Fact-checked by Grok 2 weeks ago

Catenary

A catenary is the curve assumed by a perfectly flexible, , and inextensible or cable when suspended from two points and acted upon solely by a , supporting its own weight without rigidity. This shape arises naturally from the balance of and along the filament, forming a , U-like profile that minimizes . Mathematically, the catenary is described by the equation y = a \cosh\left(\frac{x}{a}\right), where a is a positive constant determined by the linear density of the chain and the horizontal tension at the lowest point, and \cosh denotes the hyperbolic cosine function. This form reflects the curve's intrinsic properties; for example, the tangential tension at any point is T = \frac{H}{a} y, where H is the horizontal tension, and the arc length from the vertex is s = a \sinh\left(\frac{x}{a}\right). The catenary differs from a parabola, which it superficially resembles but only approximates under certain conditions, like shallow sags. Historically, the catenary was first investigated by in the early , who incorrectly conjectured it to be parabolic based on observations of hanging ropes. The correct hyperbolic form was independently derived in 1691 by , , and as part of a challenge posed by Jakob Bernoulli, marking a key advancement in the . In and , the catenary plays a crucial role; for instance, the main cables of suspension bridges like the follow a catenary under their own weight before the roadway load transforms the shape toward a parabola. Inverted catenaries provide stable, compression-resistant forms for arches and vaults, as seen in structures like the in , which uses a shape to efficiently distribute weight without tensile stress. These applications highlight the catenary's optimality in load-bearing designs under , enabling pure compression in inverted forms.

Overview

Definition and Physical Interpretation

A catenary is the curve formed by an idealized, perfectly flexible and inextensible or cable of uniform suspended from two points and acted upon solely by its own weight in a uniform . This configuration assumes the chain has negligible thickness and that the supports contribute no significant mass or rigidity to the system. Physically, the catenary represents the shape achieved when the within the chain at every point precisely counteracts the downward gravitational force tangent to the curve, ensuring no net or moments act on any . In this state, the horizontal component of remains constant throughout, while the vertical component varies to support the accumulating weight of the below each point. The resulting profile minimizes the of the system under the given constraints. The catenary appears as a , symmetric U-shaped , distinct from a parabola in its profile: it is flatter (shallower) near the at the bottom and steeper approaching the endpoints. This subtle difference arises because the weight distribution along the , rather than horizontally, governs the shape. In and , the catenary manifests in sagging overhead lines between utility poles or in loosely hanging vines supported at intervals, where the idealized conditions approximate real behavior closely enough to produce the characteristic .

Basic Geometric Properties

The arc length s of a catenary curve y = a \cosh(x/a) from the vertex at (0, a) to a point (x, y) is given by s = a \sinh(x/a). This formula arises from integrating the differential arc length element ds = \sqrt{1 + (dy/dx)^2} \, dx, where the integrand simplifies to \cosh(x/a) due to the identity $1 + \sinh^2(u) = \cosh^2(u), yielding the hyperbolic sine upon integration. The to the catenary has dy/dx = \sinh(x/a), which underscores its hyperbolic character. As |x| increases, |\sinh(x/a)| grows exponentially without bound, causing the \theta with the —where \tan \theta = \sinh(x/a)—to approach \pm 90^\circ, reflecting the curve's asymptotic steepening toward vertical orientation while the curve itself rises exponentially. The \rho for the catenary is \rho = a \cosh^2(x/a), or equivalently \rho = a / \cos^2 \theta in terms of the tangent angle \theta. This expression indicates that \rho achieves its minimum value of a at the (x=0, \theta=0), where the is greatest, and increases monotonically away from the , corresponding to progressively gentler bending as the curve extends. Rotating the catenary about its directrix (the x-axis) produces the , a recognized as a with zero . This property is vividly illustrated by films, which naturally form catenoids when stretched between two circular rings, minimizing surface area under .

