Anda di halaman 1dari 431

Energy methods in structural mechanics

A comprehensive introduction to matrix and


nite element methods of analysis

F. Guarracino and A. Walker, FREng

Thomas Telford

Published by Thomas Telford Publishing, Thomas Telford Limited, 1 Heron Quay, London E14 4JD.
URL: http://www.t-telford.co.uk
Distributors for Thomas Telford books are
USA: ASCE Press, 1801 Alexander Bell Drive, Reston, VA 20191-4400
Japan: Maruzen Co. Ltd, Book Department, 310 Nihonbashi 2-chome, Chuo-ku, Tokyo 103
Australia: DA Books and Journals, 648 Whitehorse Road, Mitcham 3132, Victoria
First published 1999

A catalogue record for this book is available from the British Library
ISBN: 0 7277 2757 5
# F. Guarracino and A. Walker, and Thomas Telford Limited, 1999
All rights, including translation, reserved. Except for fair copying, no part of this publication may be
reproduced, stored in a retrieval system or transmitted in any form or by any means, electronic, mechanical, photocopying or otherwise, without the prior written permission of the Books Publisher, Thomas
Telford Publishing, Thomas Telford Limited, 1 Heron Quay, London E14 4JD.
This book is published on the understanding that the authors are solely responsible for the statements
made and opinions expressed in it and that its publication does not necessarily imply that such statements
and/or opinions are or reect the views of the publishers.
Typeset by Academic Technical Typesetting, Bristol
Printed and bound in Great Britain by Bookcraft (Bath) Limited

Preface
The last two centuries have witnessed an impressive development in the
size and complexity of engineering structures. During the early part of
the 19th century much of the design analysis was based on empiricism
and experience. The latter half of the 1800s saw the inception of methods giving insight to the forces generated in frame structures due to
external eects, such as gravity, wind and thermal radiation. Supreme
examples of the complexity of structures that could be constructed on
the basis of these hand calculation methods are the Eiel Tower (Fig. 1)
and the Forth Bridge (Fig. 2). These structures are evidently robust
and their continued use is a tribute to the skill of the engineers to
apply the analysis methods available to them and to ensure that the
structures incorporated adequate levels of safety to compensate for
the engineers' ignorance of the detailed forces and deformations in
the elements of their structures.
The methods that were available to structural and mechanical
engineers from 1860 to 1950 were based essentially on energy principles
and on the requirements of equilibrium of forces and compatibility of
deformations. These factors, basically, still underlie all the current
methods of structural analysis. The major change in the intervening
years since the last quarter of the 19th century has been the possibility
of applying the methods to even more complex forms of structures and
the possibility of extending the linear analysis into the non-linear
regime to examine the failure modes of the structures.
The advent of structures such as aircraft and airships, and their
development during the period 19001930 emphasised the need for
the design of minimum weight structures which implied that the very
large levels of conservatism incorporated in bridge and building
design had to be reduced. The methods available to airship designers
were adequate to allow them to calculate the forces in the quite complex shapes of wires and beams that formed the frames of the ships.
An example of this is given in Fig. 3, which shows an airship called
the R27 on the right. The photograph shows the bulkhead between

iv

Preface

Fig. 1. Eiel Tower

gas bags. The bulkhead was formed by an outer ring frame supported
from an inner hub by tensioned wires, much like the construction of a
bicycle wheel. The novelist Nevil Shute1 was an engineer working on a
later airship called the R100, a sister ship of the ill-fated R101, which
had a similar bulkhead, and he describes how he and a colleague
worked together for many weeks calculating day after day the tensions
in the wires until eventually their analysis converged to satisfy force
equilibrium and deformation compatibility requirements.
This higher degree of analysis complexity was required for airships to
maintain their low weights but, in general, for bridges and framed
1

Nevil Shute, Slide Rule, Heinemann, London, 1954.

Preface

Fig. 2. Forth Rail Bridge (courtesy of ICE Library)

buildings the level of analysis was much less sophisticated which meant
that the estimation of the real forces in the members of structures could
be quite inaccurate. However, reliance on experience and factors of
safety enabled a satisfactory approach to the design and construction

Fig. 3. Structural frame in R27 (courtesy of Imperial War Museum)

vi

Preface

of bridges during the rst half of the 20th century. The analysis of
complex structures such as aircraft and ships was based on formulations derived from experience, testing and the results of research and
specialist analysis performed on elements such as plates and curved
shells.
It was widely appreciated that the availability of analysis capabilities
that would supply many more accurate details on the deformations
and loads in structural elements in complex structures, and that
could be applied in a commercial environment, would provide engineers with the facility to design structures with specied levels of
safety and with greater economy of materials. More particularly, however, the growth of the aerospace industry, with supersonic military
and passenger aircraft ying at very great altitudes, and the need to
design huge oshore structures and ships with high levels of structural
eciency and the ability to withstand extreme environmental loads,
provided a signicant spur to develop more eective structural analysis
capabilities. The major innovation that has enabled success to be
achieved came not from the mechanical or structural engineers but
from electronic engineers, mathematicians and physicists. They have
made available extremely powerful computing capabilities that
enable tasks, such as those tackled by Nevil Shute, to be completed
in a few seconds instead of weeks and at very much greater structural
detail. The present generation of structural and mechanical engineers
has access to powerful computers and programs that enable designers
routinely to evaluate the strains and displacements in very complex
structures and thus to rene the form of their structure to obtain
greater economic and technical eciency. Figs 46 are a few examples
of structures that have been designed during the last 20 years using
computer-based analysis methods.
This book presents a unied development of the analysis principles
and methods that are currently embodied in computer analysis programs. The development is founded on fundamental energy principles,
particularly the law of conservation of energy, and proceeds through a
uniform yet essentially rigorous approach by gradually introducing the
reader to variational principles and thus deriving the underlying
approaches in frame and nite element analysis. It is the authors'
experience in teaching structural analysis methods, and applying
them in practice, that in text books the presentations of the basis of
energy methods and frame and nite element analysis can too often
be confusing and lead engineers to misunderstand the results they
obtain from the application of these methods. In fact, in many cases,
because engineers do not have a sucient understanding of the principles underlying nite elements, they treat computer analysis as a `black

Preface

vii

Fig. 4. Tsing Ma suspension bridge (courtesy of Mott MacDonald)

box' and take an uncritical approach to the results from the analyses.
This, we feel, is dangerous since a computer analysis will nearly always
achieve some result from the calculations, but this does not necessarily
imply that the result is appropriate to the structure being designed.
In preparing this book we have followed a logical structure that will
enable the reader to follow the development of the nite element and
frame analysis methods from fundamental energy principles through
to the type of formulations that are embedded in commercial nite
element programs. The emphasis in the development is on strain, compatibility of deformations and energy. Stress is considered as a less
important, secondary concept calculated from the strains in a structure,
so that it practically plays no part in the theoretical basis of the analysis
methods. Our intention is to provide engineers with an understanding
that allows them to interpret results from nite element analyses and
to determine what the results mean in a physical sense. The presentation
in the text encompasses methods of structural analysis that can be used
to provide simplied formulations in particular cases and enable the
investigation of parametric variation without recourse to lengthy computation. Our experience is that these methods, such as those proposed
by Rayleigh, Cotterill, Castigliano and others, are very valuable during
preliminary design where a `feel' for the structure's capability to resist
the applied loads is being investigated. The text shows that these simplifying methods are one step towards the computer-based methods used
in the design oce. By basing the development in this text rmly on

viii

Preface

Fig. 5. Oshore structure

energy principles it is clear how the principles of CotterillCastigliano


and the trial function methods proposed by Rayleigh and Ritz lead
directly to practical computer methods without any recourse to additional and somewhat misleading aspects such as the concept, sometimes
introduced in other treatments, of cutting a structure up into totally
independent elements and then imposing equilibrium at discrete
points, i.e. the node joining elements.
This book is essentially an introductory text and therefore deals with
a linear elastic treatment of structural analysis. However, we feel it is
important that the reader is made aware of at least one of the limitations of this form of analysis. As structural forms become evermore
ecient in their use of materials the possibility of failure by buckling
becomes an important factor. This aspect is introduced in the last

Preface

ix

Fig. 6. Space shuttle craft

chapter of this book to show how geometric non-linearity can be introduced into nite element and frame analysis. Of course, this type of
non-linearity leading to failure in practical structures is only one
facet of the limit state behaviour of structures. The other facets such
as material non-linearity, local buckling and the eects of the variations in structural geometry are reserved for consideration in a
future text.
F. Guarracino
A. Walker

Acknowledgement
This book has been a long time in gestation. During that period, our
wives have shown considerable forbearance over the time we have
spent discussing the contents of the book and developing the presentation.
We are truly grateful to Angela and Barbara, without whose patience
and encouragement this book would never have seen the light of day.

Contents
1

Work and energy


Introduction
1.1 Work and energy
1.2 The principle of conservation of energy
1.3 Equilibrium of mechanical systems

Kinematics and equilibrium of systems of rigid bodies


Introduction
2.1 Equilibrium of rigid bodies
2.2 A general innitesimal displacement kinematics for
systems of rigid bodies
2.3 Reactions of constraints
2.4 Internal reactions for a straight beam
2.5 A three-dimensional example
Suggested exercise problems

11
11
11

Deformation of bodies and material properties


Introduction
3.1 Deformation properties in one dimension
3.2 One-dimensional thermal strain
3.3 Three-dimensional strain
3.4 Straindisplacement relationships
3.5 Simplications possible when deformations are
very small
3.6 Deformation in the neighbourhood of an
arbitrary point
3.7 Transformation of the components of strain with
the change of reference frame
3.8 Principal directions of deformation. Maximum
and minimum extensions
3.9 Plane strain
3.10 Straindisplacement compatibility conditions

49
49
50
58
60
63

1
1
1
5
9

20
33
35
43
46

67
70
73
75
78
79

xii

Contents

3.11 Strain energy


3.12 Yield criteria
Suggested exercise problems
4

81
87
91

Theory of elastic deformation of beams


Introduction
4.1 Deformation of axially loaded bars
4.2 Deformation of beams
4.3 The theory of very small displacements
4.4 Euler's method for the analysis of beam deformations
4.5 The boundary conditions associated with Euler's
equation of beam deformation
4.6 Example of application of Euler's method of beam
deformation for a beam with a variable depth
4.7 Approximate solution of Euler's equation of beam
deformation using the method of nite dierences
4.8 Applied displacements and constraint reactions
4.9 Eects of temperature changes
Suggested exercise problems

93
93
95
98
102
105

General principles in the analysis of linear elastic structures


Introduction
5.1 The principle of superposition of the eects in the
linear theory of elasticity (existence and uniqueness
of the solution)
5.2 Reciprocal theorems in the linear theory of elasticity

141
141

Total potential energy


Introduction
6.1 The principle of the stationary value of total
potential energy
6.2 Two important remarks regarding the principle of
stationary value of total potential energy
6.3 CotterillCastigliano's rst theorem
6.4 CotterillCastigliano's second theorem in the linear
theory of elasticity
6.5 Numerical example: frame analysis by
CotterillCastigliano's theorem
Suggested exercise problems

151
151

154
155

The method of trial functions


Introduction
7.1 The basis of the method of trial functions

169
169
170

111
116
121
126
130
137

142
147

151

158
160
166

Contents

7.2 The method of RayleighRitz


7.3 The quality of the trial function
7.4 Numerical example: the method of trial functions
applied to a cantilever with linearly varying depth
7.5 The systematic search for a trial function
7.6 Localised Rayleigh functions
7.7 Numerical example: application of Rayleigh
localised trial functions to the analysis of a frame
Suggested exercise problems
8

10

xiii

173
178
183
187
189
199
203

Matrix analysis of pin-jointed trussed structures


Introduction
8.1 Plane pin-jointed structures as systems of axially
loaded bars
8.2 Matrix formulation of the elastic equilibrium
equations for plane pin-jointed structures
8.3 Numerical example: analysis of a truss
8.4 Matrix analysis of pin-jointed space structures
Suggested exercise problems

205
205

Matrix analysis of rigid-jointed framed structures


Introduction
9.1 Plane rigid-jointed frame structures as systems
of beams
9.2 Matrix formulation of the elastic equilibrium
equations for plane rigid-jointed frames
9.3 Numerical example: analysis of a rigid jointed frame
9.4 Matrix analysis of space frames
9.5 Equivalent joint loads
9.6 Restraint displacements
9.7 Some nal remarks on the matrix analysis of
structures
Suggested exercise problems

233
233

Analysis of thin plates


Introduction
10.1 Kirchho's model of plate deformation
10.2 Equilibrium of stretched plates
10.3 Equilibrium of bent plates.
10.4 An example of the method of trial function for the
solution of bending plates.
10.5 More examples of application of simple trial
functions to plate bending

267
267
269
270
274

206
212
221
226
230

233
240
245
249
254
259
261
263

286
290

xiv

Contents

10.6 Localised RayleighRitz method applied to


plate bending
11

12

The theory of nite elements


Introduction
11.1 Sub-division of a structure into nite elements
11.2 Displacement functions and shape functions
11.3 Element mapping and intrinsic coordinates
isoparametric elements
11.4 Convergence criteria for the displacement
functions the patch test
11.5 Some general families of shape functions
11.6 Strain energy, work of applied loads and equilibrium
equations
11.7 Evaluation of stiness matrices and equivalent nodal
loads numerical integration
11.8 Numerical example: analysis of a compressed
foundation

295
303
303
305
307
314
318
323
330
335
341

Stability of equilibrium and non-linear deformations of


beamcolumns
Introduction
12.1 The concept of stability. Initial buckling of columns
12.2 Energy criterion of stability
12.3 Some illustrative examples of stability conditions.
Branching points and limit points.
12.4 Euler's equation of beam deformation in the
presence of axial loads
12.5 Analytical treatment of some examples of columns
12.6 Method of trial functions for the initial buckling
of columns
12.7 Localised trial functions applied to the initial
buckling of struts
12.8 Method of trial functions for the initial buckling
of frames
12.9 Analysis of a frame loaded by vertical and
lateral loads
12.10 Localised trial function analysis of frame buckling
Suggested exercise problems

404
408
417

Index

421

353
353
353
357
358
365
372
384
391
399

1.

Work and energy

Introduction

This chapter introduces the concepts of work and energy that are the
basic building blocks of the methods of analysis that are developed
later in the textbook. We take as self-evident the concept of force and
follow Newton's1 approach that a force is an inuence that causes
changes in displacements and velocities within a dened mechanical
system. A force may also cause deformation of a body and is represented
by a vector that impinges on a body along a specied line of action.
This chapter proposes a denition for equilibrium and establishes a
test for the equilibrium of a body in a given frame of reference.

1.1

Work and energy

The concept of work W is basically related to the idea of a force


moving its point of application along a certain path. For example,
given a force P (assumed constant) and a displacement s
A2 A1 of its point of application (Fig. 1.1), we will dene the
work done by the force as the scalar product
W P  s P  A2 A1

1:1

which can also be expressed in terms of Cartesian components


W Px sx Fy sy

1:2

Similarly, the elementary work of a force moving along a certain path s


at a certain time t can be dened by
dW Ps t ds

1:3

The coordinate ds is tangential to the path s at the time t and Ps t is


the component of the force P at the time t in the same direction (Fig.
1.2).
1

Newton, Isaac (Woolsthorpe, Lincholnshire, 1642London, 1727), English physicist


and mathematician, one of the most prominent gures in the history of modern science.

Energy methods in structural mechanics


Sx

A2

Sy
A1

P
Py
x

Px

Fig. 1.1

The work done between the positions a0 and a1 is consequently


written as
a1

Ps ds

1:4

a0

From a dimensional point of view the work W is the product of a force


multiplied by a length (that is F  L) and this concept encompasses
moment times rotation, and pressure times change of volume: these
all being products characterised by the same physical dimensions.
Work is performed when we apply a moment to the stopper in order
to open a bottle, or when we inate a balloon by increasing its internal
pressure (Fig. 1.3).
It appears thus natural to introduce the concept of energy as the
capability of a system to perform work.

a1

a0
Ps

Fig. 1.2

Work and energy

Fig. 1.3

The variation of energy E is dened as the rate of work performed


by, or on the system
E W

1:5

The energy E constitutes the measure of the capability of performing


work stored in the system itself. It is worth noting that, according to
this denition, we are concerned only with changes of energy. In
fact, as we will see later, the absolute value of energy of a system is
not relevant to our purposes.
Dierent forms of energy can be recognised in nature. A body with a
mass m moving at a certain speed v can move another resting body by
hitting it; the rst body is said to possess a form of energy relative to
the stationary body that is called kinetic energy (Fig. 1.4).
The moving body can deform an elastic spring which, in turn, can
push the impinging body away by exerting a force on it (Fig. 1.5).
The spring is thus able to possess what is dened as elastic energy.
Heat is a form of energy: this can be illustrated by heating a pot of
water and arranging the pressure of the vapour to lift a weight, which
v*2

m2
m2
m1
v*1

v1
m1

Fig. 1.4

Energy methods in structural mechanics

l0

l < l0

Fig. 1.5

was of course the primary source of power of the Industrial Revolution. This is illustrated diagrammatically in Fig. 1.6. Other examples
of energy are found in chemical and nuclear reactions.
Appropriate mathematical expressions can be derived to quantify
the energy gained or lost by a system. We say, for example, that
within a certain coordinate frame a mass m moving at a velocity v
possesses the kinetic energy described by the formula
K 12 mv2

1:6

Notice that the velocity must be measured relative to some frame of


reference which for many engineering calculations can conveniently be
taken as xed to the Earth's surface. It is convenient to use the convention that when the body is stationary in that particular reference frame
it has zero kinetic energy. However, relative to another frame, say one
xed on the Sun, the body has considerable velocity and is seen to
possess kinetic energy.
In a nuclear reaction an absolute energy can be attributed, according
to the theory of relativity, to a mass m by
E mc2

1:7

H2O

Fig. 1.6

Work and energy

where c is the velocity of light and is an absolute value irrespective of


the frame of reference.
To sum up, it is very important to bear in mind that energy is a
general property attributed to all systems and bodies in nature. The
universal nature of energy provides the foundation for the treatment
of structural mechanics which is developed here.

1.2

The principle of conservation of energy

The principle of conservation of energy, also known as the First Law of


Thermodynamics, is a general axiom founded on the presumption that,
in the absence of any interaction with the environment, the energy of a
given system remains constant in the passage from one state to
another.
In other words, energy is not created or destroyed in a system
isolated from its environment and, consequently, the only possible
variation is given by the exact amount of exchange of energy or,
what is equivalent, work with other systems.
In the following text such a system, which experiences only internal
changes of state without interaction with external agents, is called a
closed system. Strictly speaking, no perfectly closed systems are found
in nature, since there is always some interaction between all material
bodies of the universe across nite systems. Nevertheless, the concept
of a closed system is a very useful approximation, both from a theoretical and a practical point of view, and is employed extensively in the
development of our analytical methods.
Everyone has experienced the exchange of heat by conduction that
takes place between two bodies at dierent temperatures or the phenomenon of heat generation by friction. The principle of conservation
of energy requires that in a closed system in the rst case the sum of the
heat of the two bodies remains constant, as well as in the second case
the sum of the work performed against friction and the heat produced
is a constant value.
This principle can be summarised in the following mathematical
statement, valid for any closed system
E const:

1:8

As we are going to deal with simple thermodynamic systems, attention


will now be restricted to two basic ways of exchanging energy: namely,
mechanical work W and heating Q.
Henceforth, we can write the principle of conservation of energy in
the following form
E W Q

1:9

Energy methods in structural mechanics


P

l
l'

Fig. 1.7

where E is the change of the level of energy in the system, W is the


work done on or by the system and Q is the amount of heat exchanged.
We will assume that the positive sign is for the work done on the system
and the heat given to it; the negative sign holds for work done by the
system and heat taken from it. Equation (1.9) is sucient to link the
variations of energy of a given system to its changes of state, that is
the changes of temperature, position or shape of its component parts.
In fact, the state of a mechanical system can be described by a number
of independent geometrical variables together with one thermodynamic
variable, say the temperature T. In mechanics these geometrical
variables are called generalised coordinates and the number of such
coordinates n required to specify completely the state of the system is
equal to the number of degrees of freedom less, of course, the geometrical constraints imposed at the system boundary. This is exemplied
later in the text.
For the sake of clarity, let us now consider the simple structural
element shown in Fig. 1.7, i.e. a bar subject to forces P at both ends.
It is natural to assume as the parameters dening the state of the
system to be its actual length l 0 and its temperature T. The initial temperature of the element is T0 and its corresponding length is l. A zero
value of energy is attributed to this initial state which is acceptable since
we are interested in measuring, by denition, the variation of energy.
When the system undergoes an innitesimal change of state, from
the principle of conservation of energy (equation (1.9)) we have
dE dW dQ

1:10

where, by denition,
dW P dl

1:11

and dl is the variation in length of the member.


If the material is linearly elastic, the following relationship can be
stated between the deformation and both the applied force and variation of temperature
l l 0 l

P
lT T0
k

1:12

Work and energy

where k, the stiness, and , the coecient of linear thermal expansion,


are constants for the material and the element, at least for small changes
in temperature.
Further, the following relationship can be assumed between the heat
exchanged by the system and its temperature:
dQ C dT

1:13

provided the coecient of heat capacity C is not a function of temperature or stiness; which experimentally is accurate for a limited temperature range and small deformations. The requirement in equation
(1.13) is adequate for the type of mechanics considered here.
By means of equations (1.11), (1.12) and (1.13), equation (1.10)
becomes
dE C dT kl l T T0 dl

1:14

and by a simple integration we get


ET; l 0 C T T0 12 k l 2 k l l T T0 A

1:15

As we have attributed the zero value of energy to the state dened by


T T0 and l 0 l, we have in this case
ET0 ; l CT0 T0 12 k l l2 k l l l T0 T0 A
0

1:16

The value of the constant A is, therefore


A0

1:17

ET; l 0 CT T0 12 k l 2 k l l T T0

1:18

and
This expression constitutes the denition of the internal energy E of
the structural member in term of its variables of state, i.e. its actual
length l 0 and its temperature T.
In the same manner, from equation (1.9), we can derive the expression of the energy possessed by a body of mass m dened as a point
moving along a straight path at a certain velocity v, as illustrated in
Fig. 1.8.
Let us assume that the body is motionless at the beginning of the
path and attribute zero energy to this initial state. Subsequent states
of the system will be dened by a single variable; the most convenient
variable is the position x along the straight path. According to the
second of Newton's laws, a force must be applied to the body in
order to move it along its direction.

Energy methods in structural mechanics


v=0
m

Fig. 1.8

This force will produce a variation of velocity with the time t given
by the formula
dx_
1:19
dt
where the force Ft is considered as a function of time and the dot
represents the derivative with respect to the time, i.e.
Ft m

dx
1:20
dt
Equation (1.9) states, in absence of any exchange of heat energy, that
v  x_ 

dE dW

1:21

The work done, dW, is therefore given by the formula


dx_
dx
dt
Thus, using equation (1.20)
dW m

1:22

dE mx_ dx_

1:23

This equation can be integrated between the velocity limits x_ 0 and


x_ v, and yields
E 12 mv2

1:24

This is the expression that was stated earlier, see equation (1.6), and
represents the energy possessed by the body due to its velocity v, i.e.
the rate of change of its state variable x with respect to time.

Work and energy

On account of this relationship we can dene this form of energy as


kinetic energy, and the derivation of equation (1.24) provides a basis
for the denition of equation (1.6).

1.3

Equilibrium of mechanical systems

A mechanical system is said to be in static equilibrium when it is


motionless and experiences no change of state within a specied
frame of reference, i.e. all the geometrical variables required to specify
completely the position of the system in the space do not change for all
time. This is a common sense denition that provides engineers with a
valuable operational approach to the analysis and design of real structures. In the case of simple thermodynamic systems this intuitive concept can be extended to all the state variables of the system; that is to
say, the geometrical variables plus the temperature T.
We want now to establish a condition of equilibrium for mechanical
systems in terms of energy. This will form the basis of our ability to
predict the state of equilibrium of bodies subjected to any specied
system of forces. Later, the condition of equilibrium is further
extended to enable the analysis of complex structures to evaluate
their global and local deformations. Without loss of generality to
our purposes we shall restrict our attention to the case of absence of
heat exchange and assume dQ 0.
Evidently, if a motionless system has no rate of change in its geometrical variables xi with time, this means
dxi
0
dt

1:25

The kinetic energy of the system is thus constant and can be set equal
to zero. That is
X
1
_2
K
1:26
2 mi xi 0
i

where mi are the generalised masses of the system which, by denition,


are nite and positive quantities. In this case a generalised mass can be
thought of as asssociated with a certain geometrical variable xi .
Vice versa, if equation (1.26) holds then equation (1.25) must be
true, since the mass must always be positive, and the two statements
are therefore equivalent. This means that we can state the constant
zero value of kinetic energy as the condition of equilibrium for a
mechanical system.
As we assumed dQ 0, the energy of a mechanical system in equilibrium reduces itself to the strain energy U. This kind of energy is

10

Energy methods in structural mechanics

related to the deformation of the system: in equation (1.18), for example, it is represented by the term 1=2k l 2 k l0 l T T0 . This
aspect of the energy is considered in greater detail in chapter 3.
For the moment, we can state that for a mechanical system in
equilibrium we have, from the principle of conservation of energy
(equation (1.10)),
dE dU dW

1:27

This equation means that any innitesimal increment of work done on


the system tends to produce only deformation of the system itself and
does not produce any kind of motion.
Importantly, equation (1.27) can be used to obtain a practical test of
equilibrium for mechanical systems. Let us consider a mechanical
system in a state of equilibrium and suppose that on account of the
geometrical constraints it could actually be able to move. We can
then apply to it a trial displacement eld; this is dened to be a eld
of displacements that have innitesimal magnitude and are compatible
with the constraints on the system, but are imaginary and do not
actually occur. Such a trial eld will be denoted with the symbol ,
and the same symbol will be used for all the variations in any quantity
aected by the same set of displacements. These quantities will be
called virtual to underline the fact that they could take place but in
actual fact are imaginary.
As the system has been dened to be in equilibrium, equation (1.27)
holds for any trial set of displacements and we have
U W

1:28

This relationship is at the same time a condition and a test of equilibrium. Thus, if the condition in equation (1.28) is satised for any
eld of displacements then equation (1.27) is true and we have
dK 0: this means that the kinetic energy of the system is constant
and equal to zero and the system is in equilibrium.
Alternatively, we can simply specify any possible set of displacements of the system and verify that equation (1.28) is satised: this
operation constitutes a practical test of equilibrium.
This very important approach, which is central to engineering
mechanics, is explored in detail throughout this book. In the following
chapter our study starts with the equilibrium of simple systems of rigid
bodies.

2.

Kinematics and equilibrium of


systems of rigid bodies

Introduction

Although strictly speaking all bodies and engineering structures are


deformable, common sense tells us that some systems can be treated
as being composed of components that are essentially rigid, that is
undeformable. An example of such a system would be a collection of
billiard balls that move about the table and travel quite large distances
compared to their diameters. When the balls collide there will be some
deformation of the diameter, but this is so small compared to the
dimension of the diameter and to the distances moved that the deformations can be completely ignored in the calculation of the movements
of the balls.
This approach is very important in engineering mechanics since it is
not uncommon for a structure to experience both global and local
displacements. Another example of this division is the movement of
an aircraft that as part of its function may travel many thousands of
kilometres in ight and have extensive vertical movement in response
to air currents and turbulence. These same aerodynamic eects will
also cause exing of the wings. The deformations in the exing are
very small compared to the global body movements so that the
global movements are usually calculated on the basis that the aircraft
is a rigid body. The local deformations of the wing are calculated in
separate analyses with the aircraft modelled as being stationary in a
dened frame of reference.
This chapter presents the practice of using energy principles to evaluate the equilibrium of systems of rigid bodies with respect to applied
loading.

2.1

Equilibrium of rigid bodies

We will dene as rigid a body whose distance between any pair of its
points remains unchanged for any transformation and for all time.

12

Energy methods in structural mechanics

This idealisation is very useful in the analysis of many systems which


can be approximated to be rigid on account of their very limited deformations. More generally speaking, we can say that any displacements
within the body are negligible compared to the gross displacement
experienced by the body itself.
According to this denition, for a rigid body, or system of bodies,
dl 0 for any pair of its points and thus
U0

and

dU 0

2:1

No strain energy can therefore be stored in an undeformable body. The


condition of equilibrium (equation (1.28)) becomes
W 0

2:2

which requires the work done by the external forces for any eld of
virtual displacements to be zero. This equation is generally known as
the principle of stationary work. In the case of rigid bodies it is possible
to perform this test of equilibrium in a very simple manner as the components of any innitesimal displacement eld can be represented in
terms of linear diagrams.
Let us start from the analysis of a rigid body free from any constraints on its movements in the two-dimensional Cartesian space
(i.e. in our case a plane whose points are referred to coordinates relative to two mutually orthogonal axis; essentially this idea was introduced by Descartes1 ). Its positions will be completely dened by the
coordinates of two of its points or, alternatively, by the coordinates
of one point and an angle of rotation in the plane. We say that the
body possesses three degrees of freedom and we call these parameters
(i.e. the two coordinates of a xed point and the angle of rotation)
Lagrangian coordinates (after Lagrange2 ). The numbers of degrees of
freedom can be determined, at least for simple collections of bodies,
by counting the number of coordinates that are needed to draw the
xed shapes of the bodies in some arbitrary position.
To x our ideas consider a simple rigid bar (Fig. 2.1) subject to two
forces PA and PB at the ends A and B. As an eect of its denition,
a rigid bar is one that cannot deform by bending or stretching and can
only experience rigid body movements.
To describe the position of the bar in the xy space we choose as
Lagrangian coordinates, for example, the coordinates of the extremity
1

Descartes, Rene (La Haye, Touraine, 1596 Stockholm, 1650), French mathematician and philosopher.
2
Lagrange, Joseph-Louis (Turin, 1736 Paris, 1813), Italian mathematician of French
origins.

Kinematics and equilibrium of systems of rigid bodies

13

B
C
P (A)

P (B)

y
d (C)

Fig. 2.1

A and the angle of rotation  along axis orthogonal to the xy plane.


The angle of rotation is assumed positive when anticlockwise. This
description of the system means that the coordinates of any point,
say C, along the axis of the bar can be expressed with respect to the
coordinates of the point A, xA and yA, to the angle  and to the
distance dC from the end point A
xC xA dC cos

2:3

yC yA dC sin

2:4

Now consider a possible displacement and refer to Fig. 2.2.


The new coordinates x0 C and y0 C of a point at distance dC
from the extremity A are given by the relationships
x0 C x0 A dC cos0
0

2:5

y C y A dC sin

2:6

where x0 A, y0 A and 0 are the coordinates of the displaced point A


and the angle assumed by the displaced bar, respectively.
B
C

u (C)

u (A)
y

'
A

Fig. 2.2

B
C

14

Energy methods in structural mechanics

If instead of an actual nite displacement we consider a virtual displacement eld, the virtual movement can be dened by
0  
Therefore, equations (2.5) and (2.6) become

2:7
3

x0 C x0 A dCcos cos sin sin

2:8

y0 C y0 A dCsin cos cos sin

2:9

Now, since the virtual displacement eld is required to be innitesimal, for small displacements expand the terms sin and cos in
the following manner
cos0
sin0 2

   
2:10
1!
2!
sin0
cos0 2

   
2:11
cos cos0
1!
2!
and take into consideration only the rst order terms in the above series;
that is, we can assume 2 , 4 ; . . .  1 and 3 , 5 ; . . .  , and
hence
sin sin0

sin  

2:12

cos  1

2:13

Thus equations (2.8) and (2.9) become


x0 C x0 A dC cos dC sin 
0

y C y A dC sin dC cos 

2:14
2:15

or, more conveniently,


x0 C x0 A dx C dy C 

2:16

yC yA dy C dx C 

2:17

where dx C dC cos and dy C dC sin are, respectively,


the components of the distance dC along the axes x and y.
Let us denote with ux x0 x and uy y0 y the components of
the virtual displacement of a chosen point. By subtracting equation
(2.3) from equation (2.16) and equation (2.4) from equation (2.17)
we obtain

ux C ux A dy C 

2:18

uy C uy A dx C 

2:19

Note: cos cos cos sin sin and sin sin cos cos sin .

Kinematics and equilibrium of systems of rigid bodies

15
ux

uy

Fig. 2.3

It is evident that if the components of displacement are drawn for all


the points of the element we get two straight lines that constitute the
diagrams of the displacement components, as shown in Fig. 2.3.
Let us now write down the condition of equilibrium equation (2.2)
following on from the relationships in equations (2.18) and (2.19).
Dening Px and Py , respectively, as the components of the generic
force P along the axes x and y, we have
W Px A ux A Py A uy A Px Bux A ly 
Py Buy A lx  0

2:20

and grouping the terms, we have


W Px A Px B ux A Py A Py B uy A
Px Bly Py Blx  0

2:21

By denition the virtual variation of the Lagrangian coordinates


ux A, uy A and  are totally arbitrary and independent. Therefore, in order to satisfy equation (2.21), each of the terms separately
must equal zero, and equation (2.21) can be replaced by the following
three equations
Px A Px B 0

2:22

Py A Py B 0

2:23

Px Bly Py Blx 0

2:24

16

Energy methods in structural mechanics

Notice that three independent coordinates have been identied as


being sucient to describe the actual position of the body and this,
by the requirement of zero virtual work, will automatically give
three independent equations of equilibrium.
The equilibrium conditions may also be expressed as
Px A Px B

2:25

Py A Py B

2:26

ly Py B

lx Px B

2:27

which means that the forces PA and PB must have the same direction of the axis of the element and be opposite in direction for the
system to be in equilibrium, as shown in Fig. 2.4.
The application of zero virtual work shown above provides exactly
the same result as we would have arrived at using Newton's Laws. In
fact, equations (2.22) to (2.24) are equivalent to
.
.
.

Summation
P of vertical components of the forces being equal to
zero, i.e.
Py 0.
Summation P
of horizontal components of the forces being equal
to zero, i.e.
Px 0
Summation of the P
moments about some point in the plan being
equal to zero, i.e.
M 0.

However, the approach leading to equations (2.25) to (2.27) is somewhat more advantageous than the Newtonian method. A merit of the
energy approach is that in the case of complex systems it may be
relatively simpler for engineers to decide on the numbers of degrees
of freedom than to decide on how many equilibrium equations are
valid. In fact, the energy method automatically determines the correct
number and format of the equilibrium equations.
So far we have dealt with a rigid but unconstrained body. As a
further step and as an example of a constrained body let us now consider the same rigid bar hinged at one end (Fig. 2.5).
B
P(B)

ly

A
P(A)

Fig. 2.4

lx

Kinematics and equilibrium of systems of rigid bodies

17

B
P(B)

Fig. 2.5

It is evident that a vector displacement uA of the extremity A is


not possible any more, so ux A and uy A are both zero. We are
thus left with a single degree of freedom, determined by the Lagrangian
coordinate . Any virtual displacement eld will be dened by the
variation .
Applying a force PB at the point B, the work requirement in equation (2.2) can be written
W Px Bly  Py Blx  0

2:28

yielding the condition of equilibrium


ly Py B

lx Px B

2:29

This requires that the direction of the applied force is along the axis of
the bar.
A further and particular merit of the energy approach is that it automatically rules out the presence of constraint reactions in absence of
friction.
The situation becomes slightly more complicated dealing with a
system of rigid bodies mutually constrained. As usual, it is rst of all
necessary to identify the number of Lagrangian parameters required
to completely dene the position of the system in the Cartesian
space. We have already seen that in the Cartesian two-dimensional
space a rigid unconstrained body possesses three degrees of freedom.
With reference to the hinged rigid bar in Fig. 2.5 this number has
reduced to one because a point of the bar is xed in the space and
no displacements are allowed at that point.
From observation, we can intuitively state a formula that provides
the number of degrees of freedom of a rigid body in two-dimensional

18

Energy methods in structural mechanics


C
A

Fig. 2.6

space, i.e.
f 3c

2:30

where f is the total number of degrees of freedom and c is the number


of simple constraints applied. In the case of the hinged bar (Fig. 2.5),
c 2 because two components of translation are eliminated at the
point A.
If we are dealing with a system of rigid bodies mutually constrained,
equation (2.30) may be re-written as
f 3b c

2:31

where b is the number of rigid bodies forming the system. It must be


emphasised, however, that in this case we can also have relative constraints, i.e. constraints limiting the mutual positions of the bodies.
As an example of this, consider the system shown in Fig. 2.6. This
gure shows two rigid bars mutually hinged. One bar is also attached
to the foundation by another hinge and the other is constrained to slide
parallel to the foundation at the point C. The number of simple
constraints is thus given by the sum of two, due to the hinge, plus
one, due to the sliding support, plus two, due to the internal hinge
(once the position of one of the two members is dened the internal
hinge suppresses two degrees of freedom for the other). We have
therefore
f 3b c 3  2 2 1 2 6 5 1

2:32

This means that all the congurations of the system can be determined
by a single coordinate, which can be the angle of rotation of the rst
rigid bar member  or the y-coordinate of the central hinge, yB.
This can be conrmed by sketching the bodies at some arbitrary position and noting that only one angle has to be dened to enable the
sketch to be completed, as is shown in Fig. 2.7.

Kinematics and equilibrium of systems of rigid bodies


B

19

uy (B)
y

Fig. 2.7

Once the diagrams of the displacement components have been


drawn, see Fig. 2.8, the equation of equilibrium (2.2) can be very
straightforwardly written for any system of forces acting upon the
system.
It is worth noting that we have zero displacements for points A and
B, which are therefore xed centres of rotation, and consequently we
see zero values in both diagrams. Moreover, we read the same value
for the components of displacement in correspondence to the internal
hinge, which is a relative centre of rotation.
This concept is more fully claried in the following section, where
kinematics of systems of rigid bodies for innitesimal displacement
elds will be systematically developed.
ux

uy

Fig. 2.8

20

Energy methods in structural mechanics

2.2

A general innitesimal displacement kinematics for


systems of rigid bodies

So far we have assumed that a rigid body in two-dimensional Cartesian


space, i.e. a plane, possesses three degrees of freedom and that for an
innitesimal rotation we may assume
sin  

2:33

cos  1

2:34

We have subsequently discussed some very simple examples of systems


of rigid bodies. Now a more general theory for the kinematics of rigid
bodies will be developed which will allow us to deal with any system of
rigid bodies mutually connected in the two-dimensional space. Once
these concepts are fully stated, the extension to the three-dimensional
Cartesian space will be briey discussed.
We begin our discourse by introducing the following theorem,
proposed by Euler4 :
Every displacement of a two-dimensional rigid body in a plane can be reduced to a
translation or a rotation about an axis normal to the plane.

This is true in general and holds both for innitesimal and nite displacement elds. We are not going to demonstrate this theorem, but the
reader will nd it easy to verify the generality of the theorem by trial
sketches. In fact, once two points A and B have been chosen on a
rigid body, by denition the line AB will retain the same length for
any displacement of the body. Dening A0 B 0 as the position of the
line after the displacement has occurred, it is straightforward to recognise that in every instance we can bring AB on A0 B 0 (and therefore
the whole body) by means of a simple translation or a rotation.
Conversely, it is worth noting that, dealing with innitesimal displacement elds, we can take the normal to the radius for the arc, as
shown in Fig. 2.9; actually, this is the geometrical meaning of formulae
(2.12) and (2.13), i.e. taking the rst-order terms only in the series of
equations (2.10) and (2.11).
In the present section we will assume that all the displacements are
innitesimal and their components will be simply indicated as ux and
uy . Rotations from the reference conguration, which also are assumed
to be innitesimal, will be simply indicated by .
In this case the kinematics of a free body is extremely simple; once
we have dened the centre of rotation (which is a proper one in the
4

Euler, Leonhard (Basel, 1707 Petersburg, 1783), Swiss mathematician, author of


several fundamental contributions to the mechanics of structures.

Kinematics and equilibrium of systems of rigid bodies

21

A
A'

Fig. 2.9

case of a rotation and an improper one in the case of a simple translation), we can project this point on the axes where we are drawing the
components of displacement. When the centre is a proper one, its displacement is zero, otherwise the centre is said to be at innity and this
means that all the points have the same components of displacement.
In fact, a translation is nothing more than an act of rotation having its
centre at an innite distance, being  0.
The displacement of any of the points of the body, shown in Fig. 2.9,
can be given with reference to the displacement of the centre of
rotation, according to equations (2.18) and (2.19)
ux A ux C dy C

2:35

uy A uy C dx C

2:36

where dx A and dy A are, respectively, the components of the distance d from the centre of rotation C to the point A along the axes x
and y. As stated, in the case of a simple translation the centre is an
improper point at innity,  0 and we have
ux A ux B

2:37

uy A uy B

2:38

where A and B are any pair of points on the body. Note that the diagrams of the displacement components form two straight lines as was
pointed out in the previous section and is illustrated in Fig. 2.10.
With regard to the constraints a distinction is made between external
and internal constraints. External constraints have the eect of limiting

22

Energy methods in structural mechanics


ux

dy (A)

dx (A)

C
uy

Fig. 2.10

the displacement of the body with reference to the points of the plane,
which are considered xed. Internal constraints have the eect of limiting the displacement of the body with reference to the points of another
body of the system. Therefore the internal constraints aect relative
displacements. External and internal constraints both can be classied
according to the number of degrees of freedom they aect. For example the xed end of the rigid rod in Fig. 2.11 prevents the displacement

Fig. 2.11

ux (A) = uy (A) = = 0

Kinematics and equilibrium of systems of rigid bodies


A

23

B
ux (A) = uy (A) = 0

Fig. 2.12

ux (A) 0
uy (A) = 0

0
x

Fig. 2.13

of the point A and any rotation, and therefore suppresses all the three
degrees of freedom of the rod.
A hinge acts to prevent any displacement of the point to which it is
applied but leaves open the possibility of rotation (Fig. 2.12). A roller prevents any displacements of the point P along its normal axis (Fig. 2.13).
A two-pinned bar prevents, for innitesimal displacement elds, any
displacements along its axis and therefore acts like a roller (Fig. 2.14).

ux (A) 0
uy (A) = 0

0
x

Fig. 2.14

24

Energy methods in structural mechanics

Whatever their shapes, it is evident that the constraints described


above can be classied according to the degrees of freedom they suppress
in the plane and can be therefore categorised as simple, double or triple.
Having referred the displacements of all the points of a rigid body to
the position of its centre of rotation, we can now examine the kinematics of a constrained body through the conditions imposed by the
constraints on its centre of rotation. In Fig. 2.15 the most common
Name

Type

Notes

Roller

Simple

ux 6 0
uy 0
 6 0

The centre of rotation must


belong to the normal to the
roller's direction

Simple

ux 6 0
uy 0
 6 0

The centre of rotation must


belong to the axis of the bar

Double

ux 0
uy 0
 6 0

The centre of rotation is the


pinned point

Double

ux 6 0
uy 0
0

The centre of rotation


is improper and lays at the
innite on the direction of
the bars

Triple

ux 0
uy 0
0

No centres of rotation are


allowed

y
x
Two-pinned bar

y
x
Hinge

y
x
Pair of parallel
two-pinned bars

y
x
Fixed end

y
x

Fig. 2.15

Kinematics and equilibrium of systems of rigid bodies

25

Fig. 2.16

types of constraints have been collected, together with their eects on


the centres of rotation of a rigid body.
Of course, it may happen that the constraints imposed on the body
forbid completely the existence of a centre of rotation and in this case
the body is xed in the dened space. This is the case, for example, of
the body constrained by three two-pinned bars, as shown in Fig. 2.16.
This happens because the centre of rotation should belong to the axis
of each of the two-pinned bars but they do not concur in any point.
Therefore the centre of rotation cannot exist and the body is prevented
from any movement.
By denition, a simple constraint imposes a single condition on the
centre of rotation, while a double constraint imposes two conditions
and a triple constraint imposes three. As the position of a proper or
improper centre of rotation is dened by two conditions, it is clear
that at least three simple constraints are needed (or, what is equivalent,
a double plus a simple one or a triple one) to x a body in the space. In
fact, a double constraint is fully equivalent to two simple constraints
and a triple is fully equivalent to three simple ones. This is in accordance with the formula (2.30) stated earlier
f 3c
provided, however, that the constraints are all eective.
In other words, if we consider the example shown in Fig. 2.17 we will
discover that one two-pinned bar is ineective as it requires the centre
of rotation to be on the axes of the other two two-pinned bars. The
centre of rotation is therefore placed in C and the beam still possesses
a degree of freedom.
It must thus be borne in mind that the term c in equation (2.30) (as
well as in equation (2.31)) is the number of the eective simple constraints applied. We have also stated that the internal constraints
restrict the displacements of a rigid body with reference to another

26

Energy methods in structural mechanics

Fig. 2.17

one and therefore aect relative displacements instead of absolute


ones. Nevertheless the internal constraints can still be calculated as
shown in the equation (2.31)
f 3b c
since for any part of the system they suppress the same number of
degrees of freedom as if they were external.
For example, consider the system shown in Fig. 2.18. It is evident
that the internal hinge still suppresses two degrees of freedom for the
element BC whose position is dened by the position of the element
AB and by the angle of rotation 2 . Given the presence of internal
constraints, discovering if a constraint is eective or not becomes a
slightly more complicated matter for a general system of rigid
bodies. In fact, we must recognise along with the presence of
absolute centres of rotation the presence of relative centres of rotation.
Nevertheless it is quite simple to state two theorems regarding the
position of the absolute and relative centres of rotation for any innitesimal displacement eld.
Let us consider two rigid bodies A and B and their innitesimal elds
of displacement in the plane. To x our ideas we can make reference to
Fig. 2.19.
A

Fig. 2.18

Kinematics and equilibrium of systems of rigid bodies

27
ux

cB

cAB

cA

uy

Fig. 2.19

According to what we have previously said we can imagine that the


respective displacement elds can be reduced to rotations with regard
to the centres CA and CB . Now, consider the relative displacement
elds, i.e. the displacements of the points of B with respect to A and
the displacements of the points of A with respect to B. In order to
draw the relative displacements of the points of B with respect to A
we can imagine A to be xed in space and express the relative displacement eld with regards to a relative centre of rotation CAB .
Moreover, we can imagine B to be xed in space and express the
relative displacements of the points of A with respect to B with
regard to the same centre, being the two relative elds equal in magnitude but opposite in direction. As a direct consequence of this fact we
have
CAB  CBA

2:39

28

Energy methods in structural mechanics

This is because that in two-dimensional space the relative centre of rotation is a point which has the same absolute displacement either considered belonging to A or to B. Since we know that the displacement of a
point P in the rst case is orthogonal to the direction CA P and in the
second to the directions CB P we are led to the conclusion that the
relative centre CAB must lie on the line CA CB .
We can therefore state the rst of the proposed theorems
If each of two rigid bodies A and B undergoes an innitesimal displacement eld in twodimensional space, the absolute centres of rotation CA and CB and the relative centre
CAB must all lie on the same straight line.

