If the metric space X is a graph, then the edge joining the vertices u and v will be denoted by uv.

    In order to consider a graph G as a geodesic metric space, we must identify any edge uvE(G) with a real interval with length L(uv); thus, any point in the interior of any edge is a point of G. Any connected graph G is naturally equipped with a distance defined on its points, induced by taking shortest paths in G. Therefore, we consider G as a geodesic metric graph.

    If X is a geodesic metric space and x1,x2,x3 are points in X, then the union of three geodesics [x1x2], [x2x3] and [x3x1] is a geodesic triangle, and we denote it by T={x1,x2,x3} or T={[x1x2],[x2x3],[x3x1]}. The geodesic triangle T is δ-thin if any side of T is contained in the δ-neighborhood of the union of the two other sides. We denote by δ(T) the sharp thin constant of T, i.e., δ(T):=inf{δ0:T is δthin}. We say that the space X is δ-hyperbolic (or satisfies the Rips condition with constant δ) if every geodesic triangle in X is δ-thin. We will denote by δ(X) the sharp hyperbolicity constant of X, i.e., δ(X):=sup{δ(T):T is a geodesic triangle inX}. We say that X is hyperbolic if X is δ-hyperbolic for some δ0, i.e., if δ(X)<.

    Given any graph G and x,yG, we define the distance dG(x,y) (or simply d(x,y)) as the minimum of the lengths of the curves in G joining x and y.

    Throughout the paper, we deal with graphs that are connected and such that each ball of finite radius contains just a finite number of edges; we also allow edges of arbitrary length. Then, the graphs are geodesic metric spaces.

    There are several equivalent definitions of hyperbolicity [2,25,26,27]. It is natural to chose this definition by its deep geometric meaning [2]. See [27,28,29,30,31] for more details about hyperbolic graphs.

    Given two geodesic metric spaces (X,dX) and (Y,dY), a map f:XY is said to be an (α,β)-quasi-isometric embedding, with constants α1, β0, if the inequalities

    α1dX(x,y)βdY(f(x),f(y))αdX(x,y)+β.

    hold for every x,yX. We say that f is ε-full if for each yY there exists xX with dY(f(x),y)ε. The map f is a quasi-isometry if there exist constants α1, β,ε0 such that f is an ε-full (α,β)-quasi-isometric embedding.

    A fundamental property of hyperbolic spaces is the following:

    Theorem 2.1 (Invariance of hyperbolicity). Let f:XY be an (α,β)-quasi-isometric embedding between the geodesic metric spaces X and Y. If Y is hyperbolic, then X is hyperbolic.

    Furthermore, if f is ε-full for some ε0 (a quasi-isometry), then X is hyperbolic if and only if Y is hyperbolic.

    Given a geodesic triangle T={x,y,z} in a geodesic metric space X, let TE be an Euclidean triangle with sides of the same length than T. Since there is no possible confusion, denote the corresponding points in T and TE by the same letters. The maximum inscribed circle in TE meets the side [xy] (respectively [yz], [zx]) in a point z (respectively x, y) such that d(x,z)=d(x,y), d(y,x)=d(y,z) and d(z,x)=d(z,y). We call the points x,y,z, the internal points of {x,y,z}. By [2, p. 28 and 38], there is a unique map f of the triangle {x,y,z} onto a tripod (a star graph with one vertex w of degree 3, and three vertices x0,y0,z0 of degree one, such that its restriction on each side of the triangle is an isometry, d(x0,w)=d(x,z)=d(x,y), d(y0,w)=d(y,x)=d(y,z) and d(z0,w)=d(z,x)=d(z,y)). Note that d(x,z)=d(x,y)=1/2(d(x,y)+d(x,z)d(y,z)), this quantity is usually denoted by (y|z)x, the Gromov product of y and z with base point x. The triangle {x,y,z} is δ-fine if f(p)=f(q) implies that d(p,q)δ. The space X is δ-fine if every geodesic triangle in X is δ-fine.

