Research article Special Issues

Angiotensin II, dopamine and nitric oxide. An asymmetrical neurovisceral interaction between brain and plasma to regulate blood pressure

  • Received: 10 June 2019 Accepted: 24 July 2019 Published: 26 July 2019
  • Vital functions, such as blood pressure, are regulated within a framework of neurovisceral integration in which various factors are involved under normal conditions maintaining a delicate balance. Imbalance of any of these factors can lead to various pathologies. Blood pressure control is the result of the balanced action of central and peripheral factors that increase or decrease. Special attention for blood pressure control was put on the neurovisceral interaction between Angiotensin II and the enzymes that regulate its activity as well as on nitric oxide and dopamine. Several studies have shown that such interaction is asymmetrically organized. These studies suggest that the neuronal activity related to the production of nitric oxide in plasma is also lateralized and, consequently, changes in plasma nitric oxide influence neuronal function. This observation provides a new aspect revealing the complexity of the blood pressure regulation and, undoubtedly, makes such study more motivating as it may affect the approach for treatment.

    Citation: I. Banegas, I. Prieto, A.B. Segarra, M. Martínez-Cañamero, M. de Gasparo, M. Ramírez-Sánchez. Angiotensin II, dopamine and nitric oxide. An asymmetrical neurovisceral interaction between brain and plasma to regulate blood pressure[J]. AIMS Neuroscience, 2019, 6(3): 116-127. doi: 10.3934/Neuroscience.2019.3.116

    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] Rashid Ali, Rabia Imtiaz, Moin-ud-Din Junjua, Fuad A. Awwad, Emad A. A. Ismail, Ahmed S. Hendy . Multiplicity of kink and hump structures in Van der Waals gas system. AIMS Mathematics, 2025, 10(5): 12493-12518. doi: 10.3934/math.2025564
    [6] 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
    [7] 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
    [8] 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
    [9] 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
    [10] 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
  • Vital functions, such as blood pressure, are regulated within a framework of neurovisceral integration in which various factors are involved under normal conditions maintaining a delicate balance. Imbalance of any of these factors can lead to various pathologies. Blood pressure control is the result of the balanced action of central and peripheral factors that increase or decrease. Special attention for blood pressure control was put on the neurovisceral interaction between Angiotensin II and the enzymes that regulate its activity as well as on nitric oxide and dopamine. Several studies have shown that such interaction is asymmetrically organized. These studies suggest that the neuronal activity related to the production of nitric oxide in plasma is also lateralized and, consequently, changes in plasma nitric oxide influence neuronal function. This observation provides a new aspect revealing the complexity of the blood pressure regulation and, undoubtedly, makes such study more motivating as it may affect the approach for treatment.


    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 $ \mathbb{R}^2 $ such that every point in $ \mathbb{R}^2 $ 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(N \log N) $ 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 $ \gamma: [a, b] \rightarrow X $ is a geodesic if $ L(\gamma|_{[t, s]}) = d(\gamma(t), \gamma(s)) = s-t $ for every $ s, t\in [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 $ uv\in E(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 $ x_1, x_2, x_3 $ are points in $ X $, then the union of three geodesics $ [x_1 x_2] $, $ [x_2 x_3] $ and $ [x_3 x_1] $ is a geodesic triangle, and we denote it by $ T = \{x_1, x_2, x_3\} $ or $ T = \{[x_1x_2], [x_2x_3], [x_3x_1]\} $. The geodesic triangle $ T $ is $ \delta $-thin if any side of $ T $ is contained in the $ \delta $-neighborhood of the union of the two other sides. We denote by $ \delta(T) $ the sharp thin constant of $ T $, i.e., $ \delta(T): = \inf\{ \delta\ge 0: \, T \, \text{ is }\delta -\text{thin}\, \}. $ We say that the space $ X $ is $ \delta $-hyperbolic $ ($or satisfies the Rips condition with constant $ \delta) $ if every geodesic triangle in $ X $ is $ \delta $-thin. We will denote by $ \delta(X) $ the sharp hyperbolicity constant of $ X $, i.e., $ \delta(X): = \sup\{ \delta(T): \, T \, \text{ is a geodesic triangle in}\, X\, \}. $ We say that $ X $ is hyperbolic if $ X $ is $ \delta $-hyperbolic for some $ \delta \ge 0 $, i.e., if $ \delta(X) < \infty $.

