Fact-checked by Grok 2 weeks ago

Hyperboloid model

The hyperboloid model is a of two-dimensional realized as a surface in three-dimensional , consisting of the upper sheet of the two-sheeted defined by the equation x^2 + y^2 - z^2 = -1 with z > 0. In this model, points correspond to position vectors on the satisfying the q(X) = -1, while geodesics are the nonempty intersections of the with planes through the origin of . The hyperbolic distance between two points u and v is given by d(u, v) = \arccosh(-\langle u, v \rangle), where \langle \cdot, \cdot \rangle denotes the Minkowski inner product with signature (+, +, -), ensuring the induced Riemannian metric has constant -1. Developed in the late by and , the —sometimes called the Minkowski model due to its embedding in Lorentzian space—offers a symmetric and computationally convenient framework for , despite its embedding in three dimensions making visualization somewhat challenging compared to planar models. Its primary advantages lie in the fact that isometries of the hyperbolic plane correspond directly to linear transformations in the O(2,1) that preserve the , facilitating algebraic computations and connections to , where the hyperboloid represents the of constant . This model is isometric to other standard representations, such as the Poincaré disk (via from the south pole) and the Klein-Beltrami model (via central projection onto the plane z=1), allowing seamless translations between them for different analytical purposes.

Minkowski Space Foundations

Quadratic form

The Minkowski space \mathbb{R}^{n,1} underlying the hyperboloid model is the real vector space \mathbb{R}^{n+1} equipped with coordinates (x_0, x_1, \dots, x_n), where x_0 plays the role of a time-like coordinate and x_1, \dots, x_n are space-like. This space is endowed with the indefinite Q(x) = -x_0^2 + \sum_{i=1}^n x_i^2, which defines the algebraic structure essential for embedding hyperbolic geometry. The associated , known as the Minkowski inner product, is given by \langle u, v \rangle = -u_0 v_0 + \sum_{i=1}^n u_i v_i for vectors u = (u_0, \dots, u_n) and v = (v_0, \dots, v_n). The indefiniteness of Q arises from its (n,1), meaning it takes both positive and negative values, unlike the positive definite form of . This property classifies nonzero vectors based on the sign of Q(x): timelike if Q(x) < 0, spacelike if Q(x) > 0, and lightlike (or null) if Q(x) = 0. Such classifications distinguish causal structures in the geometry, with timelike vectors forming the interior of the defined by the Q(x) = 0. The level sets of Q for a nonzero constant c are hyperboloids in \mathbb{R}^{n+1}. Specifically, for c = -1, the equation Q(x) = -1 describes a two-sheeted , consisting of two connected components separated by the ; the upper sheet, where x_0 > 0, serves as the embedding surface for the hyperboloid model of n-space. For c = 1, the level set Q(x) = 1 yields a one-sheeted , which is connected and lies outside the . These surfaces illustrate the hyperbolic nature of the geometry induced by the indefinite form.

Metric signature conventions

In the hyperboloid model of , the underlying employs an indefinite with one of two primary sign conventions: the mostly minus signature (+,-,-,\dots,-), where the time-like component is positive, or the mostly plus signature (-+,+,\dots,+), where it is negative. These conventions arise from the on \mathbb{R}^{n,1}, and the choice determines the form of the Q used to define the embedding. The mostly plus signature (-+,+,\dots,+) leads to a quadratic form Q(\mathbf{x}) = -x_0^2 + \sum_{i=1}^n x_i^2, with the hyperboloid defined by Q(\mathbf{x}) = -1. Equivalently, this can be expressed as x_0^2 - \sum_{i=1}^n x_i^2 = 1, where the upper sheet corresponds to x_0 > 0, ensuring the position vectors are time-like (negative norm in this signature). In contrast, the mostly minus signature (+,-,-,\dots,-) yields Q(\mathbf{x}) = x_0^2 - \sum_{i=1}^n x_i^2 = 1 directly, again selecting the upper sheet x_0 > 0 for time-like vectors (positive norm here), which aligns the induced metric to be positive definite on the surface. The choice affects the interpretation of time-like versus space-like separations but preserves the overall geometry when consistently applied. Historically, the mostly minus convention (+,-,-,\dots,-) is prevalent in and texts, as it makes proper time intervals positive, facilitating connections to . Conversely, some and literature favors the mostly plus signature (-+,+,\dots,+), emphasizing spatial coordinates and aligning with certain algebraic conventions in Lorentz groups. This preference in geometry texts often stems from treating the model as an abstract embedded in , independent of physical time. Regardless of the signature, normalization is achieved by setting the level of Q to \pm 1, ensuring the induced Riemannian metric on the hyperboloid has constant sectional curvature -1, matching the standard hyperbolic space \mathbb{H}^n. This scaling distinguishes the model from related surfaces like de Sitter space and guarantees isometry with other hyperbolic models, such as the Poincaré disk.