    The following is an important result in the theory of hyperbolicity (see, e.g, [2, Proposition 2.21, p.41]):

    Theorem 2.2. Let us consider a geodesic metric space X.

    (1) If X is δ-hyperbolic, then it is 4δ-fine.

    (2) If X is δ-fine, then it is δ-hyperbolic.

    Given a geodesic metric space X and a geodesic triangle T in X, let us define the fine constant of T as δfine(T):=inf{δ:T isδfine}, and the fine constant of X as

    δfine(X):=sup{δfine(T):T is a geodesic triangle in X}.

    In this section, we deal with triangulations contained in the Euclidean plane.

    A Voronoi diagram V(S) of a set S of points in the plane is a partition of the plane into regions {Vs}sS, called Voronoi cells, each corresponding to a point s in S, such that for each sS, every point within its corresponding region Vs is closer to s than to any other point of S.

    We say that SR2 is a set in general position if it is not contained in a straight line, it contains at least three points and if no four points of S belongs to some circle. S is locally finite if any Euclidean ball in R2 contains just a finite amount of points in S.

    Although Voronoi diagrams can be defined for any set S in any metric space, the more usual framework is R2 with a finite set S in general position. We consider along this work locally finite sets S in general position in R2.

    In this paper, we study the dual graph T(S) of the Voronoi diagram V(S): Consider the straight-line embedding of T(S), where the vertex corresponding to the Voronoi cell Vs is the point s, and the edge connecting the vertices s1 and s2 is the Euclidean segment s1s2 (with its Euclidean length) when Vs1 and Vs2 share a side. We call this embedding the Delaunay graph T(S) of S. Note that this definition makes sense since S is a locally finite set, although T(S) is an infinite graph when S is an infinite set.

    Recall that the convex hull CH(P) of PRd is the intersection of all convex sets containing P. Thus,

    CH(P)=n=2{nj=1λjpj:pjP,λj0,nj=1λj=1},

    since this union is a convex set, and every convex set containing P must contain every point nj=1λjpj with pjP, λj0 and nj=1λj=1.

    In [23, pp. 206–207], the following results are found for finite sets. The arguments in their proofs allow to prove the same conclusions for locally finite sets.

    Proposition 3.1. Let S be a locally finite set in general position in R2.

    (1) Every vertex of the Voronoi diagram V(S) is the common intersection of exactly three edges of the diagram.

    (2) Every vertex v in the Voronoi diagram V(S) is the center of a circumference C(v) defined by three points of S, and the Voronoi graph (the 1-skeleton of V(S)) is regular of degree three.

    (3) For each vertex v in the Voronoi diagram, the circumference C(v) contains no other point of S.

    (4) Every nearest neighbor of sS defines an edge of the Voronoi polygon Vs.

    (5) The polygon Vs is unbounded if and only if s is a point on the boundary of the convex hull CH(S) of the set S.

    The following theorem of Delaunay [32] shows the importance of the Delaunay graph.

    Theorem 3.2. Let S be a finite set in general position in R2. Then the Delaunay graph T(S) is a triangulation of the convex hull CH(S) of S.

    The Delaunay graph T(S), which can be written also as

    {T(v):v is a Voronoi vertex},

    is a triangulation of CH(S), where T(v) denotes the triangle in T(S) associated to the Voronoi vertex v, see [23, pp. 209–210]. Note that item (2) in Proposition 3.1 gives that the Voronoi graph is regular of degree three.

    The following fact is direct when S is a finite set, as is the case in Theorem 3.2.

    Proposition 3.3. Let S be a locally finite set in general position in R2. Then, for each x,qCH(S) there exists just a finite amount of Voronoi vertices v1,,vk, such that T(vj) intersects the Euclidean segment xq.