    Given any graph $ G $ and $ x, y \in G $, we define the distance $ d_G(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, d_X) $ and $ (Y, d_Y) $, a map $ f: X\longrightarrow Y $ is said to be an $ (\alpha, \beta) $-quasi-isometric embedding, with constants $ \alpha\geq 1, \ \beta\geq 0, $ if the inequalities

    $ \alpha^{-1}d_X(x,y)-\beta\leq d_Y(f(x),f(y))\leq \alpha d_X(x,y)+\beta. $

    hold for every $ x, y\in X $. We say that $ f $ is $ \varepsilon $-full if for each $ y \in Y $ there exists $ x\in X $ with $ d_Y(f(x), y)\leq \varepsilon $. The map $ f $ is a quasi-isometry if there exist constants $ \alpha\geq 1, \ \beta, \varepsilon \geq 0 $ such that $ f $ is an $ \varepsilon $-full $ (\alpha, \beta) $-quasi-isometric embedding.

    A fundamental property of hyperbolic spaces is the following:

    Theorem 2.1 (Invariance of hyperbolicity). Let $ f: X \longrightarrow Y $ be an $ (\alpha, \beta) $-quasi-isometric embedding between the geodesic metric spaces $ X $ and $ Y $. If $ Y $ is hyperbolic, then $ X $ is hyperbolic.

    Furthermore, if $ f $ is $ \varepsilon $-full for some $ \varepsilon\ge0 $ (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 $ T_E $ 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 $ T_E $ by the same letters. The maximum inscribed circle in $ T_E $ 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 $ x_0, y_0, z_0 $ of degree one, such that its restriction on each side of the triangle is an isometry, $ d(x_0, w) = d(x, z') = d(x, y') $, $ d(y_0, w) = d(y, x') = d(y, z') $ and $ d(z_0, 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 $ \delta $-fine if $ f(p) = f(q) $ implies that $ d(p, q)\le \delta $. The space $ X $ is $ \delta $-fine if every geodesic triangle in $ X $ is $ \delta $-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 $ \delta $-hyperbolic, then it is $ 4 \delta $-fine.

    (2) If $ X $ is $ \delta $-fine, then it is $ \delta $-hyperbolic.

    Given a geodesic metric space $ X $ and a geodesic triangle $ T $ in $ X $, let us define the fine constant of $ T $ as $ \delta_{fine}(T): = \inf \big\{ \delta: \, T \text{ is} \delta -\text{fine} \big\}, $ and the fine constant of $ X $ as

    $ \delta_{fine}(X): = \sup \big\{ \delta_{fine}(T): \, T \text{ is a geodesic triangle in } X \big\}. $

    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 $ \{V_s\}_{s \in S} $, called Voronoi cells, each corresponding to a point $ s $ in $ S $, such that for each $ s \in S $, every point within its corresponding region $ V_s $ is closer to $ s $ than to any other point of $ S $.

    We say that $ S \subset \mathbb{R}^2 $ 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 $ \mathbb{R}^2 $ 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 $ \mathbb{R}^2 $ with a finite set $ S $ in general position. We consider along this work locally finite sets $ S $ in general position in $ \mathbb{R}^2 $.

