Fact-checked by Grok 2 weeks ago

Hyperboloid

A hyperboloid is a quadric surface in three-dimensional Euclidean space defined by a second-degree equation, existing in two distinct forms: the one-sheeted hyperboloid, which is a connected surface resembling a tube or hourglass, and the two-sheeted hyperboloid, consisting of two separate bowl-shaped components. The standard equation for a one-sheeted hyperboloid aligned along the z-axis is \frac{x^2}{a^2} + \frac{y^2}{b^2} - \frac{z^2}{c^2} = 1, where a, b, and c are positive constants determining the scale along each axis, producing elliptical cross-sections parallel to the xy-plane and hyperbolic cross-sections in planes parallel to the xz- or yz-planes. Similarly, the two-sheeted hyperboloid has the equation \frac{x^2}{a^2} + \frac{y^2}{b^2} - \frac{z^2}{c^2} = -1, yielding no real intersection with planes parallel to the xy-plane for small |z| but hyperbolas elsewhere, with the sheets separated along the z-axis. Both types of hyperboloids are ruled surfaces, meaning they can be entirely generated by the motion of straight lines, a property that distinguishes them among surfaces and enables their with minimal while maintaining structural . The one-sheeted hyperboloid, in particular, is doubly ruled, allowing two distinct families of lines to cover the surface, which facilitates applications in and . For instance, Shukhov pioneered the use of hyperboloid structures in the late , constructing the first such tower—a 37-meter —at the 1896 All-Russian Industrial Exhibition in , leveraging the form's aerodynamic stability and ease of assembly without extensive scaffolding. This innovation influenced modern designs, including cooling towers, which adopt the one-sheeted hyperboloid shape for optimal heat dissipation and wind resistance. In mathematics, hyperboloids arise as surfaces of revolution by rotating a hyperbola around one of its axes—the transverse axis for the one-sheeted form and the conjugate axis for the two-sheeted form—and play key roles in , , and , where the two-sheeted hyperboloid models in Minkowski . Parametric representations, such as x = a \cosh u \cos v, y = a \cosh u \sin v, z = c \sinh u for the one-sheeted case (with u \in \mathbb{R}, v \in [0, 2\pi)), allow for detailed study of their curvature and geodesics. These surfaces exemplify the diversity of forms, bridging with practical feats.

Definition and Classification

Canonical Forms

The one-sheeted hyperboloid is a surface that is doubly ruled by two families of straight lines, distinguishing it from ellipsoids, which are compact and non-ruled, and paraboloids, which exhibit parabolic rather than profiles. The two-sheeted hyperboloid is not ruled. These surfaces arise as special cases of the general equation and represent unbounded structures with cross-sections in principal planes. The canonical equation for the hyperboloid of one sheet, aligned with the coordinate axes, is \frac{x^2}{a^2} + \frac{y^2}{b^2} - \frac{z^2}{c^2} = 1, where a, b, c > 0 are positive parameters that define the semi-axes lengths: a and b along the x- and y-directions, and c scaling the hyperbolic opening along the z-axis. This form describes a connected surface that intersects the xy-plane in an with semi-axes lengths a and b. For the hyperboloid of two sheets, the canonical equation is -\frac{x^2}{a^2} - \frac{y^2}{b^2} + \frac{z^2}{c^2} = 1, with the same parameters a, b, c > 0 interpreting the semi-axes similarly, though here the sheets separate along the z-axis, each extending infinitely in opposite directions without connecting. The minimum distance between the sheets is $2c, occurring at the vertices (0, 0, \pm c). Leonhard Euler introduced the in the , recognizing it as the counterpart to the within his of surfaces.

Distinction from Other Quadrics

surfaces are classified according to the signature of the defining them, which corresponds to the number of positive, negative, and zero eigenvalues of the associated . This signature determines the geometric type: definite forms (all eigenvalues positive or all negative) yield , while indefinite forms (mixed signs) produce , and forms with a zero eigenvalue lead to paraboloids or cylinders. Hyperboloids specifically exhibit indefinite signatures without zero eigenvalues in their non-degenerate cases, setting them apart from the positive definite signature of bounded ellipsoids and the signatures involving zeros in unbounded paraboloids. The one-sheet hyperboloid has a of (2,1), meaning two positive and one negative eigenvalue, resulting in a connected, . In contrast, the two-sheet hyperboloid has a of (1,2), with one positive and two negative eigenvalues, producing two disconnected sheets. Eigenvalue analysis of the quadratic form matrix distinguishes these types precisely; for example, in the canonical form \frac{x^2}{a^2} + \frac{y^2}{b^2} - \frac{z^2}{c^2} = 1, the eigenvalues are \lambda_1 = 1/a^2 > 0, \lambda_2 = 1/b^2 > 0, and \lambda_3 = -1/c^2 < 0, confirming the one-sheet hyperboloid. Ellipsoids require all \lambda_i > 0, such as in \frac{x^2}{a^2} + \frac{y^2}{b^2} + \frac{z^2}{c^2} = 1, while paraboloids involve one zero eigenvalue, as in the elliptic paraboloid z = \frac{x^2}{a^2} + \frac{y^2}{b^2}. Degenerate hyperboloids arise when the matrix becomes singular or the equation factors linearly, such as when one eigenvalue is zero, leading to a cylinder with hyperbolic cross-sections extending infinitely along one . Further degeneration, where the factors into two linear terms (e.g., (x - y)(x + y) = 0), results in a pair of intersecting planes.