Model Definition

The hyperboloid sheet

The of n-space, denoted H^n, is defined as the set H^n = \{ x = (x_0, x_1, \dots, x_n) \in \mathbb{R}^{n+1} \mid \langle x, x \rangle_L = -1, \, x_0 > 0 \}, where \langle \cdot, \cdot \rangle_L denotes the indefinite inner product on the \mathbb{R}^{1,n} with mostly plus signature: \langle x, y \rangle_L = -x_0 y_0 + \sum_{i=1}^n x_i y_i. This Q(x) = \langle x, x \rangle_L distinguishes the model from embeddings by incorporating the pseudo- structure of . Geometrically, H^n forms the upper (or forward) sheet of the two-sheeted hyperboloid Q(x) = -1, with the two sheets comprising the disconnected components of the surface. The is asymptotic to the Q(x) = 0 in the limit as x_0 \to \infty, where the cone separates the space-like and time-like regions and bounds the model's extent. The selection of the upper sheet x_0 > 0 ensures a connected suitable for modeling , avoiding the lower sheet x_0 < 0 which is isometric but often excluded for convenience in applications. The Riemannian structure on H^n arises as an isometric embedding into Minkowski space, with the metric induced by restricting the ambient Lorentzian inner product to the tangent space at each point x \in H^n. Specifically, for a tangent vector v at x satisfying \langle x, v \rangle_L = 0, the metric is ds^2 = \langle v, v \rangle_L, which is positive definite on the tangent space and inherits the pseudo-Riemannian properties selectively. This induced metric equips H^n with the geometry of a simply connected space-form. The hyperboloid model realizes constant sectional curvature -1, a defining feature of hyperbolic n-space, as verified by direct computation of the Riemann curvature tensor from the embedding or through comparison with other models. This curvature value scales the geometry such that H^n serves as the standard model for \mathbb{H}^n with radius $1$. In ambient coordinates x = (x_0, \dots, x_n), points on H^n satisfy x_0 = \sqrt{1 + \sum_{i=1}^n x_i^2}, linking the time-like coordinate x_0 to the space-like components and encoding hyperbolic distances via the Lorentzian norm. These coordinates facilitate computations involving projections or transformations while preserving the intrinsic hyperbolic properties, such as the base point at (1, 0, \dots, 0).

Induced hyperbolic metric

The induced hyperbolic metric on the hyperboloid sheet H^n = \{ u \in \mathbb{R}^{n+1} \mid \langle u, u \rangle = -1, \, u_0 > 0 \} arises from restricting the Minkowski metric of the ambient space \mathbb{R}^{1,n} with inner product \langle u, v \rangle = -u_0 v_0 + \sum_{i=1}^n u_i v_i to the tangent spaces of H^n. This restriction yields a positive definite Riemannian metric of constant -1, endowing H^n with the standard structure. In geodesic polar coordinates centered at a base point, with radial coordinate r \geq 0 and angular coordinates on the unit sphere S^{n-1}, the line element takes the form ds^2 = dr^2 + \sinh^2 r \, d\theta^2, where d\theta^2 denotes the standard metric on S^{n-1}. This infinitesimal metric measures lengths of tangent vectors on H^n and follows directly from parametrizing the hyperboloid via hyperbolic functions, such as u_0 = \cosh r and the spatial components involving \sinh r times unit vectors. The geodesic distance d(u, v) between distinct points u, v \in H^n is given by d(u, v) = \arccosh(-\langle u, v \rangle), derived by parametrizing the unique geodesic connecting u and v as the intersection of H^n with the plane spanned by u and v in \mathbb{R}^{n+1}. Along this geodesic, the parameter t satisfies \langle \gamma(t), \gamma(t) \rangle = -1 and traces an arc where the Minkowski inner product yields the hyperbolic angle, leading to \cosh t = -\langle u, v \rangle at the endpoints. This formula originates from applying the hyperbolic law of cosines to the right triangle formed by u, v, and the origin in the embedding space, where the "angle" at the origin corresponds to the spatial separation. The distance d(u, v) is real and positive for u \neq v, since -\langle u, v \rangle \geq 1 with equality only when u = v, as ensured by the Cauchy-Schwarz inequality in the Lorentzian metric for future-directed timelike vectors. The d(u, w) \leq d(u, v) + d(v, w) holds due to hyperbolic addition formulas, such as \cosh(a + b) = \cosh a \cosh b + \sinh a \sinh b \geq \cosh(|a - b|), which imply the subadditivity of the function on [1, \infty). This embedding preserves the hyperbolic structure up to , as the hyperboloid model realizes H^n as a Riemannian of \mathbb{R}^{1,n} whose intrinsic geometry matches the unique (up to scaling) of constant curvature -1.