    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 $ V_s $ is the point $ s $, and the edge connecting the vertices $ s_1 $ and $ s_2 $ is the Euclidean segment $ s_1s_2 $ (with its Euclidean length) when $ V_{s_1} $ and $ V_{s_2} $ 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 $ P \subset \mathbb{R}^d $ is the intersection of all convex sets containing $ P $. Thus,

    $ CH(P) = \bigcup\limits_{n = 2}^{\infty} \Big\{ \sum\limits_{j = 1}^{n} \lambda_j p_j : \,\, p_j \in P, \,\, \lambda_j \ge 0, \,\, \sum\limits_{j = 1}^{n} \lambda_j = 1 \Big\}, $

    since this union is a convex set, and every convex set containing $ P $ must contain every point $ \sum_{j = 1}^{n} \lambda_j p_j $ with $ p_j \in P $, $ \lambda_j \ge 0 $ and $ \sum_{j = 1}^{n} \lambda_j = 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 $ \mathbb{R}^2 $.

    (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 $ s \in S $ defines an edge of the Voronoi polygon $ \partial V_s $.

    (5) The polygon $ \partial V_s $ 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 $ \mathbb{R}^2 $. 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 \,\text{ 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 $ \mathbb{R}^2 $. Then, for each $ x, q \in CH(S) $ there exists just a finite amount of Voronoi vertices $ v_1, \dots, v_k, $ such that $ T(v_j) $ intersects the Euclidean segment $ xq $.

    Proof. Since $ x, q \in CH(S) $, there exist two convex polygons $ P_x, P_q $ with points in $ S $ and such that $ x\in P_x, q\in P_q $. 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)\}\bigcup \big\{(k,0):\, k \in \mathbb{Z}^+ \big\}, $

    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 $ \mathbb{R}^2 $. 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 $ \mathbb{R}^2 $, then Proposition 3.5 gives that $ CH(S) = \mathbb{R}^2 $ and so, $ S $ is not a finite set.

    Recall that if $ V_0 $ is a subset of the vertices of a graph $ G $, the subgraph $ G_0 $ induced by $ V_0 $ is the subgraph of $ G $ with vertices $ V_0 $ and edges $ \{ uv \in E(G):\, u, v \in V_0\} $.

    Lemma 3.6. Let $ S $ be a locally finite set in general position in $ \mathbb{R}^2 $ and $ S_1 \subseteq S $. Then the subgraph of $ T(S) $ induced by $ S_1 $ is also a subgraph of $ T(S_1) $.

    Proof. Assume that $ p, q \in S_1 $ and $ pq \in E(T(S)) $. Thus, the Voronoi cells of $ p $ and $ q $ in $ V(S) $ share a Voronoi edge $ e $, and

    $ |x-p| = |x-q| \le |x-s| $

    for every $ x \in e $ and $ s\in S $. Therefore,

    $ |x-p| = |x-q| \le |x-s| $

    for every $ x \in e $ and $ s\in S_1 $, and so, $ e $ is contained in the Voronoi cells of $ p $ and $ q $ in $ V(S_1) $. This implies that $ pq \in E(T(S_1)) $. Hence, the subgraph of $ T(S) $ induced by $ S_1 $ is also a subgraph of $ T(S_1) $.

    In [33] appears the following result.

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

    $ \frac{d_{T(S)}(p, q)}{|p-q|} \le \frac{2\pi}{3 \cos \frac{\pi}6} \approx 2.42, $

    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 $ \mathbb{R}^2 $. Then, for any two points $ p, q \in S $,

    $ \frac{d_{T(S)}(p, q)}{|p-q|} \le \frac{2\pi}{3 \cos \frac{\pi}6} \approx 2.42. $

    Proof. Fix $ p, q \in S, $ and let $ B_1 $ be the Euclidean closed ball with center $ p $ and radius $ d_{T(S)}(p, q) $. Note that any geodesic in $ T(S) $ joining $ p $ and $ q $ is contained in $ B_1 $, and so, $ q \in B_1 $.

    Since $ S $ is a locally finite set, $ S_1 = S \cap B_1 $ is a finite set. Let us define $ r = \max \{ d_{T(S)}(p, v) : \, v \in S_1 \} $, $ B_2 $ the Euclidean closed ball with center $ p $ and radius $ r $, and $ S_2 = S \cap B_2 $.

