You are on page 1of 14

This document is part of the notes written by Terje Haukaas and posted at www.inrisk.ubc.ca.

The notes are revised without notice and they are provided as is without warranty of any kind. You are encouraged to submit comments, suggestions, and questions to terje@civil.ubc.ca. It is unnecessary to print these notes because they will remain available online.

Euler-Bernoulli Beams

The Euler-Bernoulli beam theory was established around 1750 with contributions from Leonard Euler and Daniel Bernoulli. The work built on earlier developments by Jacob Bernoulli. However, the beam problem had been addressed even earlier. Galileo attempted one formulation but misplaced the neutral axis. Leonardo da Vinci also seems to have addressed the problem of beam bending. The two key assumptions in the Euler-Bernoulli beam theory are that the material is linear elastic according to Hookes law and that plane sections remain plane and perpendicular to the neutral axis during bending. The latter is sometimes referred to as Naviers hypothesis. In contrast, Timoshenko beam theory, which is covered in another document, relaxes the assumption that the sections remain perpendicular to the neutral axis, thus including shear deformation. In the following, the governing equations are established, followed by the formulation and solution of the differential equation. Thereafter, the computation of stresses and cross-section constants is described. A number of sign conventions are adopted in the following: The z-axis is increases upward Displacement w is positive in the direction of the z-axis

Terje Haukaas

University of British Columbia

www.inrisk.ubc.ca

Distributed load qz is positive when it acts in the downward direction Clockwise shear force is positive Bending moment that imposes tension at the bottom of a horizontal beam element is positive Counter-clockwise rotation is positive so that it can be interpreted as the slope of the deformed beam element Tensile stresses and strains are positive, compression is negative

Equilibrium

The equilibrium equations are obtained by considering equilibrium in the x- direction for the infinitesimal beam element in Figure 1. Notice that the distributed load, q, acts in the downward direction, while the z-axis is in the upward direction. The notation qz is employed in other documents to identify the case where positive load acts in the positive z-direction.

V+dV

M+dM

dx

Figure 1: Equilibrium for infinitesimally small beam element.

Vertical equilibrium yields: Moment equilibrium about the rightmost edge yields:

q=!

dV dx

(1)

V=

dM dx

(2)

In Eq. (2) it is noted that second-order terms are neglected. This essentially means that terms with dx2 (the multiplication of an exceedingly small value by itself) are considered approximately equal to zero.

Section Integration

Integration of axial stresses over the cross-section:

M = $ !" # z d A
A

(3)

Euler-Bernoulli Beams

Page 2

Terje Haukaas

University of British Columbia

www.inrisk.ubc.ca

where the minus sign appears because it is compressive (negative) stresses in the positive z-axis domain that gives a positive bending moment, i.e., bending moment with tension at the bottom. Figure 2 is intended to explain this further.

Negative compression stress Positive z-values

M x
Minus sign is cross-section integral is necessary to get positive bending moment

Positive tension stress Negative z-values


Figure 2: The reason for the minus sign in Eq. (3).

Material Law

The material law throughout linear elastic theory is Hookes law:

! = E " #

(4)

In the context of two-dimensional theory of elasticity, the use of Eq. (4) implies a plane stress material law. It implies that there is zero stress, i.e., air on the sides of the beam. The alternative plane strain version of the two-dimensional Hookes law is more appropriate in cases where the beam is only a strip of a long rectangular plate that is supported along the two long edges. In that case the strain is restrained in the y-direction:

& " yy = $ % " xx E Which, substituted into the material law in the x-direction yields:

! yy =

" yy

#$ %

" xx =0 E

(5)

" yy " xx ($ % " xx ) = " xx (1 # $ 2 ) & " = E % ! (6) " xx #$ = #$ xx xx E E E E E 1# $2 All the derivations and results in the following are based on the material law xx=E.xx from Eq. (4). However, the plain strain version is easily introduced by replacing the Youngs modulus, E, in any equation by E/(1-2).

