Self-buckling

Last updated

A column can buckle due to its own weight with no other direct forces acting on it, in a failure mode called self-buckling. In conventional column buckling problems, the self-weight is often neglected since it is assumed to be small when compared to the applied axial loads. However, when this assumption is not valid, it is important to take the self-buckling into account.

Contents

Elastic buckling of a "heavy" column i.e., column buckling under its own weight, was first investigated by Greenhill at 1881. [1] He found that a free-standing, vertical column, with density , Young's modulus , and cross-sectional area , will buckle under its own weight if its height exceeds a certain critical value:

where is the acceleration due to gravity, is the second moment of area of the beam cross section.

One interesting example for the use of the equation was suggested by Greenhill in his paper. He estimated the maximal height of a pine tree, and found it cannot grow over 300-ft tall. This length sets the maximum height for trees on earth if we assume the trees to be prismatic and the branches are neglected.

Mathematical derivation

A column exhibiting a compressive buckling load due to its own weight. A column exhibiting a compressive buckling load due to its own weight..png
A column exhibiting a compressive buckling load due to its own weight.

Suppose a uniform column fixed in a vertical direction at its lowest point, and carried to a height , in which the vertical position becomes unstable and flexure begins. There is a body force per unit length , where is the cross-sectional area of the column, is the acceleration due to gravity and is its mass density.

The column is slightly curved under its own weight, so the curve describes the deflection of the beam in the direction at some position . Looking at any point on the column, we can write the moment equilibrium:

where the right-hand side of the equation is the moment of the weight of BP about P.

According to Euler–Bernoulli beam theory:

Where is the Young's modulus of elasticity of the substance, is the second moment of area.

Therefore, the differential equation of the central line of BP is:

Differentiating with respect to x, we get

We get that the governing equation is the third order linear differential equation with a variable coefficient. The way to solve the problem is to use new variables and :

Then, the equation transforms to the Bessel equation

The solution of the transformed equation is

Where is the Bessel function of the first kind. Then, the solution of the original equation is:

Now, we will use the boundary conditions:

From the second B.C., we get that the critical length in which a vertical column will buckle under its own weight is:

Using , the first zero of the Bessel function of the first kind of order , can be approximated to:

Euler's mistake

The column under its own weight was considered by Euler in three famous papers (1778a, 1778b, 1778c). [2] [3] [4] In his first paper, Euler (1778a) concluded that the column simply supported under its own weight would never lose its stability. In his second paper on this topic Euler (1778b) described his previous result as paradoxical and suspicious (see Panovko and Gubanova (1965); Nicolai, (1955); [5] Todhunter and Pierson (1866) [6] on this topic). In the next, third in series, paper, Euler (1778c) found that he had made a conceptual mistake and the “infinite buckling load” conclusion was proved to be wrong. Unfortunately, however, he made a numerical mistake and instead of the first eigenvalue, he calculated a second one. Correct solutions were derived by Dinnik (1912), [7] 132 years later, as well as Willers (1941), [8] Engelhardt (1954) [9] and Frich-Fay (1966). [10] Numerical solution with arbitrary accuracy was given by Eisenberger (1991). [11]

222 years later, after Euler's mistake in 1778, in the year 2000, Elishakoff [12] [13] revisited this famous problem and derived closed-form solutions for the first time, for self-buckling problems, by resorting to semi-inverse method. Numerous other problems are treated in his monograph. [14]

See also

Related Research Articles

<span class="mw-page-title-main">Lorentz force</span> Force acting on charged particles in electric and magnetic fields

In physics, the Lorentz force is the combination of electric and magnetic force on a point charge due to electromagnetic fields. A particle of charge q moving with a velocity v in an electric field E and a magnetic field B experiences a force of

<span class="mw-page-title-main">Gaussian beam</span> Monochrome light beam whose amplitude envelope is a Gaussian function

In optics, a Gaussian beam is a beam of electromagnetic radiation with high monochromaticity whose amplitude envelope in the transverse plane is given by a Gaussian function; this also implies a Gaussian intensity (irradiance) profile. This fundamental (or TEM00) transverse Gaussian mode describes the intended output of most (but not all) lasers, as such a beam can be focused into the most concentrated spot. When such a beam is refocused by a lens, the transverse phase dependence is altered; this results in a different Gaussian beam. The electric and magnetic field amplitude profiles along any such circular Gaussian beam (for a given wavelength and polarization) are determined by a single parameter: the so-called waist w0. At any position z relative to the waist (focus) along a beam having a specified w0, the field amplitudes and phases are thereby determined as detailed below.

<span class="mw-page-title-main">Navier–Stokes equations</span> Equations describing the motion of viscous fluid substances

The Navier–Stokes equations are partial differential equations which describe the motion of viscous fluid substances, named after French engineer and physicist Claude-Louis Navier and Anglo-Irish physicist and mathematician George Gabriel Stokes. They were developed over several decades of progressively building the theories, from 1822 (Navier) to 1842-1850 (Stokes).

In the calculus of variations and classical mechanics, the Euler–Lagrange equations are a system of second-order ordinary differential equations whose solutions are stationary points of the given action functional. The equations were discovered in the 1750s by Swiss mathematician Leonhard Euler and Italian mathematician Joseph-Louis Lagrange.

In continuum mechanics, the Froude number is a dimensionless number defined as the ratio of the flow inertia to the external field. The Froude number is based on the speed–length ratio which he defined as:

<span class="mw-page-title-main">Buckling</span> Sudden change in shape of a structural component under load

In structural engineering, buckling is the sudden change in shape (deformation) of a structural component under load, such as the bowing of a column under compression or the wrinkling of a plate under shear. If a structure is subjected to a gradually increasing load, when the load reaches a critical level, a member may suddenly change shape and the structure and component is said to have buckled. Euler's critical load and Johnson's parabolic formula are used to determine the buckling stress of a column.

<span class="mw-page-title-main">Hamilton–Jacobi equation</span> A reformulation of Newtons laws of motion using the calculus of variations

In physics, the Hamilton–Jacobi equation, named after William Rowan Hamilton and Carl Gustav Jacob Jacobi, is an alternative formulation of classical mechanics, equivalent to other formulations such as Newton's laws of motion, Lagrangian mechanics and Hamiltonian mechanics.

<span class="mw-page-title-main">Bending</span> Strain caused by an external load

In applied mechanics, bending characterizes the behavior of a slender structural element subjected to an external load applied perpendicularly to a longitudinal axis of the element.

<span class="mw-page-title-main">Hydraulic head</span> Specific measurement of liquid pressure above a vertical datum

Hydraulic head or piezometric head is a specific measurement of liquid pressure above a vertical datum.

<span class="mw-page-title-main">Euler–Bernoulli beam theory</span> Method for load calculation in construction

Euler–Bernoulli beam theory is a simplification of the linear theory of elasticity which provides a means of calculating the load-carrying and deflection characteristics of beams. It covers the case corresponding to small deflections of a beam that is subjected to lateral loads only. By ignoring the effects of shear deformation and rotatory inertia, it is thus a special case of Timoshenko–Ehrenfest beam theory. It was first enunciated circa 1750, but was not applied on a large scale until the development of the Eiffel Tower and the Ferris wheel in the late 19th century. Following these successful demonstrations, it quickly became a cornerstone of engineering and an enabler of the Second Industrial Revolution.

In physics, the Einstein relation is a previously unexpected connection revealed independently by William Sutherland in 1904, Albert Einstein in 1905, and by Marian Smoluchowski in 1906 in their works on Brownian motion. The more general form of the equation in the classical case is

In quantum mechanics, the probability current is a mathematical quantity describing the flow of probability. Specifically, if one thinks of probability as a heterogeneous fluid, then the probability current is the rate of flow of this fluid. It is a real vector that changes with space and time. Probability currents are analogous to mass currents in hydrodynamics and electric currents in electromagnetism. As in those fields, the probability current is related to the probability density function via a continuity equation. The probability current is invariant under gauge transformation.

In the study of partial differential equations, the MUSCL scheme is a finite volume method that can provide highly accurate numerical solutions for a given system, even in cases where the solutions exhibit shocks, discontinuities, or large gradients. MUSCL stands for Monotonic Upstream-centered Scheme for Conservation Laws, and the term was introduced in a seminal paper by Bram van Leer. In this paper he constructed the first high-order, total variation diminishing (TVD) scheme where he obtained second order spatial accuracy.

The Poisson–Boltzmann equation is a useful equation in many settings, whether it be to understand physiological interfaces, polymer science, electron interactions in a semiconductor, or more. It aims to describe the distribution of the electric potential in solution in the direction normal to a charged surface. This distribution is important to determine how the electrostatic interactions will affect the molecules in solution. The Poisson–Boltzmann equation is derived via mean-field assumptions. From the Poisson–Boltzmann equation many other equations have been derived with a number of different assumptions.

<span class="mw-page-title-main">Mathematical descriptions of the electromagnetic field</span> Formulations of electromagnetism

There are various mathematical descriptions of the electromagnetic field that are used in the study of electromagnetism, one of the four fundamental interactions of nature. In this article, several approaches are discussed, although the equations are in terms of electric and magnetic fields, potentials, and charges with currents, generally speaking.