    Let $ \mathfrak{T} $ be the set of triangles of $ T(S) $ containing at least an edge in $ B_2 $. 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, $ \mathfrak{T} $ is a finite set of triangles. Let $ S_3 $ be the union of $ S_2 $ and the set of vertices of the triangles in $ \mathfrak{T} $.

    Lemma 3.6 gives that the subgraph $ \Gamma $ of $ T(S) $ induced by $ S_3 $ (which includes any geodesic in $ T(S) $ joining $ p $ and any point in $ S_1 $) is also a subgraph of $ T(S_3) $. Hence, $ d_{T(S_3)}(p, q) = d_{T(S)}(p, q) $.

    Seeking a contradiction assumes that there exists an edge $ uv \in E(T(S_3)) \setminus E(T(S)) $ with $ u, v \in S_1 $. Since $ \Gamma \cap B_2 $ includes any geodesic in $ T(S) $ joining $ p $ and any point in $ S_1 $, there exists a curve $ g $ in $ \Gamma \cap B_2 $ joining $ u $ and $ v $. Since $ \Gamma $ is a subgraph of the planar graph $ T(S_3) $, we have $ uv \cap g = \{u, v\} $.

    Let $ F $ be the compact set in $ \mathbb{R}^2 $ bounded by $ g $ and $ uv $. Fix an edge $ e $ in $ g $. Since $ F \subseteq CH(S) $, there exists a triangle $ T_1 $ in the Delaunay triangulation $ T(S) $ such that $ e $ is a side of $ T_1 $ and the interior of $ T_1 $ intersects the interior of $ F $. Let $ s_1 $ be the vertex of $ T_1 $ in the interior of $ F $ (recall that $ T(S_3) $ is a planar graph). Let $ g_1 $ be the curve obtained from $ g $ by replacing $ e $ for the two other sides of the triangle $ T_1 $, and $ F_1 $ the compact set in $ \mathbb{R}^2 $ bounded by $ g_1 $ and $ uv $.

    Fix an edge $ e_1 $ in $ g_1 $. Since $ F_1 \subseteq CH(S) $, there exists a triangle $ T_2 $ in the Delaunay triangulation $ T(S) $ such that $ e_1 $ is a side of $ T_2 $ and the interior of $ T_2 $ intersects the interior of $ F_1 $. Let $ s_2 $ be the vertex of $ T_2 $ in the interior of $ F_1 $. Let $ g_2 $ be the curve obtained from $ g_1 $ by replacing $ e_1 $ by the two other sides of the triangle $ T_2 $, and $ F_2 $ the compact set in $ \mathbb{R}^2 $ bounded by $ g_2 $ and $ uv $.

    By iterating this argument we obtain an infinite sequence $ \{s_n\} \subset S $ contained in the compact set $ F $. This is a contradiction since $ S $ is a locally finite set, and so, $ E(T(S_3)) \cap B_1 = E(T(S)) \cap B_1 $.

    Seeking for a contradiction assume that $ d_{T(S_3)}(p, q) < d_{T(S)}(p, q) $. Hence, there exists a geodesic $ \eta $ in $ T(S_3) $ joining $ p $ and $ q $, with $ L(\eta) < d_{T(S)}(p, q) $. Note that $ \eta $ is contained in $ B_1\cap E(T(S_3)) $ and it is not contained in $ E(T(S)) $. Therefore, $ \eta $ contains an edge in

    $ B_1 \cap \big( E(T(S_3)) \setminus E(T(S)) \big), $

    a contradiction. Thus, $ d_{T(S_3)}(p, q) = d_{T(S)}(p, q) $.