! xx =

Euler-Bernoulli Beams

Page 3

Terje Haukaas

University of British Columbia

www.inrisk.ubc.ca

Kinematics

The relationship between the axial strain and the transversal displacement of a beam element is sought. It is first recognized that bending deformation essentially implies shortening and lengthening of fibres in the cross-section. Fibres on the tension side elongate, while fibres on the compression side shorten. The starting point for the considerations is to link the axial strain to the change of length of the imaginative fibres that the cross-section is made up of. The same consideration as in kinematics of truss members, namely that strain is elongation divided by original length yields:

du (7) dx Next, the axial displacement u is related to the rotation of the cross-section. In particular, consider the infinitesimal rotation d of the infinitesimally short beam element in Figure 3. In passing, it is noted that d is equal to the curvature, . Under the assumption that plane sections remain plane and perpendicular to the neutral axis during deformation, each fibre in the cross-section change length proportional to its distance from the neutral axis.

!=

z, w

d!
z x, u

Figure 3: Naviers hypothesis for beam bending.

The amount of shortening or elongation depends upon the rotation of the cross- section. A geometrical consideration of to Figure 3 shows that the shortening and lengthening, i.e., axial displacement, of each infinitesimally short fibre is
du = ! d" # z

(8)

Finally, the rotation is related to the transversal displacement. For this purpose, consider two points on a beam that is dx apart, as shown in Figure 4. The relative displacement is dw, which is measure positive upwards. Consequently, a geometrical consideration of Figure 4 shows that: (9) where the equation is simplified by assuming that the deformations are sufficiently small so that tan().
Euler-Bernoulli Beams Page 4

tan(! ) =

dw " ! dx

Terje Haukaas

University of British Columbia

www.inrisk.ubc.ca

!
dw

dx

Figure 4: Rotation of the cross-section of a beam element.

By combining equations, the kinematic equation for beam members is obtained:

d 2w ! = " 2 # z dx

(10)

This expression implies an approximation of the exact curvature of the beam. Mathematically, curvature is defined as

!"

1 R

(11)

where R is the radius of curvature of the beam. In the Euler-Bernoulli beam theory that is presented here, the curvature is approximated by

!"

d# d 2 w " dx dx 2

(12)

Notice that there are two approximation signs. The first alludes to the fact that differentiation is carried out with respect to the x-axis. Unless the deformations are negligible this is inaccurate; differentiation should be carried out with respect to the s-axis that follows the curving beam axis. The second approximation is due to Eq. (9). From that equation it is observed that the accurate expression for is:

# dw & ! = tan "1 % ( $ dx '

(13)

If this expression was utilized in the derivations above then the differentiation of the inverse tan-function yields

!"

d# = dx $ $ dw ' 2 ' ) ) &1 + & % % dx ( (

$ d 2w ' & % dx 2 ) (

(14)

Euler-Bernoulli Beams

Page 5

Terje Haukaas

University of British Columbia

www.inrisk.ubc.ca

which reduces to the expression in Eq. (12) when the slope dw/dx is small. However, the curvature expression in Eq. (14) is still approximate because the differentiation is carried out with respect to the x-axis and not the beam axis. From mathematics, the exact curvature expression is:

!=

" d 2w % $ # dx 2 ' & " " dw % % ' ' $1 + $ # # dx & &


2 3 2

(15)

Differential Equation

The governing differential equation for beam members is obtained by combining the equations for equilibrium, section integration, material law, and kinematics:

q=!

dV d2M d2 = ! 2 = 2 $ " # z dA dx dx dx A
(16)

d2 d2 d 2w 2 = 2 $ E # % # z d A = ! 2 $ E # 2 # z d A dx A dx A dx = ! EI d 4w dx 4

where the modulus of elasticity is assumed constant over the cross-section and the moment of inertia is defined:

I = ! z 2 d A
A

(17)

General Solution

Although solving the differential equation is not part of typical structural analysis it is instructive to study its solution for simple reference cases. In particular, the solution of the differential equation is the starting point for the selection of shape functions in the finite element method. Those shape functions are often approximate, while the solution of the differential equation reveals the exact shape when the member deforms. The general solution of the differential equation reveals whether the finite element shape functions are exact or not. For beam members, the general solution of the differential equation is obtained by integrating four times:

w( x ) =

1 qz 4 ! ! x + C1 ! x 3 + C2 ! x 2 + C3 ! x + C4 24 EI

(18)

Given a uniform distributed load qz, the displaced shape is a fourth order polynomial. Without any distributed load, the displacement shape of a beam member is a third-order polynomial. To obtain the solution for a specific beam
Euler-Bernoulli Beams Page 6