We call this theorem the rst rule of concatenated displacements for


rigid bodies.
The second rule of concatenated displacements for rigid bodies, i.e. the
second of the proposed theorems, follows in a straightforward manner
from the rst one. If we consider three rigid bodies A, B and C and
their innitesimal displacement elds in the plane we can recognise
the existence of three absolute centres of rotation CA , CB and CC
and of three relative centres of rotation CAB , CAC and CBC . We already
know that relative centres must be aligned with the corresponding
absolute ones. Let us consider one of the bodies, say A, to represent
the xed space. The relative centre CAB becomes the absolute centre
CB and the relative centre CAC becomes the absolute centre CC . Therefore CBC must be aligned with CB and CC and so also must be the
relative centres CAB , CAC and CBC .
Thus the second rule of concatenated displacements for rigid bodies
can be written as:
For innitesimal displacement elds, the relative centres of rotation of any triplet of
rigid bodies must lie on a straight line.

The rst and the second rule of concatenated displacements for rigid
bodies allow us to deal quite simply with the kinematics of any
system of rigid bodies. For example, consider the system of rigid
rods in Fig. 2.20.
This is a system of three rods constrained to a foundation by two
hinges and mutually connected by two more hinges. The number of
double constraints is four and we get from formula (2.31)
f 3b c 3  3 2  4 1

2:40

This means that we have at least one degree of freedom for the system,
which therefore can actually move.
In order to discover if the constraints are all eective, we can
perform the analysis of its kinematics by means of the two rules
stated. First of all locate the absolute centres of rotation C1 and C3 ,

Kinematics and equilibrium of systems of rigid bodies


B

29

C
2

Fig. 2.20

which must coincide with the pinned ends A and D, and thereafter the
relative centres C12 and C23 , which have to be located at the hinges B
and C. This is shown in Fig. 2.21.
The rst rule of concatenated displacements now enables us to dene
the position of the absolute centre C2 . This centre must lie at the intersection of the lines C1 C12 and C23 C3 , as it is shown in Fig. 2.22. No
constraints can be classied as ineective and the knowledge of these
centres is sucient to draw the diagrams of the components of displacement of the system. These diagrams depend on the parameter chosen
B C12

C23 C
2

A C1

C3 D

Fig. 2.21

30

Energy methods in structural mechanics


ux

C2

C12

C23

C3

C1

uy

Fig. 2.22

to represent the single degree of freedom available, e.g. the angle of


rotation of the rod 1.
Now, consider a nal aspect of concatenated displacement. So far
we can assert that for a system of n rigid bodies undergoing a certain
eld of innitesimal displacements we have n absolute centres of rotation. We can also arrange all the centres of rotation in a symmetrical
matrix
0
1
C11    C1n
..
.. A
@ ...
2:41
.
.
Cn1    Cnn
where the centres of rotation o-diagonal in the matrix are the relative
ones. Therefore, the number of relative centres is
nn 1
2
and the total number of centres is
nn 1
2

2:42

2:43

Kinematics and equilibrium of systems of rigid bodies


A

C
1

31

E
2

Fig. 2.23

We can now reect on the fact that we usually take into consideration systems whose elds of displacement are dependent on a single
parameter; that is, one degree of freedom activated at a time. In
order to dene any diagram of displacement components for such
systems we must draw n straight lines and we need to dene 2n 1
points. We are thus led to the conclusion that in order to draw these
diagrams we do not need to locate all the centres of rotation but we
just need to know the position of 2n 1 centres.
Finally, it is necessary to underline that a careful study of the centres
of rotation of the system must be always carried out, as this is the only
way to investigate the kinematics of rigid bodies in a consistent and
rigorous manner. In fact, as we have already pointed out, we can
often have an odd layout of the constraints themselves. For example,
let us consider the structure in Fig. 2.23.
In this system we have two rigid rods constrained by two double
constraints and three simple ones. If equation (2.31) is applied, regardless of the eectiveness of the constraints, we have
f 3b c 3  2 2 2 1 1 1 1

2:44

The fact that the number of degrees of freedom is negative might suggest that the system is not only allowed to move, but it has one redundant constraint, too. However, if we carry on the study of the centres of
rotation, as illustrated in Fig. 2.24 we discover that rod 1 cannot move
because its absolute centre of rotation must belong to the axes of the
two-pinned bars and at the same time it must be an improper centre
on the direction of the pair of parallel two-pinned bars. As the twopinned bars are not aligned with the pair of parallel ones this centre
cannot exist. Moreover, each of these two-pinned bars is sucient to
forbid the existence of the centre C1 and therefore one is redundant.
The hinge in D can therefore be considered as an absolute centre of
rotation for rod 2 and this is the case, as the roller in E requires this

32

Energy methods in structural mechanics

C1
B

A
y

Fig. 2.24

centre to be aligned with its axis. It is evident that the roller constitutes
an ineective constraint. Therefore, element 2 can actually move and its
eld of displacements is dened by a single degree of freedom, i.e. the
angle of rotation , as is shown in Fig. 2.25.
Equation (2.31) gives the correct result if we drop from the calculation the redundant constraint and the ineective one
f 3b c 3  2 2 2 1 1

2:45

To summarise with respect to the kinematics of systems of rigid bodies,


we can have two types of constraints
(a)
(b)

eective constraints, ones that are able to suppress some


degrees of freedom of the body
ineective or redundant constraints, ones that aect degrees of
freedom already suppressed by other constraints.
A

Fig. 2.25

Kinematics and equilibrium of systems of rigid bodies

33

The only way to discover the eectiveness of a constraint is to conduct


a careful study of the kinematics of the system.

2.3

Reactions of constraints

So far we have assumed that the condition of equilibrium for the


systems of rigid bodies is given by the principle of stationary work
(equation (2.2)), that is,
W 0
which requires the work done by the external forces for any eld of virtual displacements to be zero. If we remember the denition of virtual
displacement eld, i.e. a eld of displacements that is innitesimal and
compatible with the constraints of the system, but is imaginary and
does not actually occur, we are immediately led to the conclusion
that the principle of stationary work can only be applied to those systems that possess at least one degree of freedom. This is a requirement
because a system that does not possess any degrees of freedom must be
considered in equilibrium irrespective of the particular set of applied
forces, since we will always have
ui A 0

2:46

for any of its points and therefore equation (1.26), which has been proposed as the denition of equilibrium,
X
1
_2
K
2 mi xi 0
i

always holds true. However, the principle of stationary work can also
be applied to evaluate the forces that the constraints must apply to the
system to keep it in equilibrium. This approach can be dened as the
principle of substitution of constraints and establishes an equivalence
between a geometrical entity, i.e. a constraint, and a static entity, i.e.
a force.
As an example, consider the three-pinned arch shown in Fig. 2.26,
that is subjected to the vertical force P at the hinge B.
Equation (2.31) gives
f 3b c 3  2 2 2 2 0

2:47

as there are no ineective constraints, since the centres C1 , C2 and C12


are not aligned.
Therefore the system is in equilibrium regardless of the particular set
of applied forces. As the hinge at A can be imagined as being composed
of two orthogonal rollers, as pointed out in Fig. 2.27, we can conceptually suppress the hinge at A and leave the vertical roller only
(Fig. 2.28).

34

Energy methods in structural mechanics

C12 B

C1 A

C2 C

Fig. 2.26

In this situation the system can move to take up a displaced shape,


and we can evaluate, using the principle of stationary work (equation
(2.2)), the vertical force exerted on A in order to maintain the system in
equilibrium. This force can be thought of as being equivalent to the
eect of the horizontal roller.
With reference to Fig. 2.28, equation (2.2) can be written as

R 0
2:48
2
In fact, once we have assigned an arbitrary vertical displacement 
to the point A, the vertical displacement of the point B is seen to be
=2 and from equation (2.48) we have
W P

P
2:49
2
The negative sign means that the force R has the opposite direction to
the force P. We call R the reaction of the horizontal roller. The procedure is therefore very simple
R

A
A

Fig. 2.27

Kinematics and equilibrium of systems of rigid bodies

35

C12

C1 C2

R
uy

uy (A) =

uy (B) =
2

Fig. 2.28
In order to obtain the reaction of a certain constraint we can suppress that constraint
and draw the innitesimal displacement eld thus allowed (if any). The reaction is
represented by the force that, acting instead of the constraint, renders the total work
of all the applied actions to be zero. The reaction is considered to act in the same
direction of the suppressed constraint.

Of course the procedure is meaningful when there are no redundant


constraints. If the suppression of a constraint does not enable the
system to have a displacement eld, we are not able to identify any
work in equation (2.2) and cannot evaluate the reaction. Actually, it
should be noticed that in the presence of redundant constraints a
system of rigid bodies is generally in equilibrium and the fact that
we are not able to calculate the reaction of some constraints is only
due to the idealisation of rigidity. In the following text we will learn
that dropping this simplication allows the determination of the reaction of all of the constraints.
For the moment, we say that systems whose constraint reactions can
be evaluated according to the scheme of rigid bodies are called statically determinate, otherwise they are called statically indeterminate.

2.4

Internal reactions for a straight beam

In the previous section it was assumed that external constraints are


able to transmit certain forces to the bodies in order to keep them in
equilibrium. In the same manner we can assume that a certain part

36

Energy methods in structural mechanics

P
z

P
2

P
2

Fig. 2.29

of a body transmits a system of forces to the remaining part of it in


order to keep that part of the body in equilibrium. The material in
this section is illustrated, for clarity, by the simply supported beam
shown in Fig. 2.29.
Using the principle of stationary work (equation (2.2)) we can
simply determine the reactions transmitted by the supports at A and
B, i.e.
P
2:50
2
We can also contemplate an imaginary vertical cut through the beam
at the section marked YY, at a distance z from the support A. In
this situation we can consider the portion of beam between the support
A and the section YY as a free body subjected to the force P=2 at A,
as shown in Fig. 2.30.
Notice that a free body is dened to be
RA RB

a portion of a structure, or system, that is abstracted from the structure and the
continuity of the structure between the free body and the rest of the structure is
represented by forces applied at the boundary of the free body.

Thus the portion of beam shown in Fig. 2.30 will be subjected to the
resultant of the forces transmitted by the portion of beam on its
right side, which must be able to maintain it in equilibrium.
Looking at the case of a beam contained in the plane zy, let us
decompose this resultant force in terms of a horizontal component N

Kinematics and equilibrium of systems of rigid bodies

37

z
Y

Y'

Y'

P
2

M
N
S

Fig. 2.30

and a vertical component S both applied to the centroid of the cross


section YY, together with a moment M, as is customary for any
system of forces in the plane. We will call these components axial
force, shear force and bending moment, respectively, in the beam at
the section YY. Their values are calculated using the principle of
stationary work, equation (2.2), and are
P
Pz
and
M
2:51
2
2
If the principle of zero work is applied to the rigid beam on the right
hand of the cut, we nd that the resultant of the forces transmitted
to the portion of beam on the right side of the cut must have the
same magnitude but opposite direction of the resultant of the forces
applied to the left side of the cut; which is in accordance with the
third of Newton's Laws. Therefore, we have
N 0;

P
Pz
and
MR
2:52
2
2
where the superscript R denotes that the forces are applied to the right
side of the cut.
Notice that no superscripts were applied to equation (2.51). This is
because it is conventional to assume that the axial force, the shear
force and the bending moment, i.e. the so called characteristic of internal action, are the components of the force acting on the left side of the
cut. They are normally referred to a Cartesian reference frame in which
the z-axis coincides with the axis of the beam, aligned with the centroids of the normal cross section (which for simplicity here is thought
to be uniform along the beam length), and the x and y-axes coincide
N R 0;

SR

38

Energy methods in structural mechanics

Fig. 2.31

with the principal axes of inertia of the cross sections. The convention
adopted in this book is for the z axis to be positive when measured
from the left-hand end of the beam and the y axis is ordinarily arranged
to be vertical with downwards positive, as shown in Fig. 2.31.
Of course, this is purely a convenient convention and others can be
equivalently adopted, provided they enable consistent communication.
By sequentially applying the equilibrium equation (2.2) to any
chosen portion of the beam under consideration, it is easy to recognise
that
(a)

(b)

(c)

the axial force at any section is equal to the opposite of the


sum of all the horizontal components of the external forces
acting to the left side of that section or to the sum of all the
horizontal components of the external forces acting to the
right side of that section
the shear force at any section is equal to the opposite of the
sum of all the vertical components of the external forces
acting to left side of that section or to the sum of all the vertical
components of the external forces acting to right side of that
section
the bending moment at any section is equal to the opposite of
the sum of all the moments yielded by the external forces
which act to the left side of that section with reference to the
centroid of the section itself or to the sum of all the moments
yielded by the external forces which act to the right side of that
section.

It is therefore quite an easy procedure to determine the values of


the characteristics of the internal action at any section. At the same
time it can be valuable to obtain a graphical representation of the
distribution of the various forces acting along the beam by plotting
their distribution along the length of the beam. These plots are

Kinematics and equilibrium of systems of rigid bodies

39

w (z)
z

Fig. 2.32

commonly referred to as axial force, shear force and bending moment


diagrams.
To simplify the preparation of the diagrams it is very useful to establish some general relationships between distributed forces, shear force
and bending moment. Distributed force is nothing more than a system
of forces that act over at least part of the length of the beam. This type
of force is generally measured per unit length and is represented by a
function qz of the position z along the axis of the beam. For example,
a special case of a uniformly distributed force is given by the self-weight
wz of the simple supported beam of Fig. 2.32, while a non-uniformly
distributed force is exemplied by the cantilever beam shown in Fig.
2.33 (for the sake of simplicity both these forces are considered to
act normally to the beam axis).
Given a displacement eld uy z, the work of a distributed force
qz between the positions z0 and z1 is obviously provided by the
integral
z1

qzuy z dz

2:53

zo

as shown in Fig. 2.34.

q (t)
z

Fig. 2.33

40

Energy methods in structural mechanics


q (z)
z0

z1

z
uy (z)

Fig. 2.34

Let us now take into consideration an innitesimal element of beam


obtained by means of a conceptual cut at two very close cross sections
(Fig. 2.35).
The element under consideration will be subject to an axial force N,
a shear force S and a bending moment M at its left hand end and to an
axial force N dN, a shear force S dS and a bending moment
M dM at its right end. Since the length dz is very small, we can
neglect the amount by which the distributed vertical load qz varies
over the element.
The equilibrium of the innitesimal part of beam can be once more
investigated by means of the principle of stationary work, equation
(2.2). There are three degrees of freedom in the zy plane, so
 W N N dN  uz S S dS qz dz  uy
M M dM S dz  0

2:54

In this equation the second order innitesimal quantities, namely the


products of the type dS dz, are neglected.
q (z)

M + dM
N + dN

S + dS
dz

Fig. 2.35

Kinematics and equilibrium of systems of rigid bodies

41

Equation (2.54) simply leads to the following three equations


dN 0

2:55

dS
qz
2:56
dz
dM
S
2:57
dz
which must hold for the equilibrium of any portion of the beam. In
particular note that equation (2.56) implies that the slope of the
shear force distribution is equal in magnitude but opposite in sign to
the intensity of the vertically distributed lateral load at any position
along the beam. Similarly, note that equation (2.57) implies that the
slope of the bending moment distribution is equal to the value of the
shear force at any position along the beam.
However, it is obvious that the presence of an applied vertical point
force P (or couple M) at a certain section A implies that for A0 and A00 ,
the sections immediately on either side of the conceptual cut, we have
SA00 SA0 PA
00

MA MA MA

2:58
2:59

Therefore, according to equations (2.56) and (2.57), an applied vertical


point force P leads to an abrupt step change in both the shear force
distribution and the slope of the bending moment distribution, while
an applied couple M causes an abrupt step change in the bending
moment distribution but leaves the slope unchanged.
At this point a simple step by step procedure can be outlined for
plotting the shear force and bending moment diagrams in cases of
statically determined systems of beams
(a)
(b)
(c)

nd the reactions of the constraints at the ends of each beam


nd the values of the shear force at key points along the beam
(e.g. points of application of point forces)
plot the shear force diagram according to equations (2.56) and
(2.58).

Notice that equation (2.56) immediately leads to the following relationship


z2
Sy z2 Sy z1 qz dz
2:60
z1

for any couple of positions which do not include an applied vertical


point force.

42

Energy methods in structural mechanics

(d )

locate the sections where the shear force is zero. These sections, according to equation (2.57) are positions of maximum
or minimum bending moment
(e) nd the values of the bending moment at key points along the
beam
( f ) plot the bending moment diagram according to equations
(2.57) and (2.59).

Notice that equation (2.57) also leads to an useful integral relationship:


z2

M z2 M z1

S z dz

2:61

z1

for any two sections of the beam which do not include an applied
couple.
An example of shear force and bending moment diagrams for a
system of beams is shown in Fig. 2.36.
The presence of axially directed distributed or point loads and the
consequent procedure for drawing the diagram of the axial force can
be treated by means of trivial modications of the previous equations
and is left to the reader.
q

l/2

l/2

+P

+ P + ql

Pl ql
2

P
2

+ Pl
4

Fig. 2.36

Kinematics and equilibrium of systems of rigid bodies

2.5

43

A three-dimensional example

Up to now examples in this chapter have been restricted to those in


two-dimensional Cartesian space, i.e. to rigid bodies which can be
analysed with reference to a plane of symmetry that contains also
the applied forces and constraints. This may seem a severe restriction,
but actually it is not, as a very large number of structures can be
analysed as formed of substructures which can be treated as a plane.
However, the concepts which enabled the treatment of equilibrium
in the preceding sections remain exactly the same in the case of rigid
bodies in three-dimensional Cartesian space. The condition of equilibrium is obviously the principle of stationary work (equation (2.2)),
that is
W 0
for any eld of virtual displacements. Of course the virtual displacements take place in the three-dimensional space, but apart from this
fact all the simplications deriving from their being innitesimal
hold true in the same manner.
In order to illustrate this fact with a practical example, consider the
beam shown in Fig. 2.37.

x
B

2
C

y
z
l
5

Fig. 2.37

44

Energy methods in structural mechanics


b

P
3

Section A

Section B

Section C

Fig. 2.38

This is a simple steel I beam constrained by six two-pinned short


bars and subject to a force P at the middle of the span l. The exact position of the constraints and of the applied force is shown in Fig. 2.38
with reference to the cross sections at A, B and C.
It is easy to infer that a rigid body in the three-dimensional space
possesses six degrees of freedom. For example, once we have assumed
the three orthogonal axes x, y and z to be immovable on the beam, it is
clear that the position of the beam in the space will be determined by
the position of these axes. As the position of the axes is uniquely identied by, say, the coordinates of their origin O0 and by the three angles
of rotation that the axes themselves form with reference to another
reference frame xed in the space, it follows that six parameters are
sucient to dene the position of the beam in the space. The situation
is shown in Fig. 2.39.
Thus, six simple constraints may be sucient to suppress any degree
of freedom, provided they are all eective. This fact, as has been said
before, can be investigated only by analysing the kinematic of the body.
With reference to section A of Fig. 2.38, it appears clear that this
section can only rotate about the centre given by the intersection of
the directions of the pinned bars 2 and 3. The same happens to section
C, but section B is allowed to rotate about a centre which must lie along
the axis of the pinned bar 4 and therefore cannot be aligned with those
of sections A and C. We can therefore conclude that the beam cannot
undergo any rotation about any axis parallel to the axis z. As any translation of the sections A and C is forbidden, too, we see that the origin O0
of the axes x, y and z cannot move in the plane xy as well. Three
degrees of freedom consequently must be eliminated.
On account of the presence of pinned bars 2 and 5 any rotation
about the x axis must be excluded and the same happens to any
rotation about the y axis on account of the presence of pinned bars

Kinematics and equilibrium of systems of rigid bodies

45

3
2
O'

x
O

Fig. 2.39

3 and 6. Finally, no translations along the z axis are allowed by the


pinned bar 1 and therefore all the degrees of freedom must be considered as eliminated. The constraints are all eective.
In order to nd the reaction of the constraint represented by the twopinned bar 4, we can suppress it and draw the innitesimal displacement eld thus allowed. This turns to be a rotation about an axis
identied by the centres dened by the intersection of the directions
of the pinned bars 2 and 3 and by the intersection of the directions
of the pinned bars 5 and 6, respectively. Thus, all sections can rotate
about the trace of this axis in their respective planes and the situation
for section B is shown in Fig. 2.40.
The equilibrium equation (2.2) can be therefore written as
 W Pb  Rb  0

2:62

and we have
R P

2:63

where the negative sign indicates that the reaction R is opposite in


direction to the force P. The reader will nd very easily, by the same
way, that the reactions of all the other constraints are null.

46

Energy methods in structural mechanics

b
b

Fig. 2.40

W
l
P

l/2

Fig. 2.41

Suggested exercise problems


1.

2.

A simple pin-jointed frame composed of rigid bars is shown in


Fig. 2.41. The bars AB and CD are at an angle  to the vertical.
Use the principle of stationary work to derive a relationship
between the weight W of the bars and the force P required to
maintain equilibrium of the frame.
[Ans: P 4W tan ]
Two weightless rigid bars are shown in Fig. 2.42. One is pivoted at
A and the other at B. The bars contact without friction at C. Use
the principle of stationary work to derive a relationship between
the forces P and W and the angle  to maintain equilibrium.
[Ans:  cos1 W=2P]

Kinematics and equilibrium of systems of rigid bodies

47

l/2
B
l/2

C
l/2

P
W

Fig. 2.42

h
W
l/2
A

Fig. 2.43

3.

4.

A bar AB of weight W is suspended from a rigid foundation by


two extensionless but exible wires, each of length l. A second
bar has pins attached at a distance l=2, as shown in Fig. 2.43.
Use the principle of stationary work to determine the distance h
that maintains equilibrium of the two bars if it is assumed that
no friction occurs between the pins and the wire.
[Ans: h 0:711l ]
Two solid rigid cylinders of the same density lie inside the bore of a
rigid cylinder, as shown in Fig. 2.44. Use the principle of stationary work to derive the value of the angle  that corresponds to
equilibrium of the system.
[Ans:  18:38]

48

Energy methods in structural mechanics

4D

2D
D

Fig. 2.44

375 mm
200 m

50 mm
75 mm

P
2 kN

400 m
m
200 m

Fig. 2.45

5.

A clamping xture is shown diagrammatically in Fig. 2.45. The


elements of the xture can be treated as being rigid bars and the
pin connections as being completely frictionless. Use the principle
of stationary work to calculate the force P that is required to
generate the clamping force of 2 kN.
[Ans: P 0:445 kN]

3.

Deformation of bodies and


material properties

Introduction

The previous chapter has presented the application of energy principles


to the analysis of systems of rigid bodies. These are systems in which
the deformations of the bodies are insignicant when compared to
the movements of the component bodies in the system. At that stage
we illustrated this with the example of the collision of two billiard
balls in which the deformations at the contact area are negligibly
small compared to the distances moved by the balls. The rigid body
approach allows us to predict the macroscopic, or global, movements
of the components parts of the systems.
However, no body is ever completely rigid and we can consider the
analysis of most systems at two scales; one is the macroscopic that, as
described above, is concerned with the movements of the bodies as if
they never change their shape. The other scale is rather microscopic
in which we are concerned with the detailed changes of shape of the
bodies comprising the system. In this scale the colliding billiard balls
are `frozen' in macroscopic space and only the local deformations in
the balls as they impact are considered. In reality, of course, systems
operate simultaneously in the macroscopic and microscopic scales
but it simplies analysis, usually without signicant loss of accuracy,
to separate the analytical treatment of the two scales.
It must be noted that there are also many systems like, for example,
rubber bands, which experience deformations on a large scale and
whose analysis cannot be performed according to the earlier outlined
procedures. These systems require a special treatment and there are
several theories in structural mechanics which allow us to tackle
such analyses. However, on account of their complexity, these methods
of analysis are denitely out of the scope of the present book and the
interested reader can nd excellent introductory textbooks to this
fascinating area of mechanics.

50

Energy methods in structural mechanics

Before moving on to formulate the energy involved in the deformation of a body we need to examine how materials deform and to characterise the deformations, and the forces that cause the deformations.
The approach followed is to subject a sample of the material to a
simple one-dimensional test and to generalise, using mathematical
constructs, the results from the test to obtain a description of the properties of the material for deformations in three dimensions.

3.1

Deformation properties in one dimension

The approach followed for most materials is exemplied here by reference to testing of a typical structural carbon steel. Figure 3.1 shows a
specimen machined to a suitable shape for the uniaxial test. The shape
of the specimens, Figs 3.1(a) and (b), rectangular and round section,
respectively, have evolved during a century of material testing. The
objective of the shape is to ensure that the gauge length, lg , has uniform
conditions resulting from the axial force, P, applied to the specimen in
the test machine and is not inuenced by the complex conditions of
forces in the parts of the specimen gripped by the testing machine, as
shown hatched in Fig. 3.1. The test section shown in Fig. 3.1(a) has
thickness t and breadth b giving an area A.
During a typical test to evaluate the quasi-static properties of the
material the axial tensile load is applied at a slow rate that eliminates
any dynamic eects and causes the gauge length gradually to extend
from lg to lg lg . The change of length is measured by an instrument
t
P

P
b

lg
(a)

P
d

lg
(b)

Fig. 3.1

Deformation of bodies and material properties

51

1
2

lg

Fig. 3.2

that is called an extensometer and is attached to the surface of the test


specimen. Several types of extensometers have been developed during
the past century to measure the very small change of length, typically
lg 0:05 to 0:5 mm. The instrument may be based on mechanical
magnication of the changes of the gauge length or it may use the
change of electric resistance of a length of wire that is constrained to
change its length proportionally with the change of the gauge length.
Figure 3.2 shows typical results from tests on specimens with dierent
values of cross-sectional area, say by increasing the thickness.
It is evident that the initial part of the curves is fairly linear and the
load at which the curve reaches a maximum is dependent on the area.
Figure 3.3 shows comparable results from tensile tests specimens in
which the area is maintained constant and the gauge length is changed.
Moreover the load at which the curves reach a maximum is unaffected by the gauge length but the extension is aected by the changes.

lg1

lg2

lg3
lg3 > lg2 > lg1

lg

Fig. 3.3

Energy methods in structural mechanics

P/A

52

lg /lg

Fig. 3.4

It is also evident from tests that if the ordinates of the graphs of the
results are normalised, as shown in Fig. 3.4, by dividing the applied
load by the area of the specimen at the gauge length and by dividing
the extension by the gauge length, then all the curves coincide.
The normalisation denes two parameters that are used widely in
structural analysis, namely,
strain; "  lg =lg

and

stress;   P=A:

If in the above denition of strain the initial gauge length is used, then
what is normally dened as the nominal strain is obtained. A dierent
measure of strain, called the true strain, can be obtained by dividing the
elongation of the bar by the actual gauge length, which increases as the
tensile load is applied. In the development of this book we will always
make reference to the nominal strain.
Stress is a parameter that can be derived simply in the analysis of the
results of a tensile test by measuring the applied load and dividing it by
the corresponding initial value of the area of the cross-section of the
gauge length, which gives the nominal stress (a more exact value of
the axial stress, known as the true stress, can be calculated by using
the actual area of the bar, which can become signicantly less than
the initial one). The denition of stress depends on the assumption
that the load is uniformly spread across the cross section of the
gauge length. In general, the state of stress is complex in practical
structures and since it cannot be measured directly it must be inferred
from measured values of strains and deformations using assumed or
measured values of material properties. Alternatively, strains can be
measured directly on the surface of a structure by arranging the readings from extensometers to be normalised by dividing the measured

Deformation of bodies and material properties

53

Fig. 3.5

elongation of the gauge length by the gauge length of the instrument.


Modern electrical resistance strain gauges can have very short gauge
lengths, currently as small as lg 2 mm, and the electronic monitoring
of the changes in resistance of the gauges directly provide strain values,
usually up to a maximum value of strain of 0:05, i.e. 5%. An example
of the measurement of strains on a tensile test specimen using microgauges is shown in Fig. 3.5.
Other methods, such as the use of laser interferometry, also enable
the direct measurement of strains in engineering structures.
Figure 3.6 illustrates the full range of results that are obtained from
a tensile test on a structural grade carbon steel.
The linear part of the graph extends up to a strain value of about
0:002, that is 0:2%, and thereafter the stressstrain graph becomes
non-linear; this is called yielding. The gure shows the stress increasing
gradually with increasing strain up to a strain level of about 0:04, i.e.

54

Energy methods in structural mechanics


600

Stress, (N/mm2)

500
400
300
200
100
0

10000

20000
Strain, (%)

30000

40000

(a)
600

Stress, (N/mm2)

500
400
300
200
100
0

20000 40000 60000 80000 100000 120000 140000 160000 180000 200000
Strain, (%)
(b)

Fig. 3.6

4%, this is called strain hardening. In some structural steels the stress
does not increase with increasing strain following the onset of yielding,
up to a strain of about 2%. This is called the yield plateau, or Luder's
plateau after the famous German physicist. The maximum load is
reached at a strain of about 810%, as is shown in Fig. 3.6(b), at
which stage the cross-sectional area of the tensile specimen reduces
rapidly and the specimen fractures, this is sometimes called necking.
The ultimate load, at which necking begins, is usually about 10
20% greater than the load at which yielding is initiated. Fracture

Deformation of bodies and material properties

Stress,

55

loading
(ii)

unloading
(i)

Strain,
(a)

Stress,

Strain,
(b)

Fig. 3.7

occurs usually at about a strain of 1825% and is a measure of the


ductility of the steel.
Close to the yield strain the test specimen completely recovers its
original geometry when the load is removed. In other words the
material properties are reversible; this portion of the stressstrain
properties are classied as elastic. Thus for the elastic condition, as
shown in Fig. 3.7(a), provided that the load is kept to less than the
value at which linearity is lost (that is the proportional limit of
strain, "p ), the variation of load, increasing and decreasing, is directly
related to the strain.
In many structural steels the proportional limit and the loss of elasticity due to the onset of yielding are virtually coincident and thus the

56

Energy methods in structural mechanics

limiting elastic condition for design assessment is taken to be the yield


strain, or stress whichever is appropriate. An approximation to the
stressstrain curve is made that considerably simplies analysis and
is particularly relevant to the methods presented in this text. This is
that stressstrain curve up the yield strain is modelled as being
linear, and this gives rise to the linear elastic conditions that underlie
the later developments in this text.
If the load in Fig. 3.7(a) is reduced during the test, at a strain greater
than the yield strain, "0 , the graph follows the line (ii) in Fig. 3.7(a)
until the load is completely removed at point D. In other words,
after the yield strain is reached the material no longer has a unique
stressstrain relationship and the loading and unloading paths are
not reversible back to the original material condition of zero stress
and strain. If the test specimen is reloaded from D it will follow
the slope of the original elastic line, AB, until there is a loss of
proportionality and the stressstrain graph intersects with the yield
line at E. Thus, provided the change of strain due to external loading,
starting from zero load value, is less than the yield strain for the
material, the material may be regarded as elastic no matter what its
previous history of loading may have been. Tests show that for
many elastic materials the stressstrain properties are the same in compression as in tension.
Figure 3.7(b) shows the stressstrain curve for a material, typied by
an aluminium alloy, that has a much more marked non-linear stress
strain curve: from points A to B the relationship is fairly linear, but
from B to C there is a distinct curve. The limit of proportionality at
B does not necessarily indicate the limit of elastic behaviour of the
material. The material may load and unload elastically along the line
ACA. A pseudo elastic property is often constructed for such a
non-linear elastic material, in that the stressstrain relationship is
taken to be linear up to the load corresponding to point C, as shown
by AD in Fig. 3.7(b).
The analysis methods developed in this book are concerned only
with the linear elastic part of the stressstrain curve of the material
and with predicting the limiting levels of loading at which the material
no longer remains elastic. The methods therefore apply to the great
majority of engineering analyses that are performed as a basis for
design. The fact that the material does not fracture at the yield strain
and that the engineering structure can carry increased load before
material fracture actually occurs is used by structural engineers as a
measure of in-built safety for their structure. The elastic material
model that is used as the basis for the analysis methods developed
here is shown in Fig. 3.8.

Deformation of bodies and material properties

57

E = tan

Fig. 3.8

The strain is directly proportional to the load applied in the tensile


test, the constant of proportionality is called the material modulus, or
Young's modulus.1 The material modulus is given the symbol E. The
limiting allowable level of strain for the elastic analysis is the yield
strain, "0 . Thus the strain is related to the applied load in a simple
uniaxial loading test by
P


3:1
AE E
This equation is often referred to as Hooke's law, after Hooke,2 who
rst proposed this loaddeection proportionality.
The values of E for a typical carbon steel range from 195 000 to
210 000 N/mm2 and commonly a value of 205 000 N/mm2 is used in
design analyses. The yield strains for steels widely used in structural
applications vary from 0:12% to 0:25% depending of the grade of
material. Special heat-treated steels can have yield strains of up to
0:5%.
Stress conventionally plays a central role in structural analysis. Many
text books in strength of material and in structural mechanics make
great use of stress in the development of the theory of structures and
in the presentation of the limiting strength of materials. Mathematical
descriptions of the equilibrium of components or complete structures
"

1
Young, Thomas (Milverton, Somerset, 1773 London, 1829), English natural philosopher.
2
Hooke, Robert (Freshwater, Isle of Wight, 1635 London, 1703), English physicist.

58

Energy methods in structural mechanics

are often based on the components of stress in the structure, etc. A


structural analysis then has the objective of evaluating the magnitudes
of these stresses throughout the structure and comparing the greatest
magnitude with the maximum allowable stress for the particular
material from which the structure is composed. This comparison then
determines the maximum allowable value of the loading that may be
applied to the structure. In this text, which is primarily concerned
with linear elastic structural analysis using energy methods, the basis
of the analysis methods is not stress but the deformation, and therefore
strain elds, that exist in structure and their components. The arbiter of
the allowable loading is not the level of the maximum stress, but the
level of the maximum mechanical strain in comparison with the allowable maximum strain for the material. This text therefore does not use
the concept of stress widely, not for any reason other than that from this
standpoint in most cases there is no real need to make resort to it.

3.2

One-dimensional thermal strain

All structural materials tend to expand when heated, as illustrated in


Fig. 3.9.
The rollers in the gure represent that the specimen is not restricted
from axial movement. The thermal strain "t is dened as the ratio of
the extension of a gauge length to the original gauge length, that is
"t

 lg
lg

3:2

The ratio of the change of length to the original length is found


experimentally to be linearly related to the change of temperature,
for temperature ranges up to about 100 8C. Thus,
"t T

i:e: lg lg T

3:3

where is the coecient of linear expansion of the material. T is the


change of temperature.
The coecient of linear expansion varies slightly with temperature,
but is taken to be constant in the analysis methods developed in this

lg

lg

Fig. 3.9

Deformation of bodies and material properties

59

Coefficient of thermal expansion ,


106 C 1

16

14

12

100

200

300

Temperature C

Fig. 3.10

book. The value of the coecient for a structural steel in the range
0 < T < 100 8C is usually taken as
1:25  105 mm=mm=8C

3:4

The actual variation of the coecient of expansion with the temperature is shown in Fig. 3.10 for a high strength carbon steel.
Notice that the thermal strain equals the yield strain at a temperature change of
"
T 0
3:5

that is, for a medium grade structural steel
1:5  103
T :
120 8C
3:6
1 25  105
The foregoing remarks relate to changes of length that result from
changes of temperature where there are no constraints or applied
forces. The situation illustrated in Fig. 3.11 is one where the change
of temperature occurs simultaneously with the application of an
external force P.
The strain " induced in the bar is the sum of the mechanical strain due
to the force and the thermal strain due to the change of temperature.
T
P

Fig. 3.11

60

Energy methods in structural mechanics

Fig. 3.12

Thus
P
T
3:7
EA
The constraint can take the form of displacement restrictions on the
boundaries of the body. An extreme example is illustrated in Fig. 3.12
in which the ends of a bar are prevented from movement, i.e. l 0
and therefore the total strain in the bar is zero.
"

P
T 0 and P EA T
3:8
EA
The negative sign implies that the force is compressive to maintain zero
total strain for the increase in temperature.

3.3

Three-dimensional strain

Section 3.1 has presented the observations of the change of length of


the gauge length of a tensile test specimen when viewed in the direction
of the applied load. However, test observations show that the crosssection of the gauge length also changes its dimensions. That is, as illustrated in Fig. 3.13, the application of the load P causes changes in the
dimensions normal to the direction of the load.
The thickness t changes to t t and the breadth changes from b to
b b. If we dene a set of orthogonal coordinates, x, y and z as

t + t

lg

x
z

Fig. 3.13

lg + lg

b + b

Deformation of bodies and material properties

61

shown in Fig. 3.13, the strains in the three directions are


"x

 lg
;
lg

"y

b
;
b

"z

t
t

3:9

Observations show that that t and b are negative for positive


values of l. The test results show that the strain normal to the
direction of the applied load is related to the strain in the direction
of the load by a constant, called Poisson's ratio, , after S. D. Poisson.3
Thus,
"y
"z

3:10

"x
"x
The values of Poisson's ratio measured in tests for the elastic part of
the stressstrain curve of a typical carbon steel are about 0:270:31
and 0:3 is commonly used in design analyses for steel structures.
Poisson's ratio increases to about 0:5 for strains much greater than
the yield strain.
During the application of the load P the volume of the gauge length
will change due to the elastic strains. The change of volume V is
V l1 "x b1 "y t1 "z lbt
lbt"x "y "z f "2 ; "3 ; . . .

3:11

Since the strain is very small for elastic deformations, the higher order
terms in the equation can be ignored as being insignicantly small and
the volumetric strain "v that is the ratio of the change of volume to the
original volume, is
V
"x "y "z "x 1 2
3:12
V
Uniaxial tests are commonly carried out to establish relevant
material properties for use in design analyses; typical results have
been discussed above. It is possible to generalise these results to
describe the possible eects of applying simultaneously load in the
three orthogonal directions. The practical diculties of ensuring that
the loads are applied is such a manner that the test specimen has uniform conditions throughout its gauge volume have meant that there is
only a limited range of experimental conrmation of the theoretical
development that now follows. However, the application of the
theory to the analysis of real structures and the measurement of strains
"v 

Poisson, Simeon-Denis (Pithiviers, Loiret, 1781 Paris, 1840), French mathematician, physicist and astronomer.

62

Energy methods in structural mechanics


Py
Pz

Px
Ay

y
x
Az

Ax
Px

Pz

Py

Fig. 3.14

on the structures have provided a high degree of condence that in fact


the generalisation of uniaxial tests to triaxial loading is valid.
Figure 3.14 shows a test specimen subjected to three loads, Px , Py
and Pz applied in the x, y and z directions such that the conditions
are uniform throughout the gauge volume.
A uniform change of temperature, T, is also applied concurrently
with the change of loading. The strain, "x , in the x direction is
"x

Py Pz
Px

T
EAx EAy EAz

3:13

Similarly the strains in the y and z directions are4

"y

Py
Px Pz

T
EAy EAx EAz

3:14

"z

Pz
Px Py

T
EAz EAx EAy

3:15

In most books equations (3.13)(3.15) are stated in terms of the nominal stresses
xx Px =Ax , yy Py =Ay and zz Pz =Az , i.e.
yy zz


T
"x xx
E
E
E
yy xx zz
"y

T
E
E
E
yy


"z zz xx
T
E
E
E

Deformation of bodies and material properties

These relationships can be written in matrix format as


3
2
Px
3
2 3 2
36 EAx 7 2
7
T
"x
1   6
6 P 7
7
6 7 6
76 y 7 6
7 4 T 5
4 "y 5 4  1  56
6 EAy 7
7
T
"z
  1 6
4 Pz 5

63

3:16

EAz

In arriving at the relationship above, it has been assumed that the


material modulus is constant in all directions: this denes an isotropic
material and is a reasonable description for a vast class of materials
used in structural engineering. There are other engineering materials,
such as bre composites, for which the material modulus varies from
one direction in the body to another, this is classed as anisotropic
material. However, this introductory book is concerned only with
the development of analysis methods relevant to linearly elastic isotropic materials.