    Since $ S $ is a locally finite set, $ S_3 $ is a finite set. Thus, the inequality for finite sets in Lemma 3.7 gives

    $ \frac{d_{T(S)}(p, q)}{|p-q|} = \frac{d_{T(S_3)}(p, q)}{|p-q|} \le \frac{2\pi}{3 \cos \frac{\pi}6} \,. $

    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 $ \mathbb{R}^2 $. If $ d_2 $ denotes the Euclidean distance, then

    $ d_{2}(p, q) \le d_{T(S)}(p, q) \le \frac{2\pi}{3 \cos \frac{\pi}6} \, d_{2}(p, q) $

    for any two points $ p, q \in S $. Consequently, the identity map is a quasi-isometry between the metric spaces $ (S, d_{T(S)}) $ and $ (S, d_2) $.

    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 = \mathbb{R}^2 $, then

    $ R(K) = \sup \big\{ r : \, B(x,r) \subseteq K \text{ for some } x\in K \ \big\} . $

    Example 3.11. (1) If $ K_0 $ is any convex set in $ \mathbb{R}^{n-1} \subset X = \mathbb{R}^n $, then

    $ R\big(K_0 \times [0,\infty) \big) < \infty, \qquad R\big(K_0 \times \mathbb{R} \big) < \infty . $

    (2) If

    $ K = \big\{ (x,y)\in \mathbb{R}^2 : \, -1 \le x \le 1, \,\, y\ge -\sqrt{1-x^2} \; \big\} , $

    then $ R(K) = 1 < \infty $.

    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 $ \delta_{fine}(K) \le 2R(K) $ and $ R(K) \le 4 \delta(K)/3 $, and so, $ K $ is hyperbolic if and only if $ R(K) < \infty $.

    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)\cap L $, is contained in $ K $ and

    $ \delta_{fine}(T) \le 2r \le 2R(K). $

    Hence, $ \delta_{fine}(K) \le 2R(K) $.

    If $ B(x, r)\cap L \subseteq K $ for some two-dimensional affine subspace $ L\subseteq X $ and $ x\in K \cap L $, then let $ T $ be an equilateral triangle such that $ \partial B(x, r) $ is its circumcircle; thus, $ T $ is contained in $ K \cap L $. The length of any side of $ T $ is $ \sqrt{3}\, r $ and $ \delta(T) $ is the distance of the midpoint of any side of $ T $ to the union of the other sides, i.e., $ \delta(T) = 3r/4 $. Therefore, $ \delta(K) \ge \delta(T) = 3r/4 $ and so, $ \delta(K) \ge 3R(K)/4 $.

    Although most of the graphs embedded in $ \mathbb{R}^2 $ are not hyperbolic, there are many hyperbolic graphs, as the Cayley graphs of the groups $ \mathbb{Z} $ and $ \mathbb{Z} \times \mathbb{Z}_m $ $ (m \in \mathbb{Z}^+) $ embedded in $ \mathbb{R}^2 $. 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 $ \mathbb{R}^2 $, let us define

    $ \ell(S) : = \sup \big\{ L(e) : \, e \in E(T(S)) \big\} < \infty . $

    Theorem 3.13. Let $ S $ be a locally finite set in general position in $ \mathbb{R}^2 $. Then, $ T(S) $ is hyperbolic if and only if $ R(CH(S)) < \infty $ and $ \ell(S) < \infty $.

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

    Let us consider the inclusion $ i: T(S) \rightarrow CH(S) \subseteq \mathbb{R}^2 $. It is clear that $ |x-y| \le d_{T(S)}(x, y) $ for every $ x, y \in T(S) $. Fix $ x, y \in T(S) $. Let $ p $ and $ q $ be two vertices in $ S $ with $ d_{T(S)}(x, p) \le \ell(S)/2 $ and $ d_{T(S)}(y, q) \le \ell(S)/2 $. If we define