Terje Haukaas

University of British Columbia

www.inrisk.ubc.ca

problem, boundary conditions are specified. To prescribe a rotation, shear force, or bending moment, the following equations are useful, obtained by combining the governing equations that are established earlier:

!=

dw dx
d 2w dx 2 d 3w dx 3

(19) (20) (21)

M = EI V = EI

As an illustration of the solution to the differential equation for beam bending, Figure 5 shows plots of w(x), (x), M(x), and V(x) for a simply supported beam with uniformly distributed load. The illustration is made with qz=L=EI=1.


Figure 5: Example of response functions for beam element.

With reference to Figure 5, notice that the displacement w(x) is negative, i.e., downwards and that (x) correctly shows that the slope is negative left of the mid- span. Furthermore, notice that the plots of M(x) and V(x) are identical to the respective section force diagrams, with one exception: plotting M(x) yields a diagram drawn on the compression side, while in these notes the bending moment diagrams are consistently drawn on the tension side.

Euler-Bernoulli Beams

Page 7

Terje Haukaas

University of British Columbia

www.inrisk.ubc.ca

Cross-section Parameters

The only cross-section constant in fundamental 2D beam theory is the cross- sectional moment of inertia, I, defined in Eq. (17). In the formula, z is the distance from the neutral axis of the cross-section. For simple cross-sections, most prominently the rectangular one with width b and height h, the integral is evaluated analytically:
h /2

I = b!

" h /2

z2 dz =

b ! h3 12

(22)

For more complicated cross-sections the following procedure may be helpful: 1. Determine the location of the neutral axis. For homogeneous cross-sections, the neutral axis passes through the centroid of the cross-section. In other words, when the coordinate z has its origin at the centroid then the static moment z dA is zero. In practice, first select a reference axis in the cross- section that is parallel to the neutral axias and let zo denote its distance to the true neutral axis. Furthermore, let zi denote the distance from the arbitrarily selected axis to the centroid of each sub-area, Ai that the cross-section consists of. The distance to the true neutral axis is determined from:

Azo =

All parts of the cross-section

Ai zi

"

zo =

! A z
i

(23)

2. Determine the local moment of inertia, Ii, of each sub-area of the cross- section about the local centroid axis of that part. 3. Add contributions to the global moment of inertia from each cross-section part according to Steiners formula:

I=

All parts of the cross-section

I i + zi2 Ai

(24)

where zi is the distance from the neutral axis of the entire cross-section to the centroid of the part.

Principal Axes

This document describes bending about one axis, i.e., bending of 2D beams. With the most common cross-sections it is straightforward to understand which cross- section axis the beam will bend around. For example, a double-symmetric cross- section bends around the two symmetry-axes. However, additional analysis is necessary for general asymmetric cross-sections to determine the principal axes, i.e., the bending axes of the cross-section. As a starting point, suppose two axes directions y and z are arbitrary selected; they originate in the centroid but they generally do not coincide with the principal axes. The total axial strain at a location of the cross-section is expressed in the following extended version of Eq. (10):

Euler-Bernoulli Beams

Page 8

Terje Haukaas

University of British Columbia

www.inrisk.ubc.ca

d 2v d 2w ! = ! o " 2 # y " 2 # z dx dx

(25)

From this strain, the material law in Eq. (4) provides the stress, and the stress is integrated by Eq. (3) to obtain the following expression for the bending moment about the y-axis:

d 2v d 2w M = ! E " # o " $ z d A + E " 2 " $ y " z d A + E " 2 " $ z 2 d A dx A dx A A

(26)

The first integral vanishes because z originates at the centroid, while the last term is the ordinary bending moment from Eq. (20). As a result, Eq. (26) is rewritten as:

M = EI yz !

d 2v d 2w + EI ! dx 2 dx 2

(27)

where the product of inertia, Iyz, has been defined as:

I yz = " y ! z d A
A

(28)

It is relatively straightforward to establish formulas for Iyz and compute stresses in term of Iyz, etc. However, it is also always possible to rotate the axis system so that Iyz is zero; then the axis system is referred to as the principal axes. This is advantageous because the elementary formulas for beam bending remain valid. (Material to be added here.)