3.4

Straindisplacement relationships

During the change of shape of a body as a result of imposed loading,


strains are incurred. These strains vary throughout the volume of the
body and can be related to the displacements of the points of the
body. The relationship is illustrated using the planar body shown in
Fig. 3.15.
Figure 3.15(a) shows the original shape of the planar body. An element is identied on the body of length dl between two points A and B.
For some reason, which for the present purpose does not need to be
specied, the body changes shape as shown in Fig. 3.15(b). The
points A and B move to A0 and B 0 , respectively, and the length of

A'

(a)

Fig. 3.15

B'

l'

(b)

64

Energy methods in structural mechanics


y
B'
ux

A'

uy + duy
x

uy

ux + dux
B
A
x

Fig. 3.16

the element changes to l 0 . The change of length, and therefore the


strain, incurred by the element can be related to the local displacements.
A rectangular coordinate system (x; y) is identied with the element,
as shown in Fig. 3.16, to simplify the presentation.
The point A initially has coordinates x; 0 and moves to
A0 x ux ; uy ; the point B moves from x dx; 0 to
B 0 x dx ux dux ; uy duy .
The original length of the element is
l dx

3:17

and the deformed length is


l 0 dx dux 2 du2y 1=2


 2  2
quy
qux
qux 1=2
1

2
dx
qx
qx
qx
Thus the strain induced in the element is

 2  2

quy
l0 l
qux
qux 1=2
1

2
1
EAB
l
qx
qx
qx

3:18

3:19

Notice that in equation (3.19) the strain which occurred to the


element AB laying in the x direction was indicated as EAB and not as
"x as was done for equation (3.9). This is due to two facts that will
be fully claried in the following presentation: rstly, the reference
coordinate system x, y, z, for the specimen in Fig. 3.13 was oriented
along the principal directions of deformation, which in that particular
case were the same throughout the whole body; secondly, the notation
"ij refers to the components of strain in the small displacement theory

Deformation of bodies and material properties

65

of elasticity, where, as is true for a very large number of engineering


applications, substantial simplications can be assumed on account
of the fact that deformations are extremely small. Notice also that in
equation (3.19) the relationships are concerned only with relative
local displacements in the body. The global, i.e. rigid body, displacements do not incur strains.
The generalised straindisplacement relationships can be derived by
considering a general displacement of an element in a three-dimensional body subjected to a general form of loading. Consider a straight
segment between the points A and B of length ds at some position in
the undeformed three-dimensional body. The point A initially has
coordinates (x; y; z) and the point B initially has coordinates
x dx; y dy; z dz. When the body is subjected to a deformation
the position, orientation and length of the element changes and its
length becomes ds 0 such that we can write
EAB

jA0 B 0 j jABj ds0 ds

jABj
ds

3:20

After deformation has occurred, the new coordinates of the generic


point of original coordinates x, y and z, are, evidently,
x0 x ux x; y; z
y0 y uy x; y; z

3:21

z0 z uz x; y; z
where, once again, ux , uy and uz are the components of the displacement of the point along the Cartesian axes x, y and z. Without any
loss of generality, we shall consider the displacement a continuous vectorial function, together with the partial derivatives of its components.
Applying equation (3.21) to the points A and B and expanding the
right-hand side by Taylor's series and retaining only rst-order innitesimal terms, we have


qux
qu
qu
0
dx 1
dx x dy x dz
qx
qy
qz


quy
quy
quy
3:22
dy0
dx 1
dy
dz
qx
qy
qz


qu
qu
qu
dz0 z dx z dy 1 z dz
qx
qy
qz

66

Energy methods in structural mechanics

The square of the length of the element AB results


jABj2 ds2 dx2 dy2 dz2

3:23

and that of the deformed element A0 B 0 is


jA0 B 0 j2 ds02 dx02 dy02 dz02

3:24

so that, from equation (3.22) we have


jA0 B 0 j2 jABj2 ds02 ds2
2exx dx2 eyy dy2 ezz dz2
2exy dx dy 2exz dx dz 2eyz dy dz

3:25

where


  2  2 
quy
qux 2
quz

exx
qx
qx
qx
   2  2 
quy 1 qux 2
quy
quz

eyy

qy 2
qy
qy
qy
 2  2  2 
quy
qu
1 qux
quz
ezz z

qz 2
qz
qz
qz


1 quy qux qux qux quy quy quz quz

exy eyx
2 qx
qy
qx qy
qx qy
qx qy


1 quz qux qux qux quy quy quz quz

exz ezx
2 qx
qz
qx qz
qx qz
qx qz


1 quz quy qux qux quy quy quz quz

eyz ezy
2 qy
qz
qy qz
qy qz
qy qz
qu
1
x
qx 2

3:26
3:27
3:28
3:29
3:30
3:31

The terms eij , dened by equations (3.26)(3.31) with i; j 2 x; y; z, are


components of strain or Green's strains (after Green5 ) and equations
(3.26)(3.31) themselves represent generalised straindisplacement
relationships.

Green, George (Sneinton, Nottingham, 1793 ibidem, 1841), English physicist and
mathematician.

Deformation of bodies and material properties

67

In fact, by virtue of equations (3.20) and (3.25) we can relate the


strain EAB occurred to the element AB to the components of strain
eij as follows6
EAB 1 12 EAB ds2 exx dx2 eyy dy2 ezz dz2 2exy dx dy
2exz dx dz 2eyz dy dz

3:32

By dividing any term by ds2 we can also write


EAB 1 12 EAB exx 2x eyy 2y ezz 2z
2exy x y 2exz x z 2eyz y z

3:33

where x , y and z are the direction cosines of the vector AB.


In case the linear element before deformation is parallel to the x axis,
we have x 1 and y x 0. Therefore from equation (3.33) we
get
p
EAB 1 2exx 1
3:34
which, by virtue of equation (3.26), coincides with equation (3.19).
The straindisplacement relationships (3.26)(3.31) are absolutely
general. However, they form non-linear relationships of the partial
derivatives of the components of displacement. This is a major complication and even if progress in various engineering elds has necessitated
and produced detailed study of non-linear problems, a vast class of
engineering problems have been treated successfully by means of the
linear theory of elasticity. This is accomplished by linearising the
above mentioned relationships, as will be seen in the next section.

3.5

Simplications possible when deformations are


very small

The vast majority of materials used in structural engineering are


actually subjected only to extremely small deformations. Therefore it

In fact, it is
jA0 B0 j2 jABj2 jA0 B0 j jABj jABj2 jABj2
EAB jABj jABj2 jABj2
EAB ds ds2 ds2
2
2EAB EAB
ds2

EAB 2 EAB ds2

68

Energy methods in structural mechanics

may be assumed that components of displacement and their derivatives


are insignicant compared with unity and that any product involving
these terms can be ignored. This means that we can rewrite the
straindisplacement relationships (3.26)(3.31) in the following form
qux
qx
quy

qy

"xx

3:35

"yy

3:36

"zz

quz
qz



1 quy qux

2 qx
qy


1 quz qux

"xz "zx
qz
2 qx


1 quz quy
"yz "zy

2 qy
qz
"xy "yx

3:37
3:38
3:39
3:40

where the terms "ij , dened by equations (3.35)(3.40) with i, j 2 x; y; z,


are components of strain for small deformation and are the basis of the
linear theory of elasticity. It is worth noting that in compact form
equations (3.35)(3.40) can be often written in a summarised form as


1 qui quj
"ij

3:41
2 qxj qxi
with i, j 2 x; y; z and the convention that xx  x, xy  y and xz  z.
The components "xx , "yy and "zz have a simple geometrical meaning.
In fact, let us consider once more a linear element of the body which
before deformation is parallel to the x axis and has length dx. After
deformation, the projection of this linear element on the same axis is
given by the rst of equations (3.22), i.e.


qux
0
dx 1
dx 1 "xx dx
3:42
qx
Therefore, we have
dx0 dx
3:43
dx
that is "xx represents the relative increase in the projection on the x axis
of a linear element which before deformation was parallel to this
axis.
"xx

Deformation of bodies and material properties

69

Pyx

dy

Pxy

Pxy

Pyx
dx
x

Fig. 3.17

Moreover, the components "xy , "xz and "yz can be simply related to
an additional form of strain called shear strain. This form can be
demonstrated, for example, by reference to the deformation of a
small element of material in which the forces are parallel to the faces
of the element, see Fig. 3.17.
A practical example of this form of strain occurs in the deformations
of a thin-walled cylinder subjected purely to torsional loading, shown
in Fig. 3.18.
The perimeters AB and CD move angularly relative to each other
and the material between closely spaced perimeters will experience the
`lozenging' shown in Fig. 3.18. Considering the general deformation at
a point on a body, restricted here for clarity to a plane section, the
point moves from O to O0 . The bres dx and dy experience extension,
but, in general, they also experience relative rotation, as shown in
Fig. 3.19.

A
T

Fig. 3.18

70

Energy methods in structural mechanics


y

) dy

ux
y

dy

(1 +

uy
y

2
1

uy
x

dx

o'

v
u

(1 +

ux
x

) dx

Fig. 3.19

The rotations from the horizontal and vertical are, respectively,7




quy =qx
quy
qu
qu

1 x x
1 tan 1
qx
qx
qx
1 qux =qx
3:44


quy
qux =qy
qux
qux

2 tan 2
1 quy =qy
qy
qy
qy
and the relative rotation 1 2 results in
quy qux

2"xy
1 2
qx
qy

3:45

Sometimes the relative rotation 1 2 is dened as engineering


shear strain xy , that is
quy qux

3:46
xy
qx
qy
As the treatment of structural analysis in this book is almost entirely
concerned with problems which do not justify the more complicated
non-linear theory (the only partial exception being constituted by the
presentation of buckling phenomena in chapter 12), in the following
we will always make reference to the simplications introduced in
this section.

3.6

Deformation in the neighbourhood of an


arbitrary point

Always with reference to a deformed body, let us consider the displacement of a point B in the neighbourhood of a generic point A.
7

We recall that for small values of x it is 1=1 x 1 x.

Deformation of bodies and material properties

71

We again make reference to a generic Cartesian rectangular coordinate


system x; y; z. By a simple Taylor's expansion with origin at location A,
if only rst order innitesimal terms are retained, we have


qux
qux
ux B ux A
xB xA
qx A
qy A

qux
 yB yA qz zB zA
A



quy
quy
uy B uy A
xB xA
qx A
qy A

3:47
quy
 yB yA qz zB zA
A


quz
quz
xB xA
uz B uz A
qx A
qy A

quz
 yB yA qz zB zA
A
or also, for brevity,




qux BA qux BA qux BA
x

z
qx A
qy A
qz A



quy BA quy BA quy BA
uy B uy A
x

z
qx A
qy A
qz A



qu
qu
qu
uz B uz A z xBA z yBA z zBA
qx A
qy A
qz A

ux B ux A

3:48

The above expressions can also be written as


ux B ux A "xx xBA "xy !z yBA "xz !y zBA
uy B uy A "yx !z xBA "yy yBA "yz !x zBA 3:49
uz B uz A "zx !y xBA "zy !x yBA "zz zBA
where



1 quz quy

!x
2 qy
qz


1 qux quz
!y

2 qz
qx


1 quy qux
!z

2 qx
qy

3:50
3:51
3:52

72

Energy methods in structural mechanics

and all the partial derivatives in equations (3.35)(3.40) and (3.50)


(3.52) are evaluated at A. This means that all coecients occurring
in equation (3.49), i.e. ui A, "ij and !i , have to be considered constants
and equal to their value at A.
It follows that an innitesimal small region surrounding an arbitrary
point A on account of deformation is in general subject to the sum of
(a)
(b)

(c)

a simple translation given by the displacement vector uA;


a rotation with components !x , !y and !z . In fact in the theory
of small displacements, as we have already seen in chapter 2, it
is easy to realise that the product of a skew-symmetric matrix
! by the position vector, i.e.
3
32
2 r3 2
0
!z !y
ux
xBA
6
7
4 ury 5 6
0
!x 7
3:53
4 !z
54 yBA 5
r
BA
!y !x
0
uz
z
always represents a rigid rotation displacement eld about the
origin A;
changes in dimensions and shape given by the displacement
eld
32 BA 3
2 3 2
"xx "xy "xz
ux
x
76 BA 7
6 7 6"
"
"
3:54
5
4 uy 5 4 yx
yy
yz 54 y

BA
"
"
"
uz
z
zx
zy
zz

A'
z

u (A)
A

Fig. 3.20

Deformation of bodies and material properties

73

Thus, apart from a translational and rotational movement, the deformation of an innitely small region surrounding A is described by
equation (3.54). This is a linear transformation, where straight lines
remain straight, planes remain plane (and both remain parallel if
they were parallel before occurrence of deformation), and in general
any second-order surfaces transform into second order surfaces, that
is, for example, a sphere transforms into an ellipsoid, as is shown in
Fig. 3.20.

3.7

Transformation of the components of strain with


the change of reference frame

The transformation of the expression of the components of strain "ij


with the change of reference frame, that is, for example, in the passage
from a Cartesian system of orthogonal coordinates x1 , y1 , z1 to
another Cartesian system x2 , y2 , z2 with a dierent origin and with
rotated axes, can be pursued very easily with a minimum of matrix
algebra.
In fact, look at the situation in Fig. 3.21 and consider a vector v of
components vxz , vyz , vzz with reference to the second Cartesian frame
x2 , y2 , z2 .
In the rst reference frame x1 , y1 , z1 the same vector has components
vx1 , vy1 , vz1 which are related to the components vxz , vyz , vzz by the
formula
2
3 2
32
3
nx2 x1 nx2 y1 nx2 z1
vx2
vx1
4 vy 5 4 ny x ny y ny z 54 vy 5
3:55
1
2 1
2 1
2 1
2
nz2 x1 nz2 y1 nz2 z1
vz1
vz2
z2
v
z1
x2
y2
y1

x1

Fig. 3.21

74

Energy methods in structural mechanics

where nx2 x1 ; . . . ; nz2 z1 are the direction cosines of the axes of the second
system with respect to the rst one.
Hence, in the passage from x2 ; y2 ; z2 to x1 ; y1 ; z1 the pure strain
displacement eld given by equation (3.54) becomes
32  3
2  3 2
nx2 x1 nx2 y1 nx2 z1
ux2
ux1
7
6 u 7 6 n
6
 7
3:56
4 y1 5 4 y2 x1 ny2 y1 ny2 z1 54 uy2 5
uz1

nz2 x1

nz2 y1

nz2 z1

uz2

By remembering from geometry that the matrix n of the director


cosines is orthogonal, i.e. its inverse coincides with its transpose that
is n1 nT , the inverse relationship of the above equation (3.56),
results in
3 2
3
2  3 2
nx2 x1 nx2 y1 nx2 z1 T ux1
ux2
7 6  7
6 u 7 6 n
3:57
4 y2 5 4 y2 x1 ny2 y1 ny2 z1 5 4 uy1 5
uz2

nz2 x1

nz2 y1

that is, by virtue of equation


2  3 2
nx2 x1 nx2 y1
ux2
6  7 6
6 uy 7 6 ny x ny y
2 1
4 25 4 2 1

nz2 x1 nz2 y1
uz2
2 BA 3
x1
6 BA 7
7
6
4 y1 5

nz2 z1

uz1

(3.54)
3 2
nx2 z1 T "x1 x1
7 6
6
ny2 z1 7
5 4 "y1 x1
nz2 z1
"z1 x1

"x1 y1

"x1 z1

"y1 y1

7
"y1 z1 7
5

"z1 y1

"z1 z1

3:58

zBA
1
Thus, the following relationship holds true
2  3 2
3 2
nx2 x1 nx2 y1 nx2 z1 T "x1 x1 "x1 y1
ux2
6  7 6
7 6
6 uy 7 6 ny x ny y ny z 7 6 "y x "y y
2 1
2 1 5 4
1 1
1 1
4 25 4 2 1

nz2 x1 nz2 y1 nz2 z1
"z1 x1 "z1 y1
uz2
2
32 BA 3
nx2 x1 nx2 y1 nx2 z1
x2
6
76 BA 7
76
7
6
4 ny2 x1 ny2 y1 ny2 z1 54 y2 5
nz2 x1 nz2 y1 nz2 z1
zBA
2

"x1 z1

7
"y1 z1 7
5
"z1 z1

3:59

and by comparison with equation (3.54) the transformation law of

Deformation of bodies and material properties

75

the components of strain with the change of coordinate system


results in
2
3 2
3
nx2 x1 nx2 y1 nx2 z1 T
"x2 x2 "x2 y2 "x2 z2
6
7 6
7
6 "y x "y y "y z 7 6 ny x ny y ny z 7
2
2
2
2
2
2
2
1
2
1
2
1
4
5 4
5
"z2 x2 "z2 y2 "z2 z2
nz2 x1 nz2 y1 nz2 z1
2
3
"x1 x1 "x1 y1 "x1 z1
6
7
7
6
4 "y1 x1 "y1 y1 "y1 z1 5
"z1 x1 "z1 y1 "z1 z1
2
3
nx2 x1 nx2 y1 nx2 z1
6
7
7
3:60
6
4 ny2 x1 ny2 y1 ny2 z1 5
nz2 x1 nz2 y1 nz2 z1
Finally, as they will be useful in the next section, we can develop
these matrix products in extenso and write
"x2 x2 "x1 x1 n2x2 x1 "y1 y1 n2x2 y1 "z1 z1 n2x2 z1 2 "x1 y1 nx2 x1 nx2 y1

"x2 y2

2 "x1 z1 nx2 x1 nx2 z1 2 "y1 z1 nx2 y1 nx2 z1


...
...
"y2 x2 2 "x1 x1 nx2 x1 ny2 x1 2 "y1 y1 nx2 y1 ny2 y1

3:61

2 "z1 z1 nx2 z1 ny2 z1 2 "x1 y1 nx2 x1 ny2 y1 ny2 x1 nx2 y1


2 "x1 z1 nx2 x1 ny2 z1 ny2 x1 nx2 z1
2 "y1 z1 nx2 y1 ny2 z1 ny2 y1 nx2 z1
...
...

3.8

Principal directions of deformation. Maximum and


minimum extensions

In this section we are concerned with nding one or more directions in


which the extension given by equation (3.43), i.e.
"xx

dx0 dx
dx

assumes an extremal value, that is, the result is a maximum or a


minimum.

76

Energy methods in structural mechanics

This study can be performed very easily with the aid of the law of
transformation (3.61). In fact the problem amounts to nding the
values of n x1 , n y1 and n z1 for which the rst of equations (3.61)
becomes a maximum or a minimum. As by denition of direction
cosines the following relationship must hold true
n2x1 n2y1 n2z1 1

3:62

the problem is one of extremum with constraints and, as is known from


calculus, it can be dealt with by means of Lagrange multipliers (named
after the eminent mathematician Joseph-Louis Lagrange, see chapter
2). In other words, we are led to searching the stationary value of
the following function
Ln x1 ; n y1 ; n z1 ; " " "n2 x1 n2 y1 n2 z1 1

3:63

where " is a constant, i.e. the Lagrange multiplier. Therefore, if


we equate to zero the partial derivatives of expression (3.63) with
respect to n x1 , n y1 and n z1 , we obtain the following system of three
homogeneous algebraic equations
8
>
< "x1 x1 "n x1 "x1 y1 n y1 "x1 z1 n z1 0
"x1 y1 n x1 "y1 y1 "n y1 "y1 z1 n z1 0
3:64
>
:
"x1 z1 n x1 "y1 z1 n y1 "z1 z1 "n z1 0
The zero solution of this system has no interest on account of the condition (3.62). Consequently, in order to nd the non-zero solutions of
equation (3.64) we must equate the determinant of its coecients to
zero, that is


"x x "
"x1 y1
"x1 z1
1 1


"y1 y1 "
"y1 z1 0
3:65
det "x1 y1


"x z
"y z
"z z "
1 1

1 1

1 1

or, in expanded form


"3 E1 "2 E2 " E3 0

3:66

where
E1 "x1 x1 "y1 y1 "z1 z1

3:67

E2 "x1 x1 "y1 y1 "x1 x1 "z1 z1 "y1 y1 "z1 z1 "2x1 y1 "2x1 z1 "2y1 z1 3:68
E3 "x1 x1 "y1 y1 "z1 z1
"x1 x1 "2y1 z1 "y1 y1 "2x1 z1 "z1 z1 "2x1 y1 "x1 y1 "x1 z1 "y1 z1

3:69

Deformation of bodies and material properties

77

Because the principal extensions are independent of the choice of direction of the coordinate axes, also the coecients E1 , E2 and E3 result
independent from this choice of the coordinate axes and are named
strain invariants.
From a mathematical standpoint, solving equation (3.65) means
nding the eigenvalues of the problem
" "I 0

3:70

Equation (3.66) is a cubic in the Lagrange multiplier " and, as it is


obtained from a symmetric matrix, it can be demonstrated that it
admits always three real values, which are the eigenvalues of the
problem, equation (3.70). Corresponding to these values, we nd
three sets of direction cosines (that is, for example, n x1 , n y1 , n z1 ;
n x1 , n y1 , n z1 and n x1 , n y1 , n z1 ) which represent the eigenvectors of
equation (3.70). By virtue of a well known property of eigenvectors,
the three directions associated to these sets of direction cosines are
mutually perpendicular.
We are therefore led to the conclusion that for every point of the
body it is always possible to nd three perpendicular directions in
which the components of strain "ii assume stationary value, that is
are a maximum or a minimum relative.
The values assumed by the Lagrangian multiplier, i.e. the roots "1 ,
"2 , "3 of equation (3.66), represent exactly these components and are
called principal components of strain. In order to illustrate this point,
let us consider again the rst of equations (3.61). By grouping its
right-hand terms, it can be written as
" "x1 x1 n x1 "x1 y1 n y1 "x1 z1 n z1 nx1
"x1 y1 n x1 "y1 y1 n y1 "y1 z1 n z1 n y1
"x1 z1 n x1 "y1 z1 n y1 "z1 z1 n z1 n z1

3:71

Given that "1 and the set of direction cosines nx1 , ny1 , nz1 satisfy equations (3.64), then equation (3.71) becomes
" "1 n2 x1 "1 n2 y1 "1 n2 z1 "1

3:72

and, by repeating the same procedure, we nd also that " "2 and
" "3 .
An important property of the principal directions of deformation is
that along these directions the components " , " , " of strain, i.e.
the shear strain components, are zero. To show this property the

78

Energy methods in structural mechanics

fourth of equations (3.61) can be written as


" 2"x1 x1 n x1 "x1 y1 n y1 "x1 z1 n z1 n x1
2"x1 y1 n x1 "y1 y1 n y1 "y1 z1 n z1 n y1
2"x1 z1 n x1 "y1 z1 n y1 "z1 z1 n z1 n z1

3:73

Again, as "1 and the set of direction cosines n x1 , n y1 , n z1 satisfy equations (3.64), equation (3.73) becomes
" 2"1 n x1 n x1 n y1 n y1 n z1 n z1 0

3:74

since the axes ,  and  are orthogonal.


By applying the same line of reasoning to the fth and the sixth of
equation (3.61) we conclude that
" " 0

3.9

3:75

Plane strain

There are structural forms that are very long, in the z direction,
compared to their other dimensions. Two examples are shown in
Fig. 3.22.
Figure 3.22(a) represents a part of a very long dam wall that is supported at its base and at its ends. The ends are prevented from movement in the z direction and the loading is uniform along the length of
the dam face. We may deduce that, on account of the length, there is
symmetry of deformation at any section along the dam and that it
may reasonably be assumed that there is no displacement along the z
axis and therefore no variation of deformations or strains along the

dz
z

(a)

Fig. 3.22

(b)

Deformation of bodies and material properties

79

dam wall. In fact, the deformation of the wall can be assumed to be the
same at any cross section and can be dened completely by the nonzero components of displacement ux and uy , which are assumed to
be functions of x and y only. Thus, the response of the wall to the
applied loads may be analysed using a representative length dz of
the wall as shown in the gure. Similar conditions may be assumed
for a very long tunnel that has a uniform pressure along its length
from the surrounding soil, see Fig. 3.22(b).
These cases are normally referred to as plane strain and the assumption of zero variation of strain along the length can signicantly simplify the analysis. However, it must be remembered that the results
pertain only to deformations far from the ends and that near the
ends the deformations will actually be three-dimensional.
Formally, the above assumptions can be written as
ux ux x; y

uy uy x; y

uz 0

3:76

and the expressions (3.35)(3.40) of the strain components can be simplied as follows
"xx

qux
qx

3:77

"yy

quy
qy

3:78

quz
0
qz


1 qux quy

"xy
2 qy
qx


1 quy quz

0
"yz
2 qz
qy


1 qux quz
"xz

0
2 qz
qx
"zz

3.10

Straindisplacement compatibility conditions

3:79
3:80
3:81
3:82

The straindisplacement relationships in equations (3.26)(3.31)


describe a complete derivation of the strains from the displacement
eld throughout a body. However, one may wonder if an arbitrarily
chosen set of six continuous functions eij fij x; y; z dened in the
domain of the body always represent an actual deformation of the

80

Energy methods in structural mechanics

continuous medium. Evidently this cannot be the case as there is a


disparity between the number of strain components (six) and those
of the components of the displacement eld (three) and therefore not
all the strain components can be independent. Hence, it can be
shown that the functions eij fij x; y; z will dene the deformation
of a body, thus satisfying all the continuity conditions, only if they
full a set of six dierential equations, called strain compatibility conditions. These conditions may be derived in a straightforward manner
when deformations are very small, i.e. with reference to the simplied
straindisplacement relationships (3.35)(3.40). For the sake of simplicity, the derivation of the compatibility conditions is exemplied
with regard to a case of plain strain.
So, with reference to equations (3.77), (3.78) and (3.80)
"xx

qu
x
qx

"yy

quy

qy

"xy



1 quy qux

2 qx
qy

and dierentiating "xx twice with respect to y we obtain


q2 "xx
q3 ux

qy2
qx qy2

3:83

In similar manner we have


q2 "yy
q3 uy

qx2
qy qx2



q2 "xy 1 q3 ux
q3 uy

qx qy 2 qx qy2 qx2 qy

3:84

Using the derivatives of the displacements shown above, it is evident


that the condition of compatibility of the strains to ensure continuity
of the displacements is
2
q2 "xy
q2 "xx q "yy
0

2
qx qy
qy2
qx2

3:85

This condition is obviously a necessary one for the continuity of


deformation, but it is possible to show that it also is sucient. However, the discussion of the conditions of solvability of the set of equations of strain compatibility conditions is somewhat more complicated
and goes beyond the scope of this book.
The compatibility relationship (equation (3.85)) is for a case of plane
strain. Nevertheless, by applying a similar procedure we can easily get
the complete set of conditions relating to a general three-dimensional

Deformation of bodies and material properties


z

dz

Px
dx

81

Px

dy
y

Fig. 3.23

body, i.e.
2
q2 "xy
q2 "xx q "yy
0

2
qx qy
qy2
qx2

q2 "xx q2 "zz
q2 "xz
0

2
qx qz
qz2
qx2
q2 "yy q2 "zz
q2 "yz
0

2
qy qz
qz2
qy2

3:86

q2 "yz q2 "xx q2 "xz q2 "xy

qy qz qx qy qx qz
qx2
2
2
2
q2 "xz q "yy q "xy q "yz

qx qz qy qz qx qy
qy2

q2 "xy q2 "zz q2 "yz q2 "zx

qx qy qx qz qy qz
qz2
These conditions are generally known as Beltrami's equations.8

3.11

Strain energy

This text presents the development of methods for the analysis of structures using the concept of strain energy and work as a basis. The
formulations of strain energy for a linear elastic body are presented
in this section.
Figure 3.23 shows an innitesimal element of material as some position in a continuous body. The element has dimensions dx, dy and dz.
A force Px is applied to the element in the x direction. The value of the
force starts from zero and gradually reaches its nal value Px . The
8

Beltrami, Eugenio (Cremona, 1835 Rome, 1900), Italian mathematician.

82

Energy methods in structural mechanics

increment change of strain energy, U, stored by the element is equal


to the work done W by the force during the loading process on account
of the incremental change of length, from zero to dl "xx dx, of the
element in the x direction. Thus
"xx

U W

Px d"xx dx

3:87

As for a linearly elastic material the relationship between the force and
the strain is, according to the test results reviewed in section 3.3,
"xx

Px
E dy dz

3:88

The incremental change in strain energy for the elementary volume is


therefore
"xx

U

Px d"xx dx
0
"xx

E"xx d"xx dx dy dz

E 2
" dV
2 xx

3:89

This is a noteworthy result. In fact, whenever we deal with a linear


loaddeection relationship, as is the case in equation (3.88), which
is described by a relationship of the type
l "xx l Px =k

3:90

the work performed by the load Px in the loading process from 0 to its
nal value Px is
l

l

Px dl
0

kl dl
0

1
1
kl 2 Px l
2
2

3:91

i.e. it amounts to half the product of the nal value of the force Px by the
nal value of the displacement l.
This fact is of fundamental importance in the linear theory of elasticity
and constitutes the statement of Clapeyron's theorem, named after
Clapeyron9 who investigated the phenomenon in a wider context.
Clapeyron, Beno t-Paul-Emile (Paris, 1799 ibidem, 1864), French engineer and
physicist.
9

Deformation of bodies and material properties

83

Pz
z

dz

Px
Py

dx
Pz

Px

dy
y

Fig. 3.24

Given that the element under consideration is an innitesimal one,


equation (3.89) can be employed to dene the strain energy per unit
volume at a point, which is called strain energy density. This is
simply done by letting ' U=dV and assuming the constant which
denes the reference value of the strain energy to be zero in the unconstrained conguration. It follows immediately that the value of strain
energy for a body with a volume V, uniform strain "xx and constant
value of modulus E, is

E 2
U ' dV
"xx dV
3:92
2
V

We next consider the situation where also forces in the y and z directions are applied. The examination of the test results in the previous
sections showed that forces applied normal to each other inuence
the strains through a constant, called Poisson's ratio, . Figure 3.24
shows an elementary volume of material subjected to forces in the x,
y and z directions.
The strains related to these forces and directions are
Py
Px
Pz


E dy dz
E dx dz
E dx dy
Py
Px
Pz


"yy
E dx dz
E dy dz
E dx dy
Py
Pz
Px


"zz
E dx dy
E dy dz
E dx dz
"xx

3:93
3:94
3:95

By superposing the eects of these three forces and integrating them


over the volume of a body of volume V, if we assume an homogeneous
and isotropic condition for the material constants, E and , that is,

84

Energy methods in structural mechanics

they are constant throughout the body and are constant in every direction at every point in the body, we obtain

E
U
1 "2xx "2yy "2zz
21 1 2
V

2 "xx "yy "xx "zz "yy "zz dV

3:96

It is interesting to note that if we consider the situation where the


strains in the y and z directions are zero, that is "yy "zz 0, we conclude that this is possible provided that


Py

Px

1  dy dz
dx dz
3:97


Pz

Px

dx dy
1  dy dz
and the value of the strain energy is

E1 
"2 dV
U
21 1 2 xx

3:98

The development can be extended to include the situation where there


are direct and shear strains in the elementary volume, still assuming an
homogeneous and isotropic condition for the material constants E and .
Before doing so, consider the shear loading shown in Fig. 3.25,
similar to that already considered in Fig. 3.17.
In such a case it is found from testing that
xy

Pxy
Pyx

G dy dz G dx dz

3:99

y
Pyx

Pxy

Pxy

dy

Pyx

dx

Fig. 3.25

Deformation of bodies and material properties


y

85

Pyx

y
Pyx

C
PAC

Pxy

Pxy

Pxy

A
Pyx

PAC

Pyx

x
(a)

Pxy

x
(b)

Fig. 3.26

where the constant G is the shear modulus of elasticity of the material.


Equation (3.99) is normally referred to as Hooke's law for shearing.
This law applies also when more than one loading (shear or normal)
is acting, i.e. shear forces do not inuence each other, nor inuence
or are inuenced by normal forces. The constant G can be easily related
to the constants E and .
In fact, let us consider that the innitesimal element has the shape of
a square, that is dx dy. As this innitesimal element is in equilibrium, it is easy to conclude that we must have jPxy j jPyx j. Moreover, if we ideally make a cut along the diagonal line BD, as shown
in Figure 3.26(a), and write the equilibrium equations for the triangular sub-elements ABD and BCD, we are led to the conclusion that on
the surfaces of the cut
pcan
only be present normal tensile forces PAC of
magnitude equal to 2Pxy , see Fig. 3.26(b).
By repeating the same reasoning for an ideal cut along the diagonal
AC, we nd that along the direction
pBD
and normally to the cut a compressive force of modulus PBD 2Pxy is acting, see Fig. 3.27.
Thus, applying Hooke's law for axial loading to the direction AC we
have "AC , the component of strain along the direction AC,
Pxy
P
P
 pBD

1 
"AC pAC
E 2 dy dz
E 2 dx dz E dy dz

3:100

The law of transformation of the components of strain with the


change of reference axes, equations (3.61), applied to the case of
pure shear strain at hand yields
xy
"AC "xy
3:101
2

86

Energy methods in structural mechanics


Pyx

Pxy

PBD
PBD

Pxy

Pyx

Fig. 3.27

and allows us to write


xy
Pxy

1 
2
E dy dz

3:102

At this point, a simple comparison between equation (3.99) and equation (3.102) provides the relationship between the constant G and the
constants E and , that is
G

E
21 

3:103

The expression of the strain energy for the situation where all the
strains, direct and shear, are present in the elementary volume is
now simple. In fact, given that there is no interaction between Hooke's
law for direct and shear strains, we can simply compute the work done
by the shear forces on the elementary volume, that is


quy
1
qux
1
1
W
Pxy
dx Pyx
dy Pxy xy dx Pyx xy dy
2
2
2
qx
qy
3:104
given that for the element to be in equilibrium we must have Pxy dx
Pyx dy and provided we made reference to Clapeyron's theorem. By
virtue of Hooke's Law, equation (3.99), we can write equation
(3.104) as
1 2
E
W G xy
"2 dx dy dz
dx dy dz
2
1  xy

3:105

Finally, by superposing the eects of the shear forces Pxy , Pyz , Pxz ,
Pzx , Pyx and Pzy and by integrating over the domain of a body with
volume V we obtain the general expression of the change in strain

Deformation of bodies and material properties

energy related to the shear strains

E
U
"2xy "2xz "2yz dV
1 

87

3:106

which, added to equation (3.96), gives



E
1 "2xx "2yy "2zz
U
21 1 2
V

2 "xx "yy "xx "zz "yy "zz



21 2 2
2
2
"xy "xz "yz dV

1 

3:107

At the end of this section it is worth noting that the strain energy is a
scalar quantity and by its own nature it is independent of the particular
reference frame. Hence, the expression of the strain energy density at a
point must give the same value with respect to any system of coordinate axes. In particular, when we make reference to a system of axes
coincident with the principal directions of strain , ,  we can write
'

E
1 "2 "2 "2
21 1 2
2" " " " " "

3:108

because the shear strain components are zero in this particular reference frame.

3.12

Yield criteria

The uniaxial tests on ductile material described in section 3.1 show


that the linear loadextension property of the material is limited
to certain values of strain. At the value of strain, "0 , which we
can call the yield strain, the forcedisplacement properties become
non-linear. The onset of the non-linear conditions have for many
years been generally regarded as a limiting condition in structural
design. However, relatively recent developments in mechanics of
materials and the alliance of the non-linear material properties
assessment with more advanced numerical techniques of analysis
have enabled the actual limiting state of failure of structures to be
investigated and, in some cases, incorporated in design analysis.
These non-linear analyses can be extremely complex and involve
local and overall buckling of the structure as material yielding
spreads. This level of analysis is outside the scope of this text, but
the following provides criteria that have been developed which can

88

Energy methods in structural mechanics

be used to identify the limit of validity of the linear elastic analysis


methods.
The primary requirement in the development of the yield criteria is
to extend the observations from uniaxial tension tests to the complex
conditions of a three-dimensional strain eld that exists in the interior
and on the surface of a structure subjected to various forms of loading.
As a matter of fact, there are no currently available tests that can
validate uniquely or directly any specic yield criterion. The usual
approach is to particularise the three-dimensional yield criterion
using the yield strain derived from a uniaxial test on the material
from which the structure is composed, then apply the yield criterion,
with some reduction for safety, to the linear elastic analysis results to
determine the allowable level of the external loads that can be sustained in practice. Experience has shown this approach to provide a
fairly reasonable basis for structural practice.
The strain energy formulation developed in the previous sections
has been related to a general state of three-dimensional strain at
some point in the interior of a structure. The strain state encompasses direct and shear strains. The relative magnitudes of these
components of strain will vary with the choice of direction of the
axes, x, y and z. As we have seen, it is possible to choose the
axes such that there are no components of shear strain in the description of the strain state. Here we identify these strains as "1 , "2
and "3 .
The strain energy formulation, in terms of the principle strains is
given by equation (3.108), i.e.
'

E
21 1 2
2

 1 "1 "2 "3 2"1 "2 "1 "3 "2 "3


where ' is the strain energy density at a point in the structure and the
strain energy U for the whole structure is obtained by integrating the
strain energy density across the volume of the structure.

U ' dV
3:109
V

Since we are usually interested in the initial yielding at some point in


the structure we will use the strain energy density as the relevant property to be considered in the development of a yield criterion.
Several yield criteria have been proposed over the years and some of
these will be reviewed here.

Deformation of bodies and material properties

89

Maximum principle strain

This criterion is attributed to Saint-Venant10 and essentially states that


yield will occur for a three-dimensional strain state at a point in a structure when the largest of the principal strains at the point equals the
yield strain measured in a uniaxial tension test on a corresponding
material.
Thus yielding will occur when
"0 "1  "2  "3

3:110

The St Venant criterion has been found in practice to be particularly


relevant to brittle materials in which there is very little extension in
the material after the onset of yielding.

Maximum strain energy criterion

This criterion was proposed by Beltrami and postulates that yielding


will be initiated in the three-dimensional state of strain at a point in
a body when the strain energy density at that point equals the strain
energy density at the yield strain in a uniaxial tensile test on a corresponding material.
The state of principal strains at the inset of yielding of a uniaxial test
specimen is
"1 "0 ;

"2 "3 "0

3:111

and the corresponding strain energy density '0 is, according to equation (3.110),
E 2
"
3:112
2 0
Thus equating the strain energy density at some point in the structure
with the strain energy density at yield for a uniaxial tensile test specimen of corresponding material, we have
'0

"eff "0
where "eff is the eective strain dened by

1
1 "21 "22 "23
"eff 
1 1 2
1=2
2"1 "2 "1 "3 "2 "3

10

3:113

3:114

Saint-Venant, Adhemar-Jean-Claude Barre de (Villiers, Paris, 1797 Saint Ouan,


Loir-et-Cher, 1886), French mathematician and engineer.

90

Energy methods in structural mechanics

and constitutes a measure of the actual strain state with respect to


yielding according to Beltrami.

Von Mises criterion

This yield criterion, like that proposed by Beltrami, is based on the


strain energy density at some point in the structure. However, in the
von Mises11 criterion the particular component of the strain energy
is that which is associated with distortion of the material.
The strain energy density formulation in equation (3.108) can be
reformulated as

E
1
'
" "2 "3 2
2 31 2 1

1
2
2
2
" "2 "1 "3 "2 "3

3:115
31  1
By remembering equation (3.12) it is easy to infer that the rst term in
equation (3.115) can be associated with volumetric change owing to the
strain state by means of the material constant K, the bulk modulus.
K

E
61 2

3:116

The second term in equation (3.115) can be therefore considered to be


associated with the strain energy of distortion of the material due to
the strain state. The von Mises criterion postulates that yielding will
occur when the distortional strain energy density in a three-dimensional state of strain equals the distortional strain energy density at
the yield strain in a uniaxial tensile test specimen of the corresponding
material.
Thus the strain energy density 'd associated with distortion is
'd

E
1
" "2 2 "1 "3 2 "2 "3 2
2 31  1

3:117

The corresponding distortional strain energy density 'od for a tensile


specimen is
E 2 1  2
"0
2
3
Hence, the von Mises criterion is

11

'od

3:118

"eff "0

3:119

Mises, Richard von (Lemberg, Austria, 1883 Boston, 1953), German mathematician and engineer.

Deformation of bodies and material properties

91

where
1
"eff p
"1 "2 2 "1 "3 2 "2 "3 2 1=2 3:120
21 
Experience has shown that the von Mises criterion provides results
that accord reasonably with tests on biaxial tests on ductile materials,
such as structural steels and aluminium alloys. It is accepted that
this criterion is more appropriate for ductile materials than that
proposed by Beltrami and is preferred for application in structural
practice.

Suggested exercise problems


1.

A tensile test round specimen with a 15 mm diameter gauge section


is shown in Fig. 3.28. The specimen was subjected to a axial load
P and the change of length, l for the gauge length, lg 50 mm,
was measured using a mechanical extensometer that was accurate
to 0:005 mm. The change of diameter d was also measured.
The following are the results from the test:

P (kN)
9:9 20:1 28:45 35:8 45:9 52:9 53:20 53:4 52:9 53:2
l (mm) 0:02 0:04 0:06 0:07 0:09 0:11 0:15 0:20 0:29 0:52
d (mm) 0:01 0:02 0:02 0:03 0:04 0:05 0:06 0:09 0:16 0:27

2.

Plot these results on appropriate scales and calculate the averaged


values of the material modulus, the yield strain and Poisson's
ratio.
[Ans: "0 2:2  103 ,  0:44, E 2:56  105 N/mm2 ]
A section of structure is essentially operating in plane strain. The
strains in orthogonal x and y directions are "xx 1:78  103 ,
"yy 1:28  103 and "xy 1:66  103 . The negative sign
denotes compressive strain. Calculate the principal strains in the
plane and determine, using the von Mises criterion, if the strain
state in the material exceeds the elastic limit. The material has a
yield strain, "0 2:15  103 in a tensile test.
[Ans: "1 2:0  103 , "2 2:5  103 , "eff 3  103 ]
15 mm dia.
P

lg = 50 mm

Fig. 3.28

4.

Theory of elastic deformation


of beams

Introduction

Bars and beams are ubiquitous structural elements, acting as parts of


frames and grid structures. Essentially a beam, or bar, is a structural
component whose length is signicantly greater than its breadth and
width. Conventionally a bar is an axially loaded element, while a
beam has loads applied normal to its longitudinal axis. A beam may
be supported at various positions along its length and in various manners, as is seen later in the text.
Up to this stage we have examined the behaviour of elastic materials,
the interpretation of tensile tests and the concept of strain in one, two
and three dimensions. Chapter 1 also considered the concept of equilibrium for general systems. Essentially, the objective in structural
mechanics is to determine the manner in which specic structures or
components are deformed by external actions and to enable design calculations to evaluate the maximum levels of loading and deformations
that can safely be allowed in practice. The analysis of the response of a
beam to a specic pattern of applied loads requires that
.
.
.

The beam as a whole and all of its parts must be in equilibrium


with the applied loading.
The deformations of the beam must be continuous and compatible with the restraints of the supports.
The strain energy must conform with the properties relevant to
the material (in our case the conditions of elasticity).

The exact solution of the deformations of a generic structural element is, in general, very dicult to achieve for solids with arbitrary
shapes. The approach is simplied by dening classes of elements,
for example bars, beams, plates and shells, whose shapes have generic
features which enable us to make assumptions to reduce the complexity of the analysis. Of course, the introduction of such assumptions

94

Energy methods in structural mechanics


q

Fig. 4.1

implies a certain degree of approximation, but experience and testing


has shown that the assumptions result in predictions for the deformations of elastic and many plastic structural components within an
accuracy acceptable for engineering applications.
Starting here with bars and beams, which are essentially three
dimensional, with length, depth and breadth, we seek to simplify the
analysis by the assumption about the way in which these elements
deform in response to loading. We consider beams composed of
materials with linear elastic properties and loaded such that they
remain in the range of small deformations.
Simple observation shows that if one were to draw a line around the
surface of an unloaded beam or bar, say with rectangular cross section,
the line being composed of straight parts and normal to the axis of the
element, then when the beam is loaded and deforms, the line is still continuous and the originally straight lateral parts are still reasonably
straight. This is shown in Fig. 4.1.
The above observation leads to a generalisation that
a plane section imagined as cut through the beam before loading will remain plane after
loading.