    $ c_0 : = \frac{2\pi}{3 \cos \frac{\pi}6} \,, $

    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,

    $ \frac1{c_0} \, d_{T(S)}(x,y) - \frac{c_0+1}{c_0}\, \ell(S) \le |x-y| \le d_{T(S)}(x,y), $

    for every $ x, y \in T(S) $, and the inclusion is a $ \ell(S) $-full $ (c_0, (c_0+1) \ell(S)/c_0) $-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 $ T_0 = \{e, e_1, e_2\} $ in the Delaunay triangulation $ T(S) $. Note that $ T_0 $ is a geodesic triangle in $ T(S) $. If $ x $ is the midpoint of $ e $, then

    $ \frac{L(e)}2 = d_{T(S)} (x,e_1 \cup e_2) \le \delta(T_0) \le \delta(T(S)), $

    and so, $ \ell(S) \le 2 \delta(T(S)) < \infty $. We have proved that the inclusion $ i: T(S) \rightarrow CH(S) $ is a $ \ell(S) $-full $ (c_0, (c_0+1) \ell(S)/c_0) $-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)) < \infty $.

    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 $ \mathbb{R}^2 $ 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 $ \mathbb{R}^2 $, we have $ CH(S) = \mathbb{R}^2 $ and, since $ R(CH(S)) = R(\mathbb{R}^2) = \infty $, 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.



    Conflict of interest



    The authors declare no conflicts of interest.