    Proof. Since x,qCH(S), there exist two convex polygons Px,Pq with points in S and such that xPx,qPq. The third item in Proposition 3.1 gives that for each vertex v in the Voronoi diagram, the circumference C(v) contains no other point of S. The result follows from these facts since the Euclidean segment xq is a compact set and S is a locally finite set.

    Proposition 3.3 is a useful technical result, since an Euclidean ball can intersect infinitely many triangles of a Delaunay graph, as the following example shows:

    Example 3.4. If

    S={(0,1)}{(k,0):kZ+},

    then any ball centered at (0,1) intersects infinitely many triangles in T(S).

    The extended results for locally finite set of points in general position in Proposition 3.1, Proposition 3.3 and the argument in the proof of Theorem 3.2 in [23, pp. 209-210] give the following result.

    Proposition 3.5. Let S be a locally finite set in general position in R2. Then the Delaunay graph T(S) is a triangulation of the convex hull CH(S) of S.

    By Theorem 3.2 and Proposition 3.5, from now on we call Delaunay triangulation given by S to the Delaunay graph T(S).

    Note that if T(S) is a triangulation of R2, then Proposition 3.5 gives that CH(S)=R2 and so, S is not a finite set.

    Recall that if V0 is a subset of the vertices of a graph G, the subgraph G0 induced by V0 is the subgraph of G with vertices V0 and edges {uvE(G):u,vV0}.

    Lemma 3.6. Let S be a locally finite set in general position in R2 and S1S. Then the subgraph of T(S) induced by S1 is also a subgraph of T(S1).

    Proof. Assume that p,qS1 and pqE(T(S)). Thus, the Voronoi cells of p and q in V(S) share a Voronoi edge e, and

    |xp|=|xq||xs|

    for every xe and sS. Therefore,

    |xp|=|xq||xs|

    for every xe and sS1, and so, e is contained in the Voronoi cells of p and q in V(S1). This implies that pqE(T(S1)). Hence, the subgraph of T(S) induced by S1 is also a subgraph of T(S1).

    In [33] appears the following result.

    Lemma 3.7. Given a set S of n points in the plane, for any two points p,qS,

    dT(S)(p,q)|pq|2π3cosπ62.42,

    independent of S and n.

    Next, we prove a similar result for locally finite sets.

    Theorem 3.8. Let S be a locally finite set in general position in R2. Then, for any two points p,qS,

    dT(S)(p,q)|pq|2π3cosπ62.42.

    Proof. Fix p,qS, and let B1 be the Euclidean closed ball with center p and radius dT(S)(p,q). Note that any geodesic in T(S) joining p and q is contained in B1, and so, qB1.

    Since S is a locally finite set, S1=SB1 is a finite set. Let us define r=max{dT(S)(p,v):vS1}, B2 the Euclidean closed ball with center p and radius r, and S2=SB2.

    Let T be the set of triangles of T(S) containing at least an edge in B2. Since T(S) is a triangulation of CH(S) by Proposition 3.5, each edge in T(S) belongs at most to two triangles in T(S), and so, T is a finite set of triangles. Let S3 be the union of S2 and the set of vertices of the triangles in T.

    Lemma 3.6 gives that the subgraph Γ of T(S) induced by S3 (which includes any geodesic in T(S) joining p and any point in S1) is also a subgraph of T(S3). Hence, dT(S3)(p,q)=dT(S)(p,q).

    Seeking a contradiction assumes that there exists an edge uvE(T(S3))E(T(S)) with u,vS1. Since ΓB2 includes any geodesic in T(S) joining p and any point in S1, there exists a curve g in ΓB2 joining u and v. Since Γ is a subgraph of the planar graph T(S3), we have uvg={u,v}.