Basic Geometry

Geodesics

In the hyperboloid model of n-space, denoted H^n, are defined as the intersections of the \{ x \in \mathbb{R}^{n,1} : \langle x, x \rangle = -1, x_0 > 0 \} with two-dimensional linear subspaces of the ambient \mathbb{R}^{n,1} that pass through the origin. These subspaces are precisely the planes spanned by a point p \in H^n on the geodesic and a nonzero v \in T_p H^n at that point. Such intersections yield the shortest paths on the , analogous to straight lines in . A unit-speed parametrization of a \gamma passing through an initial point p \in H^n in the direction of a space-like q \in T_p H^n (satisfying \langle p, q \rangle = 0 and \langle q, q \rangle = 1) is given by \gamma(t) = \cosh t \, p + \sinh t \, q, for t \in \mathbb{R}. This curve lies entirely on the , as \langle \gamma(t), \gamma(t) \rangle = -1 holds for all t, and it traces the with the spanned by p and q. The of \gamma from t = 0 to t = s is exactly s, obtained by integrating the induced Riemannian element ds along the path, which yields the hyperbolic distance between \gamma(0) and \gamma(s). Geodesics in this model are complete, extending infinitely in both directions without , and minimizing, providing the shortest between any two points on H^n. For any two distinct points p, q \in H^n, there exists a such connecting them, determined by the they span with the . This uniqueness follows from the of the induced and the simply connected nature of . In the two-dimensional case of H^2, geodesics can be visualized as hyperbolic arcs within specific coordinate planes, such as the (x_1, x_3)-plane where the intersection with the x_1^2 + x_2^2 - x_3^2 = -1, x_3 > 0, yields branches of the x_1^2 - x_3^2 = -1. These curves illustrate the exponential divergence characteristic of , with distances growing rapidly away from the starting point.