<span class="mw-page-title-main">Timoshenko–Ehrenfest beam theory</span>

The Timoshenko–Ehrenfest beam theory was developed by Stephen Timoshenko and Paul Ehrenfest early in the 20th century. The model takes into account shear deformation and rotational bending effects, making it suitable for describing the behaviour of thick beams, sandwich composite beams, or beams subject to high-frequency excitation when the wavelength approaches the thickness of the beam. The resulting equation is of 4th order but, unlike Euler–Bernoulli beam theory, there is also a second-order partial derivative present. Physically, taking into account the added mechanisms of deformation effectively lowers the stiffness of the beam, while the result is a larger deflection under a static load and lower predicted eigenfrequencies for a given set of boundary conditions. The latter effect is more noticeable for higher frequencies as the wavelength becomes shorter, and thus the distance between opposing shear forces decreases.

In nonideal fluid dynamics, the Hagen–Poiseuille equation, also known as the Hagen–Poiseuille law, Poiseuille law or Poiseuille equation, is a physical law that gives the pressure drop in an incompressible and Newtonian fluid in laminar flow flowing through a long cylindrical pipe of constant cross section. It can be successfully applied to air flow in lung alveoli, or the flow through a drinking straw or through a hypodermic needle. It was experimentally derived independently by Jean Léonard Marie Poiseuille in 1838 and Gotthilf Heinrich Ludwig Hagen, and published by Poiseuille in 1840–41 and 1846. The theoretical justification of the Poiseuille law was given by George Stokes in 1845.

<span class="mw-page-title-main">Lagrangian mechanics</span> Formulation of classical mechanics

In physics, Lagrangian mechanics is a formulation of classical mechanics founded on the stationary-action principle. It was introduced by the Italian-French mathematician and astronomer Joseph-Louis Lagrange in his 1788 work, Mécanique analytique.

Lagrangian field theory is a formalism in classical field theory. It is the field-theoretic analogue of Lagrangian mechanics. Lagrangian mechanics is used to analyze the motion of a system of discrete particles each with a finite number of degrees of freedom. Lagrangian field theory applies to continua and fields, which have an infinite number of degrees of freedom.

<span class="mw-page-title-main">Euler's critical load</span> Formula to quantify column buckling under a given load

Euler's critical load is the compressive load at which a slender column will suddenly bend or buckle. It is given by the formula:

References

  1. "Greenhill, A. G. (1881). "Determination of the greatest height consistent with stability that a vertical pole or mast can be made, and the greatest height to which a tree of given proportions can grow." Proc. Cambridge Philos. Soc., 4, 65–73" (PDF).
  2. Euler, L. (1778a) Determinatio onerum, quae columnae gestare valent, Acta Academiae Scientiarum Petropolitanae, Vol. 1, 121-145 (in Latin).
  3. Euler, L. (1778b) Examen insignis puradoxi in theoria columnarum occurentis, Acta Academiae Scientiarum Petropolitanae, Vol. 1, 146-162 (in Latin).
  4. Euler, L. (1778c) De Altitudine columnarum sub proprio pondere corruentium, Acta Academiae Scientiarum Petropolitanae, Vol. 1, 163-193 (in Latin).
  5. Nicolai, E.V., Works in Mechanics, pp.436-454, Gostekhizdat, Moscow, 1955 (in Russian).
  6. Todhunter, I. and Pearson K., History of the Theory of Elasticity, Vol. 1, pp. 39-50. Cambridge University Press, 1886.
  7. Dinnik, A.N., Buckling under Own Weight, Proceedings of Don Polytechnical Institute 1 (Part 2), p. 19, 1912 (in Russian).
  8. Willers, F.A., Das Knicken schwerer Gestänge, ZAMM‐Journal of Applied Mathematics and Mechanics/Zeitschrift für Angewandte Mathematik und Mechanik, Vol. 21(1),(1941) 43–51 (in German).
  9. Engelhardt, H., Die einheitliche Behandlung der Stabknickung mit Beruecksichtung des Stabeigengewichte in den Eulerfaellen 1 bis 4 als Eigenwertproblem, Der Stahlbau, Vol. 23 (4),80–84, 1954 (in German).
  10. Frich-Fay, R., On the Stability of a Strut under Uniformly Distributed Axial Forces, Int. J. Solids Struct., Vol. 2, 361–369, 1966.
  11. Eisenberger, M., Buckling Loads for Variable Cross-Section Member with Variable Axial Forces, Int. J. Solids Struct., Vol. 27, 135–143, 1991.
  12. Elishakoff, I., A Closed Form-Solution for Generalized Euler Problem, Proc. Royal Soc. London, Vol. 456, 2409–2417, 2000.
  13. Elishakoff, I., Euler's Problem Revisited: 222 Years Later, Meccanica, Vol. 36, 265-272, 2001.
  14. Elishakoff, I., Eigenvalues of Inhomogeneous Structures: Unusual Closed-Form Solutions of Semi-Inverse Problems, CRC Press, Boca Raton, 2005, XIV + pp. 729; ISBN   0-8493-2892-6.