This is usually called Bernoulli's model of beam bending1 and serves


the very important purpose of reducing the actual three-dimensional
deformations to an idealised one-dimensional model in which the
variation of the deformations can be described with reference to the
behaviour of the longitudinal axis of the beam. It goes without
saying that the same model applies to the simpler case of axially
loaded bars that are also introduced in this chapter. Notice that bars
loaded axially in compression can exhibit a form of instability, or
buckling, and this is considered in detail in chapter 12.
An alternative but essentially identical approach to assuming plane
sections remain plane, is to make a simplifying assumption with regard
to the strain distribution induced in the beam during loading. For a
1

Bernoulli, Jacob (Basel, 1654 ibidem, 1705), Swiss mathematician.

Theory of elastic deformation of beams

95

normally loaded element the simplest and acceptably accurate assumption in the realm of linear elasticity is that
the strain varies linearly through the beam in the direction of any normal to the
longitudinal axis.

It will be seen that this assumption is in fact equivalent to the


Bernoulli assumption of plane sections. Of course, for beams composed of materials with more complex characteristics, or elements
made from a composite of dierent materials, the assumption of
linear distribution of strain has to be carefully validated by comparisons with the deformations predicted by more exact analysis and
with the corresponding results from carefully conducted tests.
In the following two sections we consider bars and beams of uniform
cross section: this is purely for the sake of simplicity, but the same
treatment can be applied to bar and beams which vary in shape
along their longitudinal axes as shown in an example in section 4.6.

4.1

Deformation of axially loaded bars

We start the mathematical treatment of the deformation of one-dimensional elements by considering axial loading only, i.e. the deformation
of bars. Following what has been said in the introduction to this
chapter and without loss of generality, a bar is dened here to be a
straight element with a uniform cross section and to be axially
loaded. Reference is also made to a Cartesian frame with the z-axis
coincident with the axis of the bar, represented by the straight line
through the centroids of the cross sections. The x and y-axes are
arranged to coincide with the principal axes of inertia of the cross
section, as shown in Fig. 4.2.
N

z
N
y

Fig. 4.2

96

Energy methods in structural mechanics


N

dl
dl'
l
l'

dA

(a)
dN

dl
(b)

Fig. 4.3

According to Bernoulli's model, we assume that, after deformation,


any cross section remains plane and orthogonal to the curve representing the deformed
axis of the beam.

Consider the bar shown in Fig. 4.3(a), which is in equilibrium with the
two opposite axial forces N applied at the ends.
First of all it is worth noticing that we can obtain an elementary
length of the bar by means of two cuts normal to the z axis at two
very closely spaced cross sections dl apart. If we rely upon the smallness of the deformation of such a slice of the bar and consider each
of the parts of the bar as rigid bodies, we are led to the conclusion
that for the equilibrium it is necessary to have the same axial force N
applied at any section. This fact follows from equation (2.55).
We will also assume that the axis of the initially straight bar remains a
straight line. The implications of this assumption are examined in
chapter 12. Given the adopted model of deformation, we can extract
an innitesimal sectional area from the body of the bar and derive
from its behaviour a relationship which describes the behaviour of
the whole bar, see Fig. 4.3(b).
We are primarily going to deal with linearly elastic materials and
from tensile tests we know that the relationship between the length
of the innitesimal specimen before the deformation, dl, due to the
axial force dN acting on its section dA, and the length after the

Theory of elastic deformation of beams

deformation has occurred, dl 0 , is given by the formula




1 dN
dl
dl 0 1
E dA

97

4:1

where E is Young's modulus for the material, see equation (3.1).


In the bar, every elementary length between two chosen cross
sections has the same elongation which is the consequence of the
assumption that plane sections remain plane. Therefore, as

N dN
4:2
A

and

dA

4:3

we can sum up all the contributions in the cross section to obtain

1
0


dN
C
B
N
0
A
C
B

4:4
dl @1
A dl 1 EA dl
E dA
A

provided, of course, that E is constant. This constitutes the deformation formulation for an innitesimal element of the bar between two
very near cross sections under the axial force N. As the axial force is
constant for the whole length of the bar, we can integrate over the
length l and obtain the relationship between the applied axial force
and the elongation of the whole bar





N
N
l dl 0
dl 1
1
dl
EA
EA
l


1

N
l
EA

4:5

The energy stored in the bar is simply given by the work done by the
axial forces N during the process of loading, i.e.
l

UW

Nl dl
0

where Nl is the axial force during the loading process.

4:6

98

Energy methods in structural mechanics

Equation (4.5) gives us


l 0 l
EA
4:7
l
and we can write, remembering Clapeyron's theorem for linearly elastic materials, see equation (3.91),
N

UW

4.2

1 EA 0
l l2
2 l

4:8

Deformation of beams

Having examined the behaviour of bars, we now turn our attention to


the deformation of beams subjected to a variety of types of loading.
Start by considering a beam to be essentially a straight structural
element of uniform cross section capable of resisting shear forces
and bending moments about the two principal axes of inertia in the
plane of its generic cross section. The shear forces and the bending
moments are considered here to be induced only by end forces and
transverse loads, i.e. normal to the longitudinal axis of the bar. As
before, a Cartesian reference frame is adopted with the z-axis coincident with the axis of the bar, represented by the straight line through
the centroids of the cross sections along the beam. The x and y-axis
are taken initially to coincide with the principal axes of inertia of the
cross section, see Fig. 4.4.
As we are primarily interested in establishing some relationships that
can describe the bending deformations of a beam, regardless of the
particular loading applied, let us consider, for the sake of generality,
the beam shown in Fig. 4.5.

Mx
My

x
Sx
y

Sy

Fig. 4.4

Theory of elastic deformation of beams

99

z
x
y

dz
l

Fig. 4.5

This beam is subject to the couples M and M about the x-axis at


the left and right ends of the beam, respectively. These couples are
equal in magnitude and opposite in sign and therefore in equilibrium.
This beam is classed as being in pure bending.
Following the same line of reasoning adopted for the axially loaded
bar of Fig. 4.3, i.e. conceptually cutting the beam in two parts at any
cross section, we can easily deduce that the same bending moment is
applied to every cross section of the beam.
Also in this case we will assume, according to Bernoulli's model,
that any cross section remains plane and orthogonal to the axis after the deformation
has occurred due to the bending moment.

In contrast to the case of the axially loaded bar, which remains


straight, we will assume that in this case
the axis of the beam becomes an arc of a circle. Moreover, we will also assume that this
arc of circle retains its original length, so that we can consider that no stretching of the
axis of the beam has occurred.

Now consider Fig. 4.6, where an element of the beam between two
cross sections at distance dz is shown.
As a consequence of the assumptions stated above, it is evident that
after deformation has occurred, the right side face of the element has
rotated through an angle d about the axis x relative to the left side
face. In this framework, it is evident that the bres of the beam have
increased or decreased their length proportionally to their distance
from the x axis, according to the simple rule
dl 0 dz y

4:9

100

Energy methods in structural mechanics


d

z
y
dz

Fig. 4.6

Notice that the bres below the axis have increased their length, as
the sign holds for their distance from the axis, while the bres above
the axis have decreased their length, being governed by negative values
of y. We can now refer to the elongation, equation (4.1), to obtain a
relationship between the bending moment M and the radius of curvature r of the axis of the beam after deformation has occurred. Thus,
from equation (4.1) we have
dl 0 dz
E dA
4:10
dz
and this means that an axial force dN acts on the typical innitesimal
element of the cross section extracted from the beam. The magnitude
of this force is obviously dependent on the deformation. Substituting
equation (4.9) in equation (4.10) we obtain
y d
dN
E dA
4:11
dz
We can now sum all the contributions of the elementary forces on
the cross section to obtain the resulting axial force N and bending
moment M. Thus,

y d
d
E dA
E y dA
N dN
4:12
dz
dz
A
A
A

y d
d
E dA
E y2 dA
4:13
M y dN
dz
dz
dN

Since we assumed the reference system xy to have its origin at the


centroid of the cross section, then from this condition we have

y dA 0
4:14
A

Theory of elastic deformation of beams

101

and therefore
N0

4:15

Notice that for pure moment the neutral axis, that is the position
where the longitudinal bres neither extend nor contract, passes
through the centroid of the cross section of the beam. The integral
term on the right hand part of equation (4.13) is termed the second
moment of area Ix , sometimes called the moment of inertia,

4:16
Ix  y2 dA
A

In this case the second moment of area has been evaluated about the xaxis. The magnitude of this parameter will vary according to the orientation of the axis about which the second moment of area is calculated.
Now, we can write
d
EI
4:17
dz x
which yields the desired relationship between the bending moment M
and the radius of curvature r. As we have
M

d 1

dz r
it follows

4:18

M
1

EIx r

4:19

Equations (4.17) and (4.19) constitute key relationships for beams


subjected purely to a bending moment.
It is a simple matter to calculate the relative rotation between the
ends of the beam of Fig. 4.5. As the bending moment is constant, we
can integrate equation (4.17) over the whole length l of the beam
and obtain
l

l
M
M
M

dz
dz
l
EIx
EIx
EIx
0

4:20

The strain energy stored in the beam is given by the work done by the
couple M during the process of loading and is

U W M d
0

4:21

102

Energy methods in structural mechanics

which we can also write, by means of equation (4.20) and Clapeyron's


theorem, equation (3.91),
UW

1 EIx 2

2 l

4:22

So far we have dealt with a bar subjected to a constant axial force and a
beam subjected to a constant bending moment about the axis x.
It is signicant that all the relationships obtained above are linear and therefore the
behaviour of a beam subjected to both bending moment and axial force can be obtained
very straightforwardly by adding the eects of the axial load to those of the bending
moment.

In the next section we will see why this can be allowed from a geometrical point of view in the framework of the theory of very small
displacements. Following from their derivation, it is also clear that
equations (4.4), (4.17) and (4.19) hold true for any element of beam
between two innitesimally near cross sections. Therefore they can
also be applied to the analysis of beams subjected to axial forces N
and bending moments M that vary along the axis z.

4.3

The theory of very small displacements

Following the development in the previous section we can now begin


the study of the deformations of simple systems of beams subjected
to a variety of types of loading and constraints. A foreword is, however, necessary to introduce the basis on which the theory presented
here is founded, i.e. the theory of very small displacements, also called
the theory of innitesmal displacements. In order to x our ideas, by
very small we may for example intend displacements of the beam
normal to its longitudinal axis that are less than about 1% of its
length. In this manner the strains will remain in the elastic range for
most engineering materials.
This theory of very small displacements is essentially based on two
hypotheses, namely
.

the displacements of a deformable body are essentially negligible


in comparison with its dimensions and therefore eects from
separate loading states can be geometrically added with reference to the position and form of the element before loading
on account of the smallness of the displacement eld the forces
acting on the body can be always referred to the unstrained, i.e.
undeformed conguration.

These hypotheses are normally suciently accurate for the great majority of engineering applications and allow a substantial simplication for
analysing most cases of structural deformation. From a mathematical

Theory of elastic deformation of beams

103

dz
0

z
uy
uy + duy
ds

Fig. 4.7

point of view, the rst hypothesis corresponds to the neglection of all


the second and higher order terms in the derivatives of the displacement
eld. This point can be claried, for example, with respect to the expression of the curvature of the axis of a beam. In fact, with reference to
Fig. 4.7, it is


duy
 arctan
dz
s
 2
q
duy
dz
ds dz2 du2y 1
dz
4:23
d2 uy
1 d d
1
dz2
s
s



 2 3

r
ds
dz
duy 2
duy
1
1
dz
dz
where uy is the displacement of the beam centroid in the y direction and
is generally a function of z, i.e. uy z. Remember that conventionally,
anticlockwise rotations are considered to be positive.
As stated above, typically in the practical engineering application of
beams the deection is less than 1% of the length and the slope is also
negligible in comparison with unity, that is for an initially straight bar
with the axis coincident with the z axis
 2
duy
1
4:24
dz

104

Energy methods in structural mechanics

Equation (4.24) leads to the linearisation of equation (4.23), for the


displaced axis of the beam
2

d uy
1
 2
4:25
r
dz
This approximation has also a particular geometrical meaning, as it
assigns, to the same order of approximation, the identical length l to
both the undeformed uy 0 and the deformed uy uy z axis of
the beam between the values z 0 and z l. In fact, for the undeformed state, uy 0, it follows that
0

l dz l

4:26

while for uy uy z we have (Fig. 4.7)


s
 2
l q l
duy
1
dz
dz2 du2y
l0
dz
0

4:27

which becomes, on account of the negligibility of the term in equation


(4.24),
0

l dz l

4:28

Therefore the line of equation y uy z has the same length of its projection on the z-axis and we can consider the horizontal displacement
of the left end of the cantilever of Fig. 4.8, to be zero when subject to
bending due to the applied couple M.
The second of the hypotheses stated above can be claried by saying
that we can consider a force N which is perfectly axial before the deformation to be still axial after the deformation has occurred; or if the
force N is normal to the beam before deformation it will still be
normal after the deformation has occurred as it can be always referred
to the unstrained conguration of the beam. This is shown in Fig. 4.9.

Fig. 4.8

Theory of elastic deformation of beams

105

Fig. 4.9

It is evident from this example that in this theory the non-linear


eects (which eventually lead to buckling) cannot be considered. Buckling is considered in chapter 12 using an appropriately modied
description of beam deections. Other implications will be discussed
as we further the study of deformable bodies.

4.4

Euler's method for the analysis of beam


deformations

In the previous sections expressions have been derived for the strain
energy of a beam that deforms under the inuence of axial forces and
bending moments. Now we consider the methods with which we can
extend these expressions to analyse the behaviour of actual beams
subjected to specied loading and constraints. From a general point
of view there are two distinct approaches to this analysis process, i.e.
.
.

use an energy expression directly (typically to seek an approximate solution)


use the dierential equations of equilibrium which can be
derived from the variational statement, equation (1.28).

The advantage of the latter approach is that when a solution is


obtained to the dierential equation describing the equilibrium of
the beam the answer can be considered exact, at least for the model
adopted. The disadvantage is that such solutions usually are restricted
to idealised boundary conditions and only to limited variations in the
geometry of the beam and its types of supports. Nevertheless, the range
of exact solutions that have been developed for beams is fairly extensive and forms a valuable counterpart to the direct use of the energybased methods. One particular value of the exact solution of the
dierential equations is that such results provide a standard against
which to compare and validate the accuracy and use of approximate
methods, such as nite elements, involving direct use of the energy
functional and trial functions.
Thus, the approach followed in this text is similar to that in current
structural engineering practice in which the deformations of beam with

106

Energy methods in structural mechanics


q

Fig. 4.10

simple boundary conditions, uniform geometry and loading are evaluated using simple expressions derived from the dierential equation of
equilibrium developed in this section. The deformations of more complex conditions of beam geometry and loading are usually calculated
by means of approximate methods, such as the nite element methods
developed later in this text, but the exact solution of simplied beam
conditions is always kept as a validation benchmark.
It is also possible to obtain approximate solutions from the dierential equation of beam equilibrium by using the method of nite dierences. This was a quite popular and successful approach in the years
19301950. However, the nite element method has essentially
eclipsed the nite dierence approach for structural analysis, although
the latter remains the basis for computer modelling of uid dynamics.
For the sake of completeness, section 4.7 presents the nite dierence
method as applied to the analysis of the deformations of beams. From
a purely theoretical standpoint, it may be noted2 that the nite dierence and the nite element method, introduced later in this text, can
both be considered to be essentially forms of assumed trial functions,
with the former applied to the functions involved in the dierential
equations of equilibrium derived from the energy formulation and
the latter applied directly to the energy functionals.
In order to derive the dierential equation of equilibrium for
straight beams we apply the classical variational calculus to the
energy functional. This route is essentially due to the eminent Swiss
mathematician Euler (actually, after a suggestion by Bernoulli3 )
hence the classication of this development can be stated as Euler's
method of analysis of beam deformations.
Consider the simple supported beam shown in Fig. 4.10, subject to a
uniformly distributed vertical load per unit length q.
2

Croll, J. and Walker, A. C. 1971 Finite dierence method as a localised Ritz process.
Intnl J. of Eng. Math. 3, 155.
3
Bernoulli, Daniel (Groningen, 1700 Basel, 1782), Swiss mathematician and physicist, nephew of Jacob.

Theory of elastic deformation of beams

107

In order to study its state of equilibrium and deformation, we can


make reference to the general condition, equation (1.28)
U W
and write down the equivalence between the variation of the strain
energy and the variation of the work done by the external forces,
both due to a virtual eld of displacements u.
Writing the variation of the work done by the external forces is quite
a simple matter
l
W q uy dz

4:29

where uy is, as usual, the displacement of the beam axis in the y


direction.
However, writing the variation of the strain energy requires a little
more development. Following the approach in equations (4.17) and
(4.22) for a beam subject to constant bending moment, we can write
the energy stored in an innitesimally long element of beam as
1
d2
Udz EIx
2
dz

4:30

which holds true provided the curvature, and therefore the bending
moment, is evaluated point by point along the beam. In fact, in general
the moment and therefore the curvature varies along the length of the
beam. The energy stored in the whole simply supported beam is given
by the integral
l
1
d2
U
EIx
2
dz

4:31

or, equivalently,
 2
l
1
d
EIx
dz
U
2
dz

4:32

which, on account of equations (4.18) and (4.25), becomes


 2
 2 2
l
l
d uy
1
1
1
U
EIx
dz
EIx
dz
2
r
2
dz2
0

4:33

108

Energy methods in structural mechanics

The variation of this expression with regards to the displacement eld


uy is4
l

U EIx u00y u00y dz

4:34

where, for the sake of enabling a succinct presentation, we have written


d2 uy
 u00y
4:35
2
dz
We can therefore write the condition of equilibrium (equation (1.28))
as
l

EIx u00y

u00y

l
dz q uy dz

4:36

This variational expression can be manipulated to obtain a perfectly


equivalent dierential equation. Thus, integrating the left hand

4
The application of the calculus of variations provides a means to minimise the energy
functional and is summarised as follows. Suppose we are given a functional U in the
integral form




q
q
; ::: dV G ;
; ::: dS
U F ;
qx
qx
V

where F and G are functions of x; . . . and S is the boundary of the closed region V.
Suppose also that the variations in are taken among the admissible set of functions
satisfying the general boundary conditions
H1 0

on S1

H2 0

on S2

where S1 S2 S.
Then, for a small admissible variation from to  we dene the corresponding
rst variation in U by
 


qF
qF
q


   dV
U
q
qq =qx
qx
V

 

qG
qG
q


   dS
q
qq =qx
qx

which stems from a straightforward series expansion.

Theory of elastic deformation of beams

109

member by parts, we obtain


l

EIx u00y

u00y

dz

EIx u00y

u0y l0

0
EIx u000
y uy dz

4:37

and integrating the right hand term of equation (4.37) we have


l

EIx u00y

u00y

dz

EIx u00y

u0y l0

EIx u000
y

uy l0

EIx u0000
y uy dz
0

(4.38)

Now, remembering that a virtual displacement eld is, by denition,


innitesimal and compatible with the constraints on the beam, let us
have a closer look at the two rst right hand terms. The rst term, i.e.
EIx u00y u0y l0

4:39

is identically zero, since the expression


EIx u00y

4:40

represents, on account of the equations (4.25), (4.18) and (4.17), the


bending moment Mz at the ends of the beam. In fact, for a beam
with simple supports at A and B and loaded such that no couples
are applied at A and B the bending moment at the ends is zero and
equation (4.39) is evidently zero.
The second term on the right hand side of equation (4.38), i.e.
l
EIx u000
y uy 0

4:41

is also equal to zero, as the variation must vanish on the simple supports at A and B which do not allow any vertical displacement.
Thus, the condition of equilibrium (equation (4.36)) reduces to
l

EIx u0000
y

l
uy dz q uy dz 0

4:42

which, since the virtual displacement eld uy is arbitrary, is equivalent


to the fourth order dierential equation
EIx u0000
y q 0

4:43

This equation is usually called Euler's equation of beam deformation,


as its solution represents the deformed conguration of the elastic
beam under the specied lateral load distribution q. In the present
example of the beam shown in Fig. 4.10, the term EIx , which is

110

Energy methods in structural mechanics

commonly referred as the exural stiness of the beam, is a constant,


therefore the solution to equation (4.43) is trivially obtained by integrating equation (4.43) four times to give
uy z

qz4
Az3 Bz2 Cz D
24EIx

4:44

where A, B, C and D are constants of integration. Now we impose on


equation (4.44) the geometrical and mechanical constraints (i.e. in
terms of displacements and forces, respectively) at the simple supports
A and B, which allowed us to drop the terms in equations (4.39) and
(4.41). This process enables the evaluation of the constants of integration. These constraints require that
uy z 0
Mz

for z 0

EIx u00y z

and z l
for

z0

4:45
and z l

4:46

Therefore we have a non-homogeneous linear system of four equations


in the four unknown constants A, B, C and D. The system is
8
uy 0 D 0
>
>
>
>
>
EIx u00y 0 EIx 2B 0
>
>
>
<
ql 4
4:47
u
l

Al 3 Bl 2 Cl D 0
y
>
>
24EI
x
>
>
>
2
>
>
> EI u00 l ql 6EI Al 2EI B 0
:
x y
x
x
2
or, in matrix form,
2
32 3 2
3
A
0
0 0 0 1
6 0 2 0 0 76 B 7 6
7
0
6
76 7 6
7
4:48
6 3 2
76 7 6
7
4
4l
5
4
5
4
l
l 1
C
ql =24EIx 5
6l

0 0

ql 2 =2EIx

and yields
A
C

ql
;
12EIx

ql 3
;
24EIx

B0
D0

4:49
4:50

Finally, we get
uy z

q
z4 2lz3 ql 3 z
24EIx

4:51

Theory of elastic deformation of beams

111

z
B

y
l

Fig. 4.11

This fourth-order parabolic expression represents the eld of vertical


displacements, uy z, which the axis of the beam undergoes to
assume its conguration of equilibrium under the assigned uniformly
distributed load q. As the axis was originally straight, equation
(4.51) also prescribes the deformed conguration of the beam, as
shown in Fig. 4.11. The solution of the dierential equation (4.43) is
exact and therefore within Bernoulli's assumptions, this method of
analysis predicts exactly the deections of the initially straight beam.
Up to now we have illustrated the Bernoulli model of beam bending
and shown that we can obtain the conguration of equilibrium of a
simple supported beam under a specied form of loading, according
to the general condition equation (1.28). In the next section we will
see that Euler's equation for beam deformation, equation (4.43), can
be employed for the bending of any other simply constrained beam,
provided appropriate boundary conditions are taken into account.

4.5

The boundary conditions associated with Euler's


equation of beam deformation

For the sake of clarity we have considered so far a simply-supported


beam. However, the procedure which led from the variational condition of equilibrium (equation (4.36)) to the Euler equation of equilibrium (4.43) is a general one and can be applied to any kind of
variously constrained beam. The resulting dierential equation (4.43)
does not change, but the boundary conditions of the problem can
assume dierent forms. Therefore, let us now consider a generic
beam, whose bending behaviour is described according to Bernoulli's
model of plane sections. For the moment the treatment is not restricted
to any specic applied constraints.
Consider a beam subjected to a generic distributed vertical load per
unit length qz, as shown in Fig. 4.12. The load qz can vary along the
axis of the beam but is always orthogonal to it and is taken to act in the
y direction.

112

Energy methods in structural mechanics

q (z)

y
l

Fig. 4.12

Irrespective of the form of constraints applied at both ends of the


beam, it is straightforward to write the general condition of equilibrium
(equation (1.28)) in the form shown in equation (4.36). Integrating twice
by parts the terms on the left hand side of equation (4.36) we obtain
equation (4.38) where, of course, the virtual displacement eld has to
be innitesimal and compatible with the constraints applied to the beam.
Let us now have a closer look at the rst two terms on the right hand
side of equation (4.38). We already know that
Mz EIx u00y z

4:52

and, if we consider Fig. 4.13, we can readily deduce that in the theory
of innitesimal displacements
z  tg

duy z
dz

4:53

duy

dz

Fig. 4.13

Theory of elastic deformation of beams

113

That is, the rotation of a generic cross section can be identied with the
negative value of the slope of the axis of the beam. Therefore we can
attribute a precise physical meaning to the terms which appear in
equation (4.38), as follows
EIx u00y M

4:54

u0y

4:55



It remains to elucidate the meaning of the term


EIx u000
y z

4:56

Remembering equation (4.52), from the relationship equation (2.57),


which links the slope of the bending moment diagram to the value of
the shear force, we have
dMz
EIx u000
4:57
y z S
dz
Thus it is clear that the rst two right hand terms of equation (4.38)
are zero whatever the constraints applied at the beam's ends. This is
exemplied by a simple cantilever, as shown in Fig. 4.14. We have at
the left hand end,
uy 0 0

and 0 u0y 0 0

4:58

and at the right hand end


Ml EIx u00y l 0

and Sl EIx u000


y l 0

4:59

so that both terms


EIx u00y u0y l0

l
and EIx u000
y uy 0

4:60

are zero. Therefore, also in this case Euler's equation of beam equilibrium (4.43) is valid. Of course, while integrating the dierential

q (z)

Fig. 4.14

114

Energy methods in structural mechanics

l/2

l/2

Fig. 4.15

equation of equilibrium, equation (4.43), we must take into account the


appropriate boundary conditions, i.e. equations (4.58) and (4.59), to
evaluate the constants of integration. The reader may nd it edifying
to verify that the terms in equation (4.60) vanish for any beam in the
yz plane with various constraints at the ends, provided the beam is
indeed xed in the frame of reference when considered as a rigid body.5
Another interesting example is that given in Fig. 4.15. This is a beam
clamped at the left-hand end and simply supported at the right-hand
end. It is subject to a point load P at the middle of the span. In
order to derive an expression for its deection, it is convenient to
divide the whole span into two half-spans of length l=2. Let us consider
the two half-spans separately and denote the vertical displacement eld
of the rst half-span as u1y and that of the second one as u2y . The
condition of equilibrium (1.28) can be written as follows
l=2

EIx u001y

u001y

l
dz

EIx u002y u002y dz Puy l=2

4:61

l=2

where, of course, the displacement at the mid point of the whole span is
unique and is denoted by u1y l=2 u2y l=2 uy l=2.
After integrating equation (4.61) twice by parts, as discussed above,
we get
l=2

l=2

000
l
EIx u001y u01y 0 EIx u002y u02y ll=2 EIx u000
1y u1y 0 EIx u2y u2y l=2
l=2

0
5

EIx u0000
1y

l
u1y dz

EIx u0000
2y u2y dz Puy l=2

4:62

l=2

It is worth pointing out that for the cantilever in Fig. 4.8, which actually corresponds
to the situation in Fig. 4.5, it is found that uy M=2EIz2 . This represents an arc
of a circle only under the assumptions of small displacement theory developed in
section 4.3.

Theory of elastic deformation of beams

115

Now, taking into account the following boundary conditions


u1y 0 0

and 0 u01y 0 0

4:63

u1y l=2 u2y l=2

4:64

1 l=2 2 l=2 u01y l=2 u02y l=2

4:65

M1 l=2 M2 l=2 ) EIx u001y l=2 EIx u002y l=2

4:66

000
EIx u000
1y l=2 EIx u2y l=2 P

4:67

u2y l 0

and M2 l

EIx u002y l

4:68

which impose the geometrical and mechanical aspects of the constraints at both ends of the beam and require the continuity at the
mid-point of the span, we are led to the equation
l=2

EIx u0000
1y

l
u1y dz

EIx u0000
2y u2y dz 0

4:69

l=2

This expression is clearly equivalent to the following set of dierential


equations


l
0000
4:70
EIx u1y 0 for z 2 0;
2


l
0000
4:71
EIx u2y 0 for z 2 ; l
2
provided the conditions given by equations (4.63)(4.68) are applied to
the general solutions


l
3
2
4:72
u1y z Az Bz Cz D for z 2 0;
2


l
u2y z Ez3 Fz2 Gz H for z 2 ; l
4:73
2
in order to dene the value of the eight constants AH. In detail, we
have
u1y 0 0

and 0 u01y 0 0

4:74

u1y l=2 u2y l=2

4:75

1 l=2 2 l=2 u01y l=2 u02y l=2

4:76

M1 l=2 M2 l=2 ) EIx u001y l=2 EIx u002y l=2

4:77

116

Energy methods in structural mechanics


000
EIx u000
1y l=2 EIx u2y l=2 P

u2y l 0 and

M2 l

4:78

EIx u002y l

Finally, we can write


11 P 3
3 Pl 2
z
z
u1y z
96 EIx
32 EIx

4:79


l
z 2 0;
2

4:80

for

5 P 3
5 Pl 2
Pl 2
Pl 3
z
z
z
96 EIx
32 EIx
8EIx
48EIx


l
for z 2 ; l
2

u2y z

4:81

These two expressions completely dene the lateral displacement eld


of the propped cantilever under analysis. Notice that the illustrated
procedure can be extended in a straightforward manner to any
system of concentrated loads.

4.6

Example of application of Euler's method of beam


deformation for a beam with a variable depth

We have seen above that it is fairly straightforward to obtain an analytical solution to Euler's dierential equation (4.43) for a beam with a
uniform exural stiness EIx . The analysis becomes more complicated
when we consider beams with variable loading and geometry. As an
illustration of this, the following analysis considers a cantilever with
a linearly varying depth subjected to a load at its tip, as shown in
Fig. 4.16.
The gure shows that the cantilever has a length l. At the built-in end
the depth of the section is d1 and at the tip the depth is d2 . The uniform

d1

dx

d2

y
l

Fig. 4.16

Theory of elastic deformation of beams

117

breadth of the beam is b, so that the second moment of area of the cross
section at some position z along the beam is
Ix z

bd 3 z
12

4:82

The depth at position z is


dz d1 d2 d1

z
l

4:83

By introducing the non-dimensional variables  d2 =d1 and   z=l,


we can write
d; d1 1 1

4:84

where  2 0; 1. Also notice that for a generic function f from calculus


we have
df df d 1 df

dz d dz l d

4:85

For an innitesimal element of beam between two cross sections at


distance dz we can consider the depth as constant and we can again
make reference to Bernoulli's model of beam deformation introduced
in section 4.2. Retracing the steps of section 4.4, we can therefore write
the dierential equation governing the deformation of the beam as


2
1 d EIx  d uy 
q
4:86
l 2 d 2
l2
d 2
As in the present case no distributed load q is applied (we consider
the weight of the cantilever negligible in comparison to the applied
force P), by integrating twice equation (4.86) we get


2
1 d EIx  d uy 
Al
4:87
l d
l2
d 2


2
EIx  d uy 
4:88
Al 2 Bl
d 2
l2
The mechanical boundary conditions are


2
1 d EIx  d uy 
S1
P
l d
l2
d 2


2
EIx  d uy 
0
M1
l2
d 2

4:89
4:90

118

Energy methods in structural mechanics

and, consequently, we have


Al P

Pl Bl 0

P
l

4:91

BP

4:92

A
)

so that


2
EIx  d uy 
Pl1 
l2
d 2
By virtue of equations (4.82) and (4.84) we can write


2
1 d uy 
Pl
12Pl
1
1 

EIx 
l 2 d 2
Eb d13 1 13

4:93

4:94

The slope of the beam can be obtained by integrating equation (4.94)


as follows
1 duy 
l d


12Pl 2
2 2  2

C
Eb d13 2 12 1  2



4:95

The constant C is evaluated using the following natural boundary


condition


1 duy 
0
l d

at  0

4:96

Hence, we have
A

2
2 12

4:97

and equation (4.95) becomes


;



1 duy 
12Pl 2 2  2

l d
Eb d13 21  2

4:98

The deection function for the beam is obtained from equation (4.98)
once again by integration, thus


12Pl 3 f1 ;
uy ;
D
4:99
Eb d13 f2 ;

Theory of elastic deformation of beams

119

where
f1 ; 3  2 2  4 2  2 5  2 3  2 2 2
21  2 ln1  

4:100

f2 ; 21   13
Evaluating the constant D for the boundary condition
uy 0; 0
we have
D


2 13

4:101
4:102

so that equation (4.99) is


3
2
3 4 2 5 2 2 21 
7
6
12Pl 3 4
21   ln1  5
4:103
uy ;
Eb d13
21   13
The most easily recognised result from equation (4.103) is the value
of the deection of the tip of a cantilever with a uniform depth d1 . In
fact, for  1 and 1 equation (4.103) gives
uy 1; 1

4Pl 3
Eb d13

4:104

Figures 4.17(a)(c) show typical results for the deection along the
beam for values of 1, 0.5 and 0.05, respectively.
In these gures the value of the deection has been normalised with
respect to the deection of the tip of a cantilever with a uniform depth,
that is all the calculated values have been divided by equation (4.104).
The value 0:05 has been chosen since the value of 0 causes the
tip of the cantilever to have no depth and therefore the deection
becomes unrealistically large at the tip.
The variation of the normalised deection at the tip of the cantilever
with the parameter is given by

 3
4 2 3 2 ln
u~y 1; 3
4:105
2 13
Figure 4.18 shows results from equation (4.105).
It is clear that for dierent and more complicate variations of the
depth of the section along the beam length the integration of the dierential equation can pose some problems. In such a case, if we wished
we could integrate the equation of equilibrium numerically. However,

120

Energy methods in structural mechanics


10

=1
08

u ()

06
04
02
0
0

02

04

(a)

06

08

10

06

08

10

06

08

10

20

= 05

u ()

15

10

05

0
0

02

04

(b)

= 005
5

u ()

4
3
2
1
0
0

Fig. 4.17

02

04

(c)

Theory of elastic deformation of beams

121

Fig. 4.18

following such an approach a simple formula for use in design calculations would not be readily available. The next section shows an alternative approach to obtaining a solution to the Euler dierential
equation of deformation.

4.7

Approximate solution of Euler's equation of beam


deformation using the method of nite dierences

In the next chapters much space will be dedicated to the development of


the method of trial functions, including the method of nite elements. In
fact this latter method has in recent years become the main approach to
evaluating the deformations of beams, plates and shells with practical
forms of geometry, boundary conditions and loading. Several exact
analytical solutions, usually for simplied boundary conditions and
loading, of the dierential equations of equilibrium derived from the
corresponding energy expression for beams, plates and shells can provide valuable information for validating the nite element numerical
modelling and are generally used to assess the accuracy of the results.
It is possible, nevertheless, to obtain approximate solutions for the
dierential equations of equilibrium for a variety of geometries and
loading conditions. This can be achieved using the nite dierence
method. Nowadays this method is not used widely in practical engineering design analysis, largely because of the greater versatility of the nite
element method and the large number of available commercial computer programs that incorporate the nite element method.
However, for the sake of completeness in this text, the present
section presents the basis of the dierence method and exemplies its
use for the determination of the deformations of beams based on the
Euler equation of equilibrium.

122

Energy methods in structural mechanics


f (x)

(n +1)
(n 1)

(n)

Fig. 4.19

Figure 4.19 shows the plot of a function f(x) of a variable x. The


slope of the curve at some position x is obtained analytically by dierentiation of the function, i.e.
df x
4:106
dx
The slope can be approximated as follows. The range of the function is
sub-divided to form a series of nodes along the curve describing the
function. The nodes are usually, but not necessarily, equidistant
from each other. In the gure the nodes are separated by a distance
h. The value of the function at node n is fn and the slope at that
node is approximated by
f 0 x

1
f
fn 1
4:107
2h n 1
Intuitively, as the nodes become very closely spaced, i.e. h ! 0, the
approximate value for the slope approaches the analytical value. The
higher order derivatives can be obtained using the same approach.
For example, the second derivative is written as the ratio between
the increment of the rst derivative between the mid-points of the
arcs n 1; n and n; n 1 and the distance h, namely

 



d df x
fn 1 fn
fn fn 1
00
f n x

dx dx x xn
h
h
f n0 xn

1
fn 1 2fn fn 1
h2

4:108

Theory of elastic deformation of beams

123

The fourth derivative is approximated by


1
f n00 1 2f n00 f n00 1
h2
1
4 fn 2 4fn 1 6fn 4fn 1 fn 2
h

f n0000

4:109

The nite dierence method is exemplied here by application to a


beam with uniform cross section, simply supported at both ends and
subjected to a uniformly distributed load q. The accuracy of the
result is judged by comparison of the value of the mid-span deection
with the corresponding value from the analytical solution of the Euler
deformation equation, see equation (4.51),
 
l
5ql 4
ql 4
uy

:
2
384EIx 76 8EIx

Three internal nodes

Figure 4.20 shows the arrangement of the nodes along the beam, in this
case the model includes three internal nodes, i.e. between the end
boundary conditions. The boundary conditions are chosen so that displacement in the y direction is zero at z 0 and l and that the moment,
and therefore the curvature, is also zero at these positions.
Thus
u0 u4 0

and

u1 
u1 ;

u5 
u3

4:110

The Euler equation (4.43) can be written in terms of nite dierences at

l
q

EI

z
y
1

h=

Fig. 4.20

l
4

124

Energy methods in structural mechanics

each of the internal nodes by virtue of equation (4.109). Hence,


Node 1
u1 4
u0 6
u1 4
u2 u3

qh4
EIx

and by setting
l
h ;
4

Q

ql 4
256EIx

we have, taking into account equation (4.110),


5
u1 4
u2 u3 Q
Similarly,
Node 2
u2 4
u3 Q
4
u1 6
Node 3
u2 5
u3 Q
u1 4
The equations describing the relationships between the deections at
the nodes and the applied loading can be put into matrix form, as
A
un Qn
where

5
A 4 4
1

4
6
4

4:111
2 3
3
2 3
u1
1
1
4 5 un 4 u2 5 Qn Q4 1 5
u3
1
5

4:112

The values of the deections at the nodes may be obtained from equation (4.111) by rearranging the equation as

un A1 Qn
which gives
2 : 3
2 3
25
u1
4 u2 5 Q4 3:5 5
2:5
u3
and the deection at the mid-span is
u2

3:5 ql 4
ql 4
:
256 EIx 73 1EIx

This is within about 5% of the analytical result shown above.

Theory of elastic deformation of beams

125

Five internal nodes with assumed symmetry

The accuracy of the numerical solution can be improved, if required,


by increasing the number of nodes included in the model. However,
this will increase the numerical eort and to maintain a minimum
eort it is worthwhile to use any form of geometric symmetry that
will occur in the beam. This is illustrated in this example by increasing
the number of internal nodes to ve and by incorporating symmetry of
deformations at the mid-span. The arrangement of the nite dierence
model is shown in Fig. 4.21.
The symmetry is readily incorporated by specifying
u4 u2 ;

u5 u1

Now
l
h ;
6

ql 4
1296EIx

Thus, dealing with each of the nodes in turn, a matrix equation, similar
to equation (4.111), can be obtained. In this case
2
3
5 4
1
6
7
A 4 4
7 4 5
2 8
6
and the solution is
2 : 3
8 75
6 : 7

un Q4 15 0 5
17:25
The deection at mid-span is
u3

17:25 ql 4
ql 4

1296 EIx 75:1EIx

Symmetry

Fig. 4.21

126

Energy methods in structural mechanics

which is within 2% of the exact analytical value given above. Again, if


required, an even greater degree of accuracy can be obtained by
increasing the number of nodes in the model.

4.8

Applied displacements and constraint reactions

Another class of actions capable of producing deformations on elastic


structures is now introduced. However, these actions cannot be classied as forces or thermal loads but they can be classied as applied
displacements. In order to x our ideas, let us consider the beam
with built-in ends shown in Fig. 4.22.
We assume that in its initial conguration the beam is undeformed
and we attribute the zero value to the energy of this status. Let us
now suppose that as part of the loading procedure the built-in end B
is forced to undergo a vertical displacement uB and a rotation

B,
as shown in Fig. 4.23.
Even in the absence of applied loads, the beam is thus forced to
assume a deformed conguration. The deformations can be examined,
as described above, by means of Euler's equation of beam deformation
(4.43) in its homogeneous form, that is
EIx u0000
y 0

4:113

provided the following boundary conditions are imposed


uy 0 0

and u0y l 0

at A

4:114

uy l u and u0y l  at B

4:115

Fig. 4.22

uy

Fig. 4.23

Theory of elastic deformation of beams

127

After some simple integration and evaluation of the constants, we get






 2
 3
u 3
u 2

uy z 2 3 z
z
4:116
l l2
l
l
On account of this deformation the beam will possess a certain level
of strain energy U. This can be calculated using the expression in
equation (4.33)
l
1
U
EIx u00y 2 dz
2
0

which, by means of equation (4.116) yields


 2

u 12
1
4
12
u2
2 3
U EIx
2
l
l
l

4:117

However, one could object that the expression of the internal energy
(equation (4.117)) should actually be zero, as the principle of conservation of energy requires that
U U W

4:118

and in our case there are no externally applied forces, and so it would
seem
W 0

4:119

But obviously this is not really the actual situation. In fact, a displacement applied to its constraints requires the structure to assume a
certain conguration. Remembering what has been presented in
chapter 2, section 2.3, we can have two cases, namely
(a)
(b)

the system is statically determinate, i.e. the constraints are


strictly only sucient to prevent any rigid body displacements
(there are no redundant constraints)
the system is statically indeterminate, i.e. there are more
constraints than degrees of freedom (some constraints are
redundant).

In the rst case the structure can move bodily without deforming, in
accordance to the kinematics of rigid bodies studied in chapter 2.
This is the case shown in Fig. 4.24.
Of course no strains are induced, as the reader can easily verify by
integrating Euler's equation of beam deformation with the following

128

Energy methods in structural mechanics

uy

Fig. 4.24

boundary conditions
uy 0 0

and EIx u00y 0

uy l u and

EIx u00y l

at A
at B

4:120
4:121

We know as well that, in this case, in the absence of applied forces the
reactions of constraints are zero (section 2.3) and therefore U is eectively zero. The case of the structure in Figs 4.22 and 4.23 is, on the
contrary, that of a statically indeterminate system. In this case some
of the constraints are redundant and a rigid body displacement eld
cannot take place. An imposed displacement and/or rotation of the
built-in end B can only occur by means of a deformation of the structure. Moreover, we discover that at B we have non-zero values of the
bending moment and of the shear force, which represent the reactions
of the constraint. Thus, we have


6
u 4

MB
Ml EIx u00y l EIx 2
6 0
4:122
l
l


12
u 6

RB
Sl EIx u000
6 0
4:123
l

EI

x
y
l3
l2
Notice that also at A, i.e. for z 0, we also have non-zero values for
the reactions at the xed ends. These are


6
u 2
00

MA M0 EIx uy 0 EIx 2
6 0
4:124
l
l


12
u 6
000

RA S0 EIx uy 0 EIx 3 2 6 0
4:125
l
l
We will now see that the deection of the beam can be related to the
reactions arising at the constraints. First of all it must be pointed out
that the deformed conguration of the beam is the same as would have
obtained by considering a cantilever xed at A and subject to a force R
 at the end B, where the values of the moment and force
and a couple M
are given by equations (4.122) and (4.123). This would have clearly
meant imposing the following two mechanical conditions at B instead

Theory of elastic deformation of beams

129

Fig. 4.25

of the two geometrical conditions (equation (4.115))




6
u 4
00

EIx uy l MB EIx 2
l
l


12
u 6

EIx u000
l

RB

EI

x
y
l3
l2

4:126
4:127

and the result would have been seen to be identical. Because of this
result, let us evaluate the strain energy for the beam shown in Fig.
4.25, which is therefore perfectly equivalent to the beam in Fig. 4.23.
We have, by Clapeyron's theorem
1
1 
R
u M

2
2




1
12
u 6
1
6
u 4 
EIx

2 u EIx 2
2
2
l
l3
l
l

and after a little algebra


 2

u 12
1
4
12
u2
U EIx
2 3
2
l
l
l

4:128

4:129

This expression turns out to be identical to that in equation (4.117).