Mathematical Representations

Cartesian Equations

The general equation of a surface in three-dimensional Cartesian coordinates is given by
Ax^2 + By^2 + Cz^2 + 2Dxy + 2Exz + 2Fyz + Gx + Hy + Iz + J = 0,
where A, B, C, D, E, F, G, H, I, J are constants. This represents a hyperboloid when the associated has a signature of either two positive and one negative eigenvalue (hyperboloid of ) or one positive and two negative eigenvalues (hyperboloid of two sheets), provided the surface is non-degenerate and the constant term after reduction satisfies specific sign conditions.
To derive the general Cartesian equation for a hyperboloid from its canonical form, one first applies a to shift the coordinate to the center of the surface, eliminating the linear terms Gx + Hy + Iz. The center (x_0, y_0, z_0) is found by solving the system of partial derivatives set to zero:
$2Ax + 2D y + 2E z + G = 0,
$2D x + 2B y + 2F z + H = 0,
$2E x + 2F y + 2C z + I = 0.
This yields a centered of the form \mathbf{x'}^T A \mathbf{x'} + J' = 0, where \mathbf{x'} = (x - x_0, y - y_0, z - z_0)^T and A is the
A = \begin{pmatrix} A & D & E \\ D & B & F \\ E & F & C \end{pmatrix}.
Next, an () is applied to diagonalize the \mathbf{x'}^T A \mathbf{x'}, using the eigenvectors of A to determine the principal axes and orientation. The eigenvalues \lambda_1, \lambda_2, \lambda_3 of A provide the coefficients in the rotated coordinates (x'', y'', z''), resulting in \lambda_1 (x'')^2 + \lambda_2 (y'')^2 + \lambda_3 (z'')^2 + J' = 0. Normalizing by dividing through by -J' (assuming J' \neq 0) and scaling yields the for the one-sheet hyperboloid:
\frac{(x')^2}{a^2} + \frac{(y')^2}{b^2} - \frac{(z')^2}{c^2} = 1,
where the primed coordinates are aligned with the principal axes, and a, b, c > 0 are determined from the reciprocals of the eigenvalues adjusted for the right-hand side sign. For the two-sheet hyperboloid, the equation is
\frac{(x')^2}{a^2} + \frac{(y')^2}{b^2} - \frac{(z')^2}{c^2} = -1,
or equivalently,
-\frac{(x')^2}{a^2} - \frac{(y')^2}{b^2} + \frac{(z')^2}{c^2} = 1,
distinguished by the opposite sign configuration in the eigenvalue scaling to ensure the hyperbolic type./12%3A_Vectors_in_Space/12.06%3A_Quadric_Surfaces) The orientation is given by the whose columns are the normalized eigenvectors of A, and the semi-axes lengths a, b, c are identified from the diagonalized coefficients.

Parametric and Implicit Forms

The hyperboloid of one sheet admits a parametric representation using , given by \begin{align*} x &= a \cosh u \cos v, \\ y &= b \cosh u \sin v, \\ z &= c \sinh u, \end{align*} where u \in \mathbb{R} and v \in [0, 2\pi), with a, b, c > 0. This parametrization arises from scaling the standard hyperbolic identities \cosh^2 u - \sinh^2 u = 1 along the respective axes to match the general \frac{x^2}{a^2} + \frac{y^2}{b^2} - \frac{z^2}{c^2} = 1. An alternative trigonometric parametrization for the one-sheet hyperboloid, suitable for computational visualization, is \begin{align*} x &= a \sec u \cos v, \\ y &= b \sec u \sin v, \\ z &= c \tan u, \end{align*} with u \in (-\pi/2, \pi/2) and v \in [0, 2\pi). Substituting these into the yields \sec^2 u - \tan^2 u = 1, confirming the surface. For the hyperboloid of two sheets, defined by \frac{x^2}{a^2} + \frac{y^2}{b^2} - \frac{z^2}{c^2} = -1, the parametric equations using hyperbolic functions are \begin{align*} x &= a \sinh u \cos v, \\ y &= b \sinh u \sin v, \\ z &= c \cosh u \end{align*} for the upper sheet (u \in \mathbb{R}, v \in [0, 2\pi)), and similarly for the lower sheet with z = -c \cosh u. This form leverages the identity \cosh^2 u - \sinh^2 u = 1 to satisfy the negative right-hand side of the equation. In cylindrical coordinates, the hyperboloids can be expressed explicitly by solving for z. For the one-sheet hyperboloid, z = \pm c \sqrt{\frac{x^2}{a^2} + \frac{y^2}{b^2} - 1}, valid where \frac{x^2}{a^2} + \frac{y^2}{b^2} \geq 1. For the two-sheet hyperboloid, z = \pm c \sqrt{\frac{x^2}{a^2} + \frac{y^2}{b^2} + 1}, defined for all x, y. These expressions facilitate tracing the surface by varying radial and angular parameters in the xy-plane. The implicit representation of a hyperboloid is the f(x,y,z) = k, where f(x,y,z) is a such as f(x,y,z) = \frac{x^2}{a^2} + \frac{y^2}{b^2} - \frac{z^2}{c^2} for k = 1 (one sheet) or k = -1 (two sheets). The surface normal at any point is given by the \nabla f = \left( \frac{2x}{a^2}, \frac{2y}{b^2}, -\frac{2z}{c^2} \right), which is perpendicular to the tangent plane.