Horospheres

In the hyperboloid model of hyperbolic space \mathbb{H}^n, a horosphere is defined as the intersection of the upper hyperboloid sheet \{ x \in \mathbb{R}^{n,1} : \langle x, x \rangle = -1, x_0 > 0 \} with an affine hyperplane tangent to the light cone \{ y \in \mathbb{R}^{n,1} : \langle y, y \rangle = 0 \} at an ideal point on the boundary at infinity. Such ideal points correspond to light-like directions, and the tangency ensures the hyperplane is parallel to the asymptotic direction of geodesics approaching that point. Horospheres exhibit several key properties in this model. They are flat hypersurfaces with zero Gaussian curvature, making them intrinsically isometric to Euclidean space \mathbb{E}^{n-1}. Additionally, each horosphere is equidistant from its defining ideal point, serving as a boundary-parallel surface that foliates \mathbb{H}^n into a family of parallel horospheres approaching the same point at infinity. This foliation covers the entire space without overlap, except at the ideal boundary. The metric induced on a horosphere from the hyperbolic metric is Euclidean, reflecting its flat geometry. However, when viewed extrinsically in the ambient , regions on the horosphere exhibit area growth that is exponential in the hyperbolic distance from a reference point on the horosphere; specifically, for constant curvature -1, the area scales as e^{(n-1)d} where d is the hyperbolic distance along the foliation direction. This exponential expansion underscores the diverging nature of near the boundary. Horospheres are intimately related to Busemann functions in the hyperboloid model. The Busemann function b_\xi : \mathbb{H}^n \to \mathbb{R} associated with an ideal point [\xi](/page/Xi) at is defined as b_\xi(x) = \lim_{t \to \infty} \left( d(x, \gamma(t)) - t \right), where \gamma is a ray from a base point to \xi, and d is the hyperbolic distance. Horospheres centered at \xi are precisely the level sets \{ x \in \mathbb{H}^n : b_\xi(x) = c \} for constants c \in \mathbb{R}, providing a orthogonal to the horospheres. In the case of \mathbb{H}^2, a horosphere in the model appears as the intersection of the hyperboloid sheet with a in \mathbb{R}^{2,1} to the at an ideal point, resulting in a hyperbolic arc that asymptotically approaches the boundary light rays while maintaining the intrinsically. This view emphasizes the horosphere's role as a flat frontier parallel to the ideal boundary, distinct from its curved projections in other models.

Isometries

Lorentz group

The isometries of the hyperboloid model are the linear transformations of \mathbb{R}^{n+1} that preserve the Q(x) = -x_0^2 + \sum_{i=1}^n x_i^2, forming the O(n,1) under the mostly plus convention. These transformations restrict to diffeomorphisms of the H^n = \{ x \in \mathbb{R}^{n+1} : Q(x) = -1, x_0 > 0 \} that preserve the induced Riemannian metric, thereby acting as isometries of the . The group O(n,1) consists of four connected components, distinguished by the sign of the (orientation-preserving or reversing) and the preservation of the forward (time-orientation). The proper orthochronous component, denoted SO^+(n,1), comprises the orientation-preserving isometries that maintain the time direction and acts transitively on H^n, meaning any point in H^n can be mapped to any other via an element of this subgroup. Furthermore, SO^+(n,1) acts transitively on the unit of H^n, allowing any at a point to be mapped to any other of the same length at another point. The Lie algebra of O(n,1) is \mathfrak{so}(n,1), whose infinitesimal generators correspond to spatial rotations in the Euclidean subspaces (compact generators) and Lorentz boosts, which are hyperbolic rotations in the time-space planes (noncompact generators). The exponential map from \mathfrak{so}(n,1) to SO^+(n,1) is surjective, reflecting the connected nature of the proper orthochronous subgroup.

Classification of isometries

The isometries of the can be classified based on their action and fixed points, corresponding to elements of the O(n,1) that preserve the upper sheet of the . Orientation-reversing isometries include reflections, which are orthogonal reflections across orthogonal to space-like vectors in the ambient ; these fix the entire pointwise and reverse . Orientation-preserving isometries fall into three main categories: elliptic, (or boosts), and parabolic. Elliptic isometries, often called rotations, act as rotations in space-like planes within the Minkowski space, generating compact subgroups isomorphic to SO(n); they fix a point in the hyperbolic space and rotate the orthogonal subspace around it. For example, in the hyperboloid model of H^n, such a rotation preserves a geodesic through the fixed point and acts as a Euclidean rotation on the tangent space at that point. Hyperbolic isometries, known as boosts or translations, operate as hyperbolic rotations in time-like planes, displacing points exponentially along a unique invariant (the axis); they have no fixed points in the but fix two ideal points at . In the hyperboloid model of H^2 embedded in with (2,1), a along the x-axis takes the form \begin{pmatrix} \cosh \theta & \sinh \theta & 0 \\ \sinh \theta & \cosh \theta & 0 \\ 0 & 0 & 1 \end{pmatrix} in the basis (x_0, x_1, x_2), where θ parameterizes the translation distance, preserving the Lorentz inner product and mapping the hyperboloid to itself. Parabolic isometries correspond to symmetries of horospheres, fixing exactly one ideal and inducing Euclidean motions (translations and rotations) on the horosphere; they have no fixed points in the and translate along horocycles. In the hyperboloid model, these arise from linear transformations involving null directions in , preserving the structure. Discrete subgroups of these isometries, such as Fuchsian groups in dimension 2 or Poincaré groups in higher dimensions, generate fundamental domains and tessellations in the hyperboloid model, often used to study hyperbolic manifolds.