Stresses

Although the Euler-Bernoulli beam theory is formulated in terms of axial stress, , beam bending involves both axial and shear stresses. The axial stress is directly related to the bending moment, while the shear stress is directly related to the shear force as described shortly. One way of obtaining an expression for axial stress in terms of the bending moment is to combine material law and kinematics equations, which yields:

! = "E #

d 2w # z dx 2

(29)

Then substitute the differential equation without equilibrium equations, i.e., Eq. (20), to obtain:

! ="

M # z I

(30)

It is noted that a positive bending moment, i.e., tension at the bottom, correctly yields negative stresses at the top, i.e., compression, where z is positive. This is the reason for the minus sign in Eq. (30), which also correctly gives positive tension stresses at the bottom when a positive moment acts on the cross-section. In
Euler-Bernoulli Beams Page 9

Terje Haukaas

University of British Columbia

www.inrisk.ubc.ca

summary, the beam theory presented in this document consists of the governing equations shown in Figure 6.

q
dV dx dM V= dx q=!

q = ! EI

d 4w dx 4

M = EI

d 2w dx 2

M
M = $ !" # z d A
A

M ! = " #z I

!="

d 2w #z dx 2

"

! = E "#

Figure 6: Governing equations in Euler-Bernoulli beam theory.

Shear Stress and Shear Centre for Open Cross-sections

When approaching shear stresses from bending, an anomaly in Euler-Bernoulli beam theory is first noted. The theory is based on the assumption that plane sections remain plane and perpendicular to the neutral axis. In other words, the only strain that takes place is the axial shortening or elongation of the fibres in the cross-section. Effectively, this prevents shear strain. With no shear strain there is no shear stress, which adds up to zero shear force. In other words, the shear force is not part of the theory. This is an anomaly, because shear force will develop even in simple beams that are subjected to transversal load. The anomaly is customarily addressed by recovering the shear force by equilibrium once the bending moment is computed. In fact, according to Eq. (2) the shear force is equal to the derivative of the bending moment; this is the equation that recovers the shear force. Another document on Timoshenko beam theory describes an approach to further extend the beam theory to include deformation due to shear forces. To obtain expressions for the shear stress, , in terms of the shear force, V, consider the infinitesimally short beam element in Figure 7. Furthermore, consider a cut in the cross-section and let qs denote the shear flow at that location.

Euler-Bernoulli Beams

Page 10

Terje Haukaas

University of British Columbia

www.inrisk.ubc.ca

Axial stresses

V+dV M+dM dx
Shear stresses (shear flow)

Figure 7: Shear flow by equilibrium of infinitesimal beam element.

The shear flow is the force per unit length of the beam that ensures equilibrium with the axial stresses, which are greater on one side than the other due to dM:

qs ! dx = # d " dA =
As

dM ! z dA I As

(31)

where As is the cross-sectional area outside the cut. Given that V=dM/dx, this yields where

qs =

V !S I

(32)

S=

As

! z dA
V "S I "t

(33)

The shear stress is calculated by distributing the shear flow over the thickness, t, of the cross-section at the particular location:

!=

(34)

For example, for a rectangular cross-section the maximum shear stress is at the neutral axis, with value equal to

3 V (35) " 2 A The shear centre of a cross-section, sometimes called the centre of twist, is the point where the resultant of the shear force must act to avoid rotation of the cross-section. The coordinates of the shear centre are denoted by ysc and zsc, and there are several techniques to determine them. The simplest case is double-symmetric cross- sections; for these cross-sections the shear centre coincide with the centroid. In fact, if a cross-section has an axis of symmetry then the shear centre is located on this