Historical Development

Early Observations and Uses

The catenary curve, the natural shape assumed by a freely hanging uniform chain or rope under , was observed empirically in ancient for its structural efficiency. In the AD, the Sasanian Taq-i Kisra (Arch of ) in featured an inverted catenary form in its massive unreinforced vault, spanning 25 meters wide and rising 37 meters high, constructed without centering to achieve optimal compression and stability. This design demonstrated an intuitive recognition of the curve's ability to distribute loads evenly, predating mathematical descriptions. While Roman engineers favored semi-circular arches in aqueducts and bridges like the (built in 62 BC), these structures relied on similar principles of for , though their shapes approximated rather than precisely followed the catenary; the thick voussoirs and spandrels compensated for deviations from the ideal inverted form. In practical applications, hanging ropes and chains were noted for even load distribution in early systems and operations during medieval times, where chains hoisted materials and naturally formed the to minimize without analytical intervention. During the , (1452–1519) sketched hanging ropes and chains in his notebooks as part of his mechanical studies, capturing the curve's form intuitively while exploring . Medieval and Renaissance tent designs also incorporated the shape empirically, with rope-spread pavilions exhibiting catenary sag in their roof profiles for tautness and wind resistance. The transition to scientific consideration began in the late with treatises on . Flemish engineer (1548–1620) discussed equilibrium in hanging systems in his 1586 work De Beghinselen der Weeghconst (The Art of Weighing), using a "wreath of beads" model to illustrate balanced forces in chain-like configurations, laying groundwork for later analysis without deriving the curve's equation. In 1669, Joost Jungius experimentally disproved Galileo's parabolic conjecture for the catenary using a hanging chain.

Mathematical Formulation and Key Contributors

The mathematical formulation of the catenary curve originated in the late as part of the burgeoning field of and variational methods. This work coincided with a challenge posed by his brother, , to derive the equation of the "chainette" or hanging chain curve. Independently, , , and solved the problem in 1691, obtaining the curve via of the condition for a uniform flexible chain under . Their subsequent in 1692–1693 refined the , confirming the explicit form of the as the graph of the hyperbolic cosine function: y = a \cosh\left(\frac{x}{a}\right) where a is a determined by the chain's and . Huygens had earlier introduced the term "catenaria" in a 1690 to Leibniz, emphasizing its distinct non-parabolic nature. During the , Leonhard Euler significantly expanded the catenary's mathematical framework, applying infinite series expansions and integral representations to solve related variational problems, including the derivation of minimal surfaces from rotated catenaries in his 1744 memoir. Euler's methods integrated the catenary into broader theories of curves and surfaces, using series to approximate solutions and integrals to express arc lengths and evolutions under constraints.

Core Mathematical Description

Standard Equation and Parametric Forms

The standard equation of the catenary curve, describing the shape assumed by a uniformly dense, flexible chain suspended from two points under gravity, is given by y = a \cosh\left(\frac{x}{a}\right), where the origin is placed at the lowest point of the curve, x is the horizontal coordinate, y is the vertical coordinate measured upward, and a > 0 is a scaling parameter that controls the curve's width and sag. The parameter a is physically interpreted as a = T_0 / (\rho g), with T_0 denoting the horizontal tension at the vertex, \rho the linear mass density of the chain, and g the gravitational acceleration; larger values of a correspond to shallower curves. For practical suspended chains between supports at unequal heights or offset positions, the equation is translated and scaled accordingly, such as y = a \cosh\left(\frac{x - x_0}{a}\right) + y_0, to match boundary conditions without altering the intrinsic shape. Alternative representations include and inverse forms derived from identities. One such expression utilizes the definition of the , particularly in the normalized case where a = 1: y = \frac{e^{x} + e^{-x}}{2}. This form, equivalent to y = \cosh x, explicitly shows the catenary's relation to and decay. The inverse relation, solving for x in terms of y, yields x = \pm a \arccosh\left(\frac{y}{a}\right), useful for computations involving vertical coordinates, such as determining horizontal span from height. A common hyperbolic parametric form is x = a \sinh t, y = a \cosh t, where t serves as a parameter related to the slope angle, satisfying y^2 - x^2 = a^2 and tracing the curve as t varies from -\infty to \infty. For inverted catenaries in structural applications like arches, the curve is reflected to form a downward-opening under , often expressed as y = b - a \cosh\left(\frac{x}{a}\right), where b is a vertical shift ensuring the maximum at x = 0, or in a normalized variant y = a \left( \cosh\left(\frac{x}{a}\right) - 1 \right) shifted appropriately for positive heights from the base. Similar weighted forms, such as flattened catenaries y = A \cosh(B x) with A B \neq 1, adjust the parameter ratio to approximate real-world variations in material uniformity. The dimensionless form of the catenary, obtained by normalizing coordinates such that \tilde{x} = x / a and \tilde{y} = y / a, simplifies to \tilde{y} = \cosh \tilde{x}, underscoring the curve's scale invariance: all catenaries are geometrically similar, differing only by the choice of a, which absorbs units of length related to physical scales like tension and density. This universality facilitates analysis across diverse systems, from microscopic chains to large-scale cables.

Derivation from Equilibrium Conditions

The catenary represents the equilibrium shape of a uniform, inextensible or hanging freely under , with its weight acting vertically downward. To derive the governing equation from static , consider the forces on segments of the chain, assuming constant weight per unit length w (the product of linear mass density and ) and no other external forces besides and . The component of remains along the chain, denoted T_0, due to the absence of horizontal forces. For a portion of the chain from its lowest point (the vertex) to a point at horizontal distance x, the vertical component of tension at that point balances the total weight w s of the arc length s subtended by that portion. The tension T at angle \theta to the horizontal satisfies T \cos \theta = T_0 and T \sin \theta = w s, yielding \tan \theta = (w s)/T_0. Since \tan \theta = dy/dx where y(x) describes the curve height, and ds = \sqrt{1 + (dy/dx)^2} \, dx, the arc length relates to the slope. To obtain the differential equation, examine the force balance on an infinitesimal element spanning horizontal distance dx. The net vertical force must be zero in equilibrium, so the change in vertical tension component equals the element's weight w \, ds = w \sqrt{1 + (dy/dx)^2} \, dx. The vertical tension component is T_0 \, (dy/dx), so differentiating gives T_0 \, d^2 y / dx^2 = w \sqrt{1 + (dy/dx)^2}. Defining the characteristic length a = T_0 / w, this simplifies to \frac{d^2 y}{dx^2} = \frac{1}{a} \sqrt{1 + \left( \frac{dy}{dx} \right)^2}. Solving this second-order nonlinear begins by letting z = dy/dx, yielding dz/dx = (1/a) \sqrt{1 + z^2}. Separating variables produces a \, dz / \sqrt{1 + z^2} = dx. Integrating both sides gives a \sinh^{-1} z = x + C_1, so z = \sinh((x + C_1)/a). For symmetry with a minimum at x = 0 where dy/dx = 0, C_1 = 0, hence dy/dx = \sinh(x/a). Integrating again yields y = a \cosh(x/a) + C_2. Choosing the origin such that y(0) = a sets C_2 = 0, resulting in the catenary y = a \cosh(x/a). The a physically quantifies the curve's , increasing with greater or lower weight for a flatter profile. An equivalent derivation uses a free-body diagram on a small chain element of length ds, with tensions T and T + dT at angles \theta and \theta + d\theta. Horizontal equilibrium implies d(T \cos \theta) = 0, confirming constant T_0 = T \cos \theta. Vertical equilibrium gives d(T \sin \theta) = w \, ds, and substituting ds = dx / \cos \theta with \tan \theta = dy/dx recovers the same differential equation as before. This approach highlights the local force resolution leading to the global curve shape.