    Let F be the compact set in R2 bounded by g and uv. Fix an edge e in g. Since FCH(S), there exists a triangle T1 in the Delaunay triangulation T(S) such that e is a side of T1 and the interior of T1 intersects the interior of F. Let s1 be the vertex of T1 in the interior of F (recall that T(S3) is a planar graph). Let g1 be the curve obtained from g by replacing e for the two other sides of the triangle T1, and F1 the compact set in R2 bounded by g1 and uv.

    Fix an edge e1 in g1. Since F1CH(S), there exists a triangle T2 in the Delaunay triangulation T(S) such that e1 is a side of T2 and the interior of T2 intersects the interior of F1. Let s2 be the vertex of T2 in the interior of F1. Let g2 be the curve obtained from g1 by replacing e1 by the two other sides of the triangle T2, and F2 the compact set in R2 bounded by g2 and uv.

    By iterating this argument we obtain an infinite sequence {sn}S contained in the compact set F. This is a contradiction since S is a locally finite set, and so, E(T(S3))B1=E(T(S))B1.

    Seeking for a contradiction assume that dT(S3)(p,q)<dT(S)(p,q). Hence, there exists a geodesic η in T(S3) joining p and q, with L(η)<dT(S)(p,q). Note that η is contained in B1E(T(S3)) and it is not contained in E(T(S)). Therefore, η contains an edge in

    B1(E(T(S3))E(T(S))),

    a contradiction. Thus, dT(S3)(p,q)=dT(S)(p,q).

    Since S is a locally finite set, S3 is a finite set. Thus, the inequality for finite sets in Lemma 3.7 gives

    dT(S)(p,q)|pq|=dT(S3)(p,q)|pq|2π3cosπ6.

    This finishes the proof, since p and q are arbitrary points in S.

    Corollary 3.9. Let S be a locally finite set in general position in R2. If d2 denotes the Euclidean distance, then

    d2(p,q)dT(S)(p,q)2π3cosπ6d2(p,q)

    for any two points p,qS. Consequently, the identity map is a quasi-isometry between the metric spaces (S,dT(S)) and (S,d2).

    In Proposition 3.12 below, we need the following definition. Given a normed vector space (X,), we denote by B(x,r) the ball of radius r centered at x with respect to . If K is a convex set in a normed vector space X, we define

    R(K):=sup{r: there exist a two-dimensional affine subspace LX and xKL with B(x,r)LK}.

    Remark 3.10. Note that if X=R2, then

    R(K)=sup{r:B(x,r)K for some xK }.

    Example 3.11. (1) If K0 is any convex set in Rn1X=Rn, then

    R(K0×[0,))<,R(K0×R)<.

    (2) If

    K={(x,y)R2:1x1,y1x2},

    then R(K)=1<.

    The following technical result is interesting by itself.

    Proposition 3.12. If K is a convex set in a normed finite-dimensional vector space X, then δfine(K)2R(K) and R(K)4δ(K)/3, and so, K is hyperbolic if and only if R(K)<.

    Proof. Let T be a geodesic triangle in K. Since K is a convex set in X, T is also a geodesic triangle in X. Let r be the radius of the maximum inscribed circle in T, and x its center. Let L be the two-dimensional affine subspace of X containing T. Thus, this inscribed circle to T, i.e., B(x,r)L, is contained in K and

    δfine(T)2r2R(K).

    Hence, δfine(K)2R(K).

    If B(x,r)LK for some two-dimensional affine subspace LX and xKL, then let T be an equilateral triangle such that B(x,r) is its circumcircle; thus, T is contained in KL. The length of any side of T is 3r and δ(T) is the distance of the midpoint of any side of T to the union of the other sides, i.e., δ(T)=3r/4. Therefore, δ(K)δ(T)=3r/4 and so, δ(K)3R(K)/4.

    Although most of the graphs embedded in R2 are not hyperbolic, there are many hyperbolic graphs, as the Cayley graphs of the groups Z and Z×Zm (mZ+) embedded in R2. Also, one of the authors shows in [24] a hyperbolic tessellation of the Euclidean plane with convex tiles. The following result characterizes the Delaunay triangulations contained in the Euclidean plane that are hyperbolic.