!=

Euler-Bernoulli Beams

Page 11

Terje Haukaas

University of British Columbia

www.inrisk.ubc.ca

axis. For general cross-sections, one approach to determine ysc and zsc is described in the document on warping torsion, where the omega diagram is utilized. However, a somewhat simpler approach, when the consideration of warping torsion is off the table, is offered here. The principle is simple; by definition, the moment of the shear flow about the shear centre must be zero. This leads to the following procedure to determine the coordinates of the shear centre, provided y and z are the principal axes through the centroid of the cross-section: 1. Select an arbitrary point as trial shear centre, and let ysc and zsc denote the coordinates of the shear centre relative to the centroid; in other words, let ysc and zsc denote the distances from the centroid to the shear centre 2. In accordance with Eq. (32), determine the shear flow in the cross-section due to a shear force in the z-direction 3. Write the equation that expresses the moment of the shear flow in Item 2 about the trial shear centre; in general, both ysc and zsc will appear in this expression 4. Similar to Item 2, determine the shear flow in the cross-section due to a shear force in the y-direction 5. Similar to Item 3, write the equation that expresses the moment of the shear flow in Item 4 about the trial shear centre 6. Set the equations from Items 3 and 5 both equal to zero and solve these two equations for the two unknowns ysc and zsc Only one moment equation is needed for single-symmetric cross-sections; in that cases the procedure simplifies to: 1. Select an arbitrary point along the symmetry axis as trial shear centre, and let e denote the distance from the centroid to that point 2. In accordance with Eq. (32), determine the shear flow in the cross-section due to a shear force in the direction perpendicular to the axis of symmetry 3. Write the equation that expresses the moment of the shear flow in Item 2 about the trial shear centre; e will appear in this expression 4. Set the equation from Items 3 equal to zero and solve for e

Shear Stress and Shear Centre for Closed Cross-sections

The determination of shear stress and shear centre for closed cross-sections, i.e., cross-sections with cell, can be approached in two ways. In the first approach, the shear centre coordinates are first determined, using the omega diagram as described in the document on warping torsion. Next, a cut in the cross-section is made to yield an open cross-section; the new coordinate s, which traces the cross- section around the cell, originates at the cut. The unknown shear flow at the cut is denoted qo, and the shear flow at all other locations are determined relative to this value in accordance with Eq. (32):

q = qo +

V ! S I

(36)

Euler-Bernoulli Beams

Page 12

Terje Haukaas

University of British Columbia

www.inrisk.ubc.ca

Once q is determined at all locations of the cross-section, the moment of the shear flow about the shear centre is computed:

T =! " q ! h ds = ! " qo ! h ds + ! "

V ! S ! h ds I

(37)

where the integrals are made around the cell, and h(s) is the distance from the shear centre to the tangent line of the cross-section at s. By definition the moment, T, about the shear centre must be zero, and solving for qo yields:
qo = ! V ! # S " h ds = ! V " S " h ds " # I ! 2 " Am " I ! # h ds

(38)

where the last equality expresses that the integral of h around the cross-section is twice the cell area, Am. Having the value of qo, the shear flow is determined at other locations by Eq. (36). The second approach to determine the shear flow and shear centre in closed cross-sections entails determining the shear centre at the end. Also this approach introduces a cut in the cross-section, but now compatibility is considered instead of moment equilibrium. To this end, consider Figure 8.

ds du

!1
!2
dx

d!
h

Figure 8: Two contributions to shear strain.

Figure 8 illustrates the two following contributions to shear strain in an infinitesimal element of the cross-section:

! = !1 +! 2 =

du d" + # h ds dx

(39)

where is the rotation of the cross-section. Material law states that:

Euler-Bernoulli Beams

Page 13

Terje Haukaas

University of British Columbia

www.inrisk.ubc.ca

! =

" q = G G #t

(40)

and combination of Eqs. (39) and (40) yields

du =

q d# ! ds " ! h ! ds G !t dx q d$ ds # ! ! h ds " G !t dx

(41)

Integration around the cell yields the total gap opening due to the cut:

u=! "

(42)

To ensure compatibility, this gap opening must be zero: u=0. From that equation it is of interest to solve for d /dx because the cross-section should not rotate due to shear force:

d! = dx

! # G " t ds ! # G " t ds = = 0 2A ! # h ds
m

(43)

Eq. (36) provides the expression for the shear flow, q, which yields: Solving for qo yields:

! " G ! t ds + ! " I ! G ! t ds = 0
S # G " t ds V ! qo = ! " 1 I ! # G " t ds

qo

V !S

(44)

(45)

When the material is homogeneous, the expression simplifies to:

qo = !

V " I

! # t ds
1 ! # t ds

(46)

Having the value of qo, the shear flow is determined at other locations by Eq. (36).

Euler-Bernoulli Beams

Page 14

You might also like