History

Precursors and early formulations

The foundations of , which underpins the hyperboloid model, trace back to efforts in the 18th and 19th centuries to challenge Euclid's . In 1766, explored a geometry where the sum of angles in a is less than π, conceptualizing it on an "imaginary " with radius involving √(-1), where distances become ; this provided an early analytic precursor to non-Euclidean spaces, though aimed to disprove such geometries via contradiction. The explicit discovery of occurred independently in the early , prompting a search for concrete models to demonstrate its consistency within framework. published the first systematic treatment in 1829, developing trigonometry and axioms for a geometry allowing multiple parallels, while outlined a similar "" in 1832 as an appendix to his father's work. These advances, building on earlier hints like those from and Ferdinand Karl von Prillwitz (as Taurinus in 1826), spurred mathematicians to seek embeddings in higher-dimensional or projective spaces. Wilhelm Killing advanced this quest in the late 1870s and 1880s by embedding the hyperbolic plane in using . Drawing from Karl Weierstrass's 1872 lectures on coordinates, Killing introduced Weierstrass coordinates in 1878–1880 to represent the Lobachevskian plane on a hyperboloid sheet via the quadratic form x² + y² - z² = -1, enabling computations in non-Euclidean spaces; he formalized this in his 1885 monograph Die Nicht-Euklidischen Raumformen, treating the hyperboloid as a surface of constant negative . Henri Poincaré independently formulated the hyperboloid model around 1880–1881 in unpublished notes, later publishing in 1881 on the invariance of hyperbolic structures under quadratic forms and linking it to Fuchsian groups in his 1882 paper on automorphic functions. These works connected the model to projective geometry and complex analysis, projecting the hyperboloid onto disk representations for visualizing group actions. Contemporary contributions further refined coordinate systems and applications. In 1882, Homersham Cox developed homogeneous coordinates for "imaginary geometry," applying them to force systems and implicitly supporting hyperboloid embeddings. Alfred Clebsch and Ferdinand Lindemann, in their 1891 edition of Clebsch's lectures, discussed quadratic relations like x₁² + x₂² - 4k² x₃² = -4k², exposing the model's projective properties and metric implications. Alexander Macfarlane, in his 1894 Papers on Space Analysis, treated the hyperboloid explicitly as a metric space, deriving the hyperbolic law of cosines (cosh c = cosh a cosh b - sinh a sinh b cos C) from versor algebra and area ratios, emphasizing its utility for angle definitions in non-Euclidean contexts.

Modern interpretations

In the early 20th century, the gained prominence through its integration into the framework of , particularly via Hermann Minkowski's formulation of in 1907. Minkowski linked the hyperboloid to the future in , representing the space of four-velocities where points on the hyperboloid correspond to observers with constant , providing a geometric of relativistic . Building on this, H. Jansen explicitly focused on the hyperboloid as a model for in his 1909 paper, offering the first detailed exposition of its properties for the two-dimensional case by deriving geodesics, distances, and angles directly from the metric on the hyperboloid sheet. This work emphasized the model's advantages in computations involving the , bridging pure geometry with relativistic applications. Varićak further advanced these ideas in 1912 by applying the hyperboloid model to relativistic , reinterpreting velocity addition and transformations using on the model to avoid complex numbers and align with non-Euclidean interpretations of Lorentz boosts. Post-relativity developments included historical analyses such as W.F. Reynolds' 1993 recount, which traced the model's evolution from 19th-century precursors to its relativistic embeddings while highlighting its computational tractability. The model also integrates with other hyperbolic representations, such as the , through projective mappings that preserve the metric up to conformal factors, enabling transitions between models for visualization and analysis. In contemporary applications, the hyperboloid model facilitates the rendering of hyperbolic tilings in , where its embedding in allows efficient generation of infinite tessellations projected onto Euclidean displays for games and visualizations. Additionally, in , the forward hyperboloid serves as a quantization surface in point-form approaches, enabling Lorentz-invariant formulations of scalar fields on spacetimes with compactly generated Cauchy horizons. These uses illustrate the model's evolution from a geometric to a versatile tool in physics and computation.