    [1] Oparil S, Acelajado MC, Bakris GL, et al. (2018) Hypertension. Nat Rev Dis Primers 4: 18014. doi: 10.1038/nrdp.2018.14
    [2] Prieto I, Villarejo AB, Segarra AB, et al. (2014) Brain, heart and kidney correlate for the control of blood pressure and water balance: role of angiotensinases. Neuroendocrinology 100: 198–208. doi: 10.1159/000368835
    [3] Prieto I, Segarra AB, Martinez-Canamero M, et al. (2017) Bidirectional asymmetry in the neurovisceral communication for the cardiovascular control: New insights. Endocr Regul 51: 157–167. doi: 10.1515/enr-2017-0017
    [4] Prieto I, Segarra AB, de Gasparo M, et al. (2018) Divergent profile between hypothalamic and plasmatic aminopeptidase activities in WKY and SHR. Influence of beta-adrenergic blockade. Life Sci 192: 9–17.
    [5] Parati G, Ochoa JE, Lombardi C, et al. (2013) Assessment and management of blood-pressure variability. Nat Rev Cardiol 10: 143–155. doi: 10.1038/nrcardio.2013.1
    [6] Banegas I, Prieto I, Vives F, et al. (2006) Brain aminopeptidases and hypertension. J Renin Angiotensin Aldosterone Syst 7: 129–134. doi: 10.3317/jraas.2006.021
    [7] Banegas I, Prieto I, Vives F, et al. (2009) Asymmetrical response of aminopeptidase A and nitric oxide in plasma of normotensive and hypertensive rats with experimental hemiparkinsonism. Neuropharmacology 56: 573–579. doi: 10.1016/j.neuropharm.2008.10.011
    [8] Banegas I, Prieto I, Segarra AB, et al. (2011) Blood pressure increased dramatically in hypertensive rats after left hemisphere lesions with 6-hydroxydopamine. Neurosci Lett 500: 148–150. doi: 10.1016/j.neulet.2011.06.025
    [9] Banegas I, Prieto I, Segarra AB, et al. (2017) Bilateral distribution of enkephalinase activity in the medial prefrontal cortex differs between WKY and SHR rats unilaterally lesioned with 6-hydroxydopamine. Prog Neuropsychopharmacol Biol Psychiatry 75: 213–218. doi: 10.1016/j.pnpbp.2017.02.015
    [10] Prieto I, Segarra AB, Villarejo AB, et al. (2019) Neuropeptidase activity in the frontal cortex of Wistar-Kyoto and spontaneously hypertensive rats treated with vasoactive drugs: a bilateral study. J Hypertens 37: 612–628. doi: 10.1097/HJH.0000000000001884
    [11] Cuspidi C, Tadic M, Grassi G, et al. (2018) Mancia G. Treatment of hypertension: The ESH/ESC guidelines recommendations. Pharmacol Res 128: 315–321.
    [12] Ramírez-Sánchez M, Prieto I, Wangensteen R, et al. (2013) The renin-angiotensin system: new insight into old therapies. Curr Med Chem 20: 1313–1322. doi: 10.2174/0929867311320100008
    [13] Ferdinand KC, Balavoine F, Besse B, et al. (2019) Efficacy and safety of Firibastat, a first-in-class brain aminopeptidase A inhibitor, in hypertensive overweight patients of multiple ethnic origins: A Phase 2, open-label, multicenter, dose-titrating study. Circulation 140: 138–146. doi: 10.1161/CIRCULATIONAHA.119.040070
    [14] Gao J, Marc Y, Iturrioz X, et al. (2014) A new strategy for treating hypertension by blocking the activity of the brain renin-angiotensin system with aminopeptidase A inhibitors. Clin Sci (Lond) 127: 135–148. doi: 10.1042/CS20130396
    [15] Keck M, De Almeida H, Compère D, et al. (2019) NI956/QGC006, a potent orally active, brain-penetrating aminopeptidase a inhibitor for treating hypertension. Hypertension 73: 1300–1307. doi: 10.1161/HYPERTENSIONAHA.118.12499
    [16] Wright JW, Mizutani S, Murray CE, et al. (1990) Aminopeptidase-induced elevations and reductions in blood pressure in the spontaneously hypertensive rat. J Hypertens 8: 969–974. doi: 10.1097/00004872-199010000-00013
    [17] Wright JW, Jensen LL, Cushing LL, et al. (1989) Leucine aminopeptidase M-induced reductions in blood pressure in spontaneously hypertensive rats. Hypertension 13: 910–915. doi: 10.1161/01.HYP.13.6.910
    [18] Zini S, Masdehors P, Lenkei Z, et al. (1997) Aminopeptidase A: distribution in rat brain nuclei and increased activity in spontaneously hypertensive rats. Neuroscience 78: 1187–1193. doi: 10.1016/S0306-4522(96)00660-4
    [19] Banegas I, Prieto I, Segarra AB, et al. (2017) Study of the neuropeptide function in Parkinson's disease using the 6-Hydroxydopamine model of experimental Hemiparkinsonism. AIMS Neuroscience 4: 223–237. doi: 10.3934/Neuroscience.2017.4.223
    [20] Kondo K, Ebihara A, Suzuki H, et al. (1981) Role of dopamine in the regulation of blood pressure and the renin--angiotensin-aldosterone system in conscious rats. Clin Sci (Lond) 61: 235s–237s. doi: 10.1042/cs061235s
