You are on page 1of 46

CHPT-4 DEVELOPMENT of BEAM

EQUATIONS
4.1 INTRODUCTION
In this section, we will derive the stiffness matrix for a simple beam element. A
beam is a long,slender structural member generally subjected to transverse
loading that produces significant bending effects as opposed to twisting or axial
effects. This bending deformation is measured as a transverse displacement and
a rotation. Hence, the degrees of freedom considered per node are a transverse
displacement and a rotation

Consider the beam element shown in Figure 4.1.

The beam is of length L with axial local coordinate x and transverse local
coordinate y. The local transverse nodal displacements are given by vi ’s and the
rotations by i ’s. The local nodal forces are given by fiy ’s and the bending
moments by mi ’s as shown. We initially neglect all axial effects. Figure 4–2
indicates the sign conventions used in simple beam theory for positive shear
forces V and bending moments m.
Beam stiffness Matrix Based on Euler-Bernoulli Beam theory
(Considering Bending Deformations Only)

The differential equation governing elementary linear-elastic beam behavior


(called the Euler-Bernoulli beam as derived by Euler and Bernoulli) is based on
plane cross sections perpendicular to the longitudinal centroidal axis of the
beam before bending occurs remaining plane and perpendicular to the
longitudinal axis after bending occurs. This is illustrated in Figure 4.3 . This
occurs in practice only when a pure couple or constant moment exists in the
beam. However it is a reasonable assumption that yields equations that quite
accurately predict beam behavior for most practical beams.

The differential equation is derived as follows.

Also, the curvatureκ of the beam is related to the moment by


For small slopes,

The curvature is given by

For constant EI and only nodal forces and moments,

We will now follow the steps outlined in Chapter 1 to develop the stiffness
matrix and equations for a beam element and then to illustrate complete
solutions for beams.

Step 1 Select the Element Type

Represent the beam by labeling nodes at each end and in general by labeling the
element number (Figure 4–1).

Step 2 Select a Displacement Function

Assume the transverse displacement variation through the element length to be

The complete cubic displacement function Eq. (4.1.2) is appropriate because


there are four total degrees of freedom (a transverse displacement vi and a
small rotation i at each node). The cubic function also satisfies the basic beam
differential equation—further justifying its selection.

we express v as a function of the nodal degrees of freedom v1, v2, 1, and 2 as
follows:

Solving Eqs. (4.1.3) for a1 through a4 in terms of the nodal degrees of freedom
and substituting into Eq. (4.1.2), we have

In matrix form,
N1, N2, N3, and N4 are called the shape functions for a beam element. These
cubic shape (or interpolation) functions are known as Hermite cubic
interpolation (or cubic spline) functions.

Step 3 Define the Strain/Displacement and Stress/Strain Relationships

From the deformed configuration of the beam shown in Figure 4–5, we relate
the axial displacement to the transverse displacement by

Also using Hooke’s law,


From elementary beam theory, the bending moment and shear force are related
to the transverse displacement function. Because we will use these relationships
in the derivation of the beam element stiffness matrix, we now present them as

Step 4 Derive the Element Stiffness Matrix and Equations

First, derive the element stiffness matrix and equations using a direct
equilibrium approach.

In matrix form,

where the stiffness matrix is then

It is assumed that the beam is long and slender; that is, the length, L, to depth,
h, dimension ratio of the beam is large. A general rule for rectangular cross-
section beams is that for a length at least eight times the depth, the
transverse shear deflection is less than five percent of the bending deflection.
The inclusion of the shear deformation in beam theory with application to
vibration problems was developed by Timoshenko and is known as the
Timoshenko beam.

Beam stiffness Matrix Based on timoshenko Beam theory


(including transverse shear Deformation)

The shear deformation beam theory is derived as follows. T

he shear deformation (deformation due to the shear force V) is now included.


Referring to Figure 4–6, we observe a section of a beam of differential length
dx with the cross section assumed to remain plane but no longer perpendicular to
the neutral axis (x axis) due to the inclusion of the shear force resulting in a
rotation term indicated by . The total deflection of the beam at a point x now
consists of two parts, one caused by bending and one by shear force, so that
the slope of the deflected curve at point x is now given by

where rotation due to bending moment and due to transverse shear force are
given, respectively, by  (x ) and  (x ) .
We assume as usual that the linear deflection and angular deflection (slope) are
small. The relation between bending moment and bending deformation
(curvature) is now

and the relation between the shear force and shear deformation (rotation due
to shear) (shear strain) is given by

The difference in dv/dx and  represents the shear strain  yz (   ) of the


beam as

Now consider the differential element in Figure 4–3(c) and for summing
transverse forces and then summing bending moments. Now substituting above
equations for V and M into previous equations to obtain the two governing
equations as

To derive the stiffness matrix for the beam element including transverse shear
deformation, we assume the transverse displacement to be given by the cubic
function. We choose transverse shear straing consistent with the cubic
polynomial for v(x), such that  is a constant given by