References

  1. [1]
    [PDF] Hyperboloid Model Students - FSU Math
    Aug 25, 2017 · The Hyperboloid Model uses one sheet of a hyperboloid, with lines as intersections of planes or rays, and is related to x² + y² - z² = -1.
  2. [2]
    [PDF] William F. Reynolds (1993) Hyperbolic Geometry on a Hyperboloid ...
    The hyperboloid model (H2) of hyperbolic geometry is on one sheet of a hyperboloid in Minkowski 3-space, defined by q(X) = -1 and x0 > 0.
  3. [3]
    [PDF] Lecture 3. A Brief Introduction to Hyperbolic Geometry
    The isometry from the hyperboloid model and the upper half space model can be established using Cartan's theorem in Riemannian geometry. Cartan's theorem says ...
  4. [4]
    [PDF] Math 6640 – Hyperbolic Geometry Course Notes, Fall 2023
    Dec 7, 2023 · Thus it is a natural invariant “distance” in the geometry of special relativity. 3.2 Hyperboloid model in Minkowski space. The hyperboloid model ...
  5. [5]
    [PDF] Sign Conventions in General Relativity
    Dec 11, 2023 · This article highlights some sign conventions that appear to be standard and describes some of the effects of changing the signs of the metric ...<|control11|><|separator|>
  6. [6]
    [PDF] introduction to geometry - UChicago Math
    Aug 24, 2012 · Hyperbolic space has 3 models. We will introduce all of the models, and compute the metric on the hyperboloid model. The relationship ...
  7. [7]
  8. [8]
    [PDF] Fundamental solution of the Laplacian in the hyperboloid model of ...
    Jan 20, 2012 · Hyperbolic space in d-dimensions is a fundamental example of a space exhibiting hyperbolic geometry. It was developed independently by ...
  9. [9]
    [PDF] Notes on the Laplacian and its Eigenfunctions on Bolyai ...
    May 28, 2025 · T2 − X2 − Y 2 = R2 into a Minkowski space with metric ds2 = −dT2 + dX2 + dY 2. Despite the minus sign in the Minkowski metric, the resulting ...
  10. [10]
    [PDF] Hyperbolic Geometry and Distance Functions on Discrete Groups
    define the hyperboloid model and show that it is a metric space. A discussion of its geodesics and trigonometry completes this section. 2.3.1 Lorentzian n ...
  11. [11]
    Foundations of Hyperbolic Manifolds - SpringerLink
    This book is an exposition of the theoretical foundations of hyperbolic manifolds. It is intended to be used both as a textbook and as a reference.<|control11|><|separator|>
  12. [12]
    [PDF] Hyperbolic Geometry - UC Davis Math
    Hyperbolic geometry, a non-Euclidean geometry, was created in the 19th century. It is a negatively curved geometry with applications in many fields.
  13. [13]
    [PDF] A. Geodesics in the Hyperboloid Model B. Proof of Theorem 1 C ...
    Proof. For any geodesic γx,v(t), consider the plane spanned by the vectors x and v. Then, from Thm. 1, this plane contains all the points of γx,v(t), i.e.Missing: geometry | Show results with:geometry
  14. [14]
    [PDF] horospheres in hyperbolic geometry
    We use the Lorentz space model for the Hyperbolic Geometry. In detail, the model lives in Lorentz space Rn+1. 1 with its Lorentz product hx, yi = x1y1 + x2y2 ...
  15. [15]
    [PDF] Chapter 2: Hyperbolic Geometry - UChicago Math
    Feb 15, 2019 · This defines the Klein projective model of hyperbolic space. In this model, hyperbolic straight lines and planes are the intersection of ...<|control11|><|separator|>
  16. [16]
    GEOMETRY OF HOROSPHERES - Project Euclid
    t. In the flat case (a = ...Missing: hyperboloid | Show results with:hyperboloid
  17. [17]
    [PDF] 5.3 The Lorentz Groups O(n,1), SO(n,1) and SO0(n,1) - UPenn CIS
    Since every Lorentz isometry, A ∈ SO(n,1), preserves the Lorentz inner product, we conclude that A globally preserves every hyperboloid, Hn(r), for r > 0. We ...
  18. [18]
    [PDF] 5 Lie Groups and Lie Algebras - Oregon State University