Engineering and Architectural Applications

Inverted Catenary Arches

An inverted is formed by rotating the standard catenary curve, which describes the shape of a hanging chain under its own weight, by 180 degrees; this inversion converts the tensile forces acting along the chain into compressive forces that support the structure without inducing bending moments. The principle relies on the equilibrium of forces where the curve ensures that the line of thrust—the path of resultant compressive forces—remains within the arch's cross-section, providing inherent stability under uniform loading. In 1675, English scientist first articulated this concept, proposing that the ideal shape for a arch mirrors the inverted profile of a flexible in , encapsulated in his and later phrase: "As hangs the flexible line, so but inverted will stand the rigid arch." This insight laid the groundwork for understanding arches as structures in compression, influencing subsequent architectural analyses. Architect applied inverted catenary principles extensively in the design of the basilica in , using physical models of suspended chains and weights to generate catenary curves, which he then inverted via mirrors and photography to define the forms of columns and arches. These branching columns, inspired by tree-like structures, follow catenary-derived geometries such as hyperboloids and helicoids, enabling load distribution through pure compression and eliminating the need for traditional buttresses. A prominent modern example is the in , , completed in 1965, which adopts the form of a to achieve stability under wind and self-weight loads. Its profile is governed by an of the form y = -68.8 \cosh(0.01 x) + 1, (in feet, with origin at the top and y downward), ensuring the 630-foot-tall stainless-steel structure efficiently transfers forces to its foundations. The use of inverted catenary arches offers significant advantages, particularly in or , by optimizing material efficiency through even load distribution to the abutments and minimizing stresses. This shape allows slender profiles to span large distances while maintaining structural integrity, as the compressive forces align naturally with the curve, reducing the risk of failure from eccentric loading.

Catenary Bridges and Suspension Cables

In tensile structures such as suspension bridges and overhead cables, the catenary curve arises naturally from the of a under its own uniform weight per unit length, with horizontal balancing the vertical gravitational forces. Pure catenary shapes are used in simple suspension bridges without heavy decks, common in pedestrian and small-scale applications, though rare in large vehicular bridges due to additional loads from roadways that introduce non-uniform loading. Examples include stressed ribbon bridges like the Leonel Viera Bridge in . The exact catenary form y = a \cosh\left(\frac{x}{a}\right), where a = \frac{H}{w} (with H as the horizontal and w as the weight per unit length), provides optimal stress distribution only under idealized conditions. In standard suspension bridges, the main cable deviates from a pure catenary because the suspended imposes a uniform horizontal load distribution along the span, transforming the equilibrium shape into a close approximation of a parabola. This parabolic form emerges when the load w is constant per horizontal distance rather than per , yielding y = \frac{w x^2}{2 H}, where x is the horizontal distance from the lowest point and H is the constant horizontal tension. The deviation from catenary to parabola is minimal for long spans with heavy decks, as in the , where the cable self-weight is negligible compared to the roadway load, ensuring efficient load transfer to the towers and anchors. Power lines and guy wires, which support antennas, masts, or transmission towers, adopt a catenary profile to minimize material use while maintaining structural integrity under wind and loads. Sag calculations are critical for ensuring adequate clearance and preventing contact with or structures, employing the for maximum sag d = a \left( \cosh\left(\frac{L}{2a}\right) - 1 \right), where L is the span length and a = \frac{H}{w}. factors, typically 2.0 or higher (limiting tension to divided by the factor), are incorporated to account for temperature variations and dynamic loads, allowing engineers to predict and mitigate excessive sagging that could compromise reliability. Catenary mooring lines for ships and offshore platforms utilize the curve's inherent flexibility to equilibrate horizontal forces from , , and currents against the vertical component of the chain's weight. In a catenary (CALM) system, a is secured by multiple catenary lines anchored to the , forming a shallow sag that absorbs shocks and maintains position without excessive spikes. This configuration, common in single-point s for oil tankers, relies on the catenary's to distribute loads progressively, with the horizontal pull at the balanced by the weighted curve's uplift, enhancing in dynamic environments.

Advanced Properties and Generalizations

Relations to Other Curves and Further Geometrical Traits

The catenary curve exhibits a close relation to the parabola, particularly in engineering contexts where the sag-to-span ratio is small, typically less than 5-10%. In such cases, the parabola serves as a practical approximation to the catenary, with the error in sag calculation being less than 1% for sags around 5% of the span. This approximation stems from the Taylor expansion of the hyperbolic cosine, where higher-order terms become negligible for small arguments, yielding a akin to the parabola y = \frac{x^2}{2a}. However, the catenary emerges as the exact of a polygonal under numerous equal concentrated loads spaced proportionally to the , whereas equal horizontal spacing of those loads yields a parabolic limit; visually, both curves share a U-shape near the , but asymptotically, the catenary's contrasts with the parabola's rise. The catenary is intrinsically connected to , defined as the graph of y = a \cosh\left(\frac{x}{a}\right), where \cosh z = \frac{e^z + e^{-z}}{2} provides the smooth, symmetric profile observed in hanging chains. This hyperbolic form underscores its distinction from trigonometric curves, enabling analytical solutions in variational problems. Furthermore, the catenary possesses notable properties: it traces the path of a parabola's as the parabola rolls without slipping along a straight line. The catenary also acts as the envelope of the curve's normals, linking it to pursuit curves in classical . The evolute of the catenary, which is the locus of its centers of , is another catenary congruent to the original but translated vertically by $2aand reflected. For the [parametric](/page/Parametric) formx = a t, y = a \cosh t$, the is parameterized as \begin{align*} x &= a \left( t - \frac{1}{2} \sinh 2t \right), \\ y &= 2a \cosh^2 t. \end{align*} This self-similar property arises from the catenary's constant behavior relative to its scale parameter. The principal of the catenary is the , obtained by unwrapping a taut from the curve starting at the ; this involute-envolute duality has applications in designing non-circular for variable speed ratios and in for surfaces ensuring uniform paths. Rotating a catenary about its generates the , a ruled with zero , meaning it locally minimizes area among surfaces spanning the same boundary. This zero condition, H = \frac{\kappa_1 + \kappa_2}{2} = 0 where \kappa_1 and \kappa_2 are principal curvatures, positions the as the least-area bridge between two coaxial circular rings. In physical realizations, soap films between such rings adopt the shape to equilibrate , though the configuration is stable only for ring separations less than approximately 0.6627 times the ring diameter, beyond which it destabilizes into separate planar films. Leonhard Euler first characterized the in 1744 as a non-trivial to the equation, highlighting its role in .

Variations with Non-Uniform Conditions and External Forces

In realistic scenarios, the ideal catenary shape deviates when the chain or cable experiences non-uniform conditions or additional external forces beyond uniform . These variations arise in applications where material properties change along the length or environmental loads like or currents influence the configuration. The governing equations are derived by extending the force balance or variational principles from the uniform case, incorporating the specific perturbations. For chains with variable density ρ(x), the equilibrium shape satisfies a generalized differential equation obtained by considering the local weight distribution in the tangential and normal force balances. The horizontal tension T₀ remains constant, leading to the second-order equation \frac{d^2 y}{dx^2} = \frac{\rho(x) g}{T_0} \sqrt{1 + \left( \frac{dy}{dx} \right)^2}, where g is . This form generalizes the uniform catenary by replacing the constant density with ρ(x), often solved numerically for specific profiles like linearly varying ρ(x) in tapered cables used in systems or power lines with accumulation. Examples include offshore risers where density changes due to filling, resulting in shapes that deviate from the hyperbolic cosine, with increased sag in denser regions. In suspension bridges, the cable shape approximates a parabola rather than a pure catenary because the primary load is the uniform horizontal distribution from the roadway weight, not the cable's self-weight alone. Under a uniform load w per unit horizontal distance, the vertical force balance yields the parabolic equation y = \frac{w x^2}{2 H}, where H is the constant horizontal tension and x is measured from the lowest point. This distinguishes from the catenary, as the load is proportional to horizontal span rather than , leading to a profile that simplifies design calculations for long spans like the , where self-weight is negligible compared to deck load. The approximation holds well for shallow sags, with errors under 1% for typical ratios. The catenary of equal strength addresses non-uniform cross-section to maintain constant tensile σ throughout the , with area A proportional to T (A = T / σ). This results in an taper, as varies along the curve, leading to the equations x = (T₀ / w) ( φ) and y = (T₀ / w) φ, where φ is with and w is per unit volume times σ. Originally derived by Davies Gilbert in 1826 for wide-span bridges, this shape minimizes material use while equalizing , though practical implementation is limited by manufacturing challenges; it appears in theoretical designs for suspension bridges like Telford's crossing. When is included, the elastic catenary modifies the ideal shape through resistance to curvature, approximated using Euler-Bernoulli beam theory for small deviations. The governing equation incorporates a term M = -EI d²θ/ds², where E is , I is the , and θ is the local angle, added to the tension balance: dT/ds = ρ g sin θ and dM/ds + T sin θ = 0 (neglecting ). For slight , the shape remains close to the catenary but with reduced sag; seminal by C.Y. in 1982 shows solutions via elliptic integrals for heavy elastica, applied in taut lines or rail catenaries where affects dynamic response. This approximation is valid for slenderness ratios above 100, beyond which pure catenary suffices. Under general external forces modeled as a F with potential V, the equilibrium curve minimizes the total ∫ (V + λ √(1 + y'^2)) dx, where λ is a for constraint if fixed. For non-conservative forces like or currents, the variational form extends to ∫ √(1 + y'^2) ds + ∫ F · r ds, solved via Euler-Lagrange equations yielding d/dx (∂L/∂y') = ∂L/∂y with L incorporating force terms. In railway catenaries, crosswinds introduce lateral forces, displacing the curve laterally, analyzed through finite element methods combining catenary equations with drag coefficients; for ocean currents on moorings, the shape distorts into more complex 3D configurations. These generalizations highlight the catenary's adaptability in fluid-structure interactions.

References

  1. [1]
    [PDF] THE CATENARY Physics 258/259
    A catenary is the shape taken by a uniform chain or string freely suspended from two points. The parameters of this shape for a suspended chain are measured and ...
  2. [2]
    [PDF] the catenary
    HISTORY: Galileo was the first to investigate the catenary. It is the curve, a freely hanging heavy rope describes, if the end points have the same height.
  3. [3]
    [PDF] CHAPTER 2 REVIEW
    a curve in the shape of the function y = a cosh(x/a) is a catenary; a cable of uniform density suspended between two supports assumes the shape of a catenary.
  4. [4]
    [PDF] Catenary
    The catenary curve (from the Latin for “chain”) is the shape assumed by a ... remains true after a scaling, y = a cosh(x/a). This is good because x ...
  5. [5]
    1. Introductory Statics: the Catenary and the Arch
    The word catenary is actually defined as the curve the chain approaches in the limit of taking smaller and smaller links, keeping the length of the chain ...Missing: physics engineering
  6. [6]
    Catenary
    $$n$ -gon, the Cartesian equation of the corresponding catenary is $y=-A\cosh(x/A)$ , where $A\equiv R\cos(\pi/n)$ . \begin{figure}\begin{center}\BoxedEPSF ...
  7. [7]
    [PDF] Catenaries and Suspension Bridges – The Shape of a Hanging ...
    I begin by discussing the shape of a catenary, namely, the shape of a hanging string/cable which is supporting its own weight.
  8. [8]
    [PDF] In Praise of the Catenary - UNI ScholarWorks
    Finally, we note that the scale factor a is the only parame- ter that determines the shape of a catenary through the equa- tion y = a cosh (x/a). But ...
  9. [9]
    [PDF] 18.01 Single Variable Calculus - MIT OpenCourseWare
    c) The curve y = cosh x is known as a catenary. It is the curve formed by a chain whose two ends are held at the same height. i) Sketch the curve ii) Find ...
  10. [10]
    [PDF] Strings, Chains, and Ropes - Harvey Mudd College Mathematics
    Mar 1, 2006 · All three cases involve inextensible materials, and gravity is the dominant force responsible for the shape of a catenary (hyperbolic cosine) ...
  11. [11]
    [PDF] Mathematical Documentation 3D-XplorMath
    is perferctly flexible and inextensible, has no thickness, is of uniform density. In other words the catenary is a mathe- matical abstraction of the shape ...<|control11|><|separator|>
  12. [12]
    [PDF] Notes on Classical Mechanics - K. V. Shajesh
    Apr 29, 2025 · (a) For a mass m falling under uniform gravity we have the equation of motion ... A catenary is the curve that the rope assumes, that ...
  13. [13]
    [PDF] case studies in optimization: catenary problem - Robert Vanderbei
    In more recent times, the catenary curve has come to play an important role in civil engineering.Missing: definition | Show results with:definition
  14. [14]
    Galileo's Catenary - MathPages
    For shallower cases, the value of A is greater than 1, so the factor of A3 in the denominator causes further reduction in the maximum deviation. The catenary ...
  15. [15]
    [PDF] RUS Bulletin 1724E-152 - USDA Rural Development
    Jul 30, 2003 · 2.2 Catenary and Parabolic Sag Equations: The curved shape of a completely flexible cable suspended between two rigid supports is defined as ...
  16. [16]
    Catenary -- from Wolfram MathWorld
    Catenary. The curve a hanging flexible wire or chain assumes when supported at its ends and acted upon by a uniform gravitational force. The word catenary is ...
  17. [17]
    catenary - PlanetMath
    Oct 26, 2014 · The arc length of the catenary (2) from the apex (0,a) ( 0 , a ) to the point (x,y) ( x , y ) is asinhxa=√y2−a2 a ⁢ sinh ⁡ x a = y 2 - a 2 . •.
  18. [18]
    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 ...
  19. [19]
    [PDF] In-situ observation of a soap film catenoid - arXiv
    Dec 22, 2009 · A catenoid is a surface that is formed between two coaxial circular rings and is classified mathematically as a minimal surface. Using soap film ...
  20. [20]
    Taq Kasra - Madain Project (en)
    It is the largest vault ever constructed in the world. The catenary arch was built without centring. In order to make this possible a number of techniques were ...
  21. [21]
    Structural Knowledge within the 6th Century AD Arch of Taq-iKisra
    Aug 10, 2025 · The arch of Taq-i Kisra is the largest single-span vault of unreinforced brickwork remaining in the world, and its shape is a catenary, which ...
  22. [22]
    [PDF] Engineering Structures 101
    For stability, a circular Roman arch supporting only its own weight must be thick enough to contain an equivalent “inverted catenary” arch. Therefore ...
  23. [23]
    The Catenary - The "Chain" Curve - National Curve Bank
    Leonardo da Vinci sketched hanging chains in his notebooks. · Galileo mistook the shape to be that of a parabola. · Simon Stevin constructed problems dealing with ...
  24. [24]
    Medieval Tent Structures - A Commonplace Book
    Nov 8, 2013 · Medieval tents used at least four different structural designs. The tent spread by ropes alone is recognizable by catenary sag visible in the profile of the ...
  25. [25]
    [PDF] Solving the brachistochrone and other variational problems ... - arXiv
    Mar 30, 2010 · In 1691 Jakob Bernoulli proposed as a challenge to find the curve that assumes a freely hanging chain when held down by its ends. His brother ...
  26. [26]
    Catenary - MacTutor History of Mathematics - University of St Andrews
    Its equation was obtained by Leibniz, Huygens and Johann Bernoulli in 1691. They were responding to a challenge put out by Jacob Bernoulli to find the equation ...
  27. [27]
    Leonhard Euler (1707 - 1783) - Biography - MacTutor
    In Institutiones calculi integralis (1768-70) Euler made a thorough investigation of integrals which can be expressed in terms of elementary functions. He also ...
  28. [28]
    s Menai Suspension Bridge: a commentary on Gilbert (1826)
    Gilbert, Thomas Telford, catenary, catenary of equal strength. Author for ... Poisson SD. 1811 Traité de Mechanique, Tome 1. Paris, France: Courcier ...
  29. [29]
    [PDF] The Catenary Art, Architecture, History, and Mathematics
    This paper describes how to use the catenary curve to enable students to see and appreciate connections between mathematics and other disciplines, including ...
  30. [30]
    [PDF] 1. Introductory Statics: the Catenary and the Arch
    A catenary is the curve a hanging rope approaches, defined as the limit of a chain with smaller and smaller links, keeping the length constant.
  31. [31]
    [PDF] 1 CHAPTER 18 THE CATENARY 18.1 Introduction If a flexible chain ...
    We consider the equilibrium of the portion AP of the chain, A being the lowest point of the chain. See figure XVIII.1 It is in equilibrium under the action ...
  32. [32]
    [PDF] As Hangs the Flexible Line: - Equilibrium of Masonry Arches - MIT
    From the introduction of Simon Stevin's (1548-1602) parallelogram rule, equilibrium could be described graphically using force vectors and closed force polygons ...
  33. [33]
    Catenary Arch | Exploratorium Museum Exhibit