Geometric Properties of One-Sheet Hyperboloid

Ruling Lines and Developability

The one-sheet hyperboloid is distinguished by its property as a doubly , containing two distinct families of straight lines that lie entirely on . These rulings are infinite straight lines that generate and intersect each other, providing a skeletal structure that connects the entire connected sheet. Each point on belongs to exactly one line from each family, emphasizing the doubly ruled nature. In contrast, the two-sheet hyperboloid possesses no real straight lines lying on its surface, as any potential rulings would require imaginary parameters to satisfy the defining , resulting in disconnected sheets without such linear generators. The rulings can be explicitly parameterized for the \frac{x^2}{a^2} + \frac{y^2}{b^2} - \frac{z^2}{c^2} = 1. One family is given by (x,y,z) = (a \cos \theta, b \sin \theta, 0) + t (-b \sin \theta, a \cos \theta, c), where \theta parameterizes the family and t \in \mathbb{R} traces along each line. Substituting this into the hyperboloid yields an , confirming that the entire line lies on . The second family uses a complementary direction, given by (x,y,z) = (a \cos \theta, b \sin \theta, 0) + t (b \sin \theta, -a \cos \theta, c). This parameterization demonstrates the surface's generation by two sets of non-intersecting lines within each family, with lines from different families crossing. To prove the one-sheet hyperboloid is ruled, consider a general straight line in form (x,y,z) = (x_0 + \alpha t, y_0 + \beta t, z_0 + \gamma t) and substitute into the \frac{x^2}{a^2} + \frac{y^2}{b^2} - \frac{z^2}{c^2} = 1. This yields a in t: A t^2 + B t + C = 0. For the line to lie entirely on , this must hold identically for all t, requiring A = B = C = 0. Solving these conditions on the direction (\alpha, \beta, \gamma) yields two one-parameter families of rulings. For the two-sheet case \frac{x^2}{a^2} + \frac{y^2}{b^2} - \frac{z^2}{c^2} = -1, the corresponding analysis yields no real solutions for the direction vectors, hence no real rulings. The rulings can also be obtained as degenerate conic sections where a plane intersects the hyperboloid in a pair of straight lines. Such degeneration occurs when the plane is tangent to the asymptotic cone \frac{x^2}{a^2} + \frac{y^2}{b^2} - \frac{z^2}{c^2} = 0. Each ruling line lies on the surface and is a generator of this cone. As a , the one-sheet hyperboloid exhibits zero normal along each ruling line, since the generators are geodesics that are in space, allowing local unrolling along those directions without stretching in the . However, the surface as a whole has negative and cannot be globally flattened into a without distortion, distinguishing it from truly developable ruled surfaces like cylinders.

Plane Sections and Cross-Sections

The intersection of the one-sheet hyperboloid with a plane parallel to the xy-plane at z = k yields an ellipse for any real k. The equation of this elliptic section is \frac{x^2}{a^2 \left(1 + \frac{k^2}{c^2}\right)} + \frac{y^2}{b^2 \left(1 + \frac{k^2}{c^2}\right)} = 1, with semi-axes a \sqrt{1 + \frac{k^2}{c^2}} and b \sqrt{1 + \frac{k^2}{c^2}}. As |k| increases, the ellipse expands, reflecting the hyperbolic flaring along the z-axis. In a longitudinal plane such as y = 0, the intersection consists of a \frac{x^2}{a^2} - \frac{z^2}{c^2} = 1, with both branches lying on the single connected sheet of the hyperboloid. Similar hyperbolas arise in planes x = 0 or other meridional sections parallel to the z-axis. For general planes, the intersection is a conic section: , , or parabola, depending on the plane's orientation and position relative to the asymptotic cone. The conic degenerates into a pair of straight lines (rulings) if the plane is tangent to the asymptotic cone \frac{x^2}{a^2} + \frac{y^2}{b^2} - \frac{z^2}{c^2} = 0. Unlike the two-sheet hyperboloid, all plane sections are non-empty due to the connected nature of the surface.

Geometric Properties of Two-Sheet Hyperboloid

Asymptotic Behavior

The two-sheet hyperboloid, defined by the equation \frac{x^2}{a^2} + \frac{y^2}{b^2} - \frac{z^2}{c^2} = -1, possesses an asymptotic given by \frac{x^2}{a^2} + \frac{y^2}{b^2} - \frac{z^2}{c^2} = 0, which separates the two disconnected sheets of the surface. As distances increase along the transverse axis, particularly for large |z|, each sheet extends unboundedly toward infinity, asymptotically approaching the ; the deviation from the diminishes such that the surface aligns linearly with the 's generators in the far field. The surface exhibits no real points in the region where |z| < c, with the two sheets initiating at the vertices (0, 0, \pm c) along the -axis and flaring outward thereafter. Analogous to the two-dimensional hyperbola, the two-sheet hyperboloid features foci and directrices derived from its meridional sections, where the eccentricity e of the generating hyperbola \frac{z^2}{c^2} - \frac{x^2}{a^2} = 1 (assuming b = a for simplicity in the rotational case) is given by e = \sqrt{1 + \frac{a^2}{c^2}}, with foci at (0, 0, \pm c e).

Plane Sections and Cross-Sections

The intersection of a two-sheet hyperboloid with a transverse plane parallel to the xy-plane at z = k, where |k| > c for the standard form \frac{x^2}{a^2} + \frac{y^2}{b^2} - \frac{z^2}{c^2} = -1, yields an ellipse. The equation of this elliptic section is \frac{x^2}{a^2 (k^2/c^2 - 1)} + \frac{y^2}{b^2 (k^2/c^2 - 1)} = 1, with semi-axes scaled by \sqrt{k^2/c^2 - 1}. As |k| increases, the ellipse expands, reflecting the hyperbolic divergence along the z-axis. In a longitudinal plane such as x = 0, the intersection consists of a \frac{y^2}{b^2} - \frac{z^2}{c^2} = -1, rewritten as \frac{z^2}{c^2} - \frac{y^2}{b^2} = 1, with its two branches lying on the separate sheets of the hyperboloid. Similar hyperbolas arise in planes y = 0 or other meridional sections parallel to the z-axis. Degenerate cases may yield pairs of lines when the plane aligns with rulings approaching the asymptotic cone. For general planes, the intersection can be empty if the plane misses both sheets, an ellipse on one sheet if the plane is oriented transversely enough to intersect only that sheet, or a hyperbola on one or both sheets depending on orientation. Hyperbolic sections on the sheets enlarge as the plane nears the asymptotic cone, where cross-sections approach unbounded forms.

General Properties

Symmetries and Invariants

The symmetries of both the one-sheet and two-sheet hyperboloids are determined by the group of linear transformations that preserve their defining quadratic form Q(\mathbf{x}) = x^2 + y^2 - z^2, which has signature (2,1). This group is the indefinite orthogonal group O(2,1), consisting of all matrices A such that A^T \begin{pmatrix} 1 & 0 & 0 \\ 0 & 1 & 0 \\ 0 & 0 & -1 \end{pmatrix} A = \begin{pmatrix} 1 & 0 & 0 \\ 0 & 1 & 0 \\ 0 & 0 & -1 \end{pmatrix}. The connected component of the identity, SO^+(2,1), is isomorphic to SL(2,\mathbb{R})/\{\pm I\} and acts transitively on the upper sheet of the two-sheet hyperboloid and on the one-sheet hyperboloid itself. In the aligned with the coordinate axes, the hyperboloids exhibit axial symmetries. Continuous around the z-axis is generated by the maximal compact SO(2), corresponding to matrices of the form \begin{pmatrix} \cos\theta & -\sin\theta & 0 \\ \sin\theta & \cos\theta & 0 \\ 0 & 0 & 1 \end{pmatrix}. Discrete reflection symmetries include inversion through the origin (central symmetry), reflection over the xy-plane (z \to -z), and reflections over the xz-plane (y \to -y) and yz-plane (x \to -x), all preserving the surface. Key invariants under orthogonal transformations include the of the , which is (2,1) for hyperboloids and distinguishes them from definite forms like ellipsoids (signature (3,0)). In principal coordinates, the of the associated to the (sum of eigenvalues) and its (product of eigenvalues) provide scaling information along the axes, though the eigenvalues themselves are the primary invariants for . The center lies at the , where the of the vanishes, and the principal axes align with the eigenvectors of the matrix, diagonalizing the form without cross terms via the principal axes theorem.

Curvature and Differential Geometry

The Gaussian curvature K of a hyperboloid distinguishes its local geometry: for the one-sheet hyperboloid \frac{x^2}{a^2} + \frac{y^2}{b^2} - \frac{z^2}{c^2} = 1, K < 0 everywhere, embedding hyperbolic geometry, while for the two-sheet hyperboloid \frac{x^2}{a^2} + \frac{y^2}{b^2} - \frac{z^2}{c^2} = -1, K > 0 on each disconnected sheet, locally resembling elliptic geometry. For the one-sheet case, the is given by K = -\frac{1}{a^2 b^2 c^2 d^4}, where d = \sqrt{\frac{x^2}{a^4} + \frac{y^2}{b^4} + \frac{z^2}{c^4}}. This formula arises from the representation and confirms the negative sign, with |K| maximal near the narrowest cross-section (the "") at z=0 and approaching zero asymptotically along directions away from it. In coordinates, for the circular case (a = b) with x = a \sqrt{1 + u^2} \cos v, y = a \sqrt{1 + u^2} \sin v, z = c u, it simplifies to K(u,v) = -\frac{c^2}{[c^2 + (a^2 + c^2) u^2]^2}, always negative and independent of v due to . For the two-sheet hyperboloid, the is positive and expressed implicitly as K(x,y,z) = \frac{c^6}{[c^4 - (a^2 + c^2) z^2]^2} in the circular case (a = b), vanishing asymptotically as |z| \to \infty and peaking near the vertices at z = \pm c. Each sheet has varying positive curvature, though the surface as a whole is disconnected. The H and principal curvatures \kappa_1, \kappa_2 (satisfying K = \kappa_1 \kappa_2 and H = (\kappa_1 + \kappa_2)/2) vary across both hyperboloids. For the one-sheet, H is given parametrically by H(u,v) = \frac{c^2 [a^2 (u^2 - 1) + c^2 (u^2 + 1)]}{2a [c^2 + (a^2 + c^2)u^2]^{3/2}} (circular case), changing sign across the and zero along a equator. The principal curvatures have opposite signs due to K < 0, with curves of constant \kappa_i being algebraic of degree 16. The one-sheet hyperboloid possesses no umbilical points, where \kappa_1 = \kappa_2, as the negative K precludes equal nonzero curvatures; the two-sheeted hyperboloid has four real umbilical points. On the one-sheet hyperboloid, the straight ruling lines—generators of the ruled surface—are geodesics, as their zero space curvature implies zero geodesic curvature on the surface. More generally, geodesics intersect the rulings and can be found by solving the geodesic equations in parametric coordinates, often lying on planes through the origin in the embedding space.

Extensions and Generalizations

In Higher Dimensions

In higher-dimensional Euclidean space \mathbb{R}^n, a hyperboloid of one sheet is defined as the hypersurface given by the equation \sum_{i=1}^{p} \frac{x_i^2}{a_i^2} - \sum_{j=1}^{q} \frac{x_{p+j}^2}{b_j^2} = 1, where p + q = n, the parameters a_i > 0 and b_j > 0 scale the axes, and the has indefinite (p, q) with both p \geq 2 and q \geq 1 to ensure the surface is connected and non-degenerate. This generalizes the three-dimensional case of (2,1), where the surface is a ruled connecting two nappes of a hyperbolic . The analogue for the hyperboloid of two sheets replaces the right-hand side with -1, yielding two disconnected components when the signature allows, such as (1, n-1), and corresponding to level sets where the takes a negative value. Key properties of these hypersurfaces extend from lower dimensions, particularly regarding rulings and geometric models. For signatures like (n-1, 1), the one-sheet hyperboloid remains a ruled hypersurface, generated by families of straight lines lying entirely on the surface, analogous to the regulus structure in three dimensions; this follows from the fact that such quadrics contain maximal linear subspaces of dimension q-1 = 0 or higher in degenerate cases, but the full hypersurface admits two families of rulings. In signatures (n, 1), the two-sheet hyperboloid (specifically the upper sheet \{ x \in \mathbb{R}^{n+1} : \sum_{i=1}^n x_i^2 - x_{n+1}^2 = -1, \, x_{n+1} > 0 \}) serves as the hyperboloid model of n-dimensional hyperbolic space \mathbb{H}^n, equipped with the induced Riemannian metric of constant sectional curvature -1, where geodesics are intersections with planes through the origin in the ambient Minkowski space. In the context of , hyperboloids arise as mass shells in four-dimensional of (1,3), defined by the equation t^2 - \mathbf{x}^2 = m^2 (with m > 0 the rest ), representing the worldlines of particles with fixed ; the future-directed sheet \{ (t, \mathbf{x}) : t > 0 \} is a three-dimensional hyperboloid of two sheets, modeling the of under Lorentz transformations. This structure underscores the intrinsic of constant surfaces in , preserved by the .

Relation to Spheres and Other Surfaces

The upper sheet of the two-sheet in three-dimensional , defined by the equation x^2 + y^2 - z^2 = -1 with z > 0, is isometric to the under the induced metric, providing a model for where geodesics are intersections with planes through the origin and distances are given by \cosh d(P, Q) = - \langle P, Q \rangle. This construction parallels the sphere in , where the sphere x^2 + y^2 + z^2 = 1 models with great circles as geodesics, highlighting the duality between positive and negative spaces through forms of opposite . Stereographic projection from a point on the one-sheet hyperboloid, such as through the analog in pseudo-Euclidean coordinates, maps the surface to a minus a point, enabling the embedding of structures onto spherical domains while preserving angles and facilitating computations in constant curvature spaces. This mapping underscores the conformal properties shared with the classical of the to the , allowing hyperbolic models to be visualized on nearly complete spheres. In inversion geometry within , the hyperboloid arises as the inverse image of a under inversion with respect to a orthogonal to the time-like , transforming spheres into hyperboloids while preserving angles and mapping circles to hypercycles or horocycles in the . Such inversions extend conformal mappings from spherical to hyperbolic settings, where spheres tangent to the at become planes in the half-space model.

Applications and Structures

Architectural and Engineering Uses

The pioneering use of hyperboloid geometry in architecture is exemplified by the in , constructed between 1920 and 1922 by Russian engineer Vladimir Shukhov as a 160-meter-tall radio tower based on a one-sheet hyperboloid. This structure marked one of the earliest large-scale applications of the form, leveraging its inherent stability for a lightweight, self-supporting design composed of intersecting straight steel struts. Shukhov's innovation built on his earlier 1896 at the All-Russia in , demonstrating the practical viability of hyperboloid shapes for tall, slender constructions resistant to wind and seismic forces. Hyperboloid forms have become standard in natural draft cooling towers, where the one-sheet profile optimizes airflow and structural efficiency in power plants worldwide. These towers, typically ranging from 100 to 200 meters in height, feature a narrow and widening base to enhance the chimney effect, drawing hot air upward through evaporative cooling processes. The shape's aerodynamic properties minimize material use while maximizing draft velocity, as seen in facilities like those at the in the UK. The appeal of hyperboloid structures lies in their geometry, which consists of straight-line generators that enable construction using simple, linear beams or struts, reducing fabrication complexity and costs compared to curved forms. This results in exceptional strength-to-weight ratios, with the interlocking rulings providing inherent rigidity against and lateral loads, allowing for slender profiles that distribute forces evenly. In , hyperboloid elements continue to inspire efficient designs, such as the columns of the (1959–1970) by , which employ hyperboloid ribs to create a soaring, lightweight canopy evoking natural forms while ensuring structural integrity. Similar principles appear in tensile and lattice structures, like the (1963) in , a 108-meter hyperboloid that combines aesthetic fluidity with earthquake-resistant . Recent sustainable applications include bamboo hyperboloid towers in , such as the (2024), which utilize local materials for eco-friendly, lightweight construction.

Physical and Scientific Contexts

In special relativity, hyperboloids arise as surfaces of constant proper time in Minkowski spacetime, representing the locus of events reachable by light signals from a given origin after a fixed interval. These surfaces generalize the hyperbolic worldlines of particles undergoing constant proper acceleration, providing a geometric framework for understanding time dilation and synchronization in Lorentzian geometry. For instance, in two-dimensional Minkowski space, the hyperbola serves as the relativistic analog of a circle, with hyperboloids extending this to higher dimensions for analyzing inertial observers and wave propagation. In , hyperbolic mirrors are employed to between their two foci, enabling precise manipulation in systems like confocal and telescopes, where they correct spherical aberrations in Ritchey-Chrétien designs. Aspheric lenses often approximate hyperboloid profiles to achieve superior performance over spherical , reducing aberrations and enabling compact, high-resolution ; for example, bi-convex lenses with hyperboloid surfaces have demonstrated aberration-free with resolutions surpassing traditional microscope objectives. These designs are particularly useful for collimating divergent sources by placing the source at one , directing rays toward the second focus or approximating parallel output in hybrid systems. Hyperboloid geometries contribute to acoustics through sound scattering and , where their curved surfaces influence wave propagation and can enhance focusing in resonator-like configurations, though applications remain exploratory compared to optical analogs. In , hyperboloid structures model certain molecular bonds, such as the carbon-carbon bond, where a hyperboloid-shaped captures the rotational flexibility and observed in organic molecules, aiding simulations of biomolecular . In modern , hyperboloids define anti-de Sitter () spaces within the AdS/CFT correspondence, embedding the of AdS as a hyperboloid in higher-dimensional flat space, which facilitates holographic dualities between gravity and conformal field theories. This framework, explored in limits where the AdS hyperboloid approaches a projective lightcone, underpins studies of and strongly coupled systems.

References

  1. [1]
    Hyperboloid -- from Wolfram MathWorld
    A hyperboloid is a quadratic surface which may be one- or two-sheeted. The one-sheeted hyperboloid is a surface of revolution obtained by rotating a hyperbola.
  2. [2]
    11.6: Quadric Surfaces - Mathematics LibreTexts
    Sep 5, 2018 · A hyperboloid of one sheet is any surface that can be described with an equation of the form x 2 a 2 + y 2 b 2 − z 2 c 2 = 1 . Describe the ...
  3. [3]
    4.6: Quadric Surfaces - Mathematics LibreTexts
    Apr 24, 2025 · Hyperboloids of one sheet have some fascinating properties. For example, they can be constructed using straight lines, such as in the sculpture ...Missing: geometry | Show results with:geometry
  4. [4]
    4.7: Quadric Surfaces - Mathematics LibreTexts
    Nov 19, 2021 · Recognize the main features of ellipsoids, paraboloids, and hyperboloids. Use traces to draw the intersections of quadric surfaces with the ...Missing: geometry | Show results with:geometry
  5. [5]
    Shukhov's Hyperboloids | The Engines of Our Ingenuity
    Mar 24, 2014 · Shukhov built the first hyperboloid structure - a water tower for the 1896 All-Russian Exhibition.It loomed 120 feet over the grounds and still ...
  6. [6]
    Two-Sheeted Hyperboloid -- from Wolfram MathWorld
    ### Summary of Two-Sheeted Hyperboloid (Wolfram MathWorld)
  7. [7]
    One-Sheeted Hyperboloid -- from Wolfram MathWorld
    A hyperboloid is a quadratic surface which may be one- or two-sheeted. The one-sheeted hyperboloid is a surface of revolution obtained by rotating a ...
  8. [8]
    [PDF] Classifying Quadrics using Exact Arithmetic - Geometric Tools
    Oct 10, 2022 · In many applications, we are interested only in ellipsoids, hyperboloids or paraboloids. However, a quadratic equation can define other surfaces.
  9. [9]
    [PDF] Additional notes on Quadratic forms - Manuela Girotti
    Important consequence: The type of quadric surface is again determined by the eigenvalues of the matrix Q, more specifically by the sign of the eigenvalues or ...
  10. [10]
    [PDF] Quadric surfaces - UBC Math
    1) "degenerate": ex: quadric surface a plane to pair of planes. 2 planes: Xzy ... hyperboloid.
  11. [11]
    Calculus III - Quadric Surfaces - Pauls Online Math Notes
    Nov 16, 2022 · Notice that the only difference between the hyperboloid of one sheet and the hyperboloid of two sheets is the signs in front of the variables. ...
  12. [12]
    Quadratic forms, canonical forms of quadric surfaces with centers
    If we do a suitable rotation on F(x', y', z') we can eliminate the cross product terms and reduce it to canonical form. The same rotation will eliminate the ...
  13. [13]
    [PDF] k k k
    (a) Show that the parametric equations. ,. ,. , represent a hyperboloid of one sheet. ;. (b) Use the parametric equations in part (a) to graph the hyperboloid ...
  14. [14]
    2.5 Tangent Planes and Normal Lines
    The tangent plane to a surface S at a point ( x 0 , y 0 , z 0 ) is the plane that fits S best at . For example, the tangent plane to the hemisphere S = { ( x
  15. [15]
  16. [16]
    Rulings of One Sheet Hyperboloid - Mathematics Stack Exchange
    Feb 10, 2016 · HINTS: First you should check that the points given by the parametrization do in fact lie on the hyperboloid. Next, you want to see that the ...
  17. [17]
    [PDF] Ruled Surface Generated by a Curve Lying on ... - Heldermann-Verlag
    A ruled surface is developable if at any point we have hc0(s) ∧ ~X(s),. ~X0 ... Consider the hyperboloid of one sheet as a regular surface parameterized by.
  18. [18]
    Show that the hyperboloid of one sheet is a doubly ruled surface.
    Oct 21, 2022 · Show that the hyperboloid of one sheet is a doubly ruled surface, ie each point on the surface is on two lines lying entirely on the surface.Missing: parameterization | Show results with:parameterization
  19. [19]
    [PDF] Curvature Functions on a one-sheeted Hyperboloid
    Aug 8, 2014 · The curves of constant Gaussian curvature K0 < 0 on a one-sheeted hyperboloid are the curves of contact of a developable ruled surface tangent ...
  20. [20]
    [PDF] A Quadratic surfaces
    A hyperbolic cylinder is a quadratic surface parameterized by the equation x2 a2. − y2 b2. = 1. (A.31). The equation above describes a hyperbola in the xy-plane ...
  21. [21]
    [PDF] MULTIVARIABLE CALCLABS WITH MAPLE
    Apr 22, 1999 · Its asymptotic cone is obtained by replacing the 1 on the right hand ... (e) The upper sheet of the hyperboloid of two sheets z2 = x2 + ...
  22. [22]
    Eccentricity of Hyperbola: Formula, Definition & Properties (2025)
    Rating 4.2 (373,000) The standard formula is: e = 1 + b 2 a 2 , where e is the eccentricity, a is the length of the semi-transverse axis, and b is the length of the semi-conjugate ...
  23. [23]
  24. [24]
    [PDF] The hyperboloid as ordered symmetric space - Heldermann-Verlag
    Abstract: We study the geometry of the one sheeted hyperboloid viewed as an ordered symmetric space. The geometric results are indispensible prerequesites ...
  25. [25]
    [PDF] Quadratic forms Chapter I: Witt's theory
    So we have at the moment two invariants of a quadratic form: its dimension n, ... Let V be a hyperbolic quadratic space, and let σ ∈ O(V ) be an isometry.
  26. [26]
    [PDF] Video Lecture F9: Quadratic Forms & Principal Axes Theorem
    The columns of P are called the principal axes of the quadratic form Q. We'll see soon how this helps us classify QF and handle optimization problems. MATH ...
  27. [27]
    [PDF] Enumerative Geometry of Curvature of Algebraic Hypersurfaces - arXiv
    Jun 18, 2022 · points, the one-sheeted hyperboloid has no real ... 3 = 1} has 12 complex umbilical points. The number of real umbilical points is. • ...
  28. [28]
    How can we find geodesics on a one sheet hyperboloid?
    Jan 5, 2016 · We have that a curve γ on a surface S is called a geodesic if ¨γ(t) is zero or perpendicular to the tangent plane of the surface at the point γ( ...Geodesics in hyperboloid model of hyperbolic space [closed]Geodesics in hyperboloid model of hyperbolic geometryMore results from math.stackexchange.com
  29. [29]
    [PDF] TOPOLOGICAL CLASSIFICATIONS OF HYPERQUADRICS
    In this case, the quadratic hypersurface is the graph of a second degree polynomial in n − 1 variables; specifically, t = |u|2 − |v|2 . Therefore the quadric ...Missing: hyperboloid | Show results with:hyperboloid
  30. [30]
    [PDF] Hyperbolic Geometry - UC Davis Math
    These notes are intended as a relatively quick introduction to hyperbolic ge- ometry. They review the wonderful history of non-Euclidean geometry. They give ...
  31. [31]
    [PDF] MINKOWSKI SPACE-TIME AND HYPERBOLIC GEOMETRY
    It has become generally recognized that hyperbolic (i.e. Lobachevskian) space can be represented upon one sheet of a two-sheeted cylindrical hyperboloid in ...<|control11|><|separator|>
  32. [32]
    [PDF] Hyperbolic geometry MA 448 - University of Warwick
    Mar 1, 2013 · . Hyperbolic space is modelled on one sheet of this hyperboloid, usually the upper sheet H+ where x0 > 0. (Thinking of the direction (1,0,0) ...
  33. [33]
    κ-Stereographic Projection model - geoopt's documentation!
    Stereographic projection models comes to bind constant curvature spaces. Such as spheres, hyperboloids and regular Euclidean manifold.
  34. [34]
    [PDF] Chapter 2: Hyperbolic Geometry - UChicago Math
    Feb 15, 2019 · We also use the notation SO(n + 1) and SO+(n,1) for SO+(Q) if we want to stress the signature and the dependence on the dimension n. Example 1.1 ...
  35. [35]
    [PDF] Analysis of Inversions in Spherical and Hyperbolic Geometries
    We have a point P and a circle of inversion I. 2. Take the radius OR from the center O through the point P. 3. Construct the chord AB.
  36. [36]
    [PDF] 2. SURFACES - Penn Math
    Jan 26, 2012 · One-sheeted hyperboloid x. 2. + y. 2. = 1 + z. 2. Page 7. 7. Two-sheeted ... The Catenoid is the only minimal surface of revolution. Page 14. 14.
  37. [37]
    [PDF] Minimal Surfaces
    Dec 13, 2012 · Below, we have a hyperboloid (neg- ative Gaussian Curvature), a cylinder (zero Gaussian curvature) and a sphere. (positive Gaussian curvature).
  38. [38]
    Shukhov Tower: The Eiffel of the East - BBC News
    Apr 9, 2014 · The new "hyperboloid" architecture was pioneered by Shukhov himself - a geometric design which he had employed in the building of water ...
  39. [39]
    Shukhov Tower - World Monuments Fund
    In fact, the world's first hyperboloid structure was a water tower designed by Shukhov for the 1896 All-Russia Exhibition in Nizhny Novgorod. Built as a ...
  40. [40]
    Vladimir Shukhov - Linda Hall Library
    Aug 28, 2024 · One of his most famous structures is the Adziogol Lighthouse, a lattice hyperboloid tower that is some 70 meters tall and sits on the Dneiper ...
  41. [41]
    Why Many Cooling Towers Have a Hyperboloid Shape
    Mar 13, 2022 · Many large cooling towers have a hyperboloid shape. The unique shape offers several benefits for towers and their efficiency.
  42. [42]
    Why Are Cooling Towers Shaped Like That? - Practical Engineering
    Nov 5, 2024 · The largest natural draft towers are more than 650 feet or 200 meters tall, and more than 400 feet or 120 meters in diameter. And you want the ...
  43. [43]
    Hyperboloid Structures in GSA - Oasys software
    Apart from the cost savings of avoiding curved beams or shuttering, they are far more resistant to buckling because the individual elements are straight. This ...
  44. [44]
    (PDF) Design of Hyperboloid Structures - ResearchGate
    Aug 6, 2025 · Hyperboloid structures provide many mechanical properties with fewer materials and are very economical [37] . This type of structure can improve ...
  45. [45]
    16 Examples of Stunning Modern Architecture by Oscar Niemeyer
    Mar 25, 2016 · The swooping columns of the Cathedral of Brasília form a coronal shape called a hyperboloid structure. The Roman Catholic cathedral, most of ...
  46. [46]
    [PDF] Curved, yet Straight: Stick Hyperboloids - George W. Hart
    Figure 3: Architectural examples of stick hyperboloids: (a) 1896 tower by Shukhov; (b) battleship masts; (c) Kobe Tower; (d) Corporation Street Bridge.<|control11|><|separator|>
  47. [47]
    Hyperbolic times in Minkowski space
    ### Summary of Hyperboloids as Constant Proper Time Surfaces in Minkowski Space
  48. [48]
    Explanation of mirrors from Field Guide to Spectroscopy - SPIE
    Like a spherical mirror, a parabolic mirror forms a collimated beam of light if a source is located at its focus. A hyperbolic mirror focuses light from one ...
  49. [49]
  50. [50]
    Bi-convex aspheric optical lenses - AIP Publishing
    Mar 9, 2017 · The conic constant for the both sides is found to be negative, suggesting bi-convex lenses with hyperboloid and prolate spheroid surfaces.
  51. [51]
    Modelling Hyperboloid Sound Scattering: The Challenge of ...
    Oct 4, 2011 · The interim results of the acoustic test demonstrate that the geometry of hyperboloids have a perceptual influence on sound scattering, perhaps ...
  52. [52]
    [hep-th/0010106] A new AdS/CFT correspondence - arXiv
    Oct 25, 2000 · We consider a geometric zero-radius limit for strings on AdS_5\times S^5, where the anti-de Sitter hyperboloid becomes the projective lightcone.