    [21] Zeng C, Jose PA (2011) Dopamine receptors: important antihypertensive counterbalance against hypertensive factors. Hypertension 57: 11–17. doi: 10.1161/HYPERTENSIONAHA.110.157727
    [22] Förstermann U, Sessa WC (2012) Nitric oxide synthases: regulation and function. Eur Heart J 33: 829–837. doi: 10.1093/eurheartj/ehr304
    [23] Calver A, Collier J, Vallance P (1993) Nitric oxide and the control of human vascular tone in health and disease. Eur J Med 2: 48–53.
    [24] Venturelli M, Pedrinolla A, Boscolo Galazzo I, et al. (2018) Impact of nitric oxide bioavailability on the progressive cerebral and peripheral circulatory impairments during aging and Alzheimer's disease. Front Physiol 9: 169. doi: 10.3389/fphys.2018.00169
    [25] Steinert JR, Chernova T, Forsythe ID (2010) Nitric oxide signaling in brain function, dysfunction, and dementia. Neuroscientist 16: 435–452. doi: 10.1177/1073858410366481
    [26] Maia-de-Oliveira JP, Trzesniak C, Oliveira IR, et al. (2012) Nitric oxide plasma/serum levels in patients with schizophrenia: a systematic review and meta-analysis. Braz J Psychiatry 34: S149–155.
    [27] Nakano Y, Yoshimura R, Nakano H, et al. (2010) Association between plasma nitric oxide metabolites levels and negative symptoms of schizophrenia: a pilot study. Hum Psychopharmacol 25: 139–144. doi: 10.1002/hup.1102
    [28] Shabeeh H, Khan S, Jiang B, et al. (2017) Blood pressure in healthy humans is regulated by neuronal no synthase. Hypertension 69: 970–976. doi: 10.1161/HYPERTENSIONAHA.116.08792
    [29] Chen ZQ, Mou RT, Feng DX, et al. (2017) The role of nitric oxide in stroke. Med Gas Res 7: 194–203. doi: 10.4103/2045-9912.215750
    [30] Zheng R, Qin L, Li S, et al. (2014) CT perfusion-derived mean transit time of cortical brain has a negative correlation with the plasma level of nitric oxide after subarachnoid hemorrhage. Acta Neurochir (Wien) 156: 527–533. doi: 10.1007/s00701-013-1968-6
    [31] Taffi R, Nanetti L, Mazzanti L, et al. (2008) Plasma levels of nitric oxide and stroke outcome. J Neurol 255: 94–98. doi: 10.1007/s00415-007-0700-y
    [32] Li S, Wang Y, Jiang Z, et al. (2018) Impaired cognitive performance in endothelial nitric oxide synthase knockout mice after ischemic stroke: A pilot study. Am J Phys Med Rehabil 97: 492–499. doi: 10.1097/PHM.0000000000000904
    [33] Bernard C (1878) Etude sur la physiologie du coeur. La science experimentale: 316–366. Available from: https://gallica.bnf.fr/ark:/12148/bpt6k69992b/f316.image.
    [34] Thayer JF, Lane RD (2009) Claude Bernard and the heart-brain connection: further elaboration of a model of neurovisceral integration. Neurosci Biobehav Rev 33: 81–88. doi: 10.1016/j.neubiorev.2008.08.004
    [35] Segarra AB, Prieto I, Banegas I, et al. (2012) Asymmetrical effect of captopril on the angiotensinase activity in frontal cortex and plasma of the spontaneously hypertensive rats: expanding the model of neuroendocrine integration. Behav Brain Res 230: 423–427. doi: 10.1016/j.bbr.2012.02.039
    [36] Segarra AB, Prieto I, Banegas I, et al. (2013) The brain-heart connection: frontal cortex and left ventricle angiotensinase activities in control and captopril-treated hypertensive rats-a bilateral study. Int J Hypertens 2013: 156179.
    [37] Segarra AB, Banegas I, Prieto I, et al. (2016) Brain asymmetry and dopamine: beyond motor implications in Parkinson's disease and experimental hemiparkinsonism. Rev Neurol 63: 415–421.
    [38] de Jager L, Amorim EDT, Lucchetti BFC, et al. (2018) Nitric oxide alterations in cardiovascular system of rats with Parkinsonism induced by 6-OHDA and submitted to previous exercise. Life Sci 204: 78–86. doi: 10.1016/j.lfs.2018.05.017
    [39] Alexander N, Kaneda N, Ishii A, et al. (1990) Right-left asymmetry of tyrosine hydroxylase in rat median eminence: influence of arterial baroreflex nerves. Brain Res 523: 195–198. doi: 10.1016/0006-8993(90)91487-2
    [40] Hersh LB (1985) Characterization of membrane-bound aminopeptidases from rat brain:identification of the enkephalin-degrading aminopeptidase. J Neurochem 44: 1427–1435. doi: 10.1111/j.1471-4159.1985.tb08779.x
    [41] Wright JW, Harding JW (1997) Important roles for angiotensin III and IV in the brain renin-angiotensin system. Brain Res Rev 25: 96–124. doi: 10.1016/S0165-0173(97)00019-2
  • Reader Comments
  • © 2019 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(6152) PDF downloads(641) Cited by(2)

Figures and Tables

Figures(4)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog