Deflection (engineering)

Last updated
Deflection (f) in engineering Deflection.svg
Deflection (f) in engineering

In structural engineering, deflection is the degree to which a part of a long structural element (such as beam) is deformed laterally (in the direction transverse to its longitudinal axis) under a load. It may be quantified in terms of an angle (angular displacement) or a distance (linear displacement). A longitudinal deformation (in the direction of the axis) is called elongation .

Contents

The deflection distance of a member under a load can be calculated by integrating the function that mathematically describes the slope of the deflected shape of the member under that load. Standard formulas exist for the deflection of common beam configurations and load cases at discrete locations. Otherwise methods such as virtual work, direct integration, Castigliano's method, Macaulay's method or the direct stiffness method are used. The deflection of beam elements is usually calculated on the basis of the Euler–Bernoulli beam equation while that of a plate or shell element is calculated using plate or shell theory.

An example of the use of deflection in this context is in building construction. Architects and engineers select materials for various applications.

Beam deflection for various loads and supports

Beams can vary greatly in their geometry and composition. For instance, a beam may be straight or curved. It may be of constant cross section, or it may taper. It may be made entirely of the same material (homogeneous), or it may be composed of different materials (composite). Some of these things make analysis difficult, but many engineering applications involve cases that are not so complicated. Analysis is simplified if:

In this case, the equation governing the beam's deflection () can be approximated as:

where the second derivative of its deflected shape with respect to ( being the horizontal position along the length of the beam) is interpreted as its curvature, is the Young's modulus, is the area moment of inertia of the cross-section, and is the internal bending moment in the beam.

If, in addition, the beam is not tapered and is homogeneous, and is acted upon by a distributed load , the above expression can be written as:

This equation can be solved for a variety of loading and boundary conditions. A number of simple examples are shown below. The formulas expressed are approximations developed for long, slender, homogeneous, prismatic beams with small deflections, and linear elastic properties. Under these restrictions, the approximations should give results within 5% of the actual deflection.

Cantilever beams

Cantilever beams have one end fixed, so that the slope and deflection at that end must be zero.

Schematic of the deflection of a cantilever beam. Cantilever beam deflection.svg
Schematic of the deflection of a cantilever beam.

End-loaded cantilever beams

Cantilever beam with a force on the free end Cantilever with end load.svg
Cantilever beam with a force on the free end

The elastic deflection and angle of deflection (in radians) at the free end in the example image: A (weightless) cantilever beam, with an end load, can be calculated (at the free end B) using: [1]

where

Note that if the span doubles, the deflection increases eightfold. The deflection at any point, , along the span of an end loaded cantilevered beam can be calculated using: [1]

Note: At (the end of the beam), the and equations are identical to the and equations above.

Uniformly loaded cantilever beams

Cantilever beam with a uniform distributed load Cantilever with uniform distributed load.svg
Cantilever beam with a uniform distributed load

The deflection, at the free end B, of a cantilevered beam under a uniform load is given by: [1]

where

  • = uniform load on the beam (force per unit length)
  • = length of the beam
  • = modulus of elasticity
  • = area moment of inertia of cross section

The deflection at any point, , along the span of a uniformly loaded cantilevered beam can be calculated using: [1]

Simply supported beams

Simply supported beams have supports under their ends which allow rotation, but not deflection.

Schematic of the deflection of a simply supported beam. Simple beam deflection.svg
Schematic of the deflection of a simply supported beam.