We obtain
For instance, for a rectangular shape As is taken as 0.83 times the cross section
A, for a solid circular cross section it is taken as 0.9 times the cross section.
Using these equations,  is expressed as a polynomial in x as follows:

we can now express the coefficients a1 through a4 in terms of the nodal


displacements v1 and v2 and rotations 1 and 2 of the beam at the ends x =0 and
x= L as previously done.

we obtain

In a similar manner
where the stiffness matrix, including both bending and shear deformation, is
then given by

or by defining

We can rewrite the stiffness matrix as


Baundary Conditions:

We could now solve above equations simultaneously for the unknown nodal
displacement v2 and the unknown nodal rotations 2 and 3 .
4.4 Distributed Loading
Beam members can support distributed loading as well as concentrated nodal
loading. Consider the fixed-fixed beam subjected to a uniformly distributed
loading w shown in Figure 4–21. The reactions, determined from structural
analysis theory, are shown in Figure 4–22. These reactions are called fixed-end
reactions.

Therefore, guided by the results from structural analysis for the case of a
uniformly distributed load, we replace the load by concentrated nodal forces and
moments tending to have the same effect on the beam as the actual distributed
load. Figure 4–23 illustrates this idea for a beam.
These statically equivalent forces are always of opposite sign from
the fixed-end reactions.

Work-Equivalence Method

We can use the work-equivalence method to replace a distributed


load by a set of discrete loads. This method is based on the concept
that the work of the distributed load w(x) in going through the
displacement field v(x) is equal to the work done by nodal loads fiy
and mi in going through nodal displacements vi and i for arbitrary
nodal displacements. To illustrate the method, we consider the
example shown in Figure 4–24.

The work due to the distributed load is given by

The work due to the discrete nodal forces is given by

As an example, consider the uniformly loaded beam shown in Figure 4–25(a).


For arbitrary nodal displacements, we let 1  1 , 2  0 , v1 =0, and v2 =0 and
then obtain

A table of equivalent nodal forces has been generated as follows


General Formulation

In general, we can account for distributed loads or concentrated


loads acting on beam elements by starting with the following
formulation application for a general structure:
where {F} are the concentrated nodal forces and {F0} are called the
equivalent nodal forces. Using the table above of equivalent nodal
forces { f0} expressed in terms of local-coordinate components, we
can express {F0} in terms of global-coordinate components.

Because we now assume that concentrated nodal forces are not


present ({F}= 0 ), as we are solving beam problems with distributed
loading only in this section, we can write

On solving for {d} and then substituting the global displacements {d}
and equivalent nodal forces {F0} into Eq. Previous eqn., we obtain the
actual global nodal forces {F}.

For example, for a uniformly distributed load w acting over a one-


element beam, we have

This concept can be applied on a local basis to obtain the local nodal
forces { f} in individual elements of structures

where {f0} are the equivalent local nodal forces.


the final beam theory solution expressions
Values of displacement and slope at other locations along the beam
for the finite element solution are obtained
The beam theory solution predicts a quartic (fourth-order)
polynomial expression for y [Eq. (4.5.5)] for a beam subjected to
uniformly distributed loading, while the finite element solution v(x)
assumes a cubic displacement behavior in each beam element under
all load conditions. The finite element solution predicts a stiffer
structure than the actual one.
4.7 potential Energy approach to Derive Beam
Element Equations
We will now derive the beam element quations using the principle of
minimum potential energy. The total potential energy for a beam is

where body forces are now neglected. The terms on the right-hand
side of Eq. represent the potential energy of

(1) transverse surface loading Ty (in units of force per unit surface
area,

(2) nodal concentrated force Piy moving through displacements vi

(3) moments mi moving through rotations i.

Again, v is the transverse displacement function for the beam


element of length L shown in Figure 4–36.
Consider the beam element to have constant cross-sectional area A.
The differential volume for the beam element can then be expressed
as

and the differential area over which the surface loading acts is

where b is the constant width.

The total potential energy becomes

we express the strain in terms of nodal displacements and rotations


as

where

The stress/strain relationship is given by


The total potential energy is expressed in matrix notation as

Differentiating πp in Eq. (4.7.15) with respect to v1, 1, v2, and 2


and equating each term to zero to minimize πp, we obtain four
element equations, which are written in matrix form as

Representing the nodal force matrix as the sum of those nodal


forces resulting from distributed loading and concentrated loading,
we have

Because {f} = [k]{d}, we have,


[k] is evaluated in explicit form as

4.8 Galerkin’s Method for Deriving Beam Element


Equations

We will now illustrate Galerkin’s method to formulate the beam


element stiffness equations. We begin with the basic differential Eq.
with transverse loading w now included; that is,

We now define the residual R. Applying Galerkin’s criterion

where Ni are the shape functions defined before.

We now apply integration by parts twice to the first term

integration by parts introduces the boundary conditions.


Substituting

This Equation is really four equations , one for each

You might also like