    For each locally finite set S in general position in R2, let us define

    (S):=sup{L(e):eE(T(S))}<.

    Theorem 3.13. Let S be a locally finite set in general position in R2. Then, T(S) is hyperbolic if and only if R(CH(S))< and (S)<.

    Proof. Assume first that R(CH(S))< and (S)<. Since R(CH(S))<, Proposition 3.12 gives that CH(S), with its induced Euclidean metric, is hyperbolic.

    Let us consider the inclusion i:T(S)CH(S)R2. It is clear that |xy|dT(S)(x,y) for every x,yT(S). Fix x,yT(S). Let p and q be two vertices in S with dT(S)(x,p)(S)/2 and dT(S)(y,q)(S)/2. If we define

    c0:=2π3cosπ6,

    then Theorem 3.8 gives

    dT(S)(x,y)dT(S)(x,p)+dT(S)(p,q)+dT(S)(q,y)(S)+c0|pq|(S)+c0(|xy|+(S)).

    Hence,

    1c0dT(S)(x,y)c0+1c0(S)|xy|dT(S)(x,y),

    for every x,yT(S), and the inclusion is a (S)-full (c0,(c0+1)(S)/c0)-quasi-isometry. Since CH(S) is hyperbolic, Theorem 2.1 gives that T(S) is hyperbolic.

    Assume now that T(S) is hyperbolic. Given any edge e in T(S), let us choose a triangle T0={e,e1,e2} in the Delaunay triangulation T(S). Note that T0 is a geodesic triangle in T(S). If x is the midpoint of e, then

    L(e)2=dT(S)(x,e1e2)δ(T0)δ(T(S)),

    and so, (S)2δ(T(S))<. We have proved that the inclusion i:T(S)CH(S) is a (S)-full (c0,(c0+1)(S)/c0)-quasi-isometry between T(S) and CH(S). Since T(S) is hyperbolic, Theorem 2.1 gives that CH(S) is hyperbolic. Finally, Proposition 3.12 gives R(CH(S))<.

    The following result shows that the Delaunay triangulations of the Euclidean plane are not hyperbolic.

    Corollary 3.14. Let S be a locally finite set in general position in R2 such that T(S) is a triangulation of the Euclidean plane. Then T(S) is not hyperbolic.

    Proof. Since T(S) is a triangulation of R2, we have CH(S)=R2 and, since R(CH(S))=R(R2)=, Theorem 3.13 gives that T(S) is not hyperbolic.

    In conclusion, this work investigates the concept of hyperbolicity in geodesic metric spaces, with a specific focus on Delaunay triangulations within the Euclidean plane. The study presented here explores the hyperbolicity of Delaunay triangulations and provides insights into the conditions under which Delaunay triangulations in the Euclidean plane are hyperbolic. This research contributes to our understanding of geometric structures and their hyperbolicity, in Gromov sense.

    Note that part of the arguments in the proofs of the results in this paper work for n-dimensional space. It therefore seems natural to raise the open problem of generalising our results to dimension n.

    The authors declare that they have not used Artificial Intelligence (AI) tools in the creation of this article.

    This publication is part of the grant PID2019-106433GB-I00 funded by Agencia Estatal de Investigación MCIN/AEI/ 10.13039/501100011033. The research of Jose M. Rodríguez is supported by a grant from the Madrid Government (Comunidad de Madrid-Spain) under the Multiannual Agreement with UC3M in the line of Excellence of University Professors (EPUC3M23), and in the context of the V PRICIT (Regional Programme of Research and Technological Innovation).

    We would like to thank the reviewers by their careful reading of the manuscript and their suggestions which have improved the presentation of this work.

    The authors declare that there are no conflicts of interest. Prof. Jose M. Rodriguez is a Guest Editor for AIMS Mathematics and was not involved in the editorial review or the decision to publish this article. All authors declare that there are no competing interests.