Center-loaded simple beams

Simply supported beam with a force in the center Simple beam with center load.svg
Simply supported beam with a force in the center

The deflection at any point, , along the span of a center loaded simply supported beam can be calculated using: [1]

for

The special case of elastic deflection at the midpoint C of a beam, loaded at its center, supported by two simple supports is then given by: [1]

where

  • = force acting on the center of the beam
  • = length of the beam between the supports
  • = modulus of elasticity
  • = area moment of inertia of cross section

Off-center-loaded simple beams

Simply supported beam with a force off center Simple beam with offset load.svg
Simply supported beam with a force off center

The maximum elastic deflection on a beam supported by two simple supports, loaded at a distance from the closest support, is given by: [1]

where

  • = force acting on the beam
  • = length of the beam between the supports
  • = modulus of elasticity
  • = area moment of inertia of cross-section
  • = distance from the load to the closest support

This maximum deflection occurs at a distance from the closest support and is given by: [1]

Uniformly loaded simple beams

Simply supported beam with a uniform distributed load Simple beam with uniform distributed load.svg
Simply supported beam with a uniform distributed load

The elastic deflection (at the midpoint C) on a beam supported by two simple supports, under a uniform load (as pictured) is given by: [1]

where

  • = uniform load on the beam (force per unit length)
  • = length of the beam
  • = modulus of elasticity
  • = area moment of inertia of cross section

The deflection at any point, , along the span of a uniformly loaded simply supported beam can be calculated using: [1]

Combined loads

The deflection of beams with a combination of simple loads can be calculated using the superposition principle.

Change in length

The change in length of the beam is generally negligible in structures, but can be calculated by integrating the slope function, if the deflection function is known for all .

Where:

  • = change in length (always negative)
  • = slope function (first derivative of )
  • [2]

If the beam is uniform and the deflection at any point is known, this can be calculated without knowing other properties of the beam.

Units

The formulas supplied above require the use of a consistent set of units. Most calculations will be made in the International System of Units (SI) or US customary units, although there are many other systems of units.

International system (SI)

US customary units (US)

Others

Other units may be used as well, as long as they are self-consistent. For example, sometimes the kilogram-force () unit is used to measure loads. In such a case, the modulus of elasticity must be converted to .

Structural deflection

Building codes determine the maximum deflection, usually as a fraction of the span e.g. 1/400 or 1/600. Either the strength limit state (allowable stress) or the serviceability limit state (deflection considerations among others) may govern the minimum dimensions of the member required.

The deflection must be considered for the purpose of the structure. When designing a steel frame to hold a glazed panel, one allows only minimal deflection to prevent fracture of the glass.

The deflected shape of a beam can be represented by the moment diagram, integrated (twice, rotated and translated to enforce support conditions).

See also

Related Research Articles

In physics, the cross section is a measure of the probability that a specific process will take place when some kind of radiant excitation intersects a localized phenomenon. For example, the Rutherford cross-section is a measure of probability that an alpha particle will be deflected by a given angle during an interaction with an atomic nucleus. Cross section is typically denoted σ (sigma) and is expressed in units of area, more specifically in barns. In a way, it can be thought of as the size of the object that the excitation must hit in order for the process to occur, but more exactly, it is a parameter of a stochastic process.

<span class="mw-page-title-main">Fabry–Pérot interferometer</span> Optical device with parallel mirrors

In optics, a Fabry–Pérot interferometer (FPI) or etalon is an optical cavity made from two parallel reflecting surfaces. Optical waves can pass through the optical cavity only when they are in resonance with it. It is named after Charles Fabry and Alfred Perot, who developed the instrument in 1899. Etalon is from the French étalon, meaning "measuring gauge" or "standard".

<span class="mw-page-title-main">Young's modulus</span> Mechanical property that measures stiffness of a solid material

Young's modulus is a mechanical property of solid materials that measures the tensile or compressive stiffness when the force is applied lengthwise. It is the modulus of elasticity for tension or axial compression. Young's modulus is defined as the ratio of the stress applied to the object and the resulting axial strain in the linear elastic region of the material.

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.

Linear elasticity is a mathematical model of how solid objects deform and become internally stressed due to prescribed loading conditions. It is a simplification of the more general nonlinear theory of elasticity and a branch of continuum mechanics.

In the calculus of variations, a field of mathematical analysis, the functional derivative relates a change in a functional to a change in a function on which the functional depends.

<span class="mw-page-title-main">Stiffness</span> Resistance to deformation in response to force

Stiffness is the extent to which an object resists deformation in response to an applied force.

<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">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">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.

<span class="mw-page-title-main">Background field method</span> Technique in quantum field theory

In theoretical physics, background field method is a useful procedure to calculate the effective action of a quantum field theory by expanding a quantum field around a classical "background" value B:

The second polar moment of area, also known as "polar moment of inertia" or even "moment of inertia", is a quantity used to describe resistance to torsional deformation (deflection), in objects with an invariant cross-section and no significant warping or out-of-plane deformation. It is a constituent of the second moment of area, linked through the perpendicular axis theorem. Where the planar second moment of area describes an object's resistance to deflection (bending) when subjected to a force applied to a plane parallel to the central axis, the polar second moment of area describes an object's resistance to deflection when subjected to a moment applied in a plane perpendicular to the object's central axis. Similar to planar second moment of area calculations, the polar second moment of area is often denoted as . While several engineering textbooks and academic publications also denote it as or , this designation should be given careful attention so that it does not become confused with the torsion constant, , used for non-cylindrical objects.

<span class="mw-page-title-main">Voigt effect</span>

The Voigt effect is a magneto-optical phenomenon which rotates and elliptizes linearly polarised light sent into an optically active medium. The effect is named after the German scientist Woldemar Voigt who discovered it in vapors. Unlike many other magneto-optical effects such as the Kerr or Faraday effect which are linearly proportional to the magnetization, the Voigt effect is proportional to the square of the magnetization and can be seen experimentally at normal incidence. There are also other denominations for this effect, used interchangeably in the modern scientific literature: the Cotton–Mouton effect and magnetic-linear birefringence, with the latter reflecting the physical meaning of the effect.

<span class="mw-page-title-main">Flexural modulus</span> Intensive property in mechanics

In mechanics, the flexural modulus or bending modulus is an intensive property that is computed as the ratio of stress to strain in flexural deformation, or the tendency for a material to resist bending. It is determined from the slope of a stress-strain curve produced by a flexural test, and uses units of force per area. The flexural modulus defined using the 2-point (cantilever) and 3-point bend tests assumes a linear stress strain response.

The slope deflection method is a structural analysis method for beams and frames introduced in 1914 by George A. Maney. The slope deflection method was widely used for more than a decade until the moment distribution method was developed. In the book, "The Theory and Practice of Modern Framed Structures", written by J.B Johnson, C.W. Bryan and F.E. Turneaure, it is stated that this method was first developed "by Professor Otto Mohr in Germany, and later developed independently by Professor G.A. Maney". According to this book, professor Otto Mohr introduced this method for the first time in his book, "Evaluation of Trusses with Rigid Node Connections" or "Die Berechnung der Fachwerke mit Starren Knotenverbindungen".

<span class="mw-page-title-main">Timoshenko–Ehrenfest beam theory</span> Model of shear deformation and bending effects

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.

<span class="mw-page-title-main">Geographical distance</span> Distance measured along the surface of the Earth

Geographical distance or geodetic distance is the distance measured along the surface of the Earth, or the shortest arch length.

In civil engineering and structural analysis Clapeyron's theorem of three moments is a relationship among the bending moments at three consecutive supports of a horizontal beam.

<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:

<span class="mw-page-title-main">Span (engineering)</span> Distance between supports of an arch, bridge, etc.

In engineering, span is the distance between two intermediate supports for a structure, e.g. a beam or a bridge.

References

  1. 1 2 3 4 5 6 7 8 9 10 Gere, James M.; Goodno, Barry J. (January 2012). Mechanics of Materials (Eighth ed.). pp. 1083–1087. ISBN   978-1-111-57773-5.
  2. Roark's Formulas for Stress and Strain, 8th Edition Eq 8.1-14