It is now clear that
the strain energy stored in the statically indeterminate beam can be related to the work
done by the reactions of the constraints along the deformation process.

In other words, the strain energy stored in the beam results from the
work that we must perform against the stiness of the beam in order
to apply a given displacement at the built-end B. Therefore U is not
zero, and the principle of conservation of energy is not violated as
W is also not zero.
It should be noted that the magnitudes of the reactions, equations
(4.122) and (4.123), are linear functions of the bending stiness EIx .
As a consequence, very rigid structures induce very high values of
the bending moments and shear forces, even for quite small displacements of the constraints. At the end of this section we shall derive

130

Energy methods in structural mechanics


B

uy

Fig. 4.26

the magnitudes of the restraint reactions when a vertical displacement



are applied at the end A of the beam shown in
uA and a rotation A
Fig. 4.26.
This is a situation really complementary to that shown in Fig. 4.23
and has been dealt with in the preceding presentation. However trivial
its solution actually is, it is of considerable importance for the matrix
treatment of frames which we will develop in chapter 9 and is it therefore valuable to that development to present here the above mentioned
results in some detail.
We have from the above analysis


6u 4

MA M0 EIx 2
4:130
l
l


12u 6

RA S0 EIx
2
4:131
l3
l


6u 2

4:132
MB
Ml EIx 2
l
l


12u 6

RB
Sl EIx 3 2
l
l

4.9

Eects of temperature changes

4:133

In this, the nal section of this chapter, we wish to deal briey with the
eects of temperature changes on bars and beams. In section 1.2 we
presented the expression of the energy of a bar which we assume is in
equilibrium under the actions of equal and opposite axial forces N and
N and a temperature increase T T T0 . This is, see equation

Theory of elastic deformation of beams

131

(1.13),
ET; l 0 CT T0 12 k l 2 k l lT T0
We attribute the zero value of energy to the state dened by T T0
and l 0 l. The relationship between the elongation l l 0 l and
the applied forces and variation of temperature is given by equation
(1.12)
l

F
lT T0
k

where k is the axial stiness of the bar and is the coecient of linear
thermal expansion, which for small changes in temperature can be
considered constants for the material and the element. We can now
give a more precise meaning to the axial stiness k of the bar and,
remembering equation (4.5), write
1 EA 2 EA
l
l lT T0 4:134
2 l
l
Elucidating equation (4.134) it is noted that the rst term on the right
hand side of this equation, that is
ET; l 0 CT T0

CT T0

4:135

represents the heat stored in the bar and is directly related to the variation of temperature T T T0 . The second and third terms, i.e.
1 EA 2 EA
l
l lT T0
4:136
2 l
l
represent the dierence between the equation (4.8) for the elastic
energy U which would have been stored in the bar if the elongation
l l 0 l had been a consequence of the application of only the
axial forces N and N and a correction term. This term can be rearranged as
EA
l lT T0 N lT T0
4:137
l
and is seen to represent the work done by the axial load N as an eect
of that part of the elongation due to the thermal expansion only. This
work is not actually stored as elastic energy of the bar and is subtracted
from equation (4.8). The system is dened to be in equilibrium and
therefore the kinetic energy K has zero value. Therefore, from a
purely kinematic point of view the eect of the temperature increase
T T T0 can be reduced to the elongation
lT T0

4:138

132

Energy methods in structural mechanics


T = T T0

Fig. 4.27

As seen in section 3.2, an elongation which is related to equation


(4.138) constitutes a strain eld which does not depend on the applied
loads. Such strains are generally called inelastic strains. In the case of
thermal strains we can easily characterise them using the assumption
that the material is isotropic and has the same thermal expansion properties in all directions. Thus we can say that the meaning of the coecient of linear thermal expansion is that for a unit rise in temperature
the normal strain produced in all directions is equal to , that is
"xx "yy "zz T

4:139

If the body is free to expand this state of strain is not associated with
elastic energy and the only energy stored in it is represented by the heat
expressed in equation (4.135). The unit change in volume "v , is given by
the expression
"v "xx "yy "zz 3 T

4:140

However, we may now also consider that the body is not free to
expand. For example, let us refer to the bar shown in Fig. 4.27. This
bar is clamped at each end and is subject to a temperature increase
T T T0 . The end constraints prevent any changes in length
and thus react to produce a state of strain which, superposed on the
thermal strain, makes the total strain equal to zero. From a mathematical viewpoint we have
N
l0
4:141
EA
where N is the axial force induced in the bar due to the constraint reactions, as shown in Fig. 4.28.
The value of the axial force is seen to be
l l T

N T EA

4:142

that is a compressive force which produces an axial deformation of the


same value but opposite in sign to that induced by the thermal expansion. The energy of the bar ET is
ET CT T0

1 EA
T T0 l 2
2 l

4:143

Theory of elastic deformation of beams


N

133

l
l + l T

Fig. 4.28

and is evidently a function only of the temperature T (the length l is


xed). The magnitude of the reactive forces that are activated by temperature changes may be very high; this is exemplied by a steel bar
0.5 m long, whose sectional area is 2500 mm2 . If we dene this bar to
be rigidly xed between two walls, the force N given by equation
(4.142) is equal to 244 kN (for steel bodies between 0 8C and 100 8C
the coecient can be taken equal to 11:6  106 8C1 and the modulus can taken as 2:1  105 N/mm2 ). For this reason the forces induced
by thermal strains must always be carefully evaluated in all those structures that are exposed to direct sunlight, like bridges and tall buildings.
As we stated in chapter 1, our study is limited purely to the mechanical behaviour of bodies. Therefore we do not concern ourselves with
the heat exchanged between the environment and the body under analysis, provided all the transformations take place very slowly and the
body assumes at any stage the same temperature as the environment.
This condition is satised for practically all the structures in normal
engineering practice and we can limit ourselves to the kinematic eects
of temperature changes, i.e. the expansion and thermal strains introduced above. In this manner we can simply determine the deformed
state of a simple supported beam which is exposed to direct sun
rays, such as is shown in Fig. 4.29.
For the sake of simplicity, consider a rectangular cross section and
suppose that the temperature of the upper face of the beam, which is
directly exposed to the sun, is T2 , while the temperature of the lower
T2
T1
A

B
l

Fig. 4.29

134

Energy methods in structural mechanics


T2

x
H

T1

Fig. 4.30

face is T1 , such that


T2 > T1 > T0

4:144

T0 is, as usual, the reference temperature.


We suppose also that the temperature gradient is constant through
the depth of the cross section, that is the temperature varies linearly
between the bres at y H=2 and the bres at y H=2, as shown
in Fig. 4.30.
It is convenient to think of such a state of thermal strain as being
equivalent to the sum of a uniform temperature variation, that is
T T2
T 1
4:145
2
 i.e.
and a linear temperature variation between the values T and T,
T T2 T2 T1

T T2 1
4:146
2
2
T T2 T1 T2
T T1 1
4:147

2
2
at H=2 and H=2, respectively. The situation is pictured in Fig. 4.31.
T2

T1

Fig. 4.31

+T

T=

T1 + T2
2

T=

T2 T1
2

T =

T1 T2
2

Theory of elastic deformation of beams

135

T2

T1
A

Fig. 4.32

 we nd also
With reference to the uniform temperature variation T,
that the resulting thermal strains are uniform through the depth of the
beam and we have
"z T
4:148
On the other hand, with reference to the linear temperature variation
 we nd that the thermal strains also vary linearly
between T and T,
through the depth of the beam and we have
2y
"z T
4:149
H
It is clear that this expression corresponds to a curvature of the
longitudinal axis of the beam given by
1
1
T
4:150
r
H
which is constant along the longitudinal axis of the beam. Therefore
the simply supported beam being considered here deforms in response
to the temperature variation as shown in Fig. 4.32, and the displacements at A and B are
l


uz B T dz Tl

uz A 0

4:151

0
l=2

A
0
l=2

B
0

1
l
dz T
H
2H

4:152

1
l
T dz T
H
2H

4:153

T

It is worth noting that in the case where the same variation of temperature is applied to a xed-ended beam (Fig. 4.33) the displacements
described in equations (4.151)(4.153) are not allowed any more.

136

Energy methods in structural mechanics


T2
A

B
T1
l

Fig. 4.33

In order to evaluate the reactions that arise at the ends, we can once
again imagine the thermal strains as given by the sum of the contributions in equations (4.149) and (4.150). With reference to the uniform
axial strain eld, equation (4.149), we have already found that the constraint reactions are represented by the forces
Ry A T EA
Ry B T EA
4:154
With reference to the linear strain eld, equation (4.149), by recalling
equation (4.33) we can write the elastic energy of the beam as


l
1
1
1 2

T
U
EIx
dz
2
r
H
0



l
1
1 2

EIx u00y T
dz
2
H

4:155

where the term between brackets is the elastic curvature, i.e. the
algebraic dierence between the total curvature and the thermal one.
The condition of equilibrium is


1
00

u00y dz 0
U EIx uy T
H
l

4:156

and, following the same procedure as presented in section 4.4, we are


led to the conclusion that this condition is equivalent to the fourth
order dierential equation


1 00
EIx u00y T
0
4:157
H
The general solution of this equation is in the form


1 2
z Cz D
uy z Az3 B T
H

4:158

Theory of elastic deformation of beams

137

and the fullment of the boundary conditions


uz 0

for

z0

z u0y z 0

and z l

for

z0

4:159

and z l

4:160

produces
uy z 0

4:161

This means that apparently the beam does not undergo any deformation. In reality the elastic strains compensate exactly the thermal
ones at any point of the beam. The bending moment resulting from
equation (4.157) is seen to be constant and has the same value of the
reactive moments at the constraints A and B, that is
1
MA M0 EIx u00y 0 EIx T
H
1
MB Ml EIx u00y 0 EIx T
H

4:162
4:163

Suggested exercise problems


1.

A section through the constant thickness wall of a dam which is


composed primarily of reinforced concrete is shown in Fig. 4.34.
The wall is subjected to the pressure of the water. Ignoring the
eects of the end conditions of the dam we can make an initial estimate of the deection of the top of the dam wall using a unit length
of the wall. Calculate this deection  assuming the equivalent
uniform modulus of the reinforced concrete as Ec 40 kN/mm2 .
[Ans:  11:4 mm]

3m

40 m

Fig. 4.34

35 m

138

Energy methods in structural mechanics


400 mm

q
20 mm
C

20 mm 15 m

20 mm
t

D
10 m

75 m

30 m
15 m
(a)

(b)

Fig. 4.35

2.

3.

A bridge consists of two longitudinal girders joined by a transverse beam carrying the load. The loading on the girders can be
represented by the uniform distributed load shown in Fig.
4.35(a). The girders have the cross-section geometry shown in
Fig. 4.35(b) and are made from steel with a modulus of 205 kN/
mm2 . The bottom anges of the girders are reinforced by increasing the thickness from 20 mm to 20 t mm so that the longitudinal strain in the ange at the mid-length, position C in Fig.
4.35(a), is equal to the strain at the quarter length, position D in
Fig. 4.35(a).
Calculate the required thickness t of the reinforcement and also
the reduction in the deection that will result from adding the
reinforcing plate.
[Ans: t 23:9 mm, the mid-span deection of the reinforced beam
is 76% of the beam without reinforcement]
The main beam of a bridge is supported at mid-span by a column
sitting on an elastic foundation, as shown in Fig. 4.36(a). This is
equivalent to the idealisation shown in Fig. 4.36(b) where the
column is represented by a spring with a stiness of k [N/mm].
q

EI

EI

(a)

Fig. 4.36

(b)

Theory of elastic deformation of beams

139

Direction of fluid flow


30 m

30 m

30 m

Fig. 4.37

4.

5.

Develop an expression for the vertical deection of the mid-span


of the bridge deck, relating the deection , the exural stiness
EI, the length l, and the uniform distributed load q.
[Ans:  5ql 4 =384EI 8kl 3 ]
A pipe carrying uid with a unit weight, f 11 000 N/m3 , has an
outside diameter of 1 m and a wall thickness of 15 mm. The pipe is
supported at 30 m intervals as shown in Fig. 4.37. The uid ows
through the pipe such that it half lls the pipe. The uid has a temperature of 80 8C and it may be assumed that the conduction of the
heat from the uid to the atmosphere through the pipe is such that
the temperature varies linearly from 80 8C at the bottom of the pipe
to ambient temperature of 20 8C at the top, at end A.
Calculate the deection of the pipe at mid-span when it is rst
installed empty, the unit weight of steel is s 77 000 N/m3 .
Calculate also the change of deection as the uid ows through
the pipe and the temperature reached at the equilibrium conditions specied above.
[Ans: 1 6:3 mm, 2 13 mm]
A mechanical component comprises two straight bars joined
rigidly at B. The bar BC is subjected to a uniformly distributed
load, q 2000 N/m, as shown in Fig. 4.38. Both bars are made
A

50 mm diameter
2m
q
B

C
3m

Fig. 4.38

140

Energy methods in structural mechanics

from 50 mm diameter round steel bar and are connected to rigid


foundations at A, B and C. The modulus for steel is 205 kN/
mm2 . Develop a solution to enable the evaluation of the rotation,
c of the joint at C. Evaluate also the maximum strain "m experienced by the elements in the structure.
[Ans: c 0:029 rad, "m 1:91  103 ]

5.

General principles in the analysis


of linear elastic structures

Introduction

This text is primarily concerned with the presentation of methods of


analysis of structures and structural components, such as beams,
that have a linear relationship between the levels of the applied loads
and the eects of these loads with regards to the strains and deformations of the structure and all its components. It is common experience
to expect that if a beam is loaded by a force normal to its axis the
deformation of the beam will be in the direction of the load. If we
double the magnitude of the force we would expect that the level of
the deformation would increase by a factor of two. Essentially the
common experience forms the basis for linear behaviour of structures.
We should be aware, however, that structures do not always obey this
linear correspondence between load and eect. In chapter 12 we consider the situation of beam and frame buckling where the application
of a load aligned with the axis of the beam can result in deformations
normal to the axis. The level of the deformations is in such a case nonlinearly related to the level of the load. In other words, if we increase
the level of the load the magnitude of the deformations are increased
disproportionally. Another source of non-linear relation between loading and consequent deformation would result from non-linear material
properties, either a non-linear stressstrain path or yielding of the
material at a constant stress. This source of non-linearity is not
included in the presentations of this text.
The linear behaviour of structures and its analysis constitutes a
central feature of structural engineering and this text considers a
wide range of types of structures and components that exhibit this
form of behaviour. There are some general principles that form the
fundamental basis of this linear behaviour and of the analysis methods
developed to explain and predict the response of structure to applied
loading. This chapter presents these principles of superposition and

142

Energy methods in structural mechanics

reciprocity, in a general manner. Although the principles are illustrated


with regard to simple beams, it should be remembered that the principles are completely general and apply to all forms of structures, such as
trusses, frames and plates.

5.1

The principle of superposition of the eects in the


linear theory of elasticity (existence and uniqueness
of the solution)

So far we have introduced the rst examples of elastic structures, i.e.


bars and beams, and dealt with their treatment using energy principles.
It is now timely to give some attention to the framework in which we
are operating, i.e. the linear theory of elasticity. In order to make completely clear the foundations of this theory, let us recall what we have
already determined in the previous chapter, namely
.

equations (4.7) and (4.19) describing the elastic behaviour of


Bernoulli's model of beam deformations are both linear. This
means that eects resulting from axial forces or bending moments
are proportional to the intensity of these actions on the element
operating in the theory of innitesimal (very small) displacements allows the displacement and strain elds corresponding
to dierent load conditions simply to be added.

These are the two pillars on which the linear theory of elasticity is
founded, as indicated in Fig. 5.1.
The rst of the points above, which involves the linearity of the constitutive relationships of the model, is clearly a physical assumption,

Fig. 5.1

Theory of infinitesimal displacements

Linearised constitutive laws

Linear theory of elasticity

General principles in the analysis of linear elastic structures

143

P
q
A
B

l/2

l/2

Fig. 5.2

while the second, related to the linearisation of the displacement and


strain elds, is essentially a mathematical axiom. As may be inferred
from its description, in this theory all the equations are linear, with
comprehensible and acceptable simplications in the treatment of a
vast class of structural problems.
First of all it is clear that in this framework the principle of superposition
of eects holds true. In other words we are allowed to add and subtract
the eects due to dierent separate loading conditions. In its simplest
form this means that if we, say, were to double the magnitude of the
external forces, i.e. by adding to a certain set of loading a second set of
loading equal to the rst one, we would expect the consequent deformations also to increase by a factor of two. As a further example, let us
consider the beam shown in Fig. 5.2, loaded by a uniformly distributed
load q and by a concentrated load P at the middle of the span.
With respect to any eect, we can analyse the beam model shown in
Fig. 5.3 and add the resulting solution to the one already obtained for
the beam shown in Fig. 4.15 and analysed in chapter 4.
The eld of lateral displacements, i.e. normal to the longitudinal
axis of the beam, for the beam in Fig. 5.3, subject to the uniformly distributed load q is given by the solution of Euler's equation of beam
deformation, (4.43), under the following boundary conditions
uy 0 0

and 0 0 at A

uy l 0

and Ml 0

5:1

at B
q

Fig. 5.3

144

Energy methods in structural mechanics

From equation (4.44) we have


uy z

qz2
2z2 5 l z 3l 2
48EIx

5:2

The solution for the beam in Fig. 5.2, subject to both the evenly distributed load q and the concentrated load P at the mid-point of the span, is
thus given by the simple addition of equation (5.2) to equations (4.80)
and (4.81), i.e.
u1y z

z2
9Pl 6ql 2 11P 10qlz 4qz2
96EIx

5:3

for z 2 0; l=2 and


u2y z

l z
2Pl 2 10Plz 5P 6qlz2 4qz3
96EIx

5:4

for z 2 l=2; l.
Of course we would have obtained the same expressions by writing
the condition of equilibrium for the beam subjected to both these
loads simultaneously and developing all the calculations accordingly.
However, as will become clearer in the following, the principle of
superposition of the eects is a very powerful tool not only from a
practical, but also from a theoretical point of view.
For example, let us x our attention on two apparently very simple
questions
.
.

Does the general condition of equilibrium (1.28) always lead to a


solution of the problem?
Is this hypothetical solution unique?

In the linear theory of elasticity the answer to both these questions is


yes.
With regards to the rst question, the demonstration of such a statement is quite a complicated matter from a general point of view and it
has been fully established only in relatively recent times. Nevertheless,
it is very easy to reach the correct conclusion in many specic
problems, given the linearity of the equations that describe the
models. This has been the case, for example, of Euler's equation of
beam deformation, (4.43).
With regards to the second question, the answer was provided by
Kirchho 1 about the middle of the 19th century and its demonstration
1

Kirchho, Gustav Robert (Konigsberg, 1824 Berlin, 1887), German mathematician


and physicist.

General principles in the analysis of linear elastic structures

145

Pi

V
u=0

Fig. 5.4

constitutes an instructive application of the principle of superposition


of eects. In order to demonstrate the assertion of uniqueness, let us
start our reasoning from the most general possible standpoint.
Let V be a generic linearly elastic body with certain geometrical constraints on the boundary and subject to a generic system of forces P i ,
as shown in Fig. 5.4.
Let us initially assume the existence of two distinct solutions of this
problem, say A and B. Also we assume that these solutions are dened
by the elds of displacements uA x; y; z and uB x; y; z. As we know,
the variation of the strain energy of the body is dependent on its
state of deformation only, that is

U f "ij dV
5:5
V

Moreover, in chapter 3 we noted that in the linear theory of elasticity


this strain energy is a quadratic function of the strain components "ij ,
see equation (3.107),

E
1 "2xx "2yy "2zz
U
21 1 2
V

2"xx "yy "xx "zz "yy "zz



21 2 2
2
2

"xy "xz "yz dV


1 

5:6

Since these components are functions of the displacement eld


ux; y; z, we can derive a dierent strain eld for each of the two

146

Energy methods in structural mechanics

hypothetical solutions of the elastic problem at hand


quA
x
qx

 A

1 quy
quA
x

2 qx
qy

quB
x
qx

 B

1 quy
quBx

2 qx
qy


"A
xx
"A
xy
"Bxx
"Bxy

5:7

5:8

Finally, as the principle of superposition of eects holds true, we can


write the principle of conservation of energy with reference to the difference between the solutions A and B of the problem. This dierence
is characterised by the strain eld
B
"ij "A
ij "ij

5:9

and by the total zero sum of external actions, that is


Pi Pi 0

5:10

The principle of conservation of energy, equation (1.9), can be therefore written as


E U W 0

5:11

and requires U 0, because of the zero value of the external work,


W 0.
But, on account of equation (5.6), we have

E
B 2
A
B 2
U
1 "A
xx "xx "yy "yy
21 1 2
V

"A
zz

"Bzz 2

B
A
B
2  "A
xx "xx "yy "yy

B
A
B
A
B
A
B
"A
xx "xx "zz "zz "yy "yy "zz "zz


21 2 A
B 2
A
B 2
"xy "Bxy 2 "A

"

"

"

dV
xz
xz
yz
yz
1 
5:12

General principles in the analysis of linear elastic structures

147

B
where U is zero only if "A
ij "ij . With reference to the relationships in
equations (5.7) and (5.8), this means that the displacement elds
uA x; y; z and uB x; y; z, which represent two dierent solutions of
the problem of the elastic equilibrium, must dier at most by a constant, corresponding to a rigid displacement of the whole body. Therefore, if, as is the usual condition, the geometrical constraints are such
as to prevent any rigid displacements, the two solutions A and B must
coincide and the uniqueness of the solution is demonstrated.

5.2

Reciprocal theorems in the linear theory of


elasticity

Apart from the self-consistency of the theory, from a technical point of


view the success of the mathematical theory of linear elasticity lies
undoubtedly on the extensive availability of powerful and practical
tools for the analysis of engineering structures. Among these tools a
very important place is occupied by the so-called reciprocal theorems.
In order to introduce these theorems, let us consider the simple supported beam shown in Fig. 5.5, which we consider to be loaded in
sequence rst by the group of forces P1 and then by the group of
forces P2 .
The work done by all these forces is
 W11 W22 W12
W
5:13
where W11 is the component work from the group P1 , W22 is the component work from the group P2 (remember that for both these components of work Clapeyron's theorem holds true) and W12 is the so-called
displacement work of P1 , i.e. the work of P1 due to the displacement
P1

P1

Fig. 5.5

P2

148

Energy methods in structural mechanics


P2

P2

P1

Fig. 5.6

eld produced by P2 . Sometimes the component work, i.e. the work


done by a force along its own deformation, is also called eigenwork
(the word eigen is a German word and means `own').
If we reverse the order of loading by rst applying P2 and then P1 ,
we have, as shown in Fig. 5.6,
 W W W
W
5:14
22

11

21

where W21 is the displacement work of P2 .


Of course, since the principle of superposition of the eects holds
true, the nal conguration of the beam will be identical in both
cases and we have from the principle of conservation of energy

 W
U W
5:15
By comparison of equations (5.13) and (5.14) we are led to the conclusion that
W12 W21

5:16

This is called Betti's theorem, which can be stated as follows:


Given two forces or two groups of forces acting on a linearly elastic structure, the work
done by the rst force, or group of forces, on account of the displacements due to the
second force is equal to the work done by the second force, or group of forces, on
account of the displacements due to the rst force.

A second theorem, due to Maxwell, can be straightforwardly derived


by Betti's theorem. It states that for the same load applied at dierent
points A and B we must have the condition
uAB uBA

5:17

General principles in the analysis of linear elastic structures

149

where uAB is the generalised displacement 2 at A due to the force at B


and uBA is the generalised displacement at B due to the same force at
A. In fact, if P is a generic load applied both at A and B, we have
from equation (5.16)
WAB WBA

P uAB P uBA

5:18

and we immediately get the relationship (equation (5.17)).

Here as generalised displacements we intend the components of displacement enabling


the applied forces to perform work. In this sense we can introduce generalised forces as
well. Thus, generalised forces and corresponding generalised displacements (or vice
versa) are quantities characterised by the fact that the sum of their products always
represents a true work.

6.

Total potential energy

Introduction

Up to this point we have concentrated on the development of methods


that enable the solution for systems of rigid bodies and for beam deformations in a closed form. The latter is based on Euler's method to
derive the dierential equation of equilibrium. In some circumstances
of restricted forms of loading and boundary conditions the dierential
equation may admit a solution in a closed form. In general, however,
it has been found dicult to obtain an exact solution for several of
the practical ranges of beam geometry and loading types. For this
reason we often have to turn to approximate methods of analysis to
provide information for practical designs in structural engineering.
The foundation of the energy principles that underlie many of
the approximate methods developed in later chapters is introduced
here.

6.1

The principle of the stationary value of total


potential energy

At this point in the presentation we can introduce what is usually called


the principle of stationary value of total potential energy. From a theoretical point of view, the reader will nd this principle to be nothing
more than a direct and formal derivation of what has previously
been presented in chapter 1. Nonetheless, this principle has a broad
eld of applications in mechanics and is central to methods to obtain
solutions to several problems in structural mechanics. Particularly,
the principle provides the basis for establishing approximate solutions
by means of trial functions, that, in turn, lead directly to the nite
element method of analysis.
With reference to a closed mechanical system in equilibrium and in
the absence of heat exchange, we have from the principle of conservation of energy, equation (1.21),
dE dW

152

Energy methods in structural mechanics

and for a virtual displacement eld, i.e. a eld of displacements which


are innitesimal and compatible with the constraints on the system,
but are imaginary and do not actually occur
E W

6:1

From this equation we derived, see equation (1.28)


U W

K 0

as a condition as well as a test of equilibrium, where, of course, K is


the kinetic energy of the system.
We are now going to show that it is straightforward to build a special
functional whose variation within the eld of kinematically admissible
functions coincides with equation (1.28) and is therefore stationary
when the system is in equilibrium. In other words, if we write
UW

6:2

we have
 0

U W 0

6:3

Therefore, if we call the functional  the total potential energy of the


system, we can enunciate the principle of stationary value of total
potential energy
Given a closed mechanical system, isolated from its environment such that no exchange
of heat energy occurs across the system boundary, the stationary values of the total
potential energy represent congurations of equilibrium for the system itself.

This holds true from a mathematical standpoint. From a physical


point of view the principle of stationary value of total potential
energy can be claried with reference to a very simple pictorial representation. Thus, let us consider Fig. 6.1, where a landscape of hills and
valleys is represented.
Our mechanical system is simply simulated by a ball subjected to a
gravitational eld. Everywhere in the landscape the ball is therefore
subjected to a vertical force P having the following constant value
P mg

6:4

where m is the mass of the ball and g is the gravitational acceleration.


The force P is directed downwards into the plane. If we suppose the
landscape surface to be perfectly smooth and frictionless, we can
assume that the reaction of the surface can only act normally to the
surface itself. As no strain energy U can be stored in the system, the
condition of equilibrium, equation (1.28), becomes
W 0

6:5

Total potential energy

153

(s)
E
D

A
B
C

h (s)

Fig. 6.1

where the work W is performed only by the force acting on the system,
i.e. the gravitational force P. The reaction R of the supporting surface
cannot perform work, as by its denition its direction is orthogonal to
any virtual displacement, which, by denition, must be tangential to
the landscape surface.
Since the gravitational force P is directed downwards, its virtual
work has value only in the case where a virtual displacement possesses
a vertical component. We are consequently led to the self-evident conclusion that the ball in the landscape represented in Fig. 6.1 cannot be
in equilibrium except at the top of a hill, the bottom of a valley or at a
point of inection.
In order to express the condition of equilibrium (equation (1.28))
with reference to the total potential energy of the system, we can write
s Phs

6:6

where h(s) is the height of the landscape path, described by the parameter s. The condition of stationary value of the total potential
energy is
s P hs P  us W 0

6:7

and coincides with zero value for the virtual work performed by the
gravity force P, as shown in Fig. 6.2. It is worth noting that the
height function h(s) can be measured with reference to some base
level C. We already know that the value C of the constant cannot
aect the equilibrium, as we are interested only in variations.
From a physical point of view this means that only relative heights
are relevant and it does not matter if the whole landscape is raised of a
constant quantity everywhere.

154

Energy methods in structural mechanics


u
u

h (s)

P
P
Equilibrium

No equilibrium

Fig. 6.2

In equation (6.6) the total potential energy of the system coincides


with the load potential Phs, which represents the capability of
the gravitational force P to perform work on account of the height
hs of the ball. If no work tends to be performed by the gravitational
force, the principle of conservation of energy assures that no kinetic
energy (which by denition of this scheme is the only form of energy
that the ball can possess on account of its mass) can be gained by
the system and it is in equilibrium. The question as to whether the equilibrium state is stable or not is considered in chapter 12, it suces here
to note that the conditions and examples relative to structural systems
presented in this book are actually stable if not otherwise stated.

6.2

Two important remarks regarding the principle of


stationary value of total potential energy

Immediately after the introduction of the principle of stationary value


of total potential energy, it is valuable to review two important aspects
of the principle
(a)

(b)

The principle of the stationary value of total potential energy


is a formal representation of a state equation (i.e. the condition of equilibrium, equation (1.28)) by means of extremum
conditions on a specically designed functional. It is a mathematical device and, despite the fact that we can often give a
pictorial representation of the terms of this functional, there
is actually no need for it to have a precise physical meaning.
As a generality, not all systems can have a total potential
energy functional. Such a functional is a property of socalled conservative systems, i.e. those system in which the
work, equation (1.4) performed by the forces acting on the
system depends only on the end points and not upon the
path joining them.
However, in the theory developed in this book, all the structural systems taken into consideration are conservative by
denition. Unless otherwise explicitly stated, we consider

Total potential energy

155

only closed mechanical system in which there are no sources or


sinks of energy and the processes of deformation of bodies are
totally reversible without transformation of mechanical work
into heat and vice versa. Under these circumstances the existence of relevant potential functionals is always assured.
Of course, once again the reader has be aware that no such structures actually exist in nature; but in engineering we deal with models
of real situations. The rules of behaviour for these models are set by
ourselves, and from an engineering point of view we are satised if
they produce, as they do, results adequately accurate for our needs
of design and safety in practical applications. The satisfactory nature
of the models has to be obtained by comparing the predictions of structural behaviour from the models with corresponding results from tests.

6.3

CotterillCastigliano's rst theorem

This section introduces one of the most straightforward developments


from the principle of stationary value of total potential energy, i.e.
CotterillCastigliano's rst theorem1;2 on the derivatives of the strain
energy function U. By means of this theorem it is possible to evaluate
the loads acting on an elastic system in equilibrium by means of a function of its strain energy. This is the rst of many practically valuable
methods derived from the principle of stationary value of total potential enegy and others are introduced in following sections.
The development is illustrated by taking into consideration the cantilever shown in Fig. 6.3. The total potential energy of this system is given
by
uy Uuy P

6:8

and is a functional of the eld of vertical displacements uy z, according to Bernoulli's model of beam. As we have seen in chapter 4, the
strain energy functional Uuy can be expressed as
l
1
EIx u002
Uuy
y dz
2

6:9

Also, the displacement eld uy z is a function of the deection at the


end , and the expression for this function can be easily obtained by
requiring satisfaction of the following geometrical and mechanical

1
2

Cotterill, James Henry (Norfolk, 1836 Parkstone, Dorset, 1922), English engineer.
Castigliano, Carlo Alberto (Asti, 1847 Milan, 1884), Italian engineer.

156

Energy methods in structural mechanics

P
z
A

(B)

Fig. 6.3

constraints to the general integral of Euler's equation of beam deformation


uy 0 0

and 0 u0y 0 0

Ml EIx u y00l 0

and uy l 

6:10
6:11

Remember that the general integral of Euler's equation of beam deformation in the absence of point loads along the beam is the third-order
expression
uy z Az3 Bz2 Cz D

6:12

as can be easily deduced from equation (4.44). The required expression


for the displacement eld uy z as a function of the tip deection  of
the cantilever is therefore

3
uy z 3 z3 2 z2
6:13
2l
2l
By making use of equation (6.13), the functional for the total potential
energy, equation (6.8), can be rewritten as a simple function of the
deection 
 U P

6:14

The condition of extremum condition for this function is, of course,


d dU

P0
6:15
d d
and allows us to state mathematically CotterillCastigliano's rst
theorem:
dU
P
d

6:16

Total potential energy

157

Therefore, in order to evaluate the force P that produces a specic


deection  at the end of the cantilever, according to equation
(6.16) all we need to do is to substitute the second derivative of equation (6.13) with respect to z in equation (6.9) and successively take the
rst derivative of the resulting function with respect to . Thus, we
have


l
1
3
3 2
U
EIx 3 z 2 dz
2
l
l

6:17

and, after a little of algebra, we obtain


dU 3EIx 

d
l3

6:18

which provides the appropriate value of the force P.


Of course CotterillCastigliano's rst theorem has a much broader
eld of application than a simple cantilever subjected to one point
load. We can illustrate this by considering an elastic structure loaded
by n generalised forces Pk k 1; . . . ; n. From equation (6.2) we have
U

n
X

Pk k

6:19

k1

where k are the generalised displacements corresponding to the


generalised forces Pk .
Provided we can express the strain energy U as a function of these
generalised displacements, we have
n
X
1 ; . . . ; n U1 ; . . . ; n
Pk  k
6:20
k1

and the n conditions for the total potential energy  to be stationary


are
qU
Pk
qk

6:21

CotterillCastigliano's rst theorem can be consequently stated in


complete generality as follows
The partial derivative of the strain energy of an elastic system with respect to a
generalised displacement at a certain point gives the value of the corresponding
generalised force at the same point.

It is worth noticing that the CotterillCastigliano rst theorem, as well


as the principle of stationary value of total potential energy, does not

158

Energy methods in structural mechanics

require any hypothesis with regard to innitesimal displacements or


the linearity of the elastic forcedisplacement relationships.

6.4

CotterillCastigliano's second theorem in the


linear theory of elasticity

Even if they have no fundamental basis in energy principles, there are


several theorems in structural mechanics that can be stated in the linear
theory of elasticity by means of purely mathematical relationships.
This is the case of CotterillCastigliano's second theorem, which is
formally a dual relationship of CotterillCastigliano's rst theorem.
This second theorem requires that the problem at hand is studied
within the framework of the theory of innitesimal displacements
and restricted to linearly elastic forcedisplacement relationships. In
fact, only in such circumstances can we make reference to the principle
of superposition of eects.
If we consider an elastic structure loaded by n generalised forces Pk
(k 1; . . . ; n), we can divide the loading into two groups. The rst
group is composed only of the force Pi , and the second group contains
all the remaining forces, Pj j 1; . . . ; i 1; i 1; . . . ; n.
Then the strain energy U of the structure is
U WI WI;II WII 12 Pi i;I Pi i;II WII

6:22

where WI is the component of work of the force Pi , i.e. the work done
by the force Pi along the line of the generalised displacement i;I and
due to the application of Pi only; WI;II is the displacement work done
by the force Pi along the generalised displacement i;II due to the
second group of loading; and WII is the component of work of the
second group of loading, Pj , on account of the generalised displacements due to the second group of loading itself.
As all the forcedisplacement relationships are linear, we can write
P
i;I i
6:23
K
where K is a stiness parameter. Therefore we can express the work WI
as
1
1 P
WI Pi i;I Pi i
2
2 K
and we get

6:24

1 P
U Pi i Pi i;II WII
6:25
2 K
At this point, provided we can express the strain energy U as a function
of the n generalised forces Pk k 1; . . . ; n, we can write down the

Total potential energy

159

partial derivative of equation (6.25) with respect to Pi as


qU Pi
i;II
qPi K

6:26

Finally, taking into account equation (6.23) once again, we have


qU
i;I i;II i
qPi

6:27

This relationship constitutes the formal expression of Cotterill


Castigliano's second theorem. This theorem can be stated as follows
The partial derivative of the strain energy of an elastic system with respect to a
generalised force at a certain point gives the value of the corresponding generalised
displacement at the same point.

By the manner in which we have obtained this result, it is clear that


it holds true only in the framework of the linear theory of elasticity
and from a physical point of view it does not rely upon a real energy
principle.
In conclusion it is instructive to apply CotterillCastigliano's second
theorem to the cantilever of Fig. 6.3 in order to nd the deection 
due to the load P at B. In order to do so we have to express the
strain energy U of the cantilever as a function of the force P.
As we know that
EIx u00y Mz

6:28

we have from equation (6.9)


1
U
2

l
0

M 2 z
dz
EIx

6:29

Making reference to Fig. 6.4, it is clear that


Mz Pl z

6:30

and we can write


1
UP
2

l
0

P2 l z2
dz
EIx

6:31

performing the required mathematical operations, we have


UP

P 2l 3
6EIx

6:32

160

Energy methods in structural mechanics

M (z) = P (l z)
lt

B
z

y
l

Fig. 6.4

and from CotterillCastigliano's second theorem, equation (6.27)




6.5

dU
Pl 3

dP 3EIx

6:33

Numerical example: frame analysis by


CotterillCastigliano's theorem

A very simple frame is shown in Fig. 6.5 to which a vertical load P


is applied at mid-span of the horizontal member. We consider the
elements to be axially undeformable.
This is called a portal frame and the horizontal and vertical members
have cross sections shown in Fig. 6.5(b), respectively. Given that the
members are made from a steel with a yield strain of "0 0:12%
and a modulus of E 2:1  105 N=mm2 , we wish to calculate the
maximum value of the load P that can be applied to the frame such
that the maximum strain does not exceed 0:7"0 .
The rst step in this example is to calculate the distribution of
moments along each of the members for a general portal frame,
shown in Fig. 6.6.
The free-body diagram is shown in Fig. 6.7. The symmetry conditions for the forces at C results in
Vc 0
Equilibrium of the forces at sections C and A gives
P
and
HC HA
VA
2

6:34
6:35

Total potential energy


P

25 m

161

25 m

1
B
1

5m

A
(a)

10 mm
X

X
125 mm

10 mm
100 mm

8 mm

12 mm

150 mm

200 mm

Section 2 2

Section 1 1
(b)

Fig. 6.5

The variation of bending moments along the horizontal and vertical


members, i.e. zones 1 and 2, respectively, are
Zone 1
M1 z1 MC

P
z
2 1

6:36

162

Energy methods in structural mechanics


P

l/2

l/2

EI1

EI2

Fig. 6.6

l/2
P/2

MC

Zone 1

HC

z1

z2

VC

Zone 2

HA
MA
VA

Fig. 6.7

Total potential energy

163

Zone 2
M2 z2 MC

Pl
HC z2
4

6:36b

where, of course, at the present stage MC and HC are unknowns.


The strain energy U is
1
U
2EI1

l=2

1 6
4
2EI1

l
1

M22 z2 dz2
2EI2
0

M12 z1 dz1

l=2

7
M12 z1 dz1 I M22 z2 dz2 5

6:37

where I  I1 =I2 .
Substituting equation (6.36) into equation (6.37) and after integration the strain energy is obtained. This is

1
l
l2
l3
U
MC2 MC P P2
2EI1
2
8
96

3
3 
l2
l3
2
2 l
2
2 l
MC HC l PHC HC
I MC l MC P P
2
16
4
3
The deformation conditions for the frame are, because of symmetry
Zero rotation at the position C, i.e. C 0
Zero horizontal displacement at position C, i.e. wC 0
Thus from the second of CotterillCastigliano's theorems,
C

qU
0
qMC

Hence, from the strain energy





qU
1 1 2
1
Pl MC l I 2MC l Pl 2 HC l 2

0
qMC 2EI1 8
2
6:38
For the horizontal deection at C, it is
wC

qU
0
qHC

164

Energy methods in structural mechanics

Therefore,

 

qU
1
2
1 3
3
2

I
0
H l MC l Pl
qHC 2EI1
3 C
4

6:39

From equations (6.38) and (6.39)


HC

3P
82 I

MC

I 1
Pl
42 I

6:40

The moments along the members can be calculated using equations


(6.40) and (6.36), thus
1
1
Pl
MB
Pl
MA
82 I
42 I
The variation of the moment is linear between these values as shown
in Fig. 6.8.
The maximum strain due to the applied load P can occur at position
B in member AB or at position C in member BC and rst it is necessary
to evaluate the moments at these positions.
The second moments of area of members AB and BC about their XX
axes, i.e. I2 and I1 , respectively, are
I1

1
12

1
 125  2243 12
 115  2003 4:04  107 mm4

I2

1
12

1
 100  1703 12
 92  1503 1:51  107 mm4
1

MB = 4 (I + 2) Pl
B
1
MB = 4 (I + 2)

Pl
B

l/3

A M = 1 Pl
A
8 (2 + I)

Fig. 6.8

C
(I + 1)

MC = 4 (I + 2) Pl

Total potential energy

165

hence
I

I1
2:68
I2

Taking rst the condition at position B for member AB, the moment is
MB 0:0534Pl 267P
the strain " is
"

My
EI

where y is the distance from the XX axis to the outer surface of the
ange. Thus,
267  P  85
" :
 0:7  1:2  103
2 1  105  1:51  107
The maximum allowable value of the load, P, if the strain in member
AB at position B is not to exceed the specied level is
P 11:7  104 N
Considering now the strain at position C on the member BC,
983  P  112
 0:7  1:2  103
" :
5
7
:
2 1  10  4 04  10
hence
P 6:47  104 N
It is evident that the lesser of the loads calculated above is that which
governs the permissible loading condition.
The vertical deection of the member BC at the point of application
of the load P, i.e. vc , can be calculated from the strain energy using
again CotterillCastigliano's second theorem
vC

qU
qP=2




qU
1 1
1 3
1 3 1
1
2
2
3

M l Pl I Pl MC l HC l
qP=2 2EI1 8 C
48
8
2
4
Hence, the vertical deection at C is
vC

1 2I Pl 3
2 I 96EI1

166

Energy methods in structural mechanics

Thus
vC

1 2  2:68
6:47  104  50003
13:49 mm
2 2:68 96  2:1  105  4:04  107

The deection is generally assessed on the basis of a proportion of the


length of the structural member, thus
vC
1

370
l
This result can be considered acceptable as codied guidelines give
admissible levels of deection as
vC
1

180
l

Suggested exercise problems


1.

An arch is shown in Fig. 6.9; it has a mean radius R, is pinned at A


and C. A horizontal load P is applied at C. It is free to move horizontally at C. The cross section of the arch is rectangular as shown
in Fig. 6.9. Develop expressions for the maximum moment, Mmax
the vertical deection at B, vB and the horizontal deection at C,
hC .
[Ans: Mmax 0:311PR, vB 1:73PR3 =Ebd 3 ,
hC 8:66PR3 =Ebd 3 ]
B
d

R
X

d
b
Section X X

Fig. 6.9

Total potential energy


P

167

A
B

D
2P cos

Fig. 6.10

2.

3.