    [1] J. Alonso, T. Brady, D. Cooper, T. Delzant, V. Ferlini, M. Lustig, et al., Notes on word hyperbolic groups, in: E. Ghys, A. Haefliger, A. Verjovsky (Eds.), Group Theory from a Geometrical Viewpoint, World Scientific, Singapore, 1992.
    [2] E. Ghys, P. Harpe, Sur les Groupes Hyperboliques d'après Mikhael Gromov, Progress in Mathematics 83, Birkhäuser Boston Inc., Boston, MA, 1990. https://doi.org/10.1007/978-1-4684-9167-8
    [3] M. Gromov, Hyperbolic groups, in "Essays in group theory", Edited by S. M. Gersten, M. S. R. I. Publ., Springer, New York, NY, 1987, 75–263. https://doi.org/10.1007/978-1-4613-9586-7-3
    [4] K. Oshika, Discrete groups, AMS Bookstore, 2002.
    [5] R. Charney, Artin groups of finite type are biautomatic, Math. Ann., 292 (1992), 671–683. https://doi.org/10.1007/BF01444642 doi: 10.1007/BF01444642
    [6] E. Tourís, Graphs and Gromov hyperbolicity of non-constant negatively curved surfaces, J. Math. Anal. Appl., 380 (2011), 865–881. https://doi.org/10.1016/j.jmaa.2011.02.067 doi: 10.1016/j.jmaa.2011.02.067
    [7] R. Krauthgamer, J. R. Lee, Algorithms on negatively curved spaces, FOCS 2006.
    [8] Y. Shang, Lack of Gromov-hyperbolicity in colored random networks, Pan-American Math. J., 21 (2011), 27–36.
    [9] Y. Shang, Lack of Gromov-hyperbolicity in small-world networks, Cent. Eur. J. Math., 10 (2012), 1152–1158. https://doi.org/10.2478/s11533-012-0032-8 doi: 10.2478/s11533-012-0032-8
    [10] Y. Shang, Non-hyperbolicity of random graphs with given expected degrees, Stoch. Models, 29 (2013), 451–462. https://doi.org/10.1080/15326349.2013.838510 doi: 10.1080/15326349.2013.838510
    [11] K. Verbeek, S. Suri, Metric embeddings, hyperbolic space and social networks, Discret. Math., 59 (2016), 1–12. https://doi.org/10.1016/j.comgeo.2016.08.003 doi: 10.1016/j.comgeo.2016.08.003
    [12] M. Abu-Ata, F. F. Dragan, Metric tree-like structures in real-life networks: An empirical study, Networks, 67 (2016), 49–68. https://doi.org/10.1002/net.21631 doi: 10.1002/net.21631
    [13] Y. Shavitt, T. Tankel, On internet embedding in hyperbolic spaces for overlay construction and distance estimation, INFOCOM 2004.
    [14] D. Krioukov, F. Papadopoulos, M. Kitsak, A. Vahdat, M. Boguña, Hyperbolic geometry of complex networks, Phys. Rev. E, 82 (2010), 036106. https://doi.org/10.1103/physreve.82.036106 doi: 10.1103/physreve.82.036106
    [15] E. Jonckheere, Contrôle du traffic sur les réseaux à géométrie hyperbolique–Vers une théorie géométrique de la sécurité l'acheminement de l'information, J. Europ. Syst. Autom., 8 (2002), 45–60.
    [16] E. Jonckheere, P. Lohsoonthorn, Geometry of network security, Amer. Control Conf., ACC (2004), 111–151.
    [17] O. Narayan, I. Saniee, Large-scale curvature of networks, Phys. Rev. E, 84 (2011), 066108. https://doi.org/10.1103/physreve.84.066108 doi: 10.1103/physreve.84.066108
    [18] F. Aurenhammer, Voronoi diagrams—a survey of a fundamental geometric data structure, ACM Comput. Surv., 3 (1991), 345–405. https://doi.org/10.1145/116873.116880 doi: 10.1145/116873.116880