    Catenary arches are very strong because they redirect the vertical force of gravity into compression forces that press along the curve, holding the arch's ...
  34. [34]
    Hooke's cubico–parabolical conoid | Notes and Records ... - Journals
    In 1675 Robert Hooke published, as one of his 'Inventions', a Latin anagram concerning the 'true...form of all manner of arches for building'.
  35. [35]
    The Idea Behind La Sagrada Familia: Gaudi’s Hanging Chain Model
    ### Summary: Gaudí’s Use of Inverted Catenaries for Sagrada Família Columns and Arches
  36. [36]
    Five construction techniques used by Gaudí every architect must know
    An inverted catenary curve built as a masonry arch is capable of carrying great weights whilst being made out of light materials such as brick or tile. Gaudí ...
  37. [37]
    Gateway to Mathematics Equations of the St. Louis Arch
    Why is a hanging chain described by the catenary equation? We glibly stated that a hanging chain follows the catenary curve, but why? The proof of that ...
  38. [38]
    Catenary Cables and Arches – Basic Concepts of Structural Design ...
    A catenary is a funicular shape for an unloaded cable and is determined solely by the self-weight of the cable, which is uniformly distributed along its length.
  39. [39]
    How does the parabolic curve help in the integrity of a suspension ...
    Jan 1, 2021 · The parabolic shape is the closest natural geometry a cable assumes under uniformly distributed horizontal load ( not the self-weight, that is ...
  40. [40]
    How much will a cable sag? A simple approximation
    Mar 9, 2024 · The sag is the difference between the height at the end points and the height in the middle. The equation of a catenary is. y = a cosh(t/a). and ...
  41. [41]
    Fundamentals of Conductor Sag - Tension Calculation
    Sep 25, 2024 · Discover the complexities of sag-tension calculation in transmission lines, balancing elongation and slack amid real-world conditions.
  42. [42]
    Catenary Anchor Leg Mooring - Wärtsilä
    The single point mooring system. A catenary anchor leg mooring system consists of a large buoy anchored by catenary mooring lines.
  43. [43]
    A review on mooring lines and anchors of floating marine structures
    Mooring line components mainly consist of mooring chain, steel wire, synthetic fiber rope, clump weight, buoy, and connector. Besides, tendon and tether are ...
  44. [44]
    [PDF] Small in-plane oscillations of a slack catenary using assumed modes
    Oct 2, 2023 · Routh [1] wrote equations for the uniform chain but provided analytical solutions for a non-uniform chain whose equilibrium shape is a cycloid.
  45. [45]
    XV. On the mathematical theory of suspension bridges, with tables ...
    ... catenary of equal strength. A curve not merely of speculative curiosity, but of practical use, where a wide horizontal extent may chance to be combined with ...
  46. [46]
    An amateur's contribution to the design of Telford's Menai ... - Journals
    Apr 13, 2015 · The 'catenary of equal strength' is the form of a cable whose cross-sectional area is made proportional to the tension at each point, so that ...
  47. [47]
    Variational approaches to the elasticity of deformable strings with ...
    Jul 2, 2024 · Some results showing the behavior of the elastic catenary can be found in Figure 4. ... : The effect of bending stiffness on inextensible cables.
  48. [48]
    Wind deflection analysis of railway catenary under crosswind based ...
    This paper evaluates the railway catenary's wind deflection under crosswind based on wind tunnel experiments and a nonlinear finite element model.Missing: generalized variational principle