    Rotations correspond to ordinary trig, with compact orbits; boosts correspond to hyperbolic trig, with noncompact orbits. The correct generalization of this ...
  19. [19]
    [PDF] Geometry of hyperbolic space - Matrix Editions
    The most natural – if not the easiest to visualize – is the hyperboloid model, which imitates the definition of the unit sphere. We use the symbol Hn to denote ...<|separator|>
  20. [20]
    [PDF] Hyperbolic geometry MA 448 - University of Warwick
    Mar 1, 2013 · The hyperboloid model In this model, hyperbolic space is geometry on a. 'sphere of radius i'. Consider R3 with the indefinite Lorentz metric.
  21. [21]
    [PDF] Hyperbolic geometry in the work of Johann Heinrich Lambert - HAL
    Mar 5, 2015 · Because of that, Proclus is also considered sometimes as a precursor of hyperbolic geometry. ... constructed for the hyperbolic plane) as a sphere ...
  22. [22]
    [PDF] Hyperbolic geometry: history, models, and axioms - DiVA portal
    Only with the work of later mathematicians, hyperbolic geometry found acceptance, which occurred after the death of both Bolyai and Lobachevsky. So far we have ...
  23. [23]
    [PDF] Hyperbolic Geometry - UNL Digital Commons
    Jul 18, 2007 · In the nineteenth century, however, hyperbolic geometry was extensively studied by János Bolyai and Nikolai Ivanovich Lobachevsky. (Because of ...
  24. [24]
    [PDF] The hyperboloid model of hyperbolic geometry - SciSpace
    Abstract. The main goal of this thesis is to introduce and develop the hyperboloid model of Hyperbolic Geometry. In order to do that,.
  25. [25]
    Vorlesungen über geometrie : Clebsch, Alfred, 1833-1872
    Jun 4, 2009 · Publication date: 1876 ; Topics: Geometry, Analytic, Forms (Mathematics), Transformations (Mathematics) ; Publisher: Leipzig : B.G. Teubner.Missing: hyperbolic | Show results with:hyperbolic
  26. [26]
    Alexander MacFarlane Principles of Elliptic and Hyperbolic Analysis ...
    Alexander MacFarlane Principles of Elliptic and Hyperbolic Analysis 1894 - Free download as PDF File (.pdf) or read online for free.Missing: metric | Show results with:metric
  27. [27]
    [PDF] MINKOWSKI SPACE-TIME AND HYPERBOLIC GEOMETRY
    Projection of the hyperboloid to a disc or spherical ball gives an associated Beltrami-Klein representation of velocity space. This geometrical representation ...
  28. [28]
    [PDF] Minkowski's Modern World - HAL-SHS
    Nov 26, 2015 · The first section sketches the background to physical geometry at the time of Minkowski's first lecture on relativity in 1907, and in the second ...
  29. [29]
    Hyperbolic Geometry on a Hyperboloid - jstor
    In 1909, Jansen [12] gave what appears to be the first detailed exposition ... Jansen, Abbildung der hyperbolischen Geometrie auf ein zweischaliges Hyperboloid, ...
  30. [30]
    [PDF] the hyperbolic theory of special relativity - arXiv
    definite quadratic form (1) of the Minkowski space. Since from equation (1) (x y z) is time- like, (dx dy dz) must be space-like since no two time-like ...
  31. [31]
    [PDF] Hyperbolic length, lines, and distances - Department of Mathematics
    Oct 5, 2020 · Hyperboloid model. Poincaré disc model is related to the hyperboloid model projectively. If a point [t,x1, ...,xn] is on the upper sheet of ...
  32. [32]
    [PDF] HyperRogue: Playing with Hyperbolic Geometry
    HyperRogue is a computer game whose action takes place in the hyperbolic plane. We discuss how HyperRogue is relevant for mathematicians, artists, teachers, and ...
  33. [33]
    Point-form quantum field theory - ScienceDirect.com
    Summary and outlook. Point-form quantum field theory was first developed in the 1970s by taking the forward hyperboloid xμxμ = τ2 as the quantization surface.