    [19] F. Aurenhammer, R. Klein, Voronoi diagrams, In Handbook of computational geometry, (J. Sack and G. Urrutia, Eds.), North-Holland, Amsterdam, 2000,201–290.
    [20] G. Voronoi, Nouvelles applications des parametres continus à la theorie des formes quadratiques, J. Reine. Angew. Math., 134 (1908), 198–287. https://doi.org/10.1515/crll.1908.134.198 doi: 10.1515/crll.1908.134.198
    [21] R. Sedgewick, J. Vitter, Shortest paths in Euclidean graphs, Algorithmica, 1 (1986), 31–48. https://doi.org/10.1109/sfcs.1984.715943 doi: 10.1109/sfcs.1984.715943
    [22] P. Chew, There is a planar graph almost as good as the complete graph, In Proceedings of the Second Symposium on Computational Geometry, Yorktown Heights, NY, 1986, pp. 169–177. https://doi.org/10.1145/10515.10534
    [23] F. P. Preparata, M. I. Shamos, Computational Geometry: An Introduction, Springer-Verlag, New York, 1985. https://doi.org/10.1007/978-1-4612-1098-6
    [24] W. Carballosa, Gromov hyperbolicity and convex tessellation graph, Acta Math. Hungar., 151 (2017), 24–34. https://doi.org/10.1007/s10474-016-0677-z doi: 10.1007/s10474-016-0677-z
    [25] B. H. Bowditch, Notes on Gromov's hyperbolicity criterion for path-metric spaces in Group theory from a geometrical viewpoint, ed. E. Ghys, A. Haefliger and A. Verjovsky; World Scientific, River Edge, NJ, 1991, 64–167.
    [26] J. C. Hernández, G. Reyna, O. Rosario, Results on hyperbolicity in graphs: a survey, Discret. Math. Lett., 4 (2011), 60–72.
    [27] Á. Martínez, Quasi-isometries between visual hyperbolic spaces, Manus. Math., 137 (2012), 195–213 https://doi.org/10.1007/s00229-011-0463-8 doi: 10.1007/s00229-011-0463-8
    [28] S. Bermudo, J. M. Rodríguez, J. M. Sigarreta, J. M. Vilaire, Gromov hyperbolic graphs, Discret. Math., 313 (2013), 1575–1585. https://doi.org/10.1016/j.disc.2013.04.009 doi: 10.1016/j.disc.2013.04.009
    [29] A. Cantón, A. Granados, D. Pestana, J. M. Rodríguez, Gromov hyperbolicity of periodic graphs, Acta. Math. Sin. English Ser., 30 (2014), 79–90. https://doi.org/10.1007/s10114-013-2370-2 doi: 10.1007/s10114-013-2370-2
    [30] J. M. Rodríguez, J. M. Sigarreta, Y. Torres-Nuñez, Computing the hyperbolicity constant of a cubic graph, Int. J. Comput Math., 91 (2014), 1897–1910.
    [31] Y. Wu, C. Zhang, Chordality and hyperbolicity of a graph, Electr. J. Comb., 18 (2011), 43. https://doi.org/10.37236/530 doi: 10.37236/530
    [32] D. Delaunay, Sur la sphère vide: A la mémoire de Georges Voronoï, Bull. Acad. Sci. URSS, Classe Sci. Math. Natur., 6 (1934), 793–800.
    [33] J. M. Keil, C. A. Gutwin, Classes of graphs which approximate the Complete Euclidean Graph, Discr. Comput. Geom., 7 (1992), 13–28. https://doi.org/10.1007/BF02187821 doi: 10.1007/BF02187821
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1261) PDF downloads(64) Cited by(0)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog