Fact-checked by Grok 2 weeks ago

Catenoid

A catenoid is a in three-dimensional , generated by rotating a about its axis of symmetry, resulting in a with zero . It can be parameterized as x(u,v) = (c \cosh(v/c) \cos u, c \cosh(v/c) \sin u, v) for parameters u \in [0, 2\pi) and v \in \mathbb{R}, where c > 0 scales the surface, forming a catenoid of constant width that narrows to a "neck" and flares out asymptotically to catenary ends. Discovered by Leonhard Euler in 1744 and rigorously proven to be a by Jean-Baptiste Meusnier in 1776, the catenoid represents the first known non-planar example of a , predating broader classifications in . As the unique axially symmetric in \mathbb{R}^3, it spans two parallel coaxial circles when truncated, analogous to a minimizing surface area between rings, and exhibits finite total of -4\pi with a Morse index of 1. The catenoid is topologically a punctured (genus zero) with two ends, and its Gauss map is anti-holomorphic, covering the unit except for the north and south poles, underscoring its role in the Weierstrass-Enneper representation of s. It is the only properly embedded, non-planar of finite topology in \mathbb{R}^3 with these properties, and by theorems such as those of Collin and Nitsche, it uniquely intersects every horizontal plane in a Jordan curve when oriented appropriately. Beyond , the catenoid models physical phenomena like shapes in capillarity and serves as a benchmark in for simulating s via .

Definition and Geometry

Catenary Curve Basis

The catenary curve describes the shape assumed by a perfectly flexible, inextensible or of uniform when suspended from two points and acted upon solely by . Its standard , with the at the , is given by
y = a \cosh\left(\frac{x}{a}\right),
where a > 0 is a scaling parameter that determines the curve's width and sag, related to the horizontal tension at the lowest point and the of the multiplied by (specifically, a = T_0 / (\mu g), with T_0 the horizontal tension, \mu the mass per unit length, and g the ).
The catenary's equation arises from either force balance on chain segments or variational principles minimizing subject to fixed length. In the force balance approach, consider a small segment of the chain: the horizontal tension component remains constant, while the vertical component increases with the weight supported, leading to the
\frac{d}{ds} \left( \frac{dy}{dx} \right) = \frac{1}{a},
where s is the along the curve; integrating this yields the hyperbolic cosine form. Alternatively, using the to minimize the \int y \, ds (with fixed total length) produces the Euler-Lagrange equation (y + h) y'' = 1 + (y')^2, which integrates to the same solution, confirming its equilibrium shape.
Historically, approximated the as a parabola in his 1638 , based on empirical observations of hanging chains, though this was an inexact model valid only for shallow sags. The true equation was rigorously derived in 1691 by , , and , in response to a challenge posed by to find the curve of a hanging chain. had earlier coined the term "catenary" (from Latin catena, meaning chain) in a 1690 letter to . Key geometric properties include the from the to a point (x, y), given by s = a \sinh(x/a), which reflects the curve's nature. The intrinsic equation relating the \rho and is \rho a = s^2 + a^2, highlighting the 's non-constant . A common misconception, stemming from Galileo's approximation, equates the catenary to a parabola y = k x^2; while similar for small angles, the catenary grows exponentially at large distances, unlike the parabola's . When rotated about its axis of symmetry, the catenary generates the catenoid surface.

Surface of Revolution

The catenoid is formed as a by rotating a curve about its directrix, the of . Specifically, the generating catenary in the radial-axial plane is given by r = a \cosh\left(\frac{z}{a}\right), where a > 0 is a scaling parameter, r is the radial distance from the z-, and z is the axial coordinate serving as height. This rotation occurs around the z-, producing a rotationally symmetric surface that connects smoothly across all azimuthal angles. Leonhard Euler first derived this construction in 1744 while investigating surfaces of minimal area. The resulting geometry yields an implicit Cartesian equation x^2 + y^2 = a^2 \cosh^2\left(\frac{z}{a}\right), or equivalently in cylindrical coordinates, r = a \cosh\left(\frac{z}{a}\right). Visually, the catenoid resembles an , with the narrowest "neck" occurring at z = 0, where the radius r = a. As |z| increases, the radius expands exponentially, flaring outward to infinity in both directions along the , creating two funnel-like ends joined at the . This shape arises directly from the hyperbolic nature of the profile. To describe distances on the surface, the provides the . Parametrizing the surface as \mathbf{X}(z, \theta) = \left( a \cosh\left(\frac{z}{a}\right) \cos \theta, \, a \cosh\left(\frac{z}{a}\right) \sin \theta, \, z \right) for z \in \mathbb{R} and \theta \in [0, 2\pi), the coefficients are E = \cosh^2\left(\frac{z}{a}\right), F = 0, and G = a^2 \cosh^2\left(\frac{z}{a}\right). Thus, the line element is ds^2 = \cosh^2\left(\frac{z}{a}\right) \, dz^2 + a^2 \cosh^2\left(\frac{z}{a}\right) \, d\theta^2. The associated surface area element follows as dA = \sqrt{EG - F^2} \, dz \, d\theta = a \cosh^2\left(\frac{z}{a}\right) \, dz \, d\theta, which integrates over the azimuthal direction to yield the arc-length-integrated form $2\pi a \cosh^2\left(\frac{z}{a}\right) \, dz for meridional slices. These expressions capture the basic geometry without delving into curvature properties.

Parametric Representation

The catenoid admits a standard parametrization in terms of angular and meridional coordinates, expressed as the position vector \mathbf{X}(u,v) = \left( a \cosh v \cos u, \, a \cosh v \sin u, \, a v \right), where a > 0 is a scaling constant that sets the minimum radius (neck size) to a, u \in [0, 2\pi) parameterizes the rotational angle, and v \in \mathbb{R} corresponds to the signed height along the axis of revolution. This form arises from rotating the catenary curve x = a \cosh(z/a) about the z-axis. The partial derivatives yield the tangent vectors \mathbf{X}_u = \left( -a \cosh v \sin u, \, a \cosh v \cos u, \, 0 \right), \quad \mathbf{X}_v = \left( a \sinh v \cos u, \, a \sinh v \sin u, \, a \right). These vectors are orthogonal, as their \mathbf{X}_u \cdot \mathbf{X}_v = 0, and the induced has coefficients g_{uu} = a^2 \cosh^2 v, g_{vv} = a^2 \cosh^2 v, and g_{uv} = 0. Consequently, the is diagonal with equal nonzero entries, reflecting the surface's . This parametrization is isothermal (conformal), meaning the metric takes the form ds^2 = e^{2\phi(u,v)} (du^2 + dv^2) with \phi(u,v) = \ln(a \cosh v), which preserves angles and facilitates computations in theory. Alternative representations may rescale the parameters, such as setting a = c and replacing v with v/c in the to adjust the , or employ the Enneper-Weierstrass formulation using analytic functions for broader generation.

Properties as a Minimal Surface

Zero Mean Curvature

The mean curvature H of an oriented surface in \mathbb{R}^3 is defined as the average of the principal curvatures: H = \frac{\kappa_1 + \kappa_2}{2}, where \kappa_1 and \kappa_2 are the normal curvatures at each point. A surface is minimal if and only if H = 0 everywhere, meaning the principal curvatures satisfy \kappa_1 = -\kappa_2. For the catenoid, parametrized in the form \mathbf{r}(u,v) = (c \cosh(v/c) \cos u, c \cosh(v/c) \sin u, v) with u \in [0, 2\pi) and v \in \mathbb{R}, the mean curvature vanishes identically. This follows from direct computation using the second fundamental form of the parametrized surface, which yields H = 0 at every point, or equivalently, from the condition that the coordinate functions satisfy the minimal surface equation (Laplace-Beltrami equation) \Delta_{\Sigma} \mathbf{r} = 2H \mathbf{n} = 0, where \Delta_{\Sigma} is the Laplace-Beltrami operator on the surface and \mathbf{n} is the unit normal. Such verification traces back to Meusnier's 1776 analysis confirming the catenoid as a solution to the minimal surface problem. In the context of surfaces of revolution generated by rotating a curve y = y(x) (with y > 0) about the x-axis, the condition H = 0 reduces to a second-order ordinary differential equation for the generating curve: y'' = \frac{1 + (y')^2}{y}. This equation arises from the Euler-Lagrange equation applied to the surface area functional A = 2\pi \int y \sqrt{1 + (y')^2} \, dx, whose minimizers are precisely the surfaces with zero mean curvature. The catenary curve y(x) = a \cosh\left( \frac{x - b}{a} \right) (for constants a > 0 and b) satisfies this differential equation, generating the catenoid upon rotation. Among complete minimal surfaces in \mathbb{R}^3 that are rotationally symmetric about an , the is the only non-trivial example, up to scaling and translation; the plane is the trivial case. This uniqueness follows from the analysis of embedded minimal surfaces with finite total and at most two ends, where forces the profile to be a .

Gaussian Curvature Analysis

The Gaussian curvature K of the catenoid, parametrized in standard form as \mathbf{r}(u, v) = (a \cosh(v/a) \cos u, a \cosh(v/a) \sin u, v), is given by K = -\frac{1}{a^2 \cosh^4(v/a)}, where a > 0 is the scaling parameter and (u, v) \in [0, 2\pi) \times \mathbb{R}. This expression reveals that K < 0 everywhere on the surface for finite v, with K approaching 0 as |v| \to \infty, reflecting the catenoid's asymptotic flatness at its ends. The negative sign of the Gaussian curvature implies that the catenoid possesses locally hyperbolic geometry, akin to spaces of constant negative curvature where geodesics diverge more rapidly than in Euclidean space./07%3A_Geometry_on_Surfaces/7.01%3A_Curvature) As |v| increases toward infinity, the diminishing magnitude of K indicates a transition to nearly Euclidean behavior, consistent with the surface's cylindrical topology at large distances from the neck. Since the Gaussian curvature is nowhere zero, the catenoid is not developable, meaning it cannot be isometrically mapped onto the plane without tearing or stretching. Geodesics on this surface of revolution, which represent the shortest paths between points, satisfy Clairaut's relation: the product of the radial distance from the axis of rotation and the cosine of the angle between the geodesic and the meridian is constant along each geodesic. This relation facilitates the explicit solution of geodesic equations, revealing paths that may spiral or asymptote toward the ends without crossing the minimal neck radius. The total Gaussian curvature of the complete catenoid, obtained by integrating K over the entire surface, equals -4\pi. By the Gauss-Bonnet theorem, this finite value aligns with the surface's topology as a genus-zero surface with two punctures (Euler characteristic 0, adjusted for the ends), confirming its classification among minimal surfaces of finite total curvature.

Stability Conditions

The stability of the catenoid as a minimal surface with fixed boundary conditions is determined by the second variation of the area functional. For a compactly supported normal variation field f vanishing on the boundary, the second variation is \delta^2 A = -\int_\Sigma f (\Delta_\Sigma + |A|^2) f \, dA, where |A|^2 is the squared norm of the second fundamental form and the Jacobi operator is L = \Delta_\Sigma + |A|^2. The catenoid is stable if all eigenvalues of L (with Dirichlet boundary conditions) are non-negative, ensuring the quadratic form is non-negative; negative eigenvalues indicate instability directions. Due to the catenoid's axial symmetry, the Jacobi operator separates into Fourier modes in the angular direction and a one-dimensional ordinary differential operator in the meridional coordinate v, yielding the Schrödinger-type equation J_b \psi = -\partial_v^2 \psi - 2 \cosh^{-2}(v + b) \psi = \lambda \psi on the interval corresponding to the surface height, where b parameterizes the neck size. The eigenvalues \lambda of this Darboux-Pöschl-Teller operator govern stability: positive \lambda for all modes imply stability, while a negative \lambda signals instability, with the lowest mode often corresponding to axisymmetric pinching. For coaxial circular boundaries of equal radius a separated by distance h, catenoid solutions exist only for h \leq h_c \approx 1.325 a (equivalently, the normalized half-separation h/(2a) \approx 0.6627), where h_c is the maximum height at which the two branches coincide. Below h_c, the outer (larger neck radius) is stable with all positive eigenvalues, while the inner catenoid (smaller neck) is unstable. At h = h_c, the Jacobi operator acquires a zero eigenvalue with eigenfunction corresponding to infinitesimal pinching, beyond which no catenoid exists and the catenoid ceases to be area-minimizing. In this regime, when the neck radius falls below the critical value (corresponding to h > h_s < h_c, where the area of the inner catenoid equals that of two disks), the Goldschmidt discontinuous solution—two flat disks of radius a spanning the boundaries, joined by a vertical line segment of measure zero—becomes the global area minimizer, as its area $2\pi a^2 is strictly smaller than any continuous spanning surface. This solution emerges as the weak limit of sequences of pinching catenoids and highlights the non-regularity of minimizers for close boundaries. The Morse index of the catenoid, the number of negative eigenvalues of the Jacobi operator (or the dimension of the maximal subspace where \delta^2 A < 0), is 1 for the unstable inner branch between the area-equality threshold h_s and h_c, positioning it as a saddle point in the moduli space of area-stationary surfaces with fixed topology.

Physical Realizations and Applications

Soap Film Experiments

In the mid-19th century, Belgian physicist conducted pioneering experiments with soap films to explore minimal surfaces, including the catenoid formed between two coaxial parallel circular rings of equal radius. By dipping the rings into a soap solution and adjusting their separation, Plateau observed that the film assumes a catenoidal shape when the distance between the rings is less than a critical value of approximately 1.325 times the ring radius, driven by surface tension seeking to minimize the film's area. This configuration aligns with the zero mean curvature condition for equilibrium, as the soap film's shape balances the pressures across its thin boundary layers. When the ring separation exceeds this critical distance, the catenoid becomes unstable and collapses, with the film rupturing at the narrow waist and retracting to form two separate flat circular disks spanning each ring. This discontinuous transition corresponds to the Goldschmidt solution, a minimal area configuration identified by mathematician Carl Wolfgang Benjamin Goldschmidt in 1831, where the total surface area is lower than that of the unstable catenoid despite the topological change. Plateau's observations provided empirical validation for Leonhard Euler's 1744 theoretical prediction that the catenoid, as a surface of revolution generated by a catenary curve, achieves minimal area among such forms bounded by two circles. These experiments have enduring value in education and research, confirming foundational principles of minimal surfaces through accessible physical demonstrations. Modern variants often employ rigid wire frames to maintain ring geometry during separation adjustments, allowing precise in-situ measurements of film thickness and dynamics near instability. Digital simulations, such as those using the Surface Evolver software, further visualize the catenoid's evolution and collapse without physical soap solutions, highlighting surface tension's role in area minimization and facilitating parametric studies of boundary conditions.

Architectural and Engineering Uses

The catenoid shape, derived from the inversion of the catenary curve, has been employed in architectural arches to achieve structural efficiency under compression, minimizing material use while distributing loads evenly. German architect and engineer pioneered the application of such forms in tensile structures, inverting catenary principles to inform lightweight membrane designs. A seminal example is the roof of the (1972), where Otto utilized soap film models—analogous to small-scale catenoid experiments—to generate a vast, undulating acrylic-covered cable-net canopy spanning over 74,800 square meters with minimal material, enabling transparency and integration with the landscape. This approach leveraged the natural equilibrium of inverted catenary curves for wind resistance and aesthetic fluidity. In engineering, catenoid-inspired forms approximate efficient shell structures, particularly in cooling towers where the shape optimizes airflow and compressive strength. Early 20th-century hyperboloid lattice towers by Russian engineer Vladimir Shukhov, such as the 1919 Shabolovka Radio Tower in Moscow, served as precursors, using ruled hyperbolic surfaces in steel gridshells up to 160 meters tall, which halved material requirements while enhancing rigidity against buckling. These designs exploited the double curvature for self-bracing, influencing later concrete hyperboloid cooling towers that dominate industrial landscapes for their thermal efficiency and reduced construction costs. Recent analyses propose pure catenoid geometries as alternatives to hyperboloids for chimneys, demonstrating slightly lower bending moments (~3%) but increased radial stresses, with potential for material efficiency in optimized designs exceeding 100 meters in height. Contemporary parametric design tools have expanded catenoid applications in saddle-shaped roofs, enabling complex geometries that enhance environmental performance. Software like Rhino and Grasshopper facilitates the generation of catenoid-based surfaces through algorithmic modeling of minimal surfaces, as seen in experimental pavilions and canopies where the form provides superior wind resistance—up to 30% better load distribution than planar roofs—and material savings through optimized paneling. These designs prioritize sustainability, with the catenoid's zero mean curvature ensuring even stress, as in conceptual urban roofs that integrate photovoltaic elements on curved profiles for improved aerodynamics. A notable example is the National Taichung Theater (2016) in Taiwan, designed by Toyo Ito, which incorporates catenoid forms in its curved structural volumes to create interconnected performance spaces. Despite these advantages, scaling catenoid structures to large dimensions poses challenges related to inherent instability, necessitating reinforcements to prevent collapse under external loads. Pure catenoids exhibit a critical aspect ratio beyond which they become unstable, as observed in physical models where taller forms revert to planes or without boundary supports; in engineering, this requires additional bracing or hybrid materials. Such limitations underscore the need for computational simulations in design to balance the shape's efficiency with practical durability.

Helicoid Duality

Bonnet's theorem establishes a one-parameter family of minimal surfaces, known as the associate family, obtained by rotating the coordinate functions of a given minimal surface by an angle θ in the complex plane via the Weierstrass representation. For the catenoid, parameterized as \mathbf{X}(u,v) = (a \cosh v \cos u, a \cosh v \sin u, a v), the associates form a continuous deformation where θ = 0 yields the catenoid itself and θ = π/2 produces the \mathbf{X}(u,v) = (a \sinh v \cos u, a \sinh v \sin u, a u). All members of this family are isometric to the original catenoid, preserving lengths and angles while maintaining minimality. The explicit parametric equations for the associate family starting from the catenoid are given by \begin{aligned} x(u,v) &= a (\cos \theta \cdot \cosh v \cos u - \sin \theta \cdot \sinh v \sin u), \\ y(u,v) &= a (\cos \theta \cdot \cosh v \sin u + \sin \theta \cdot \sinh v \cos u), \\ z(u,v) &= a (v \cos \theta + u \sin \theta), \end{aligned} where θ parameterizes the family and (u, v) are the original coordinates; a variant form adjusts the hyperbolic functions to emphasize the limit. These equations demonstrate how the catenoid "unwinds" into the as θ increases to π/2, with intermediate surfaces bridging the two geometries. Geometrically, the duality arises from a 90-degree rotation in the normal plane of the surface, transforming the catenoid's conjugate directions to yield the helicoid as its associate at θ = π/2. All associates in the family share the same Gauss map, meaning they have identical normal vectors at corresponding points, which underscores their isometric relation and common minimal properties. Meusnier identified both the catenoid and helicoid as minimal surfaces in 1776. Their connection as associates in a one-parameter family of minimal surfaces was established by Pierre Bonnet in 1867. In modern minimal surface theory, this associate family plays a crucial role in solving the Plateau problem by enabling isometric deformations that span boundary conditions between catenoid-like and helicoid-like configurations.

Bonnet Transformation Effects

The Bonnet transformation, also known as the associate family construction, generates a one-parameter family of minimal surfaces from a given minimal surface by rotating its Weierstrass data through an angle θ ∈ [0, π/2]. For the catenoid, this deformation produces a continuous family of immersed minimal surfaces that interpolate between the catenoid and the , preserving the minimal surface property (zero mean curvature) across all members. These associate surfaces are obtained by applying the transformation to the catenoid's Weierstrass representation, where the Gauss map g(z) = z and the height differential dh = dz/z on the punctured complex plane ℂ \ {0}. All surfaces in the Bonnet family of the catenoid share the same Gauss map and the same conformal metric structure, ensuring that the induced metric differs only by a conformal factor independent of θ. This preservation extends to the total Gaussian curvature, which remains -4π for the entire family, reflecting the catenoid's topology of genus zero with two ends. The vertical flux is also conserved up to rotation, maintaining key geometric invariants while allowing the surface to deform isometrically in the parameter domain. In the limiting cases, as θ approaches 0, the associate surface converges to the original catenoid, while as θ approaches π/2, it approaches the , which serves as the conjugate surface in the family. This duality highlights the catenoid-helicoid pair as endpoints of the deformation. Computationally, the family is generated using the , where the immersion is given by integrals involving the rotated differentials: for parameters z = u + iv, the coordinates are X(u,v) = Re ∫ (φ₁, φ₂, φ₃) dz, with φ₁ - iφ₂ = (1 - g²)/2 * (dh / g), φ₃ = dh, adjusted by e^{iθ} for the associates; this approach leverages complex analysis to visualize and construct the intermediate minimal surfaces.

Advanced Mathematical Aspects

Critical Catenoid Conjecture

The Critical Catenoid Conjecture posits that the critical catenoid is the unique embedded free boundary minimal annulus in the unit ball B^3 \subset \mathbb{R}^3. The critical catenoid is a specific scaling of the catenoid that meets the unit sphere orthogonally, serving as a free boundary minimal surface with two boundary components on the sphere. This conjecture, part of broader questions in free boundary minimal surface theory, highlights the catenoid's role in classifying low-topology minimal surfaces with boundary constraints. The conjecture remains a significant open problem, with partial results establishing uniqueness under additional assumptions, such as σ-homothetic deformations or overdetermined eigenvalue conditions on the support function. In 2023, Espinar and others characterized the critical catenoid as the only such annulus whose support function satisfies certain infinite-order conditions at the boundary. A 2025 result by Liu and Wang proved uniqueness for every embedded free boundary minimal annulus in the ball using reflection principles. These advances draw on tools from elliptic PDEs, Steklov eigenfunctions, and symmetry methods, connecting to Fraser-Schoen's work on extremal eigenvalues. Numerical and variational evidence supports the conjecture, excluding other configurations like perturbed annuli. Related to the Plateau problem for two coaxial rings of radius r separated by height h, the catenoid provides a minimal surface solution for h < h_c(r) \approx 1.325 r, the critical height beyond which no catenoid exists. For sufficiently small h, geometric measure theory results imply the catenoid is the unique area-minimizing surface among connected spanning surfaces, via symmetry (moving plane method) and regularity theorems. However, as h increases, the catenoid becomes unstable, and the global minimizer is the Goldschmidt discontinuous solution consisting of two flat disks, which has smaller area than the catenoid near h_c. Stability analysis via the second variation shows positive eigenvalues for small h, confirming local minimality. This framework bridges classical variational problems—solving the Euler-Lagrange equation for surfaces of revolution—with modern tools like varifolds and currents for handling singularities and minimality.

Asymptotic and Uniqueness Results

The catenoid exhibits distinct asymptotic behavior at its ends as the axial coordinate |z| \to \infty. In its standard parametrization, the surface is generated by rotating the catenary curve r(z) = a \cosh(z/a) around the z-axis, where a > 0 is a scaling parameter. For large |z|, \cosh(z/a) \sim \frac{1}{2} e^{|z|/a}, so the radius grows exponentially as r(z) \sim \frac{a}{2} e^{|z|/a}. This implies that each end flares outward, approaching a geometry where the approaches zero. The area of the catenoid diverges as the ends extend to , with linear growth in the extrinsic characterizing the ends due to the finite total of -4\pi. Specifically, the area up to a large R on one end is asymptotically $2\pi R, akin to cylindrical growth, distinguishing it from the quadratic growth of planar ends. In end expansions, logarithmic terms appear in the asymptotic description of Jacobi fields associated with deformations, reflecting the slow convergence to the asymptotic in the varifold sense. A key uniqueness result establishes the catenoid as complete of revolution in \mathbb{R}^3, up to scaling and the plane itself. This follows from Schoen's 1983 , which proves that among complete minimal surfaces of finite total with two ends, only the plane (one end, effectively) and the catenoid satisfy the conditions, regardless of (higher ruled out for two ends). The proof relies on the property of the Gauss map, positive mapping properties, and monotonicity formulas to enforce and exclude other topologies. Bernstein-type results further contextualize the catenoid's role in theory. The Bernstein theorem asserts the non-existence of non-flat entire minimal graphs in \mathbb{R}^3, implying that the only complete minimal graphs over \mathbb{R}^2 are planes, proved via elliptic regularity and maximum principles. The catenoid serves as a fundamental illustrating that this uniqueness fails for non-graphic minimal surfaces, as it is complete, , and non-flat but cannot be expressed as a global graph over any plane due to its and flaring ends. Recent developments highlight the catenoid's influence on broader conjectures, particularly in . The Lawson conjecture, posed in 1970, proposed that the Clifford torus is the unique embedded minimal torus in the S^3. Proved by in 2007 using partial differential equations and curvature estimates, the result draws methodological parallels to the catenoid's uniqueness via Gauss map analysis and symmetry arguments, underscoring the catenoid's prototypical role in classifying low-genus minimal surfaces. Post-2000 extensions, including works on higher-genus surfaces, continue to leverage catenoid end constructions in S^3 via .

References

  1. [1]
    [PDF] Characterizations of Complete Embedded Minimal Surfaces: Finite ...
    Mar 15, 2005 · The catenoid is a surface obtained by rotating a catenary around the z-axis. It is the only minimal surface of revolution, and can also be ...<|control11|><|separator|>
  2. [2]
    [PDF] The Morse Index of a Minimal Surface - Penn Math
    The catenoid (discovered by Euler in 1744 and proven to be minimal by Meusnier in 1776) is the unique axially symmetric minimal surface in ℝ3. Figure 2. The ...
  3. [3]
    Minimal Surfaces and Soap Bubbles - Berkeley Math
    For example, a minimal surface that spans two rings is given by the catenoid surface below: Notice that this is not the only possible minimal surface: another ...
  4. [4]
    Catenary -- from Wolfram MathWorld
    The curve is also called the alysoid and chainette. The equation was obtained by Leibniz, Huygens, and Johann Bernoulli in 1691 in response to a challenge by ...Missing: derivation | Show results with:derivation
  5. [5]
    [PDF] Catenary
    We'll present four solutions. The first one involves balancing forces. The other three involve various variations on a variational argument. First solution: Let ...
  6. [6]
    Catenary - MacTutor History of Mathematics - University of St Andrews
    The catenary is the shape of a perfectly flexible chain suspended by its ends and acted on by gravity. Its equation was obtained by Leibniz, Huygens and Johann ...Missing: derivation | Show results with:derivation
  7. [7]
    Methodus inveniendi lineas curvas maximi minimive proprietate ...
    May 25, 2018 · Methodus inveniendi lineas curvas maximi minimive proprietate gaudentes, sive, Solutio problematis isoperimetrici latissimo sensu accepti.
  8. [8]
    [PDF] calculus of variations: minimal surface of revolution - UChicago Math
    We will first present a classical catenoid solution using calculus of vari- ation, and we will then discuss the conditions of existence by considering the ...Missing: sources | Show results with:sources<|control11|><|separator|>
  9. [9]
    [PDF] Minimal surfaces in Euclidean spaces by way of complex analysis
    Sep 3, 2021 · The subject has a luminous history, going back to 1744 when Leonhard Euler. [32] showed that pieces of the surface now called catenoid (see ...
  10. [10]
    [PDF] differential geometry hw 4 - Colorado State University
    Proof. Let X(u, v) = (acoshv cosu, acoshv sinu, av) be the parametriza- tion of the catenoid and let Y (z, w)=(w cosz,w sinz, az) be the parametriza-.
  11. [11]
    [PDF] Differential Geometry
    For the catenoid ϕ(u, v)=(a cosh v cos u, a cosh v sin u, av), we have E = G = a2 cosh2 v, F = 0, so ϕ is isothermal. Moreover, ∇2ϕ = 0, so this is a ...
  12. [12]
    Catenoid -- from Wolfram MathWorld
    A catenary of revolution. The catenoid and plane are the only surfaces of revolution which are also minimal surfaces.
  13. [13]
    Minimal Surface -- from Wolfram MathWorld
    Minimal surfaces are defined as surfaces with zero mean curvature. A minimal surface parametrized as x=(u,v,h(u,v)) therefore satisfies Lagrange's equation.
  14. [14]
    Minimal Surface of Revolution -- from Wolfram MathWorld
    which is the equation for a catenary. The surface area of the catenoid product by rotation is. A, = 2piintxsqrt(1+y^('2))dx=2piintxsqrt(1. (83). = 2piintx/(sqrt ...
  15. [15]
    UNIQUENESS, SYMMETRY, AND EMBEDDEDNESS OF MINIMAL ...
    In 1962, J. C. C. Nitsche [8] showed that the catenoid is the only complete minimal surface in R3 which intersects each plane parallel to a given plane ...
  16. [16]
    9.7.1 Differential geometry of developable surfaces - MIT
    2). Since the Gaussian curvature of a developable surface is zero everywhere [412,76], the maximum and minimum principal curvatures (3.49) and (3.50) of a ...
  17. [17]
    [PDF] m435: introduction to differential geometry
    In fact Clairaut's relation holds for more general surfaces of revolution r(u, v)=(f(u) cosv, f(u) sinv, g(u)), where f,g parametrise a smooth curve in the x-z ...
  18. [18]
    Stability of the Classical Catenoid and Darboux–Pöschl–Teller ...
    Oct 29, 2022 · Indeed, the stability operator of the catenoid, arising from the second variation of the area functional, is of Schrödinger-type with a ...<|control11|><|separator|>
  19. [19]
  20. [20]
    Booklet on Minimal Surfaces - Freie Universität Berlin
    Apr 23, 2013 · The Belgian physicist Joseph Antoine Ferdinand Plateau (1801-1883), a Professor at Ghent, carried out extensive experiments with soap films in ...
  21. [21]
    [PDF] In-situ observation of a soap film catenoid - arXiv
    Dec 22, 2009 · When the surface area of a stable catenoid equals the surface area of the two disks, the rings are separated by h0 ≃ 1.05R. We are interested ...
  22. [22]
    Geometry of catenoidal soap film collapse induced by boundary ...
    Sep 8, 2021 · These analyses are used to show how much of the geometry of collapsing catenoids is accurately captured by a few active modes triggered by ...
  23. [23]
    DISCONTINUOUS SOLUTIONS IN THE CALCULUS OF ...
    In 1831 Goldschmidtf discovered the very interesting and well known example in which the curve generating a surface of revolution of minimum area is composed ...
  24. [24]
    The Catenoid - Minimal Surfaces
    The Catenoid stands at the beginning of the theory of minimal surfaces. Leonhard Euler, in 1744, showed that among surfaces of revolution, it has minimal area.<|control11|><|separator|>
  25. [25]
    Influence of boundary conditions on the existence and stability of ...
    Sep 1, 2014 · We note that the critical distance hc of a diabloid is relatively smaller than that of the associated catenoid for which Δc ≈ 1.33.Missing: exact | Show results with:exact
  26. [26]
    Catenoid Soap Films - Ken Brakke
    The soap film that forms between two parallel circular rings has the shape of a catenoid, which has the equation r = cosh(z) in cylindrical coordinates.
  27. [27]
    Minimal Surfaces and Architecture: New Forms
    Apr 30, 2013 · (In the sixties the German architect Frei Otto started to experiment with soap films. He had in mind to use the models of minimal surfaces to ...
  28. [28]
    Design and construction of catenary-ruled surfaces - ScienceDirect
    This paper proposes a new type of curve-ruled surfaces, termed catenary-ruled surfaces, which can be conveniently designed using catenary rulings and ...
  29. [29]
  30. [30]
    Contemporary and Parametric Geometries in Architecture
    Nov 5, 2020 · A list of basic parametric geometries used in contemporary architectural design processes ... Examples of minimal surfaces include catenoids ...Calculus To Topology · Spline · Minimal Surfaces (zero Mean...
  31. [31]
    [PDF] Minimal Surfaces with Catenoid Ends - arXiv
    Apr 26, 2008 · In this paper, we use the conjugate surface construction to prove the existence of certain non-periodic symmetric immersed minimal surfaces.
  32. [32]
    [PDF] chapter 3: minimal surfaces - UChicago Math
    Feb 10, 2019 · Second fundamental form and mean curvature. A parametric surface Σ in Rn is just a smooth map from some domain Ω in R2 to Rn. Let's denote ...
  33. [33]
    [PDF] The classical theory of minimal surfaces - Universidad de Granada
    Feb 2, 2011 · In 1776, Meusnier verified that the catenoid is a solution of. Lagrange's equation (1). This surface has genus zero, two ends and total ...
  34. [34]
  35. [35]
    [PDF] Bernstein type theorem for entire weighted minimal graphs in ...
    x(u, v) = cos θ sinh v sin u + sin θ cosh v cosu,. (3) y(u, v) = −cos θ sinh v cos u + sin θ cosh v sin u,. (4) z(u, v) = u cosθ + v sin θ;. (5) where −π<u ...
  36. [36]
    Minimal surfaces with catenoid ends - MSP
    The method of the proof of Theorem 3.1 can be used to prove the existence of other immersed finite total curvature surfaces with catenoid ends and genus greater ...
  37. [37]
    Helicoidal minimal surfaces in the 3-sphere: An approach via ... - arXiv
    Mar 1, 2024 · The helicoid was first proved to be minimal by Meusnier in 1776. Catalan showed in 1842 that the helicoid, together with the plane, are the ...
  38. [38]
    [PDF] minimal surfaces of least total curvature
    ... surface, see [Cos1]. For finite minimal surfaces, there is a Gauß-Bonnet formula relating the total curvature to genus and winding numbers: ∫RKdA = 2π ...<|control11|><|separator|>
  39. [39]
    THE GEOMETRY OF MINIMAL SURFACES OF FINITE GENUS I
    So the bottom end of M is a catenoid end of negative logarithmic growth λ1. It is clear from the Ordering. Page 43. THE GEOMETRY OF MINIMAL SURFACES. 43.
  40. [40]
    A proof of the Lawson conjecture for minimal tori embedded in $\S3
    Mar 5, 2007 · In 1970 Lawson conjectured that the Clifford torus is, up to congruences, the only torus minimally embedded in \S3. We prove here Lawson ...Missing: catenoid | Show results with:catenoid