A section of pipe is subjected to continuous line loads. The section


through the pipe is shown in Fig. 6.10, where the loads P are the
value of the line load per unit length of the pipe. Treating the pipe
as a plane strain problem, develop an expression for the maximum
moment Mmax in the ring for  608, and the position of the
maximum moment.
[Ans: Mmax 0:091PR at 1208 from A]
A rigid jointed frame is connected to a rigid foundation, as shown
in Fig. 6.11. The members of the frame are made from tube with
outside diameter D, and wall thickness t. Develop an expression
for the allowable load qa such that the maximum strain in the

q
D

t
l

60

Fig. 6.11

168

Energy methods in structural mechanics


10 m

10 m
B

10 kN
C

1m

05 m

Fig. 6.12

4.

frame is not to exceed the material yield strain "0 of the tube
material.
[Ans: qa ED2 t"0 =16:8l 2 ]
A diving board consists of a beam attached to a frame, as shown in
Fig. 6.12. The board is to be proof tested by applying a 10 kN load
at the end C. Analysing the frame as being pin-jointed, calculate
the vertical and horizontal deections at C, hC , vC , that
would be expected to occur during the proof test. The beam has
a second moment of area of 5  104 mm4 and the bars have a
cross-sectional area 200 mm2 . E 200 kN/mm2 .
[Ans: hC 0:13 mm, vC 0:87 mm]

7.

The method of trial functions

Introduction

It is evident from the development of energy principles in structural


analysis and the examples presented in chapter 6 that these methods
can be very powerful in developing solutions to practical situations
in structural mechanics. The applicability of the methods, however,
depends on having available information, such as the variation in
the deection or of the moment of the beam or the frame being analysed. In many practical circumstances the precise variation of these
parameters is not available, but it would be very valuable if the
energy methods could still be applied. In this chapter we shall show
how this can be achieved.
It is appreciated that the solutions obtained by the use of approximate methods are themselves approximations to the exact situation
of deformation or forces in the structural model; but in any case in
engineering the prediction of structural behaviour will always be
approximate to some extent since knowledge on material properties,
loading and construction dimensions is never exact and, consequently, precisely represented in the mathematical model. Therefore,
provided we have some exact solutions available, usually related to
idealised structural forms and loading, with which to calibrate and
validate the approximate methods, the use of energy methods with
approximate variations in deections and internal forces is a very
powerful approach that underlies current structural engineering
practice. This practice makes considerable use of approximate
energy methods, usually in the form of the nite element method,
that are available in commercially produced computer programs.
The present chapter outlines the method of trial functions, outlines
how suitable trial functions may be identied and will enable the
reader to understand how the nite element method is eectively an
automated way of incorporating special trial functions in structural
analysis.

170

7.1

Energy methods in structural mechanics

The basis of the method of trial functions

From the outset we must be aware that the method of trial functions is
an approximate approach that enables engineers to obtain predictions
for the response of complex structures to practical loading where no
exact solution to the behaviour of the structural model can be
obtained. The degree of approximation and how to estimate this is
considered in the following text.
So far we have dealt with basic ideas in the theory of elastic structures and in particular with the treatment of the Bernoulli model for
a beam. This is a particularly simple model of the behaviour of a
one-dimensional structure, with an approximation to the strain distribution that is quite satisfactory from an engineering point of view. As
we have seen in chapter 4, the general expression for the displacement
eld of a generic beam with constant bending stiness EIx is yielded by
the fourth-order dierential equation (4.43).
However, even in the simple case in which we need to take into consideration a straight beam whose section is variable along the axis and
subjected to a distributed load qz, the value of which is dependent on
z, we have seen in section 4.6 that we are led to a dierent and more
complicated equation, equation (4.86), that can be also written as


d2 uy z
d2
EIx z
qz
7:1
dz2
dz2
The solution of this equation cannot now be expressed as a fourthorder polynomial as was the case for equation (4.44), and the possibility of obtaining a closed form is strictly dependent on the particular
expressions of the functions EIx zand q(z) for the problem at hand.
This in the general case is not possible.
Thus even this very simple case clearly illustrates the necessity for
establishing an approximate method of solution. As was pointed out
in chapter 6, the ability to provide approximate solutions is one of
the most important applications of the principle of stationary value
of total potential energy, i.e. determining approximate solutions to
the problem of elastic equilibrium. This important application of the
principle of stationary potential energy is made by assuming equations
(6.2) and (6.3) as starting points and making use of trial functions. This
application is described below.
The trial function is essentially a possible representation of the
unknown quantity that has been used to describe the total potential
energy functional; in general this quantity is the displacement eld of
the structure or its elements. Under some circumstances, the quantity
is also called the shape or coordinate function.

The method of trial functions

171

In order to clarify our ideas, let us consider the simply supported


beam already studied in chapter 4 and depicted in Fig. 4.10. We
found that the exact solution to this problem is given by equation (4.51)
q
uy z
z4 2lz3 l 3 z
24EIx
and is a fourth-order polynomial expression. The conguration of
equilibrium of this beam can also be described in terms of the total
potential energy functional, i.e.
uy 0

7:2

where the potential energy uy according to equation (6.2), is given by


l
l
1
002
uy
EIx uy dz quy dz
2
0

7:3

This expression is clearly a functional (i.e. a function of a function) of


the lateral displacement uy z of the axis of the beam. If treated with
the usual methods of the calculus of variations, the condition in equation (7.2) would lead us to the Euler equation of beam deformation, see
equation (4.43), together with the boundary conditions
uy 0

for

z 0 and

Mz EIx u00y 0

for

zl

7:4

z0

and z l

7:5

However, let us now imagine that there is actually no simple solution


to this structural model posed by equations (4.43), (7.4) and (7.5). In
this circumstance we could use our engineering experience of the displacement of beams subjected to lateral loading and guess a suitable
displacement shape that could act as a trial displacement eld. The
analysis process is initiated by substituting the trial displacement eld
uy z in equation (7.3). In order to be admissible, the trial displacement
eld must satisfy the geometrical boundary conditions in equation
(7.4). We may choose, i.e. guess, a simple displacement eld which is
dened in shape but not in amplitude, for example
z
uy a sin
7:6
l
as is shown in Fig. 7.1.
Substituting the trial function, equation (7.6), into the potential
energy function, equation (7.3), we have
l
l

z002
z
1

EIx a sin
dz
7:7
uy
dz qa sin
2
l
l
0

172

Energy methods in structural mechanics


q

uy* = a sin

z
l

Fig. 7.1

This process achieves a very important simplication in the analysis


we have reduced the functional of the total potential energy to a simple function of the
parameter a, which regulates the amplitude of the shape function, equation (7.6).

This means that the condition of equilibrium (equation (7.2)) can be


satised by minimising a simple function rather than a functional,
i.e. the integral of the function of the parameter a. This is a situation
very much more amenable to algebraic manipulation than is the
integral equation for equilibrium, equation (7.2).
Thus in the present case we have
l
l

z002
z
1
a
dz
EIx a sin
dz qa sin
2
l
l
0

l
l
4
 
z
1
2
2 z

dz qa sin
dz
EIx a 4 sin
2
l
l
l
0

4
ql
EIx a2 2 a
3

4l

7:8

and the condition of equilibrium, equation (7.2), yields


da
0
da

a4

ql 4
EIx 5

7:9

This relationship denes the value of the parameter a in terms of


the applied load and the geometric and elastic material parameters
of the beam and is the amplitude of the trial displacement eld uy .

The method of trial functions

173

We have thus obtained the solution of our problem without making


reference to the calculus of variation or solving dierential equations.
Of course, we must bear in mind that this is only an approximate
solution for the structural model of the beam; nevertheless, it may
furnish surprisingly good results. For example, if we compare the
vertical deection at the middle of the span obtained by the procedure
of trial function to the exact solution, given by equation (4.51) for
z l=2, we have
ql 4
uy l=2 0:0130208
EIx

(exact solution)

ql 4
uy l=2 0:0130711
EIx

(trial function)

7:10

That is the guessed trial function predicts a deection at the mid-span


of the beam with an error of 0.39% compared to the exact value.
With regard to the bending moment we have,
Ml=2 EIx u00y l=2 0:025000ql 2
M  l=2 EIx uy 00 l=2 0:025801ql 2

7:10b

That is, the approximate method overestimates the maximum


moments with an error of 3.2%. Of course this is a very simple
example and a fortunate case, due to the fact that the shape function,
equation (7.6), happens to t the exact displacement eld very well and
it is therefore a good choice. Generally, engineers through their observations have a good facility to describe the probable shape of the
deformations that will occur in their structure when the loading is
applied. This facility is an excellent basis for `guessing' an appropriate
trial function for the energy method described above. However, in
general, we may not always be so fortunate in our choice of trial function as was the case for the example analysis described, and we should
pay a good deal of attention to the procedure that is explained in the
following section.

7.2

The method of RayleighRitz

In the previous section we have taken a rst approach to the method of


trial function for the solution of the problem of elastic equilibrium in
the case of a simply supported beam. We now generalise the method by
presenting a more comprehensive procedure that can be developed
from the simple idea introduced above: this is called the method of

174

Energy methods in structural mechanics

RayleighRitz.1;2 Section 7.1 has shown that the goal of the method of
trial functions consists of commuting the minimisation of a functional,
which requires a variational treatment and leads to a dierential
formulation, to the minimisation of a function, which is achieved by
means of elementary derivatives and leads to an algebraic problem.
Algebraic formulations are very much easier to solve than ones involving functionals.
The RayleighRitz procedure consists of the following steps
(a)
(b)

(c)
(d)

Write the functional  of the total potential energy of the


system.
Choose a trial function for the displacement eld; the function
must satisfy the geometrical boundary conditions (in the case
of a beam those relative to deection and slope) but it may or
may not satisfy the mechanical boundary conditions (i.e.
regarding moment and shear). This shape function must be
dependent on a nite number of parameters a1 ; . . . ; an .
Insert the trial function in the potential energy functional 
and integrate over the dimension of the structure.
Minimise the obtained function a1 ; . . . ; an with respect to
the parameters a1 ; . . . ; an . We now have a linear system of n
equations in the n unknowns a1 ; . . . ; an given by
q
q

0
qa1
qan

(e)

The solution of this system of equations denes the actual


value of the shape function and constitutes the approximate
solution of the problem of elastic equilibrium at hand.

Actually, the outlined procedure retraces the steps of the example


already given in the previous section, with the dierence that we can
now have a trial function depending on several parameters or degrees
of freedom, also termed generalised coordinates.
The procedure outlined above is exemplied now by the case of a
beam encastre at one end and simply supported at the other, as
shown in Fig. 7.2. We consider this beam subjected to a distributed
load q of constant value, which will enable us to compare the results
given by the RayleighRitz method with the exact solution of the
model.

1
Lord Rayleigh (Strutt, John William) (Langford Grove, Essex, 1842 Witham, Essex,
1919), English physicist.
2
Ritz, Walther (Sion, Valais, 1878 Gottingen, 1909), Swiss physicist.

The method of trial functions

175

B
l

Fig. 7.2

For the sake of clarity of presentation, we will proceed step by step


following the sequence presented above.
(a) The functional of the total potential energy of the system is
l
l
1
002
uy
EIx uy dz qu dz
2
0

7:11

(b) Select a shape function with several degrees of freedom, for example we choose here ve variables of the type
 
 
z
z

b1 cos
uy a0 a1 sin
l
l




2z
2z
b2 cos
7:12
a2 sin
l
l
where the variables are a0 , a1 , b1 , a2 , b2 .
This shape function must satisfy the geometrical boundary conditions at A and B in Fig. 7.2, i.e.
uy 0 and u0y 0
uy z 0

for

z0

for z l

7:13
7:14

As we have

 
 

z

z
b1 sin
uy 0 a1 cos
l
l
l
l
 
 
2
2z
2
2z
b2 sin
a2 cos
l
l
l
l

7:15

the geometrical conditions (7.13) and (7.14) lead to the system


8 
>
< uy 0 a0 b1 b2 0
uy l a0 b1 b2 0
7:16
>
: 0
uy 0 a1 2a2 0

176

Energy methods in structural mechanics

This system allows us, for example, to express a0 , a1 and b1 as functions


of a2 and b2 and we are left with a shape function that is dened by two
independent degrees of freedom only. That is
ao b2
a1 2a2

7:17

b1 0
and consequently the suitable trial function is

 

 
  
2z
2z
z

1 a2 sin
2 sin
uy b2 cos
l
l
l

7:18

(c) We substitute the trial function, equation (7.18), into the energy
functional, equation (7.11) and integrate over the length l of the
beam
uy

l
l
1
 00 2

EIx uy dz qu dz
2
0

 
  2
l
1
4
2z

EIx b2 2 cos
2
l
l
0

 
 2
42
2z
22
z
a2 2 sin
2 sin
dz
l
l
l
l



 
l  
2z
1
q b2 cos
l
0

a2 sin

2z
l


2 sin

z
l


dz

7:19

obtaining, once all the mathematical operations have been performed,


a2 ; b2

5EIx 4 2 4EIx 4 2 16EIx 3


4ql
a qlb2
 a2 3  b2
 a2 b2
 2
l5
l
3l 3
7:20

We have therefore accomplished the replacement of the functional,


equation (7.19) by the approximating function of the two parameters
a2 and b2 , shown in equation (7.20).

The method of trial functions

177

(d ) We minimise the obtained function a2 ; b2 with respect to the


parameters a2 and b2 , getting the linear system
q 10EIx 4
16EIx 3
4ql
0
3  a2
 b2
3
qa2

l
3l

7:21

q 16EIx 3
8EI

 a2 3 x 4 b2 ql 0
3
qb2
3l
l

This system of two simultaneous algebraic equations can also be stated


in matrix form, as

 


10
16=3 a2
4=
ql 4
4
 EIx 1
16=3
8
b2
(e) The solution of the system of equations (7.21) yields the values of
the parameters a2 and b2
1
15ql 4
EIx 3 16 452


ql 4
12
45
b2

EIx 2
162 454 128 3602
a2

7:22
7:23

Back substituting the results in equations (7.22) and (7.23) in the trial
function, equation (7.18) we have the expression for the approximate
solution of the problem of the elastic equilibrium for the beam
shown in Fig. 7.2.
It is instructive to compare the values of the deection, of the slope
and of the bending moment at the middle of the span furnished by the
two term trial function, equation (7.18), with those derived from the
exact solution, see equation (5.2). With regard to the deection at
the mid-span we have
ql 4
uy l=2 0:0052083
EIx

(exact solution)

uy l=2

(trial function)

ql 4
0:0043468
EIx

7:24

That is the trial function approximates the deection with an error of


16.5%. Regarding the slope we have
3

ql
l=2 u0y l=2 0:0052083
EIx
0

 l=2

uy 0 l=2

ql 3
0:0070998
EIx

(exact solution)
7:25
(trial function)

178

Energy methods in structural mechanics

That is, the trial function over-estimates the slope at the mid-span with
an absolute error of 36.3%. Finally, with regard to the bending
moment we have
Ml=2 EIx u00y l=2 0:0625000ql 2

(exact solution)

M  l=2 EIx uy 00 l=2 0:0634988ql 2

(trial function)

7:26

showing that the trial function provides a good estimate for the
moment at mid-span and has an error of 1.6%.
It is worth noting that, despite the fact that we have employed a trial
function with ve degrees of freedom, we have obtained an approximate deection that is 16.6% larger that the exact one, while in the
previous section we got an error of 0.39% using a trial function with
only one degree of freedom. Moreover, in the example shown above
the approximation for the slope is much worse than that for the deection and the approximation for the bending moment is far better than
those for the deection and the slope. This variability of the degree of
accuracy of the approximate results is due to the capability of the
particular chosen trial function to t the exact solution and its derivatives. This aspect of the method underlies much of the problems associated with approximating techniques in structural mechanics and will
be discussed to more detail in the next section.

7.3

The quality of the trial function

So far we have presented the results obtained for the beam shown in
Fig. 7.2 by means of the trial function described in equation (7.12) in
comparison to the exact solution given by the expression (equation
(5.2)). In order to make some simple observations about the quality
of the results yielded by the RayleighRitz method, let us take into
account another trial function, once again of the type
 
 
 
z
z
nz

uy a0 a1 sin
b1 cos
   an sin
l
l
l
 
nz
7:27
bn cos
l
It is natural to expect that the quality of the approximation tends
to increase along with the number of degrees of freedom since the
additional terms may be thought of as giving greater exibility to the
functions to meet the requirements for the minimum energy conditions. To check this hypothesis we assume a seven degrees of freedom

The method of trial functions

179

Table 7.1
N

uy l=2

 l=2

My l=2

5
7
exact

0:004347ql 4 =EIx
0:005097ql 4 =EIx
0:005208ql 4 =EIx

0:007100ql 3 =EIx
0:004711ql 3 =EIx
0:005208ql 3 =EIx

0:063499ql 2
0:059914ql 2
0:062500ql 2

trial function, i.e.


uy


 
 
z
z
2z
b1 cos
a2 sin
a0 a1 sin
l
l
l
 
 
 
2z
3z
3z
a3 sin
b3 cos
b2 cos
l
l
l


7:28

Following exactly the same procedure of the previous section we are


able to show in Table 7.1, the values of the deection, slope and bending moment at the middle of the span of the beam in Fig. 7.2. These
parameters are written as functions of the overall number N
N 2n 1 of degrees of freedom that comprises the trial function,
where the last row gives the exact analytical values, uy l=2, y l=2
and My l=2. For the sake of simplicity Table 7.1 can be also expressed
in terms of percentage dierences between the approximate solutions
and their exact counterpart, as shown in Table 7.2.
Table 7.2 allows us to make some interesting considerations about
the quality of the approximations corresponding to the dierent
values of N. First of all, let us x our attention on the deection. It
appears that taking into account two more degrees of freedom leads
to a better approximation to the exact value. However, as we have
already noticed before, in both these cases the approximation is
worse than in the case of the simple supported beam of the rst section,
where we employed only a single degree of freedom trial function!
We can therefore conclude that in the RayleighRitz method a
major role is played by the shape of the trial function chosen. If this
shape is very near to the exact solution, the quality of the obtainable
approximation tends to be satisfactory even with a small number of
degrees of freedom.
Table 7.2
N

uy uy %

y y %

My My %

5
7

16.54%
2.13%

36.32%
9.54%

1.60%
4.14%

180

Energy methods in structural mechanics


da (y0)
dz

a
y0

db (y0)
dz

y
z

Fig. 7.3

Secondly, let us consider the percentage dierences relating to the


bending moment. Here the situation is quite dierent because increasing the number of degrees of freedom does not lead to a better approximation. By means of a trial function with ve degrees of freedom we
have an error of 1.60% only, while by means of a trial function with
seven degrees of freedom the error changes sign and increases in absolute value to 4.14%. This behaviour can be easily explained as follows.
The fact that the trial function satises the geometrical boundary conditions and is generally near in shape to the exact solution does not
necessarily imply everywhere a good approximation to the mechanical
parameters (i.e. those concerning the internal forces), such as the
bending moment, which depends on the derivatives of the deected
shape.
In other words, as shown in Fig. 7.3, we can have trial functions
which do not dier very much from the exact curve on average, but
whose gradients can dier signicantly from the exact one at many
points.
As a consequence of this discourse it must be pointed out that
another source of error is introduced by the mechanical boundary conditions, that is the conditions relating to the forces at the boundaries. It
was emphasised earlier that the method of RayleighRitz does not
require the respect of the mechanical boundary conditions.
In the case at hand, i.e. the Bernoulli model of the beam, these conditions involve the derivatives of the displacement eld of the beam
axis. For example, it can be seen that every term in the chosen shape
function, equation (7.27) violates the mechanical condition Ml 0
at B. Actually, we have M  z EIx uy 00 z and the second derivative
of equation (7.27) is not zero at z l for a generic set of values of the
parameters a0 , a1 , b1 , . . . , an , bn . However, we could try to satisfy the

The method of trial functions

181

mechanical conditions together with the geometrical ones. Intuitively,


this should result in a better approximation but at the expense of
greater algebraic work. However, as we are going to see, this is not
necessarily the case.
In order to investigate this point, let us go back to equation (7.12)
and impose the requirement of satisfaction of the mechanical condition
M  EIx uy 00 0

for

zl

7:29

as well as the geometric requirements in equations (7.13) and (7.14).


Now we have the following system of four equations in the ve parameters a0 , a1 , b1 , a2 , b2
8 
u 0 a0 b1 b2 0
>
>
< y
uy l a0 b1 b2 0
7:30
0
0 a1 2a2 0
u
>
>
: y  00
uy l b1 4b2 0
which can be solved in terms of a0 , a1 , b1 , b2 . Thus, we get
a0 0
a1 2a2
b1 0

7:31

b2 0
As a consequence, the trial function becomes
 
  
2z
z

2 sin
uy a2 sin
l
l

7:32

If we substitute equation (7.32) in potential energy functional, equation (7.11) and integrate, as before, over the length l of the beam, we
obtain
5EIx 4 2 4ql
a
a2
7:33
 2
l3
It should be noted that by imposing an additional condition, i.e. the
moment requirement, another degree of freedom has been suppressed
and so the substitution of the approximating function, equation (7.32),
into the energy functional, equation (7.11) results in a function only of
the parameter, a2 .
We minimise the function a2 with respect to the parameter a2 ,
and get
a2

d 10EIx 4
4ql
0

a2
da2

l3

7:34

182

Energy methods in structural mechanics

which gives the value of the parameter a2 as


a2

2ql 4
5EIx 5

7:35

Finally, back substituting the result in equation (7.35) in the trial


function, equation (7.32) we have the solution corresponding to this
case in which all the geometrical and mechanical boundary conditions
are satised.
However, if we take into account for comparison the values of the
deection, slope and bending moment at the middle of the span, we
have to recognise that the approximations obtained, i.e.
4

ql
uy l=2 0:0026142
EIx
3

ql
 l=2 0:0082128
EIx

7:36
7:37

M  l=2 EIx uy 00 l=2 0:0258012ql 2


7:38
result in errors of 49.8%, 57.7% and 58.7%, respectively.
It is clear that these results are far worse than those in the case in
which, with the same trial function, only geometrical boundary conditions have been fullled.
Even if this is a limiting case (i.e. we have assumed too few degrees of
freedom), nevertheless it must be underlined that the error due to the
shape of the trial function can outweigh that due to non-compliance
with the boundary conditions. Indeed it is not inconceivable that a
trial function satisfying also the mechanical boundary conditions can
give inferior results compared with a function that satises the geometrical boundary conditions only but is generally nearer in shape to the
exact solution.
The conclusion from this assessment of the value of trial functions
is that one must choose the functions with great care and using as
guidance the maximum information on the probable deformations of
the structure. Even then it is possible that the results will not correspond closely with the exact solution. Nevertheless, even with one
degree of freedom trial function the method provides a quick initial
solution to structural problems that can be used as the basis for
later, more rened methods. However, it is evident that it would be
valuable if a systematic approach could be obtained to guide the
choice of trial functions. This is the subject of sections 7.5 and 7.6,
in which an automated approach is introduced by means of the concept of localised functions.

The method of trial functions

7.4

183

Numerical example: the method of trial functions


applied to a cantilever with linearly varying depth

Up to now all applications have been concerned with beams having


uniform cross sections and material properties that do not vary
along the beam. We now go back to an example of the type of equation
(7.1) and expand on this to illustrate the capability of the trial function
method to provide a formulation also for the deection of a cantilever
beam in which the depth of the section varies linearly along its length.
The analytical solution for this beam was presented in section 4.6, here
we use the method of trial functions. This method is such that a simple
function will always result from the algebra. However, as we have seen,
the accuracy of the prediction for the deformations may not always be
acceptable and the results must be analysed critically.
Therefore, let us now present in detail the analysis of the beam with
the tapered depth using the method of trial function. First, a one
degree of freedom function is used.
We recall that the natural boundary conditions are given by the
equations (4.96) and (4.101), i.e.
uy 0

and

duy
0
dz

at z 0

In terms of the non-dimensional variable   z=l, a one degree of


freedom trial function that satises these conditions is

 


7:39
uy  a1 1 cos
2
where a1 uy 1 is the deection of the tip of the cantilever beam.
The exural strain energy of the beam is
 2

 2

1
EIx z d uy z 2
EIx  d uy  2
U
dz

d
2
dz2
d 2
2l 3
l
0

7:40

and note again that the exural stiness EIx  must be contained
within the integral sign since for this beam the stiness varies along
the beam.
From equations (4.82) and (4.83) we can write the second moment of
area for the beam as
 3
bd1
Ix
1 1 3 Ix 1 1 3
7:41
12
where Ix is the second moment of area corresponding to the crosssection geometry of the cantilever at the built-in end and  d2 =d1 .

184

Energy methods in structural mechanics

The strain energy of exure can now be written as


 2 
1
d uy
E Ix
U 3 1 1 3
d
2l
d 2

7:42

Substituting the trial function in the strain energy gives



E Ix 2 2
 121 2 3
U 3
128
2l


3 2
3
 1 21 6 1 a21 7:43
16

The total potential energy  of the deformation of the cantilever comprises the strain energy due to the exure of the beam and the loss of
potential of the load P applied at the tip of the cantilever. Thus
 U Puy l U Pa1

7:44

The equilibrium condition is obtained from the stationary value of the


potential energy with respect to the degree of freedom a1 , that is
d
0
da1

7:45

Thus, the condition of stationarity of the potential function gives



E Ix 2 2
 121 2 3
128
l3

3 2
3
 1 21 6 1 a1 P 0
16
(7.46)
Substituting the relevant value for for a particular beam geometry in
this equilibrium equation yields the value for the deection at the tip of
the cantilever and the shape of the deection along the length of the
beam. Example values are
1

Pl 3
uy 1 :
3 02EI1

0:5

Pl 3
uy 1 :
1 95EI1

Pl 3
uy 1 :
1 31EI1

7:47

The method of trial functions

185

Fig. 7.4

It may be seen that in comparison with the exact values, obtained


from the solution of the dierential equation, the deection at the
tip of a cantilever with uniform depth 1 obtained by means of
the trial function provides a very good result, but that as the value
of is reduced the results from the approximate method diverge
from the exact value. This is shown graphically in Fig. 7.4 where
the value for the tip deection from the one degree of freedom trial
function is compared to the exact result for corresponding values
of .
The approximation is reasonable, i.e. to within 5%, up to a value of
0:5.
Again, we can try to improve the accuracy of the result from the
approximate method by increasing the number of degrees of freedom
in the trial function. This allows greater exibility to the trial function
to t the exact deected shape of the cantilever. The following illustrates this by using a two degree of freedom trial function, i.e.



 


3

uy  a1 1 cos
a2 1 cos
7:48
2
2
where a1 and a2 are the degrees of freedom, or generalised coordinates
of the trial function. The shapes of the functions corresponding to
these coordinates are shown in Fig. 7.5.
This function is substituted in the total potential energy functional
 and the equilibrium equations are obtained, as before, from the

186

Energy methods in structural mechanics


20
[1 cos /2]
[1 cos 3/2]
15

10

05

0
0

02

04

06

08

10

Fig. 7.5

stationary condition of the potential energy. In this case two equations


are obtained from the conditions
q
0;
qa1

q
0
qa2

7:49

The equilibrium equations can be described in matrix notation as




f1

f2

f3

f4



a1
a2

 
Pl 3 1
0

2E Ix 1

7:50

where f1 ; . . . ; f4 are functions only of the variable describing


the degree of taper in the beam. These functions are
f1

3 2
16  1

2 3 34 1 3 3 2 3

1 4
 1 2 3
64

f2

135 2
64  1

2
3
23 2 23 3 27
4 1 3 3

f3

135 2
64  1

2
3
23 2 23 3 27
4 1 3 3

f4

27 2
16  1

2 3 94 3 2 3 3

4
2
3
81
64  1

7:51

The method of trial functions

187

25
Exact solution
20

Two degrees of freedom


One degree of freedom

uy (1)

15

10

05

0
0

02

04

06

08

10

Fig. 7.6

The values for the generalised coordinates in the trial function are
simply calculated using


Pl 3 1 f2
a1
f1
2EI1
7:52


Pl 3
f3 f1
a2
2EI1 f2 f3 f4 f1
and the deection at the tip of the cantilever is obtained from
uy 1 a1 a2

7:53

Figure 7.6 shows the eect of including two degrees of freedom in


the trial function; the agreement to within 5% of the exact value for
the deection at the tip of the cantilever now extends in the range
0:2 > > 1.
If greater accuracy was required across the full range of the parameter it could be achieved by increasing the numbers of degrees of
freedom of the trial function. Of course, we underline again that this
would be at the expense of greatly increased algebraic complexity in
the analysis.

7.5

The systematic search for a trial function

On the basis of what we have noticed in the previous sections in this


chapter, it should be clear that in general terms to nd a trial function
is, in the rst instance, a matter of good judgement, experience and an

188

Energy methods in structural mechanics

appreciation of the physical meaning of the problem. Generally, engineers develop with practice a good `feel' for the deected shapes of
structures. However, one may wonder if there is a systematic way of
nding such a function and, more importantly, if this function tends
to converge to the exact solution along with the increase of the
number of degrees of freedom taken into consideration.
The answer is yes, but unfortunately this is a quite complicated
matter and cannot be reasoned on the basis of simple mathematics,
to which this book is restricted. We will simply state that in order to
establish some criteria of convergence for the trial function, in the
rst instance we must deal with a special set of approximation functions, also named coordinate functions.
These coordinate functions fi are characterised by the following
properties
.
.
.

each fi satises the geometrical boundary conditions


for any N, the elements f1 ; f2 ; . . . ; fN are linearly independent
for any solution g of the problem, we can always nd an approximation function a1 f1 a2 f2    aN fN which diers,
according to a certain measure, innitely little from g.

Under certain assumptions, it is possible to demonstrate that all the


choices of a set of coordinate functions ensure the convergence of
the RayleighRitz approximation to the exact solution in the limit.
However, it is clear that for a nite number of coordinate functions
fi , one choice of functions might give better accuracy over the another.
Moreover, from a technical point of view, it must again be underlined
that we can obtain satisfactory results (with care!) even from a choice
of approximation functions that are not coordinate functions.
Going back to the examples presented above, we could have demonstrated that for the beam shown in Fig. 7.1 we indeed have the convergence to the exact solution of the approximate function
 
N
X
nz

uy z
where n is odd
7:54
an sin
l
n1
since


 
nz
an sin
l

7:55

is a set of coordinate functions. Each of the terms in the series, equation (7.54) fulls the geometrical boundary conditions specied in
equation (7.4). On the other hand, with regard to the beam shown in
Fig. 7.2 the choice of the approximate function described in equation

The method of trial functions

189

(7.27), i.e.
uy

 
 
 
z
z
nz
b1 cos
   an sin
a0 a1 sin
l
l
l
 
nz
bn cos
7:56
l

is not constructed from a set of coordinate functions. In fact, not one


of the terms independently satises the geometrical boundary conditions specied in equations (7.13) and (7.14).

7.6

Localised Rayleigh functions

It is fairly evident from the foregoing examples and analyses that the
RayleighRitz technique cannot be simply and directly automated for
general use in practical structural analysis. Nevertheless, the power of
the method lends value to the search for an automated application.
This section presents the basis whereby the method of trial functions, as applied to beams, may incorporate simple functions that,
through the stationary conditions of the potential energy of the
structure, can combine to describe the deected shape of beams with
a variety of boundary and loading conditions to a high degree of accuracy. The approach is introduced here as the localised RayleighRitz
technique, but it may be seen essentially as the progenitor of the
nite element method.
Suppose we sub-divide a beam into a number of regions. The end
points of these regions are called nodes and are numbered sequentially
along the beam, as shown in Fig. 7.7.
Now, we consider a trial function centred on one of these nodes, i, as
shown in Fig. 7.8.
The trial function has the following attributes
.
.
.

unit deection at node i


zero deection at all other nodes
zero slope at all nodes.

l
1
1

2
2
region 2

Fig. 7.7

3
3

4
4

5
5

6
6

node 5

7
7

8
8

9
9

10
10

11

190

Energy methods in structural mechanics


li1

i 2

li

i 1

fi

i +1

i+ 2
z

zi1

zi

Fig. 7.8

The analysis is simplied if we identify a local coordinate system for


the region on either side of node i as shown in Fig. 7.8, the coordinate
system is
0  z i 1  li 1

0  z i  li

7:57

The presentation is further simplied if we again non-dimensionalise


the coordinate system by dening
z
z
i 1  i 1 and i  i
7:58
li 1
li
Thus, the coordinate system is (as shown in Fig. 7.9)
0  i 1  1

and 0  i  1

7:59

The trial function between nodes i 1 and i is obtained very easily.


The boundary conditions that ensure the trial function has a degree
Element i 1

i 1

ui 1

i 1

Fig. 7.9

ui

i+ 1

The method of trial functions

191

of freedom at node i and nowhere else, are


fi 1 i 1 0

at i 1 0

fi 1 i 1 1

at i 1 1

dfi 1 i 1
0
di 1

7:60

at i 1 0 and 1

A simple polynomial function satisfying these requirements is


fi 1 i 1 3i2 1 2i3 1

7:61

The trial function fi between nodes i and i+1 is obtained in the same
way, and the corresponding polynomial function is
fi i 1 3i2 2i3

7:62

that satises the requirements


fi i 1 at

i 0

fi i 0 at

i 1

dfi i
0 at
di

7:63

i 0 and 1

With this trial function the deection of the beam is obtained by multiplying the trial function by the deection, ui , at node i as shown in
Fig. 7.9.
It is clear in the gure that the deection of the beam in the region
between nodes i 1 and i is governed only by the values of the deections at these nodes.
This region between the nodes is called an element and the deection
function for element i 1 can be obtained by combining equations
(7.61) and (7.62), i.e.
uiy 1 1 ui 1 1 3i2 1 2i3 1 ui 3i2 1 2i3 1

7:64

The exural strain energy in this element of the beam is


Ui 1

EI
x
2

li 1

EI
3x
2li 1


d2 uiy 1 zi 1 2
dzi 1
dz2i 1


1  2 i 1
d uy i 1 2
di 1
di2 1
0

EI
3 x 12
ui 1 ui 2
2li 1

7:65

192

Energy methods in structural mechanics


P
1

Elements

Nodes

Fig. 7.10

The potential for automating this approach is now evident since


each element has the same form of exural strain energy with respect
to the length of the element and the deections at the nodes bounding
the element. We can exemplify the application of this approach by
using the beam shown in Fig. 7.10, where for convenience the element
lengths are all equal.
The strain energy of the whole beam is
U

EIx
12
u1 u2 2 
u2 u3 2
3
2l=4

u3 u4 2 
u4 u5 2

7:66

and the potential energy  on account of the concentrated load P at


mid-span is
7:67

 U P
u3

Now, the boundary conditions for the simply supported beam are
u1 0

and

u5 0

7:68

Hence the potential energy of the beam in these terms is




384EIx 2

u2 
u2 u3 2 
u3 u4 2 
u4 2 P
u3
l3
7:69

The equilibrium conditions result from the stationary value of the total
potential energy, that is
q
0;
q
u2

q
0;
q
u3

q
0
q
u4

7:70

The method of trial functions

193

Hence the equilibrium conditions may be put in matrix form as


32 3
2 3
2
u2
0
2 1
0
3
Pl 6 7
76 7
6
7:71
2 1 54 u3 5
415 0
4 1
768EIx
0
0 1
2
u4
From this we obtain the solution
u3

Pl 3
768EIx

u2 u4

and

u3
2

7:72

It is evident that this value for the deection at the mid-length of the
beam is quite dierent from the exact analytical solution of Euler's
equation of beam deformation, i.e.
uy l=2

Pl 3
48EI

7:73

At this point if we wish to improve the accuracy of the predicted


value of the deection we have a choice, we can either
.
.

increase the complexity of the combined trial functions through


increasing the degrees of freedom by augmenting the number of
elements, or
increase the complexity of the trial functions and the number of
degrees of freedom within each element.

Let us take the rst approach. It is evident that if we double the


number of elements in the beam, as shown in Fig. 7.11, the equilibrium
conditions are
32 3
2
2 3
u2
2 1
0
0
0
0
0
0
76 7
6
6 7
6 1
607
6 7
2 1
0
0
0
07
76 u3 7
6
6 7
76 7
6
6 7
7
7
6 0 1
607
6
2 1
0
0
0 76 u4 7
6
6 7
3
76 7
6
6 7
Pl
76 u5 7
6 0
6 7
0
1
2
1
0
0
76 7 1536EI 6 1 7 0
6
x6 7
76 7
6
6 0
607
6 7
0
0 1
2 1
07
76 u6 7
6
6 7
76 7
6
6 7
7
7
6 0
607
6
0
0
0 1
2 1 54 u7 5
4
4 5
0

u8

(7.74)

194

Energy methods in structural mechanics


P
1

Elements

Nodes

Fig. 7.11

The result is that the deection at mid-span is


u5

Pl 3
768EIx

u
u2 u8 5 ;
4

and
u
u3 u7 5 ;
2

3
u
u4 u6 5
4

7:75

which does not dier from equation (7.72). It is therefore clear that
this approach is not very fruitful. The reason lies in the fact that the
choice of the localised trial function was such as not to allow any rotations at the nodes, as a result of the third of the boundary conditions,
equations (7.60) and (7.63). In other words, the representation of the
deection makes the beam unrealistically over-stiened.
As mentioned before, the alternative approach to improve the accuracy of the localised method is to increase the complexity of the trial
function within an element. Thus, we now extend the trial function
for an element of the length of the beam to include two degrees of freedom at each node. These degrees of freedom relate to the deection at
the node i, ui , and the slope at that node, i .
Following the approach shown above for the derivation of the trial
function for the deection, the shape for the trial function for the slope
at node i is shown in Fig. 7.12.
The corresponding boundary conditions are
gi 1 i 1 0

at i 1 0; 1

dgi 1 i 1
0
di 1

at i 1 0

d gi 1 i 1
1
di 1

at i 1 1

7:76

The method of trial functions


g (i 1)
i 2

i 1

g (i)
i

gi

i+ 1

i+ 2
z

i 1

195

Fig. 7.12

gi i 0

at i 0; 1

dgi i
1
di

at i 0

dgi i
0
di

at i 1

7:76 cont:

The simplest polynomial expressions that t these boundary conditions are


gi 1 i 1 i2 1 i3 1
gi i i 2i2 i3

7:77

The deected shape along a particular element is now a function of the


deections and slopes at the nodes bounding the element. Thus, for
element i 1, bounded by nodes i 1 and i, we have, by combining
equations (7.61) and (7.77) (notice that as  duy =d  1=ln , for
simplicity in the presentation we dene n  n ln )
uiy 1 i 1 ui 1 1 3i2 1 2i3 1
i 1 i 1 2i2 1 i3 1
ui 3i2 1 2i3 1 i i2 1 i3 1

7:78

196

Energy methods in structural mechanics

As before, the exural strain energy Ui 1 for this element is


Ui 1

EI
3x
2li 1

1 
0

d2 uiy 1 i 1
di2 1

2

di 1

EIx
12
u2i 1 12
ui 1 i 1 12
ui 1 i 42i 1
2li3 1
12
ui i 1 4i 1 i 12
u2i 12
ui i 42i

It is worth noticing that the results from evaluating the partial derivatives of the strain energy with respect to the degrees of freedom ui 1 ,
i 1 , ui and i , in matrix notation are
2
3
qUi 1
6 q
7
6 ui 1 7
32
2
3
6 qUi 1 7
ui 1
24 12 24 12
6
7
7
6
6 q
7
8
12
47
EIx 6
76 i 1 7
6 12
6 i1 7
7:79
76
7
6
7 3 6
6 qUi 1 7 2li 1 4 24
12
24
12 54 ui 5
6
7
6 q
7
12
4
12
8
i
6 ui 7
4 qU
5
i1
qi
Essentially this is the basis for the stiness matrix for a typical element. In the analysis of the simply supported beam, given the increased
complexity of the trial function, we may restrict ourselves with a subdivision of the beam into only two regions, as shown in Fig. 7.13.
The total potential energy is
 U P
u2

7:80

and again, once the boundary conditions for the simply supported
beam, i.e. u1 0 and u3 0, have been imposed, the equilibrium
P
1

Fig. 7.13

Elements

Nodes

The method of trial functions

197

conditions result from the stationary value of the total potential


energy, that is
q
0;
q1

q
0;
q
u2

q
0;
q2

q
0
q3

7:81

In matrix form the reader will nd it easy to conrm that we have


32  3
2 3
2
0
8
12 4
0
1
3 6 7
7
6 12
7
6
48 0 12 76 u2 7
Pl 6 1 7
6
7:82
76  7
6 70
6
4 4
0 16
4 54 2 5 4EIx 4 0 5
0
0 12 4
8
3
The solution of this matrix equation is
2 :
3
2 3
0 125
1
7
6 u 7
Pl 3 6
6 0:08333 7
6 27
6
7
6 7
5
4 2 5 4EIx 4 0
0:125
3
and the deection at mid-span is
u2

Pl 3
48EIx

7:83

This corresponds exactly with the result from the analytical solution of
Euler's beam deformation equation, primarily because that solution
shows the beam deformation can be described by a cubic function of
z and the localised trial functions used here include cubic terms.
In conclusion, it is worth noting that a further step to introduce the
automation of the procedure can consist in applying the natural
boundary conditions after the stationarity of the total potential
energy has been required. In the case at hand it would have meant
rst writing the results of equation (7.81) as
2
32 3
2 3
u1
24 12
24 12
0
0
0
6 12
7
6
7
6
7
8
12
4
0
0 76 1 7
6
607
6
76 7
7
3 6
6 24
617
12
48
0 24 12 76 u2 7
6
76 7 Pl 6 7 0
6 12
6 7
6 7
4
0
16
12
47
6
76 2 7 4EIx 6 0 7
6
76 7
6 7
4 0
405
0 24
12
24
12 54 u3 5

0
0 12
4
12
8
0
3
7:84

198

Energy methods in structural mechanics


P
1

Elements

Nodes

Fig. 7.14

and then setting u1 0 and u3 0 by eliminating the rows and
columns relating to u1 and u3 , which gives the matrix shown in
equation (7.82).
In this manner the formulation can very easily be adjusted to enable
the analysis of beam with other boundary conditions than simply
supported. This can be exemplied with regard to a cantilever beam,
encastre at node 1 and having a concentrated load P at node 3, see
Fig. 7.14.
The natural boundary conditions are
u1 0
and
1 0
at node 1
7:85
Thus, we can simply eliminate the rst two rows and columns from the
matrix (7.84) and set the contribution of the load P at node 3. The equilibrium equations are
32 3
2
2 3
u2
48 0 24 12
0
3 6 7
7
6 0 16
7
6

12
4 76 2 7
Pl 6 0 7
6
7:86
76 7
6
6 70
4 24 12
24
12 54 u3 5 4EIx 4 1 5
12 4
12
8
0
3
and the result from this matrix equation is
u3

Pl 3
3EIx

7:87

To sum up, we see that with an accurate choice of the shape functions,
even with only two elements this approach of localised functions provides the exact answer for the deection of the cantilever and the central
deection of a simply supported beam. However, perhaps even more
impressive is the ease with which the resulting matrix formulation can
be particularised to incorporate the relevant boundary conditions for
the structure. It is this capability of the localised RayleighRitz

The method of trial functions

199

method, that has enabled it to be programmed for computers and used


in such an easy manner by engineers for a great variety of types of structures. Chapter 11 in this book introduces formally the nite element
method which directly derives from the localised trial function approach
and formulations presented here.

7.7

Numerical example: application of Rayleigh


localised trial functions to the analysis of a frame

This example illustrates the application of the method of localised trial


functions to the analysis of the frame that previously has been analysed
using the CotterillCastigliano theorem, see section 6.5. The frame is
divided into a number of regions and, similar to the analysis in section
7.6, localised functions are proposed as trial functions.
The portal frame to be analysed here is shown in Fig. 7.15. Each
element is subdivided into two regions, or elements, each of equal
length, i.e. ln l=2.
The symmetry at position C is taken into account in this analysis
to reduce the number of degrees of freedom. Again, we consider the
elements to be axially undeformable.
We recall from equation (7.79) that for a typical element the partial
derivatives relative to the strain energy, which lead to the denition of

l/2
Symmetry
P
Node 3
B

l/2

Region 3

Region 2

I2

I1

Node 4

I2

Node 2
l=5m
l/2

I=

Region 1

Node 1
A

Fig. 7.15

I1
I2

= 268

200

Energy methods in structural mechanics

the stiness matrix, can be written as


2
32
3
ui 1
24 12 24 12
6
7
8
12
47
EI 6
6 12
76 i 1 7
K n 3x 6
76
7
2ln 4 24
12
24
12 54 ui 5
12
4
12
8
i
where ui is the deection at the node i normal to the axis of the element
and i is the change of slope at that node.
The junction of the regions 1 and 2 at node 2 is such that the elements have common values of deection and change of slope. The
junction of the vertical and horizontal members at B is such that the
deection of region 2 at node 3 is zero and the change of slope for
regions 2 and 3 is common at node 3. Thus combining the stiness
matrices of the three regions to provide the stationary condition for
the strain energy of the portal frame
32 3
2
24 12 24 12
0
0
0
u1
76 7
6
6 12
6 7
8
12
4
0
0
0 7
76 1 7
6
76 7
6
6 24
6 7
12
48
0
12
0
0 7
76 u2 7
6
76 7
6
4EI
2
6 12
6 7

4
0
16
4
0
0 7
K
76 2 7
l3 6
76 7
6
6 0
6 7
0 12
4 81 I 12I 4I 7
76 3 7
6
76 7
6
6 0
6 7
0
0
0
12I
24I 12I 7
54 u4 5
4
0
0
0
0
4I
12I 8I
4
where I  I1 =I2 .
The boundary conditions are
u1 1 0 at A;

u3 0 at B

4 0 at C

and

Applying these boundary conditions and taking into account the term
relative to the potential of the vertical load P the condition of equilibrium of one half of the symmetric frame is
2 3
2
32 3
48 0
12
0
u2
0
6
6
7
6
7

4
0 76 2 7 P 6 0 7
4EI2 6 0 16
7
6
76  7 6 7 0
3
4
4
5
4
5
2
12 4 81 I 12I
l
05
3
0

12I

24I

u4

The matrix equation can be solved to give the deected shape for the
portal frame corresponding to various values of I.

The method of trial functions

201

As an example, the value of I 0:0001 is analysed and the matrix


equation gives the solution
u4

1
Pl 3
191:97 EI1

This value of I implies that the exural stiness of the horizontal


member of the frame is very much less than that of the vertical
members. Thus the horizontal member should react to the application
of the load as if it were a beam fully encastre at both ends. Note that a
value of I 0 would render the matrix equation singular.
The analytical formulation for the deection at mid-span for such a
beam is
u4

1 Pl 3
192 EI1

Thus the analysis compares nearly exactly with the analytical formulation. This is because the analytical solution which describes exactly
the deformations of the members of the frame is a cubic polynomial.
The trial function actually is also a cubic polynomial and therefore
completely satises the requirement.
Consider now the analysis of the portal frame shown in Fig. 6.5,
which has been shown in section 6.5 to have I 2:68. We have the
result from the matrix analysis as
3
2 :
2 3
u2
0 0134
7
6  7
Pl 3 6
6 0:0134 7
6 27
7
6 :
6 7
4 3 5 8EI2 4 0 0534 5
0:0423
u4
Hence the deection u4 at mid-span of the horizontal member, for a
load, P 64:7 kN, is
Pl 3
u4 0:0423
8EI2

0:0423  6:47  104  50003


13:49 mm
8  2:1  105  1:51  107

This compares exactly with corresponding value calculated using the


CotterillCastigliano theorem.
The strain at the mid-span can be calculated using the curvature of
the member BC at that point. In non-dimensional coordinates, the

202

Energy methods in structural mechanics

curvature of an element is obtained from

d 2 uy
d2
2 
ui 1 1 3 2 2 3 i 1  2 2  3
2
d
d
ui 3 2 2 3 i  2  3

ui 1 6 12 i 1 4 6
ui 6 12 i 2 6

The curvature of member BC is

2
d 2 uy
1 d uy

dz2
ln2 d 2

4

u3 6 12 3 4 6
l2
u4 6 12 4 2 6

where
 0 at node 3,  1 at node 4 and with u3 4 0:
The result from the analysis for I 2:68 gives the curvature at C as

d 2 uy
Pl
Pl

20:0534 60:0423 :
2
2EI2
13 605EI2
dz

The strain at the outer face of the member BC at point C is


"c

6:47  104  5000  112


0:84  103
5
7
:
:
:
13 605  2 1  10  1 51  10

This compares exactly with the corresponding value from the


CotterillCastigliano analysis, see section 6.5.
It is evident that the method of trial functions can be easily applied
to the analysis of frames. A more specic form of matrix-based analysis for frames is presented in chapter 9. This is based on the actual solution of Euler's equation for beam deformation and is the method that
is usually used in structural engineering practice. However, there may
be circumstances, such as the components of the frame having variable
geometry that would complicate the usual frame analysis. In such a
conditions the method of localised trial functions, i.e. the nite element
method, can be used to obtained accurate solutions for the strains and
deformations in the frame.

The method of trial functions

Suggested exercise problems


1.

2.

3.

203

Figure 7.16 shows a beam with a uniform exural stiness of EI


and built-in at both ends. The beam is subjected to a uniformly
distributed load of q/unit length. Use the method of trial function
to derive an expression for the deection shape and the value of
the deection wc at mid-length. Compare the values obtained
with the exact ones from the analytical solution of Euler's equation for beam deformation.
[Ans: Trial function wc ql 4 =390EI, Analytical value wc
ql 4 =384EI]
The beam shown in Fig. 7.17 has a exural stiness of EI for half
its length and 2EI for the other half. The beam is subjected to a
uniformly distributed load q. Use the method of trial functions
to develop an expression for the deection at mid-length. Apply
the method of localised trial functions to provide a comparable
value for the deection.
[Ans: One term trial function wc ql 4 =16:1EI, localised trial function with two elements along the beam, wc ql 4 =26:3EI, localised
trial function with four elements along the beam, wc ql 4 =16:0EI]
The beam shown in Fig. 7.18 is subjected to a distributed load,
varying from Q/unit length at the end A which is simply supported, to 0 at the end B which is built-in. The beam has a uniform
q
A

B
EI

Fig. 7.16

EI

2EI
l

Fig. 7.17

204

Energy methods in structural mechanics


Q

A
EI

Fig. 7.18

exural stiness EI. Use the method of trial functions to calculate


the deection of the beam at mid-span. Check the answer against a
two element localised trial function model.
[Ans: one term trial function, wc Ql 4 =399:0EI, two element
localised trial function, wc Ql 4 =347:8EI]

8.

Matrix analysis of pin-jointed


trussed structures

Introduction

Foregoing chapters have been concerned with developing the basis of


the analysis of bars and beams, both from the exact solution of the
Euler equation of deformation and from an approximate approach.
The present chapter and the next one extends the analytical solution
of Euler's approach to encompass collections of bars and beams that
constitute truss and frame structures. These types of structure are
ubiquitous in engineering practice, including the skeletons of buildings,
lattice bridges and a great variety of cranes and lifting equipment.
We begin this chapter with the analysis of pin-jointed structures.
These structures are collections of bars essentially subjected to axial
forces which therefore constitute the primary loading in the element.
However, as the joints in practical frame structures are fabricated by
welding or by using groups of bolts, some bending of the individual
members will be induced because of the rigidity of the joints, as well
as because of the bars' self weight. This bending is generally seen as
a secondary eect. Nevertheless, if the bars are arranged so that
their axes meet in one point at each joint, it is found that the secondary
bending can be considered negligible with comparison to the axial
forces. Indeed, in some bridges built in the 19th century the joints actually were eected using pins so that the forces in the bridge members
were in accord with the analysis.
Essentially the function of such pin-jointed structures, or trusses, as
they are commonly known, is that of carrying loads across a wide span.
It is common to introduce the analysis of pin-jointed frames through
consideration of the force equilibrium at the pin-joints. That method
requires that there is assurance that the truss has a geometry, i.e. an
arrangement of members, that allows that method of analysis to be
valid. The validity rests on the establishment that the truss is statically
determinate which means that the forces in the members can be

206

Energy methods in structural mechanics

calculated purely on the basis of static force equilibrium. As we know


from chapter 2, if we remove a member from a statically determinate
truss the structure becomes a mechanism and the equilibrium is not
assured under any loading. Conversely, if additional members are
added to a statically determinate truss geometry the truss is classied
as having statically redundant members and the forces in the members
cannot be calculated solely from the force equilibrium of the pin-joints.
Thus the traditional approach for introducing the analysis of pinjointed frames is applicable only to a limited range of truss geometries
and, moreover, is not capable of extension to the analysis of rigidjointed frames.
The method presented in this chapter is completely general and is in
fact the basis of practical design analysis of frames. In the method presented here there is no requirement to investigate if the truss is statically
determinate or structurally redundant, and the method is such that it
can encompass all forms of truss geometry and will enable the evaluation of member forces and the deformed geometry of the truss.

8.1

Plane pin-jointed structures as systems of axially


loaded bars

So far we have examined the behaviour of one-dimensional deformable


bodies, like bars and beams and introduced the principle of stationary
value of total potential energy as a basis for approximate energy methods, like the method of trial functions. In the present case the analysis
methods are extended to the systems of several bars and beams and we
start by looking at the example of a plane pin-jointed structure, also
called a truss, shown in Fig. 8.1.
This structure is typical of the arrangement of bars to support a
pulley for lifting heavy pieces of equipment and is composed of three
straight bars AB, AC and AD, pin-jointed together at the point A
and attached to a vertical rigid wall by pin-joints at the ends B, C
and D, respectively.
We assume that the pin-joints do not oer any resistance to rotation
of the bars, either with respect to the wall or with respect to the relative
rotations of the members connected at a joint. This assumes, of course,
that we neglect any frictional forces acting at the contact surfaces
between the pins and the members that could occur in a practical pinjoint, but such forces could be made small so that we can be assured
that the above approximation is reasonable and thus no moments
can arise at the ends of any bar. The only force applied to the structure
is the force P at the point A, representing the load suspended from that
joint. For the sake of generality we suppose at this stage that the loading
is inclined to the vertical.

Matrix analysis of pin-jointed trussed structures

207

45
30

A
P

Fig. 8.1

In order to deal with a structure composed of several members, it is


rst necessary to consider its individual members, whose behaviour we
are able to analyse in a relatively simple way. In the present case all the
members are simple straight bars, such that for the equilibrium of each
of them, taken separately, we can immediately deduce that they can
only bear compression or tension, as shown in Fig. 8.2.
The deformation, i.e. the kinematic behaviour, of a typical member
of the pin-jointed frame is therefore quite simple and can be described
by the relationship (equation (4.5))


N
0
l 1
l
EA
that represents the relationship between the initial and the nal lengths
of the bar, l and l 0 , respectively, and the applied axial loading, N. The
elastic energy stored in the individual typical member of the truss on
account of its deformation is related to its change of length by equation
(4.8),
U

1 EA 0
l l2
2 l

208

Energy methods in structural mechanics


N

Fig. 8.2

Given the behaviour of its component members, the study of the


equilibrium of the whole structure can be carried out as usual with
reference to the general condition of equilibrium (equation (1.28))
U W
or, as we have seen in chapter 6, with reference to the principle of
stationary value of the total potential energy, i.e.
UW
where obviously the elastic energy of the whole structure is given by the
sum of the energy stored in each of its members, i.e.,
X i
U
U
8:1
i

where i indicates the number of members in the structure. The work


done by the external forces is given by
W P  uA

8:2

where uA is the vectorial displacement of the point A. The work done


by the external forces can also be expressed in terms of components of
the vectorial displacement uA and we have
W Px ux A Py uy A

8:3

It is evident, because we are dealing with a pin-jointed truss entirely


contained in one plane, that the structure can have only two degrees of

Matrix analysis of pin-jointed trussed structures

l' i

209

ux (K )
uy (K )
K

ux (J)

uy (J)
J

li

Fig. 8.3

freedom, that is, the degrees of freedom are the two components of
displacement of the point A, which in our example, see Fig. 8.1, is
the only one free to move. Once we have dened its position under
the applied load P, then we can immediately determine for each
member how much it is stretched and the value of the axial force it
is carrying (see equation (4.7)). In order to achieve this, let us relate
the initial and nal lengths of each bar to the coordinates of its ends
and to the displacement of the point A. With reference to Fig. 8.3,
we can write for a generic element, whose end points are J and K
q
l i xK xJ2 yK yJ2
8:4
s
fxK ux K xJ ux Jg2
l 0i
8:5
fyK uy K yJ uy Jg2
The following relationships hold true between the angle i and the
coordinates of the end points J and K
cos i

1
xK xJ
li

8:6

sin i

1
yK yJ
li

8:7

210

Energy methods in structural mechanics

thus, equation (8.5) becomes


v
u
u 1 2 fu K u J cos i u K u J sin i g
u
x
y
y
li x
u
l 0i l i u
t 1
i2 fux K ux J2 uy K uy J2 g
l
8:8
We are operating in the framework of innitesimal displacements and
we can therefore simplify the expression in equation (8.8) by taking
into account only theplinear
terms of a simple Taylor's series expansion
(that is, of the type 1   1 =2). Thus, we obtain
l 0i l i ux K ux J cos i uy K uy J sin i

8:9

With regard to the bars composing our structure we can write


Element BA
i

sin i

cos i

ux J

uy J

ux K

uy (K)

ux A

uy A

l 0BA l BA ux A
U BA

8:10

1 EABA
ux A2
2 l BA

8:11

Element CA
i

sin i

308

12

cos i
p
23

ux J

uy J

ux K

uy K

ux A

uy A

p
3
1
ux A uy A
l
l
2
2
p
2
CA 
1 EA
1
3
CA
u A uy A
U
2 l CA
2
2 x
0CA

CA

8:12
8:13

Element DA
i
458

sin i
p
22

cos i

p
2
2

ux J

uy J

ux K

uy K

ux A

uy A

Matrix analysis of pin-jointed trussed structures

p
p
2
2
u A
u A
l
l
2 x
2 y
p

p

2
1 EADA
2
2
DA
u A
u A
U
2 l DA
2 x
2 y
0DA

DA

211

8:14
8:15

The total potential energy for our example structure is therefore


 U W U BA U CA U DA W
p
2
3
1 EABA
1 EACA
1
2
ux A
u A uy A

2 l BA
2 l CA
2
2 x
p


2
1 EADA
2
2
u
u

A
2 l DA
2 x
2 y
Px ux A Py uy A

8:16

It is evidently a function only of ux A and uy A. The magnitudes of the


variables ux A and uy A that render the total potential energy stationary are yielded by the fullment of the following stationary conditions
p


3
q
EABA
EACA 3
u A
BA ux A CA
u A
qux A
4 x
4 y
l
l


EADA 1
1
ux A uy A Px 0
DA
8:17
2
2
l
p


3
q
EACA 1
u A
CA
u A
quy A
4 y
4 x
l


EADA 1
1
u A ux A Py 0
8:18
DA
2 y
2
l
These conditions of course represent the equilibrium equations of the
structure under consideration.
Once we have obtained the values of ux A and uy A we know the
displacements of the ends of each member, its elongation, the corresponding axial strain it has undergone and, from equation (4.7), the
axial loading it is actually carrying. We can also evaluate the degree
of safety of the structure. In other words we know everything of the
plane truss from a mechanical point of view.
It is worth noticing that the key point of the example analysis has
been the choice of the number of variables that are able to describe
completely the kinematics of the elastic structure. We call these
variables the degrees of freedom of the structure because they allow

212

Energy methods in structural mechanics

us to determine the exact position in the plane of the structure after the
deformation has taken place.
In addition, quite distinct from what happens for the systems of rigid
bodies examined in chapter 2 that can experience a displacement eld
only and only if the constraints are not sucient to x their position in
the space, elastic structures always tend to deform under loading, even
if rigid body displacements are prevented by the restraints. Strictly
speaking, any elastic body can be thus thought to be characterised
by an innite number of degrees of freedom, as each of the degrees
is necessary to dene the displacement of any of its points. However,
if we can describe by means of appropriate functions the behaviour
of an individual elastic element, as in the case of the elongation of
axially loaded bars or the solution to Euler's equation of beam deformation, we can always analyse very complex structures by dividing
them into their members and studying the overall systems with respect
to a nite number of variables, which we call once again degrees of
freedom of the system.
As we have just seen, in the case of the pin-jointed structures these
variables can be assumed to be the displacements of the joints, or
nodal points, that link the structural components together.

8.2

Matrix formulation of the elastic equilibrium


equations for plane pin-jointed structures

In the present section we will show how the equilibrium equations that
we have derived in section 8.1 can be formulated in a much more compact form, one that is particularly suited to automatic processing by
means of computers. Many engineering structures consist of a collection of bars and beams assembled together to provide an ecient
way for resisting applied forces and carrying required loads. From a
theoretical point of view these structures can always be treated as
has been shown in the previous section, but following those procedures
step-by-step would generally involve incredibly lengthy and timeconsuming hand calculations even, as we have seen, in the case of a
fairly small number of members. However, the use of computers
oers the possibility of analysing very complex structures in a very
short time and requires minimal eort on the part of the engineers,
provided the problems are formulated in a suitable manner.
The use of linear algebra, i.e. matrix notation, stems naturally when
performing calculations on a computer, as it allows very large group of
numbers to be manipulated in a systematic and well-established
manner. At the same time a matrix approach to the analysis of structures makes possible a comprehensive method that applies to a wide
variety of types of frame structures and requires only some eort to

Matrix analysis of pin-jointed trussed structures

213

input the data dening the frame geometry, the material properties and
the loading, thus minimising the possibility of errors. Many standard
commercially developed programs are now available for use on even
quite small computers. These programs provide a capability for
calculating the displacements and strains for frames ranging from
the simplest, with say three or four members, to very complex
frames with hundreds of members. The intention here is to explain
the mechanics that underlies the computation in the programs.
As an aid to our presentation of how it is possible to formulate the
equilibrium equations in a matrix form, let us once again make reference to the pin-jointed structure of Fig. 8.1. In section 8.1 the key point
of our presentation was the analysis of the truss in terms of its individual members, whose behaviour has been related to the displacements
of the nodal joints. Following exactly the same line of reasoning, let us
try to express the same steps in a matrix format. First of all, we can recast equation (8.9) in matrix notation, which provides the relationship
between the nal and the initial lengths of each bar as function of the
displacements of its end points and of the angle  of the bar in the
plane. Thus, we have
2
3
u J
"
# x
6 u J 7
0
0
cos i sin i
6 y
7
l i l0i 1 1
7
i
i 6
4
5
u
K
0
0
cos  sin 
x
uy K
8:19
This notation enables us to identify a matrix containing terms that provide information about the position of the i-th member in the frame,
i.e.
"
#
0
0
cos i sin i
i
R
8:20
0
0
cos i sin i
A central component in equation (8.19) is the vector containing the
displacement of the member's end nodes in the general reference
frame xy, i.e.
3
2
ux J
6 u J 7
7
6 y
8:21
Di 6
7
4 ux K 5
uy K
Notice that the product between Ri and Di results in the vector d i

214

Energy methods in structural mechanics


u (K )
u (K )
u (K )
u (J)
J

u (J)

u (J)

Fig. 8.4

containing the axial components of the nodal displacements in the


local reference frame   shown in Fig. 8.4.
Therefore the relationship


u J
d i Ri Di
8:22
u K
may be properly regarded as a compatibility equation that links the displacements of the end nodes of the i-th bar in the general reference
frame xy to the axial components of displacement of the same
nodes in its local reference frame  . It is worth noticing that
Ri T Ri I. It is now clear that the elongation of the bar can be
expressed in matrix form as
l i l 0i l i 1

1 d i 1

1 Ri Di

8:23

which is equivalent to equation (8.19).


The elastic energy stored in each bar can also be expressed in matrix
notation. In fact we obtained the expression for the elastic energy,
equation (4.8), by means of Clapeyron's theorem, namely by taking
half of the work done by the axial loading N corresponding to the
length variation l. As the expression for N, equation (4.7), is readily
stated in matrix notation, i.e.
Ni

EAi i EAi
l i 1
li
l

1 Ri Di

8:24

we can write the strain energy in matrix format as


1
1 EAi
1
U i N i l i
2
2 li

1 Ri Di 1

1 Ri Di

8:25

In order to reach a much more meaningful expression to our purposes, we notice that, as the transpose of a scalar (in our case a real
number) is the scalar itself, we can apply the transpose chain rule to

Matrix analysis of pin-jointed trussed structures

215

the rst three matrices from the left in equation (8.25) (whose product
is a scalar quantity) and write


1 i i 1 EAi i T i T 1
i
1 1 Ri Di
D R
8:26
U N l
2
2 li
1
By expanding the middle product between vectors and rearranging the
terms, we obtain


1 1
1 i i 1 i T i T EAi
i
U N l D R
Ri Di
8:27
2
2
l i 1
1
The term



1 1
EAi
K i
l
1
1
i

8:28

is commonly referred to as the stiness matrix of the bar. It links the


axial displacements of the end points of the bar, as yielded by equation
(8.22), to the axial loads applied at the same ends, namely




NJ
1 1
EAi
i
d i
8:29
l
NK
1
1
Inspection of the stiness matrix reveals that it is symmetric, that is the
rst and the second rows are identical, respectively, to the rst and
second columns. This is a general property of any linearly elastic system
and can be derived from the reciprocal theorems presented in section 5.2.
In fact, this property of symmetry is equivalent to stating that a given displacement
applied at point K of an elastic body causes a force at point J equal to that which
would have been caused at point K were the same displacement to be applied at coordinate J rather than K.

Alternatively, this result can be viewed as a direct consequence of the


principle of conservation of energy. In fact, in order to be an exact
dierential, the elastic energy function U i of the generic bar must
full the following condition
qU i
qU i

qu J qu K

8:30

which by virtue of equations (8.22) and (8.27) implies that the stiness
matrix K i must be symmetric.
It is worth noting also that the term


1 1
EAi
K i Ri T i
Ri
8:31
l
1
1

216

Energy methods in structural mechanics

can also be regarded as a stiness matrix, provided we recognise that it


relates the vector of the nodal displacements Di to the vector of the
nodal forces F i in the general reference frame. In fact, the vector
F i is expressed as
3 2
3
2
Fx J
cos i
0




6 F J 7 6 sin i
0 7
7 6
7 NJ
6 y
i
i T NJ
R
F 6
76
7
4 Fx K 5 4 0
NK
cos i 5 NK
i
Fy K
0
sin 
8:32
By substituting equation (8.29) in equation (8.32) we obtain

i
1 1
i
i T EA
d i
F R
l i 1
1

8:33

and remembering equ (8.22), we have


F i K i Di

8:34

At this point the total potential energy of the whole pin-jointed structure, which we have already expressed in equation (8.16), can easily be
re-cast in matrix notation.
In order to keep the procedure as general as possible (as is desirable
for a procedure suitable for computer implementation), we align all the
nodal displacements in the vector D, namely
DT ux A uy A ux B uy B ux C uy C ux D uy D
8:35
The required matrix form of the total potential energy is then accomplished by assembling a joint stiness matrix that is obtained by summing contributions corresponding to each node from the member
stiness matrices K i . This is obviously equivalent to summing all
the expressions of the elastic energy relating to each bar, being the
associative law of multiplication valid for the matrix product.
In fact, let us write the expression of the generic member stiness
matrix K i in complete detail
2

K i

cos2 i

i
i
EAi 6
6 sin  cos 
6
l i 4 cos2 i

sin i cos i

sin i cos i

cos2 i

sin i cos i

sin2 i
sin i cos i

sin i cos i
cos2 i

sin2 i
sin i cos i

sin2 i

sin i cos i

sin2 i

3
7
7
7
5

8:36

Matrix analysis of pin-jointed trussed structures

217

This generic expression is written in the general reference frame and it is


evident that in order to obtain the stiness matrices for the bars BA,
CA and DA in a form that can be related to the joint displacement
vector D, all we have to do is identify the joints J and K in the general
numbering system on an individual basis and set the relative terms in
the appropriate position. Therefore, we can write
2

cos2 BA
6 sin BA cos BA
6
6
6 cos2 BA
6
BA 6
BA
BA
EA
6 sin  cos 
AB
K BA 6
6
l
0
6
6
6
6
4
0

sin BA cos BA

cos2 BA

sin BA cos BA

sin BA cos BA

cos2 BA

sin BA cos BA

sin2 BA

sin BA cos BA

sin2 BA

07
7
7
07
7
07
7
7
07
7
07
7
7
05

sin 

BA

sin 

BA

cos 

BA

BA

sin 

(8.37)
2

cos2 CA
6 sin CA cos CA
6
6
6
0
6
CA 6
EA
0
6
CA
K CA 6
6
l
cos2 CA
6
6 sin CA cos CA
6
6
4
0
0

sin CA cos CA


2

sin 

CA

cos2 CA

sin 

CA

sin CA cos CA

cos 

CA

sin 

CA

sin CA cos CA

cos2 CA

sin CA cos CA

sin2 CA

sin CA cos CA

sin2 CA

0
0

0
0

0
0

0
0

0
0

0
0

07
7
7
07
7
07
7
7
07
7
07
7
7
05
0

(8.38)
2

cos2 DA
6 sin DA cos CA
6
6
6
0
6
DA 6
EA
0
6
DA
K DA 6
6
l
0
6
6
0
6
6
4
cos2 DA
sin DA cos DA

sin DA cos DA


2

sin 
0

DA

0 0

cos2 DA
DA

sin 

sin DA cos DA


DA

cos 

DA

0
0

0
0

0 0
0 0

0
0

0
0

0
0

0 0
0 0

0 0

sin DA cos DA


sin2 DA

0
0

0
0

0 0
0 0

cos2 DA
sin DA cos DA

sin DA cos DA


sin2 DA

0
0
0

sin 
0
0
0

3
7
7
7
7
7
7
7
7
7
7
7
7
7
5

(8.39)
Thus, the following expressions are valid:
U BA 12 DT K BA D

8:40

U CA 12 DT K CA D

8:41

U DA 12 DT K DA D

8:42

and it follows immediately that the elastic energy stored in the frame is

218

Energy methods in structural mechanics

given by
U U BA U CA U DA
12 DT K BA D 12 DT K CA D 12 DT K DA D

12 DT KD
 is the combined stiness matrix, that is
where K
 K BA K CA K DA
K

8:43
8:44

This is true on account of the above-mentioned associative property of


the product between matrices. The total potential energy of the frame
in matrix notation can now be obtained by formulating the components of the applied force as a load vector, i.e.
PT Px A Py A

0 0

0 0

8:45

which allows us to write the work of the applied loads as


W DT P

8:46

and, consequently

 U W 12 DT KD
DT P

8:47

At this point it is worth noting that if the reader takes the trouble to
develop the expression in equation (8.47) to its full extent by means of
the values given in the tables in section 8.1, he will nd a somewhat different result from that in equation (8.16). This is because that up to this
point we have considered as active all the possible displacements of the
nodes A, B, C and D. Even if this fact may not seem to be reasonable,
since the displacements of the joints B, C and D are actually prevented
by the attachment to a rigid wall, the reason is that we are developing
here a method suitable for an automatic procedure and our steps must
conveniently correspond to several distinct phases of operation of a
computer program. Similarly to that shown in section 7.6 in the case
of the method of localised Rayleigh functions, there is no necessity
to identify the active degrees of freedom at an early stage, as this
would impose a particularisation on the stiness matrices of the elements and prevent the possibility of implementing a totally automatic
manipulation of the data.
Moreover, to enhance further the generality of the approach we
can extend the applied load vector to include the reactions at the
constraints. Thus the loading vector in this example becomes
PT Px A Py A Rx B Ry B Rx C Rx D Ry D
8:47b

Matrix analysis of pin-jointed trussed structures

219

In fact, once one has constructed equation (8.47), it is always


possible to obtain equation (8.16) by simply setting the displacements
ux and uy of the joints B, C and D to zero. However, as we are going to
see, it is better to perform this operation at the next stage.
Now, returning to the elastic equilibrium problem, it is clear that the
stationary value of the total potential energy expression, equation
(8.47), must be required with respect to the degrees of freedom set in
the displacement vector D. In eect we have to impose that
q
0
qDj

8:48

where Dj is the generic component of the displacement vector D.


It is immediately evident that equation (8.48) is a linear system of n
equations in n unknowns, namely the components of the vector D. It
can thus be written in matrix notation as

KD
P 0

8:49

as a direct consequence of equation (8.47). However, the system in


equation (8.49) does not admit a solution since the combined stiness
 is singular and therefore it cannot be inverted. From a
matrix K
physical point of view, this is because the movement of the structure
as a rigid body is not prevented and, consequently, the solution of
the problem of equilibrium cannot be obtained. It is therefore necessary to impose on the system the requirements of the geometric constraints at B, C and D. In order to do so we can write the system,
equation (8.49), in the following form
 
 


KI
KI;II
PI
DI

8:50
DII
PII
KI;II T KII
where we rearranged the equations in order to distinguish two subarrays in the nodal displacement vector D, namely DI and DII ,
corresponding, respectively, to the degrees of freedom that are active
and to the degrees of freedom which are actually suppressed by the
restraints. At the same time we distinguish two sub-arrays in the
load vector P, PI and PII ; the rst corresponds to the loads applied
to the free joints, the second to the reactions which arise at the
restrained ones. It goes without saying that the set PI is assigned,
while the set PII is unknown, as the displacements DI are unknown
and the displacements DII are set to zero. This happens because we
can obviously assign kinematic or mechanical conditions at any
degree of freedom, but never both at the same time.

220

Energy methods in structural mechanics

In the present example, the sub-array DII corresponds to the


inactive degrees of freedom ux B, uy B, ux C, uy C, ux D and
uy Dand we have
DII 0

8:51

We can sub-divide the system, equation (8.50) into two expressions


8:52
KI DI PI KI;II 0
PII KI;II T DI KII 0

8:53

Eectively, the system equation (8.52) is disjoint from equation (8.53)


and the matrix format in equation (8.49) is now reduced to a set of two
linear equations in the two unknowns active degrees of freedom, ux A
and uy A, that is
8:54
KI DI PI
which the reader will nd to be identical to equations (8.17) and (8.18),
and obviously admits a unique solution. Once we have solved equation
(8.54) and obtained the magnitudes of the displacements DI , we can
straightforwardly obtain the restraint reactions PII from equation
(8.53).
It is worth noting that the 2  2 joint stiness matrix KI is positive
denite (and therefore non-singular), which follows immediately from
the condition
U 12 DI T KI DI > 0

8:55

once any rigid body displacement has been suppressed. From a physical point of view the set of equations (8.54) requires that the sum of the
elastic reactions from the end joints of each element balances the
applied forces at any joint of the structure. In the present example
this is composed only of the force P at A.
At this stage one is allowed to wonder why we have gone through all
these lengthy and apparently cumbersome calculations in order to
obtain equation (8.54).
The reason lies in the fact that once we have established that the conditions of equilibrium (equation (8.48)) can be expressed in the form shown in equations (8.49)(8.54),
we can directly and automatically build a similar system for any plane pin-jointed
structure without bothering to repeat all the underlying reasoning in terms of energy.

As a matter of fact, the only data required to produce the equilibrium equations (8.49) are the geometry of the structure, the denition
of the load and the form of the generic element stiness matrix, equation (8.36). The imposition of the restraint conditions leads subsequently and without eort to equation (8.54).

Matrix analysis of pin-jointed trussed structures

221

This constitutes the basis of the so-called direct stiness method that
in the next chapter we will apply to rigid-jointed frames.

8.3

Numerical example: analysis of a truss

The example concerns the analysis of a simple truss, shown in Fig. 8.5.
The structure is pinned to a rigid foundation at A and D and has a
horizontal load, P1 , applied at the joint B and a vertical load, P2 ,
applied at C.
The loads are applied proportionally such that
P1 2P2
The frame is composed of circular hollow sections, with a mean
diameter of dm 250 mm and a wall thickness of t 7:5 mm. The
cross-section area of the members is

A 257:52 242:52 5:890  103 mm2
4
The Young modulus is
E 2:1  105 N=mm2
The axial stinesses of members AB, BC and CD are equal and is
EA 2:1  105  5:890  103

4:12  105 N=mm


k1
l
3000
P2
P1

3m

+ve
A

3m

Fig. 8.5

222

Energy methods in structural mechanics

The stiness of the member AC, k2 , is


1
k2 p k1
2
This example can be tackled using the method presented in section 8.1
and the reader is encouraged to apply that method and thus become
familiar with it. However, the main purpose of the example is to illustrate the more powerful and automated matrix approach described in
section 8.2.
The objective of the analysis is to evaluate the maximum allowable
level of the load P1 such that the maximum strain does not exceed 30%
of the material yield strain, "0 , where "0 1:5  103 .
The direct stiness method is essentially concerned with assembling
the stiness matrix of the complete structure from the component stiness matrices. Thus from equation (8.36) we have
Member AB
i AB


2
2

60
6
K AB k1 6
40

1
0

0 1 7
7
7
0
05

this stiness matrix is associated with the displacement vector


2

3
ux A
6 u A 7
6 y
7
DAB 6
7
4 ux B 5
uy B
Member BC
i BC 0
2

6 0
6
K BC k1 6
4 1
0

1 0

0 07
7
7
1 05

0 0

Matrix analysis of pin-jointed trussed structures

this stiness matrix is associated with the displacement vector


2

ux B

6 u B 7
6 y
7
DBC 6
7
4 ux C 5
uy C
Member CD
i CD
2


2

1
0

0
0

1 7
7
7
05

0 1

60
6
K CD k1 6
40

this stiness matrix is associated with the displacement vector


2

ux C

6 u C 7
7
6 y
DCD 6
7
4 ux D 5
uy D
Member AC
i AC

K AC


4
2

1 1

k2 6
6 1
6
2 4 1

1 7
7
7
15

1 1

this stiness matrix is associated with the displacement vector


2

3
ux A
6 u A 7
6 y
7
DAC 6
7
4 ux C 5
uy C

223

224

Energy methods in structural mechanics

Following equation (8.44), the assembled stiness matrix, taking


account of the displacement vectors, is
3
2
a
a
0
0
a
a
0
0
6 a 1 a
0 1
a
a
0
07
7
6
7
6
6 0
0
1
0
1
0
0
07
7
6
6 0
1
0
1
0
0
0
07
7
6

K k1 6
7
7
6 a
a
1
0
1

a
a
0
0
7
6
6 a
a
0
0
a
1 a 0 1 7
7
6
7
6
4 0
0
0
0
0
0
0
05
0
0
0
0
0
1
0
1
p
where a  1=2 2.
The restraint boundary conditions now are required to particularise
the stiness matrix to the truss in Fig. 8.5. These conditions are
ux A uy A ux D uy D 0
Thus, eliminating the appropriate rows and columns of the stiness
matrix we have
3
2
1 0
1
0
6 0 1
0
0 7
7
 k1 6
K
7
6
4 1 0 1 a
a 5
0

1 a

The work done by the applied loads, see equation (8.3), is


W P1 ux B P2 uy C
Thus the matrix equation for the equilibrium state of the truss is, see
equation (8.49),

KD
P1 P
where
2

ux B

6 u B 7
6 y
7
D 6
7 and
4 ux C 5
uy C

3
1
6 0 7
6
7
P 6
7
4 0 5
0:5

Matrix analysis of pin-jointed trussed structures

225

The solution of the matrix is obtained as


3
2
4:328
7
6 0
7
 1 P1 P P1 6
D K
7
6
k1 4 3:328 5
0:5
i

The strain levels, " , in each member can be obtained using equation
(8.9),
"i

l i0 l i
li

thus,
Member AB
"AB

P1
u B 0
kl y

Member BC
"BC

P1
P
ux C ux B 1
kl
EA

Member CD
"CD

P1
P
uy C 0:5 1
kl
EA

Member AC
P
1
P
"AC p1 p ux C uy C 1:414 1
EA
k 2l 2
It is evident from the above that the maximum strain occurs in
member BC and the allowable maximum load will be governed by
the maximum strain in that member. Therefore,
P
1:414 1 0:3"0
EA
and
0:3  1:5  103  2:1  105  5:89  103
394 kN
1:414
The forces in the various members can be evaluated from the strains
P1

as

N i "i EA

226

Energy methods in structural mechanics

hence
N AB 0 kN;

N BC 394 kN tens:

N CD 197 kN tens:

N AC 557 kN comp:

Now, taking the calculated value of P1 , the displacement of the nodes


can be evaluated from the matrix solution. Thus,
uy B 4:14 mm;
ux C 3:18 mm
uy C 0:48 mm

8.4

Matrix analysis of pin-jointed space structures

The presentation in sections 8.1 and 8.2 has been concerned with the
analysis of plane pin-jointed trusses, i.e. structures which were envisaged to be entirely contained in a plane together with the applied
loads. This is the case when the axes of the elements are actually contained in a plane and any geometric or mechanical feature of the
elements themselves is symmetric with respect to the same plane.
Now we intend to illustrate the extension of the direct stiness
method to deal with structures that exist in three-dimensional spacetrusses. The approach is based directly on the direct stiness method
because all the energy concepts that underlie this method remain
exactly the same as in the case of plane structures. A space pin-jointed
structure is dened to be a collection of pin-jointed bars in the threedimensional Euclidean space, as exemplied in Fig. 8.6.
The geometry and loading of the space-truss are described by reference to a general three-dimensional system of Cartesian axes xyz.
The local axes of the system are denoted by , as shown in Fig. 8.7.
Similar to the method for the two-dimensional truss, we relate here
the behaviour of the individual members of the spacetruss, which are
subjected to tension or compression loading, to the displacement of the
nodal joints. The vector containing the displacements of the end nodes
P

y
x

Fig. 8.6

Matrix analysis of pin-jointed trussed structures


K

227

Fig. 8.7

of the generic i-th element in the general reference frame xyz is


3
2
ux J
6 uy J 7
7
6
6 u J 7
7
6 z
i
8:56
D 6
7
6 ux K 7
7
6
4 uy K 5
uz K
while in the local reference frame  we write
3
2
u J
6 u J 7
7
6
6 u J 7
7
6 
i
d 6
7
6 u K 7
7
6
4 u K 5
u K

8:57

As in the previous presentation, the relationship between d i and Di


can be expressed as
d i Ri Di

8:58
i

where the transformation matrix R is given by


3
2 i
0
0
0
x iy iz
7
6 i
6 x iy iz
0
0
0 7
7
6
i
i
7
6 i


0
0
0
x
y
z
7
6
Ri 6
i
i
i 7
6 0
0
0 x y z 7
7
6
7
6
0
0 ix iy iz 5
4 0
0
0
0 ix iy iz

8:59

228

Energy methods in structural mechanics

ik are direction cosines of angles between the local axes , ,  and the
general axes x, y, z.
It easy to verify that, similar to the plane truss, in the three-dimensional space the transformation matrix (8.59) is orthogonal, i.e.
Ri T Ri I

8:60

Since we are dealing with structures composed of bars, the only internal forces that occur on account of the deformation of the elements are
the axial forces NJ and NK. However, it may be useful to collect
the nodal forces in the local reference frame in a six-elements array,
f i , that is
3
3 2
2
0
0
7 6 0 7
6
0
7
7 6
6
7
7 6
6
6 F J 7 6 NJ 7
i
7
6
7
6
8:61
f 6
76 0 7
0
7
7 6
6
7
7 6
6
5 4 0 5
4
0
F K
NK
so that in the local reference frame the relationship between the nodal
forces f i and the nodal displacements d i can be written in the
following form
f i K i d i
with

60
6
i6
60
EA
K i i 6
l 6
60
6
40
0

8:62
0

0
0

0
1

0
0

0
07
7
7
0 1 7
7
0
07
7
7
0
05

0 1

8:63

The expression in equation (8.63) represents the stiness matrix of a


generic bar in the local reference frame and it corresponds precisely
to the expression of the stiness matrix, equation (8.28), obtained in
for plane pin-jointed structures. Inspection of matrix equation (8.63)
reveals that it is symmetric, as we would expect.
The vector F i which collects the nodal forces of the generic bar in
the general reference frame is expressed as
F i Ri T f i

8:64

Matrix analysis of pin-jointed trussed structures

229

so that from equations (8.58) and (8.62) we have


F i Ri T K i Ri Di K i Di

8:65

i

where K is the stiness matrix of the generic bar in the general reference frame, which again corresponds to the expression in equation
(8.36) obtained in the earlier two-dimensional treatment. In detail we
can write
2

ix

6
6 i i
6 x y
6
i6
6 ix iz
EA
i
K i 6
2
l 6
6 ix
6
6
6 ix iy
4
ix iz

ix iy

ix iz

ix

ix iy

iy

iy iz

ix iy

iy

iy iz

iz

ix iz

iy iz

ix iy

ix iz

ix

ix iy

iy

iy iz

ix iy

iy

iy iz

iz

ix iz

iy iz

ix iz

7
7
iy iz 7
7
2
7
iz 7
7
7
ix iz 7
7
7
iy iz 7
5
i2
z

(8.66)
The solution of the spacetruss problems proceeds then exactly as in
the case of planetruss problems shown in section 8.2: the overall joint
 is obtained by assembling the contributions from
stiness matrix K
all the bars and the elastic equilibrium equation takes the now-familiar
form

KD
P 0
8:67
D is a 3n dimension array which contains all the degrees of freedom of
the space truss under consideration, n being the overall number of
 is obviously
joints. The dimension of the joint stiness matrix K
3n  3n and P is the load vector, i.e. a 3n dimension array containing
all the components of the forces applied to the joints which, for the
sake of generality, can be arranged to include the reactions of the
 is symmetric for the energy
constraints. The joint stiness matrix K
reasons explained in section 8.2.
At this stage in the analysis the constraint conditions can be applied
exactly as in the earlier work , i.e. by distinguishing two sub-arrays in
the nodal displacement vector D, DI and DII . These sub-arrays
correspond to the degrees of freedom that are active and to the degrees
of freedom which are prevented by the restraints, respectively. The
remainder of the procedure consists in condensing the system of equations (8.67) accordingly (see section 8.2). Once the reduced system has
been solved, we can in a straightforward manner calculate the restraint
reactions, the strains and the actually carried forces in any member of
the spacetruss.

230

Energy methods in structural mechanics

A
15 kN
1m

45

45

45

45

30 kN

Fig. 8.8

Suggested exercise problems


1.

The structure shown in Fig. 8.8 is to be analysed as a pin-jointed


frame. Calculate the vertical and horizontal deections at A and
1m
Pb
A

1m

C
D

Fig. 8.9

Matrix analysis of pin-jointed trussed structures

2.

231

D, Av , Ah , Ev , and Eh due to the applied loads. All the
structural members have a cross-sectional area of 200 mm2 . The
modulus for steel is 205 kN=mm2 .
[Ans: Av 15:5 mm, Ah 45:7 mm, Ev 28:0 mm,
Eh 40:5 mm]
A frame made from aluminium alloy, analysed as having pinjoints, is shown in Fig. 8.9. Calculate the forces and strains in
the members BC, BD due to the application of the vertical load
Pb at B. All members have the same cross-sectional area of
150 mm2 and the modulus of aluminium is 80 kN/mm2 .
[Ans: "BC 0:55  103 , NBC 6:62 kN (comp.),
"BD 1  103 , NBD 11:8 kN (comp.)]

9.

Matrix analysis of rigid-jointed


framed structures

Introduction

Chapter 8 has provided an introduction to the topic of the analysis of


pin-jointed frames. In that kind of structure the presence of pin-joints
allows the component members of the structure to undergo relative
rotation as the structure deforms under the action of the applied external loads. That possibility of relative rotation eliminates bending in the
components and they experience only axial tension or compression
loading.
The present chapter extends the methodology established for pinjointed trusses to encompass the analysis of rigid-jointed frames. The
chapter essentially follows the order of presentation in chapter 8 and
rst introduces the mechanics and energy formulation underlying the
analysis of a typical frame. The equations are then formalised in matrix
notation in section 9.2, thus establishing the basis for the automated computer analysis of frames that can vary from two to several thousand components members. The procedure is exemplied by application to a frame
with the geometry similar to that considered in section 8.3, but now with
rigid joints that prevent any relative rotation of the members of the frame.
The methodology for plane frames is extended to three-dimensional
frames and the chapter concludes with some general aspects regarding
the inclusion of distributed loading, such as member self-weight, and
restrained displacements in the frame analysis method.

9.1

Plane rigid-jointed frame structures as systems


of beams

In this section we shall extend the concepts introduced in the previous


section and apply them to rigid-jointed frame structures, with the aid
of the example two-dimensional system shown in Fig. 9.1.
The geometry of this plane rigid-jointed frame is similar to that of the
pin-jointed structure shown in Fig. 8.1. However, in contrast with that

234

Energy methods in structural mechanics


D

45
30

A
P

Fig. 9.1

truss, in this case we note that the members are not free to undergo relative rotation at connecting joints, thus the denition of rigid joint ( joints
are sometimes called nodes). In particular any rotation is totally prevented at the nodes connected to the wall at B, C and D, while the connection at A preserves the relative angular spacing of the attached
members AB, AC and AD at A. In other words, with rigid joints all
the ends of members connected at the same joint must experience the
same rotation. From a practical point of view, the nodes at B, C and
D can be thought of as welded to rigid elements, while the node at A
can be fabricated by welding the ends of steel elements together.
In contrast to the bar members of the previous section, the elements
into which we can divide the rigid-jointed structure are able to carry
shear forces and moments, as well as axial forces. In fact, since the
members are not free to rotate at the joints, corresponding restraint
moments will occur at the ends.
Notwithstanding the apparent complexity of the rigid-jointed structure, we can again analyse its behaviour in accordance with Bernoulli's
model, introduced in chapter 4, by taking into account both the exural and axial stiness of the members. This requires that the slopes
at the nodal points are assumed as unknown variables together with
the displacements at the nodes.

Matrix analysis of rigid-jointed framed structures

235

uy
K

ux

Fig. 9.2

We should also remember that in the present framework, that is the


linear theory of elasticity, the axial behaviour does not interact with
the bending behaviour.
The unstable eects that may be caused by the compressive axial
forces in members will be examined in chapter 12.
To analyse the bending behaviour of any individual member, modelled as a beam, (which will be superimposed on the axial deformation
summarised by equations (4.5) and (4.8)), let us refer to Fig. 9.2.
Here the relationship between the components of the vectorial displacement of the end point K and their contribution to the displacement u K that occurs orthogonal to the axis of the beam is clearly
represented. From a mathematical standpoint, for the end points J
and K we have
u J ux J sin i uy J cos i
i

u K ux K sin  uy K cos 

9:1
i

9:2

where we assume that the sign convention for the transverse displacement u is according to the local reference frame  shown in Fig. 9.2.
In this local reference frame the bending deection of the beam can be
described as a function of the variables u J, u K, J and K by
means of the same procedure that led to equation (4.116). Therefore we
have


2u J J 2u K K 3
u 


3
2
3
2
li
li
li
li


3u J 2J 3u K K 2

i2 i
i
2
l
l
l
li
J u J

9:3

236

Energy methods in structural mechanics

and the bending elastic energy stored in the beam on account of u J,


u K, J and K is
l
1
EIu002
U
  d
2
i

6EI i
li

u2 J u2 K

12EI i
li

u Ju K

2EI i 2
 J 2 K JK
li
6EI i
li

u KK u KJ u JJ u JK


(9.4)

We can now superpose the axial and the exural behaviour of the beam
element under consideration and write the total elastic energy stored in
it on account of its deformation. By adding equations (4.8) and (9.4)
we have
Ui

1 EAi 0i
6EI i 2
i 2
l

u J u2 K
3
2 li
li

12EI i
li

6EI i
li

u Ju K

2EI i 2
 J 2 K JK
li

u KK u KJ

u JJ u JK

9:5

This expression for the strain energy relates to a typical member in the
frame. With particular reference to the elements composing the example structure shown in Fig. 9.1 we can write
Element BA
i

sin i

cos i

ux J

uy J

ux K

uy K

J

K

ux A

uy A

A

l 0BA l BA ux A

9:6

u J 0

9:7

u K uy A

9:8

Matrix analysis of rigid-jointed framed structures

U BA

237

1 EABA
6EI BA 2
2
u
A

uy A
x
3
2 l BA
l BA

2EI BA 2
6EI BA

A

uy AA
2
l BA
l BA

9:9

Element CA
i

sin i

cos i
p
3

308 12

0CA

ux J

uy J

ux K uy K

J

K

ux A

A

uy A

p
3
1
u A uy A

2 x
2

CA

9:10

u J 0

9:11

p
3
1
u K ux A
9:12
u A
2
2 y
 p
2
1 EACA
1
3
CA
u A uy A
U
2 l CA
2
2 x
p
2

6EI CA
1
2EI CA
3
uy A CA 2 A
CA3
ux A
2
2
l
l
p




6EI CA
1
3
u A A
ux A
9:13
CA2
2
2 y
l
Element DA
i

sin i
p
2

458

0DA

cos i
p
2
2
DA

ux J

uy J

ux K uy K J

K

ux A uy A) 0

A

p
p
2
2
u A
u A

2 x
2 y

9:14

u J 0

9:15

p
p
2
2
ux A
u A
u K
2
2 y

9:16

238

Energy methods in structural mechanics

DA

p
2
 p
2
2
u A
u A
2 x
2 y
p
2
 p
6EI DA
2EI DA
2
2
DA3
ux A
uy A DA 2 A

2
2
l
l
p




6EI DA
2
2
ux A
uy A A

9:17
DA2
2
2
l

1 EADA

2 l DA

Following the approach used in the analysis of the pin-jointed structure of the previous section, we will now write the total potential
energy of the present frame which, using equations (9.9), (9.13) and
(9.17) above, is evidently
 U W U BA U CA U DA W

1 EABA
6EI BA 2
2EI BA 2
2
u
A

u
A

 A
x
y
3
2 l BA
l BA
l BA
p
2
3
6EI BA
1 EACA
1
BA2 uy AA
u A uy A
2 l CA
2
2 x
l
p



2
3
6EI CA
1
2EI CA
uy A CA 2 A
CA3 ux A
2
2
l
l
p




6EI CA
1
3
CA2
uy A A
ux A
2
2
l
p
p
2
1 EADA
2
2
u A
u A

2 l DA
2 x
2 y
p
 p
2
6EI DA
2EI DA
2
2
ux A
uy A DA 2 A
DA3
2
2
l
l
p




6EI DA
2
2
u A
u A A

DA2
2 x
2 y
l
Px ux A Py uy A

9:18

The energy is seen to be a function of the variables ux A, uy A and


A only. As before, the values of the variables ux A, uy A and A
that render the total potential energy stationary are yielded by the

Matrix analysis of rigid-jointed framed structures

239

fullment of the following extremum conditions


p


3
q
EABA
EACA 3
ux A
BA ux A CA
uy A
qux A
4
4
l
l
p





3
6EI CA 1
6EI CA
1
ux A uy A CA2 A
CA3
2
2
2
l
l


EADA 1
1
6EI DA
ux A uy A DA3 ux A uy A
DA
2
2
l
l
p


DA 
6EI
2
A Px 0
9:19
DA2
2
l
q
6EI BA
6EI BA
BA3 2uy A BA2 A
quy A
l
l
p


3
EACA 1
uy A
ux A
CA
4
4
l
p



p

3
3
6EI CA 3
6EI CA
uy A
ux A CA2
A
CA2
2
2
2
l
l


EADA 1
1
6EI DA
uy A ux A DA3 uy A ux A
DA
2
2
l
l
p



6EI DA
2
A Py 0
9:20
DA2
2
l
q
2EI BA
6EI BA
2EI CA
BA 2A BA2 uy A CA 2A
qA
l
l
l
p

CA 
3
6EI
1
2EI DA
CA2 ux A
uy A DA 2A
2
2
l
l
p



6EI DA
2
2
ux A
u A 0
9:21
DA2
2
2 y
l
The extremum conditions once again represent the equilibrium equations of the structure under consideration. The unique solution of this
linear system of algebraic equations provides the magnitudes of the
displacements of the node A along the x and y axes and its rotation.
The substitution of these magnitudes in equations (4.122)(4.125),

240

Energy methods in structural mechanics

(4.130)(4.133), (4.7), (8.9), (9.1)(9.3) give a complete description of


the strains and the forces induced in each beam by the applied load.

9.2

Matrix formulation of the elastic equilibrium


equations for plane rigid-jointed frames

Following the same line of reasoning employed in chapter 8, we now


apply the direct stiness method of structural analysis to the rigidjointed frame shown in Fig. 9.1.
The rst step in the development of the analysis involves identifying the overall degrees of freedom of the structure. In this case
they are represented by both displacements and rotations of each
joint with reference to the general frame xy. The general displacement vector for the example frame is therefore a 12-element array,
that is
DT ux A uy A A . . . . . . . . . ux D uy D D
9:22
while the vector containing the displacements of the end joints of the
generic element is a 6-element array, i.e.
2
3
ux J
6 u J 7
6 y
7
6
7
6
7
J
i
6
7
D 6
9:23
7
u
K
6 x
7
6
7
4 uy K 5
K
As we have already seen in section 9.1, in order to take into account
the elastic behaviour of the individual element it is necessary to superimpose the exural stiness on the axial stiness. Therefore, in the
local reference frame all the displacements of the joints are relevant
to the model and we can write
3
2
u J
6 u J 7
7
6 
7
6
7
6
J
7
9:24
d i 6
6 u K 7
7
6 
7
6
4 u K 5
K

Matrix analysis of rigid-jointed framed structures

241

It can be easily shown for the generic beam that the relationship
between the joint displacement vector Di in the general reference
frame and the joint displacement vector d i in the local reference
frame is governed by the following transformation matrix
2

cos i

6
6 sin i
6
6
6 0
i
R 6
6
6 0
6
6 0
4
0

sin i

cos i

0
i

sin i

cos 

sin i

cos i

7
07
7
7
07
7
7
07
7
07
5

9:25

so that
d i Ri Di

9:26

Equation (9.26) has been derived bearing in mind equations (8.22),


(9.1) and (9.2) and remembering that the slopes J and K are
not aected in the transformation from the general reference frame
xy to the local reference frame . Notice that matrix (9.25) is
now a 6  6 square array because it determines the transformation
between two 6-element arrays, namely equations (9.23) and (9.24).
(Matrix (8.20) was a 4  2 rectangular array in order to relate the 4element array in equation (8.21) to the 2-element array in equation
(8.22).)
A noteworthy property of matrix (9.25) is that it is orthogonal, i.e. its
inverse coincides with its transpose as we have
Ri 1 Ri Ri T Ri I

9:27

where I is the unit matrix. Matrix (8.20) satises a similar condition


but from a formal point of view we normally refer to inverse matrices
only when dealing with square matrices.
At this point we can relate the vector of the forces f i , that arise at
the end points of the generic beam on account of the displacements
d i , to the displacements themselves by means of the stiness matrix
of the beam K i . This is accomplished by re-casting in matrix notation
equations (4.122)(4.125) and (4.130)(4.133), and adding the axial
stiness contributions contained in equation (8.28).

242

Energy methods in structural mechanics

We have therefore
2

2
EAi
7 6 li
7 6
7 6
7 6
7 6 0
7 6
7 6
7 6
7 6
7 6
7 6 0
7 6
76
7 6
7 6 EAi
7 6 i
7 6
l
7 6
7 6
7 6
7 6 0
7 6
7 6
7 6
5 4
0
3

6 F J
6
6
6
6 F J
6 
6
6
6
6
6 MJ
6
6
6
6
6 F K
6
6
6
6
6 F K
6 
6
6
4
MK

12EI i
l

i3

6EI i
l

i2

6EI i
l

i2

EAi
li

6EI i

i3

i2

6EI i
l

i2

4EI i
li

12EI i
l

EAi
li

2EI i
li

0
0

12EI i
li

6EI i
li

0
12EI i
li

6EI i
li

32

76 u J
76
76
6
6EI i 7
6
i2 7
76 u J
l 76
76
6
2EI i 7
76
76 J
i
76
l
76
76
76
6
0 7
76 u K
76
76
6EI i 76
76 u K
2
76 
li
76
76
4EI i 54
K
li

7
7
7
7
7
7
7
7
7
7
7
7
7
7
7
7
7
7
7
7
7
7
7
7
5

(9.28)

or, in compact form


f i K i d i

9:29

As the transformation matrix Ri also governs the relationship


between the nodal forces vector f i and F i in the local and general
reference frames, respectively, that is
f i Ri F i

9:30

Equation (9.30) can be rewritten as


F i Ri T f i Ri T K i d i Ri T K i Ri Di

9:31

Once again we can regard the term


K i Ri T K i Ri

9:32

as the stiness matrix of the generic element in the general reference


frame xy. Its expression is displayed as equation (9.33).
At this point the solution of our problem becomes straightforward:
we can once more particularise equation (9.32) to obtain the stiness
matrices of the elements BA, CA and DA and, by identifying the
position of the joints J and K in the general ordering system, we can
assemble all the concurrent contributions in order to get the expression
 namely
of the overall joint stiness matrix K,
 K BA K CA K DA
K

9:34

20 i
1
EA
2 i
6 B i cos 
C
6B l
C
6 @ 12EI i
A
6
2 i
6 i3 sin 
6
l
6
6
6
6
6
6
sym
6
6
6
6
6
6
6
6
sym
6
6
i

6
K 6
6
6
6
6
sym
6
6
6
6
6
6
6
6
6
6
sym
6
6
6
6
6
6
4
sym
sym

sym

sym

sym

sym

cos i

sym

sin i

4EI i
li

li

2
li

6EI i

6EI

sym

0 i
1
EA
i
i
sin

cos

i
B l
C
B
C
@ 12EI i
A
i
i
sin

cos


i3
l
0 i
1
EA
2 i
sin

B li
C
B
C
@ 12EI i
A
2 i
i3 cos 
l
sin i

sym

sym

0 i
1
EA
2 i
cos

B li
C
B
C
@ 12EI i 2 i A
i3 sin 
l

2
li

6EI i

1
EAi
i
i

sin

cos

B li
C
B
C
@ 12EI i
A
i
i
i3 sin  cos 
l

1
EAi
2 i
i cos 
B l
C
B
C
@ 12EI i 2 i A
sin


3
li

li

6EI i

cos i

sym

0 i
1
EA
i
i
sin

cos

B li
C
B
C
@ 12EI i
A
i
i
i3 sin  cos 
l
0 i
1
EA
2 i
sin

B li
C
B
C
@ 12EI i
A
2 i
i3 cos 
l

1
EAi
i
i
i sin  cos 
B l
C
B
C
@ 12EI i
A
i
i
sin

cos


i3
l
0
1
EAi 2 i
sin


B li
C
B
C
@ 12EI i
A
2 i
i3 cos 
l

0
6EI i

2
li

7
7
sin i 7
7
7
7
7
7
7
7
7
6EI i
i 7
cos

7
2
7
li
7
7
7
7
i
7
2EI
7
7
i
7
l
7
7
7
7
7
6EI i
i 7
sin

7
2
7
li
7
7
7
7
7
7
i
7
6EI
i7
i2 cos  7
7
l
7
7
7
7
5
4EI i
i
l
9:33

Matrix analysis of rigid-jointed framed structures


243

244

Energy methods in structural mechanics

The load vector must be enlarged to a 12-element array to contain


couples directly applied to the joints (these, however, in our particular
example are absent). We have
PT Px A Py A

0 0

...

...

0 0

9:35

and from a formal point of view we have the recognisable system



KD
P 0

9:36

that in the present example is formed by 12 linear equations in the 12


unknown degrees of freedom ordered in the array D.
Finally, we must take into consideration the restraints represented
by the xed ends at B, C and D by setting to zero, as in the previous
section, the terms corresponding to the inactive degrees of freedom,
that is suppressing them. The reader will nd that the core of the
system in equation (9.36) is then reduced to a system of three linear
equations in the unknowns ux A, uy A and A, i.e.
3 2
3
2
Px A
ux A
7 6
7
6
9:37
KI 4 uy A 5 4 Py A 5
A

which turns out to be identical to the system formed by equations


(9.19), (9.20) and (9.21). On the same basis as in section 8.2, we nd
that the matrix KI is symmetric and positive denite.
Once again the usefulness of the equilibrium equations in the form of
equation (9.36) is evident: it can be directly applied to any number of
non-aligned members making up a plane structural system (or, as we
shall see later, a spatial structural system as well). The overall joint
stiness matrix, equation (9.34) may always be obtained by referring
the local stiness matrix of each member, equation (9.28), to the
common system axes xy and adding the corresponding contributions.
By assigning appropriate boundary conditions at each joint, the system
in equation (9.36) allows us to determine all the magnitudes of the
unknown degrees of freedom and, consequently, strains and forces
in each member. All this by means of a straightforward and practically
automatic procedure.
As in the case of the pin-jointed plane frame previously examined,
from a physical point of view the set of equations under consideration
requires that the sum of the elastic reactions from the end joints of each
element balances the applied forces at any joint of the structure. In the
present example, this is the force P at A.

Matrix analysis of rigid-jointed framed structures

9.3

Numerical example: analysis of a rigid jointed


frame

245

The pin-jointed structure shown in Fig. 8.3 is an idealisation of the


situation in which the connections of the members cannot transmit
moment from one member to another. However, the majority of
actual constructions would involve attaching the members to each
other either by welding or bolting so that they indeed have the capacity
to transfer moment. This condition is illustrated in Fig. 9.3 in which
the members are welded to cover plates that ensure rigidity to the
joint so that all the members meeting at that joint experience a
common rotation. The structure can now be classied as a plane
frame, according to the denition in section 9.1.
The objective in the present example is to exemplify the application
of the direct stiness method and to compare the forces and deformation induced in the frame by the loads to the corresponding values for
the truss.
The direct stiness method, section 9.2, is, as for the truss, essentially
concerned with assembling the stiness matrix of the complete structure from the component stiness matrices. Thus we use equations
(9.25) and (9.28) to provide the component stiness matrices. This is
exemplied below for member AB.
Member AB
i AB


2
P2

P1

B
C

A = 589 x 10 3 mm
I = 461 x 10 3 mm

Fig. 9.3

246

Energy methods in structural mechanics


2

0:02
0
6 0
:
1 96
6
6
:71
6
30
0
K AB E 6
6 0:02
0
6
6
4 0
1:96
30:71
0

30:71
0
6:14  104

0:02
0
30:71

0
1:96

0
30:71

1:96

30:71

3:07  104

0:02

0
0

3
30:71
7
0
7
7
4
3:07  10 7
7
30:71 7
7
7
5
0
4
:
6 14  10

this stiness matrix is associated with the displacement vector


DAB T ux A

uy A A

ux B uy B BT

Preparation of similar stiness matrices for members BC, CD and AC


can provide that basis for the stiness matrix for the complete frame.
The equilibrium equations for the frame are then particularised by
imposing the boundary constraint conditions, i.e.
ux A uy A A 0
For the sake of brevity in the presentation of this example, these constraint conditions are taken into account in writing the component
stiness matrices. Thus, for member AB, the stiness matrix, given
the constraints at node A, is
2 :
3
0 02
0
30:71
6
7
K AB E 4 0
1:96
0
5
4
:
:
30 71
0
6 14  10
This component of the frame stiness matrix is now associated with
the displacement vector
DAB T ux B uy B

BT

Member BC
i BC 0
2

1:96
6 0
6
6
6 0
BC
K E 6
6 1:96
6
6
4 0
0

Member CD
i CD


2

0
0:02
30:71
0
0:02
30:71

0
30:71
6:14  104
0
30:707
3:07  104

3
1:96
0
0
0
0:02
30:71 7
7
7
0
30:71 3:07  104 7
7
7
1:96
0
0
7
7
:
0
0:02
30 71 5
4
0
30:71 6:14  10

Matrix analysis of rigid-jointed framed structures

247

2 :
3
0 02
0
30:71
6
7
K CD E 4 0
1:836
0
5
28:83
0
5:766  104
Member AC

4
2

i AC

3
0:70
0:69
10:86
7
6
0:70
10:86 5
K AC E 4 0:69
10:86 10:86 4:34  104
Following equation (9.34), the stiness matrix for the frame, with the
boundary constraints incorporated is
 K AB K BC K CD K AC
K
The loading vector is similar to that for the truss taking account of the
formulation of the stiness matrix
3
2
1
6 0 7
7
6
7
6
0 7
P1 6
7
6
P
7
E6
6 0 7
7
6
4 0:5 5
0
Thus the vector of the displacements of the unconstrained nodes can be
obtained from
 1 P
D K
The result has been calculated using the value of the load P1 evaluated
for the truss, to give
3
3 2
2
3:95 mm
ux B
6 u B 7 6 0:021 mm 7
7
7 6
6 y
7
7 6
6
6 B 7 6 0:93  103 7
7
76
6
6 u C 7 6 3:02 mm 7
7
7 6
6 x
7
7 6
6
4 uy C 5 4 0:45 mm 5
0:70  104
C

248

Energy methods in structural mechanics

The axial loads in the members can be obtained from these displacements using equation (8.24), i.e.
N AB uy B 

EA 0:021  2:1  105  5:89  103

l0
3000

8:7 kN

EA
l
:
:
:
3 02 3 95  2 1  105  5:89  103
383 kN

3000
EA
N CD uy C 
l
:
:
0 45  2 1  105  5:89  103
185 kN

3000
1
EA
N AC p ux C uy C  p
2
2l
:
:
:
3 021 0 45  2 1  105  5:89  103
530 kN

6000
In comparison with the corresponding results for the truss it may be
seen that the deections at the nodes and the axial forces in the members
are slightly less for the frame than for the truss. This eect is due to the
bending action in the members in the frame adding restraint to the
actions of the forces to induce the deections. However, for the slender
members in this frame the eect of the bending actions are fairly small
and it is evident that the analysis assuming the structure would act as a
truss would be suciently accurate for engineering purposes.
A more signicant eect of the structure acting as a frame is illustrated below. It is evident from the geometry of the structure modelled
as acting as a truss that if the bracing member AC were to be damaged
or severed the truss would collapse under the action of the loads. In
eect the truss would deteriorate to act as a mechanism. However,
the actual structure because of the bending stinesses of the members
could withstand the eects of the loading, but at the penalty of
increased deections.
Suppose for the frame the brace AC were removed, the stiness
matrix for the structure would then be
 K AB K BC K CD
K
N BC ux C uy B 

with the boundary constraints again incorporated in the formulation


 The loading vector is unaected by this
of the stiness matrix K.

Matrix analysis of rigid-jointed framed structures

249

change to the stiness matrix, i.e.


3
2
1
6 0 7
7
6
7
6
7
6
0
P
7
P 1 6
6
E6 0 7
7
7
6
4 0:5 5
0
Thus the vector of the modied displacements is obtained from
 1 P
D K
The result has been calculated using the value of the load P1 evaluated
for the truss, to give
3
2
3 2
ux B
74:07 mm
6 u B 7 6 0:54 mm 7
7
6 y
7 6
7
6
7 6
:
6 B 7 6 0 013 7
7
6
76
6 u C 7 6 73:66 mm 7
7
6 x
7 6
7
6
7 6
4 uy C 5 4 0:062 mm 5
0:022
C
Evidently, the structure is now much more exible and deects signicantly more due to the removal of the bracing member. The reader
can conrm that the loads in the members are
N AB 223 kN;

N BC 169 kN

N CD 25:6 kN

Removal of the brace means that the actions of the loading is resisted
primarily by the bending stiness of the members and this means that
the axial loads induced in the members are signicantly reduced.
Generally, for less slender structural members than considered in
this example the eect of including the bending stiness will reduce
the axial loads in the members compared to those calculated as if the
structure is a truss.

9.4

Matrix analysis of space frames

Once one has understood the matrix treatment of space pin-jointed


structures, i.e. spacetrusses, dealing with the most general case of
space frames becomes quite straightforward. In general a space- frame
structure consists of a three-dimensional collection of rigidly-jointed
beams, each of which is subjected to axial and transverse loading, as
exemplied in Fig. 9.4.

250

Energy methods in structural mechanics


q
P

q
P

Fig. 9.4

The key point of the direct stiness method consists, as before, in


characterising the mechanical behaviour of a generic element in its
local reference frame and successively in establishing the relationships
that allow us to express the terms in the stiness matrices in the general
reference frame. Once this is done, we can assemble the contributions
from all the elements of the structure in the combined stiness matrix
and solve the general system of equations with reference to the nodal
loads (both directly applied and equivalent), provided the constraint
conditions have been taken into account.
In contrast to the case of pin-jointed structures, it is necessary to
take into account the exural and torsional behaviour of the generic
i-th element. The vector collecting the displacements of the end
nodes in the general reference frame is therefore enlarged to a 12elements array in order to encompass also the rotations about the
axes xyz. Thus, we have
D i T ux J uy J uz J x J y J z J . . . . . . z K
9:38
In the local reference frame  it is
d i T u J u J u J  J  J  J . . . . . .  K
9:39
The relationship between d i and Di is governed by a 12  12 transformation matrix, Ri , according to the usual product
9:40
d i Ri Di

Matrix analysis of rigid-jointed framed structures

The transformation matrix Ri is given by


3
2
L 0 0 0
6 0 L 0 0 7
7
6
Ri 6
7
4 0 0 L 0 5
0

251

9:41

where the sub-matrices L and 0 are dened by


2
3
3
2
ix iy iz
0 0 0
6 i
7
7
6
i
i 7
L 6
0 4 0 0 0 5
4 x y z 5
0 0 0
ix iy iz

9:42

ik are direction cosines of angles between the local axes , ,  and the
general axes x, y, z.
It simple to verify that once again the transformation matrix (9.42) is
orthogonal, i.e.
Ri T Ri I

9:43

The vector that contains the nodal forces in the local reference frame is
a 12-element array. It includes the internal forces that arise at each end
of the beam on account of the deformations, namely the axial forces,
the shear forces along the axes  and , the bending moments about
the axes  and  and the twisting moment about the axis , as shown
in Fig. 9.5.
Thus we have
f i T F J F J F J M J M J M J . . . . . . M K
9:44
M

M M

N
S

S
M

N
M

Fig. 9.5

252

Energy methods in structural mechanics

and in the local reference frame the relationship between the nodal
forces f i and the nodal displacements d i can be written in the
following form
f i K i d i
with

"
K i

ki11

ki21

9:45
ki12

The sub-matrices kihk are dened by1


2
EAi
0
0
6 i
6 l
6
12EI i
6
0
6 0
3
6
li
6
6
12EI i
6 0
0
3
6
li
ki11 6
6
6 0
0
0
6
6
6
6EI i
6
6 0
0
i2
6
l
6
4
6EI i
0
0
2
li
2
EAi
0
0
6 i
6 l
6
12EI i
6
0
6 0
3
6
li
6
6
12EI i
6 0
0
3
6
li
ki22 6
6
6 0
0
0
6
6
6
6EI i
6
6 0
0
2
6
li
6
i
4
6EI
0
i2
0
l
1

9:46

ki22
3
0

0
GJ i
li

6EI i
li

4EI i
li

0 7
7
6EI i 7
7
7
i2 7
l 7
7
0 7
7
7
7
0 7
7
7
7
7
0 7
7
i7
4EI 5

9:47

li
3

0
0
0
GJ i
li

0
0
6EI i
li

4EI i
li

7
7
6EI 7
7
i2 7
l 7
7
7
0 7
7
7
7
0 7
7
7
7
7
0 7
7
7
4EI i 5
li
i

9:48

Note that the relationship between the twisting moment, Mi , and the corresponding
change of rotation, i , can be written as Mi GJ=l i i . J is the polar moment of area
of the section of the structural member.

Matrix analysis of rigid-jointed framed structures

253

EAi
6 li
6
6
6
6 0
6
6
6
6 0
6
6
i
i
k12 k21 6
6
6 0
6
6
6
6 0
6
6
6
4
0

3
0

12EI
li

6EI i
li

12EI i

0
0

l
0

i3

6EI i
l

i2

GJ i
li

0
0
6EI i
li

2EI i
li

7
7
7
6EI 7
i2 7
l 7
7
7
0 7
7
7
7
7
0 7
7
7
7
0 7
7
7
7
2EI i 5
li
9:49
i

Inspection of the member stiness matrix (9.46) reveals that it is symmetric, like all the stiness matrices encountered up to now.
The subsequent procedure remains the same as in the previous cases:
we can express the vector F i that collects together the nodal forces of
the generic beam in the general reference frame in the following form
F i Ri T f i

9:50

and from equations (9.40) and (9.45) we have


9:51
F i Ri T K i Ri Di K i Di
i

where K is the stiness matrix of the generic beam in the general
reference frame. At this point we can obtain the overall combined sti by assembling the contributions from all the beams and
ness matrix K
the elastic equilibrium equation takes the usual form, equation (9.36)

KD
P 0
Of course, as we have now included all the rotations of the joints in the
degrees of freedom of the structure, D is a now 6n dimension array, n
being the overall number of joints, and the dimension of the joint sti is 6n  6n. P is the load vector, i.e. a 6n dimension
ness matrix K
array containing both the components of the forces directly applied
to the joints and the equivalent joint loads. The concept of equivalent
joint load will be claried in the next section.
 is,
It goes without saying that the combined stiness matrix, K,
once more, symmetric (see section 9.2). Once the expression (equation
(9.36)) is formulated, constraint conditions can be applied and the
system of equations can be solved.

254

Energy methods in structural mechanics

It must be stressed that for space structures the level of computational eort tends to increase considerably with the number of
elements and the use of a computer based automatic procedure
becomes essential for the vast majority of practical cases. However,
the analysis of many real cases can be successfully conducted by reducing three-dimensional problems to two-dimensional ones. This can be
achieved by taking account of symmetries in the geometry and loading
in the structure and treating the structure as an assemblage of plain
frames.

9.5

Equivalent joint loads

At this point of the presentation we point out that in the examples of


the sections 8.18.3 and 9.19.3 the self-weights of the individual
members have been neglected. This is fairly common practice in the
analysis and design of steel frames and trusses, where the loading
and strains induced by the distributed self-weight are usually very
small in comparison with the eects of the applied forces.
However, for other types of structures these weights are not negligible and we should include them in the analysis. Moreover, not all
external loads are actually applied at the joints and we may nd it
necessary to take into account several types of lateral loading, as
well as the eects of temperature changes and restraint displacements.
As the matrix stiness method introduced in sections 8.2 and 9.2
uses the joint displacements as the degrees of freedom of the system,
in order to reach the general matrix expression, equation (9.36),

KD
P 0
we have to replace the loads acting on the members by equivalent loads
acting at the joints. These loads are called equivalent joint loads and are
added to the actual joint loads to obtain the combined joint load vector
in the same form of equations (8.45) or (9.35). Thereafter, the structure
can be analysed using the matrix stiness method to evaluate the eect
of the combined joint loads.
Quite naturally, we can evaluate the vector of the equivalent joint
loads making reference once again to the standard energy approach.
According to the general condition of equilibrium, equation (1.28),
this is easily accomplished by equating the work of the actual loads
to the work of the required equivalent nodal forces for any virtual displacement of the degrees of freedom, i.e. the end nodes of the element.
For example, let us consider the beam element shown in Fig. 9.6.
This beam is subject to an evenly distributed load q acting in the direction  of the local reference frame  and we want to determine the
equivalent joint loads, i.e. those forces that are equivalent in energy

Matrix analysis of rigid-jointed framed structures

255

Fig. 9.6

to the actual load q with respect to the joint displacements. In other


words, if f^ is the vector of the equivalent nodal forces acting on
the beam in the local reference frame , we require that
^

W q W f 0

9:52

that is
l

q u  d d T f^ 0

9:53

As the displacement eld u  at the left hand side of equation


(9.53) is a function of the nodal displacement vector d by means of
equation (9.3), and the same is obviously true if we take the variation
of both terms, i.e. u  and d, then equation (9.53) can be immediately solved with respect to f^ and yields
2
3
0
6 ql=2 7
6
7
6 2
7
6 ql =12 7
7
f^ 6
9:54
6
7
0
6
7
6
7
4 ql=2 5
ql 2 =12
Of course, equation (9.54) is the vector that contains the equivalent
forces acting on the beam, and the equivalent joint loads vector p^ is
readily obtained by reversing these actions in direction. We have
3
2
0
6 ql=2 7
7
6
7
6
6 ql 2 =12 7
7
9:55
p^ 6
7
6
0
7
6
7
6
4 ql=2 5
ql 2 =12

256

Energy methods in structural mechanics

ql 2
12

ql 2
12
ql
2

ql
2

ql 2
12

ql 2
12
ql
2

ql
2

Fig. 9.7

With reference to Fig. 9.7, it is worth noting that the equivalent


nodal forces f^ coincide with the xed-end actions of the element,
that is with the restraint reactions on the beam subject to the distributed load q and considered rigidly xed at both ends.
This property applies to all types of lateral loading and it is useful to
view the table in Fig. 9.8, where the actual and equivalent loading for
some common types of applied loading are shown.
It is also worth noting that, once we have added the equivalent joint
loads to the actual joint loads in the combined joint load vector, the
direct stiness analysis of the frame under consideration produces the
same nodal displacements that we nd in the frame under the actual
lateral loads. However, the displacements and the internal forces of
the generic element must be obtained by superposing the displacements
and the related internal forces that derive from this analysis to those
corresponding to the xed-end element under the actual lateral loading.
This follows from the fact that the procedure developed above for constructing the equivalent joint load vectors has been carried out by establishing the equivalence with respect to the equilibrium of the only
degrees of freedom of the frame, i.e. the displacements of its joints.
In order to illustrate the procedure with reference to a simple frame,
consider the plane structure shown in Fig. 9.9.
This frame is composed of two straight beams. The extremity A of the
rst beam is xed, the node B is horizontally supported and the end C is
free. Suppose that a constant distributed load qy acts on the beam AB
in the direction shown and that two vertical forces P1 and P2 act at the
mid point of the span BC and at the free end C, respectively.
According to the direct stiness method, it is immediately recognised
that overall we have nine degrees of freedom, i.e. the displacements ux

Matrix analysis of rigid-jointed framed structures

FA

FA
w

MA

257

MB

FA = wa 1 ( al )2 + 12 ( al )3
FB = wa ( al )2 12 ( al )3

l=a+b

MA = wa
12 16 ( al ) + 6 ( al )2
24
2

MB = wa
8 ( al ) 6 ( al )2
24

Fb2 (3a + b)/l3

Fab2/l2

Fba2/l2

Fa2 (3b + a)/l3

pa (a + 2b)/2l

pa2/2l

Fig. 9.8

and uy and the rotation  of each joint. Only ve of them, i.e. uy B,


B, ux C, uy C, C, are not prevented by the applied constraints.
However, only the force P2 is applied directly to a joint. It is therefore
necessary to nd out what the eects of the other loads are with respect
to the degrees of freedom of the system. In order to do so we can restrain
the structure against joint displacement by introducing imaginary joint
restraints where necessary, as shown in Fig. 9.10.
Next, the constraint reactions produced by the actual member
loads (i.e. the loads that are not directly applied to the joints) are calculated. These reactions constitute the required set of xed end actions. By
reversing these xed end actions in direction we obtain a set of forces
and couples that we can consider equivalent to the member loads,
that is the equivalent joint loads. This step is shown in Fig. 9.11.

258

Energy methods in structural mechanics


q
A

C
B
P2

P1
l/2

l/2

Fig. 9.9

The combined joint loads vector is obtained by adding the actual joint
loads, which in this case is only the force P2 , to the equivalent joint
loads, as shown in Fig. 9.12.
It is straightforward to show that, once we have applied the direct
stiness method to solve the situation pictured in Fig. 9.12, everything
is known about the structure subject to the actual loads in Fig. 9.9. In
fact, the superposition of the loads in Fig. 9.11(a) and the combined
loads in Fig. 9.12 produces the loads in Fig. 9.9. At the same time
the superposition of the displacement elds does not aect the joint
displacements given by the stiness analysis of Figs 9.10 and 9.11(a),
as the joint displacements of the frame shown in Fig. 9.10 must be zero.

Fig. 9.10

Matrix analysis of rigid-jointed framed structures


qh2
12

259

qh
2
qh2
12

qh
2

q
P1

P1

qh2
12

qh
2
qh2
12

P1l
8

P1l
8

P1

Pl

81

P1l
8

qh
2
P1

P1

(a)

(b)

Fig. 9.11

qh
2
qh2
12

qh2
12

qh
2

P1l
8

P1l
8

P1

P1

+ P2

Fig. 9.12

9.6

Restraint displacements

The eects of restraint displacements can be readily incorporated


into the analysis of a framed structure by the direct stiness method.
Generally, restraint displacements consist of known translations or
rotation of the restraints. In order to illustrate this point with an
example, we again refer to the framed structure of Fig. 9.1 and apply
a vertical displacement u, at the xed node B. Such a displacement
could be due to an inelastic yielding of the built-in support, as shown
in Fig. 9.13.
The analysis of such a condition can be eected easily by recalling
the partitioned form, equation (8.50), of the general system equation

260

Energy methods in structural mechanics


D

B
u

Fig. 9.13

(8.49), i.e.
 
KI
KI;II T

KI;II
KII



DI
DII

PI

PII

This is an approach we have already considered in section 8.2. We


recall that we can always rearrange the full set of equations of our
problem to include in the sub-array DI the degrees of freedom of
those joints that are not restrained, and the degrees of freedom of
the restrained joints are placed in the sub-array DII . PI and PII
are the sub-arrays in which the applied loads and constraints reactions
are included, respectively. In the present example we have
2

ux A

6
7
7
DI 6
4 uy A 5

9:56

A
T

DII ux B uy B B ux C uy C C ux D uy D D

and all the displacements of the nodes C and D are set to zero, while for
the node B we have ux 0, uy 
u and  0.
Consequently, the system in equation (8.50) can be divided into the
following two expressions
KI DI PI KI;II DII
9:57
PII KI;II T DI KII DII

9:58

Matrix analysis of rigid-jointed framed structures

261

ui

Ri

Fig. 9.14

The system (9.57) is, as before, disjoint from equation (9.58), as the
sub-array DII comprises known values of the displacement. Therefore, the problem of elastic equilibrium for the structure shown
in Fig. 9.13 is reduced to the system, equation (9.57), of three
equations for the three unknown degrees of freedom contained in
the sub-array DI , which are the components of displacement and
the rotation of the loaded joint A. The substitution of these values in
the relationships in equation (9.58) yields the value of the restraint
reactions PII .
We can also take into consideration a linearly elastic displacement of
the restraints, that is a displacement of the restraint that is proportional to the restraint reaction, for example
ui J cRi J

9:59

where c is a constant and Ri is the reaction of the restraint of a generic


node J in the direction of the i-th axis, as shown in Fig. 9.14.
This case can also be treated in a straightforward manner. In fact, it
suces to notice that the reaction Ri J is represented in the array P
as corresponding to the i-th component of the node J and perform the
substitution expressed by equation (9.59). In this manner the i-th component of displacement of the restrained node J becomes an active
degree of freedom of the structure and can be readily incorporated
in the sub-array DI in the partitioning of the general system equation
(8.50). The rest of the procedure remains identically the same as presented above.

9.7

Some nal remarks on the matrix analysis of


structures

The direct stiness method, introduced in chapter 8 and in this chapter


for the analysis of trusses and frames, is highly suitable for automated
computer procedures. Generally, a computer program can be conveniently divided into several steps, some requiring the input of the structural data from the operator, others concerning the manipulation of
data and the computation of several quantities until the nal results

262

Energy methods in structural mechanics

are obtained. This book is intended to be a general introduction to


energy methods that lead to the matrix methods, and therefore we
do not deal with the subject of computer programming, not even
with a simple example (plenty of excellent textbooks are totally
devoted to this important topic). However, as a conclusion to the
topic of frame and truss analysis it may be instructive to summarise
the steps which are followed in an automatic procedure, as they are
substantially the same as we will encounter later in our presentation
of the nite element method.
These steps are summarised as
.

identication of structural data. The geometrical and mechanical


properties of the structure must be clearly identied. In particular it is necessary to record the number of joints and their
geometric coordinates, the number of members and their
cross-section properties, the elastic properties of the material,
the applied constraints and the inferred number of degrees of
freedom.
identication of load related data. The joint loads and member
loads acting on the structure under analysis must be clearly identied. The member loads are subsequently modied to obtain
the value of the equivalent joint loads. These latter are then
added to the directly applied joint loads and the problem is
reduced to the equilibrium of a structure subject to nodal
forces only.
construction of stiness matrix. First, the individual member
stiness matrices are calculated in the local reference frame of
any element. Subsequently, they are transformed to the general
reference frame and the contribution of each element is
assembled in the combined stiness matrix of the problem.
constraint application. The degrees of freedom that are suppressed by the constraints are eliminated in the general system
of equations of equilibrium by deleting and rearranging the
aected rows and columns in the matrices.
calculation of results. The displacements of the joints are computed from the general system of equilibrium equations. The
constraint reactions can then be simply obtained. Successively,
the evaluation of the element strains and forces induced in the
members proceeds on an individual basis: in particular, the
member end-actions can be directly computed by means of
the member stiness matrix, once the displacements of the
joints are known, and the eventual member forces can be
taken into account.

Matrix analysis of rigid-jointed framed structures


2m

263

1m
100 kN

15 m

Fig. 9.15

Suggested exercise problems


1.

Calculate the forces in the frame shown in Fig. 9.15 and the vertical deection at B caused by the applied load if
(a)
(b)

the frame is considered rst as having pin-joint connections


at A, B and C, see Fig. 9.16, and
analysed as having all the members rigidly connected at
these nodes, see Fig. 9.15.
2m

1m

B
A

100 kN
15 m

Fig. 9.16

264

Energy methods in structural mechanics


A

2P
B

2m

2m

2m

2m

3m

Fig. 9.17

2.

3.

4.

All members are fabricated using steel tubes 300 mm outside


diameter and 10 mm wall thickness. The modulus for steel is
205 kN/ mm2 .
[Ans: pin-jointed Av 0:303 mm, NAB 33:6 kN,
NBC 67.2 kN, rigid joints Av 0:193 mm,
NAB 42 kN, NBC 84 kN, Mab 1:97 kNm]
The continuous beam shown in Fig. 9.17 has two spans, AB with a
exural rigidity of I and BC with a exural rigidity of 2I. It is
welded to a rigid foundation at A and C and supported by an
axially innitely rigid strut at B. The strut provides no rotational
restraint to the horizontal beams at B. Calculate the deection at
the mid-length of both spans and calculate the maximum strain in
the beam span AB if that member is composed of a length of
hollow square channel, depth 200 mm and thickness 15 mm. The
member BC is a hollow square section, depth 300 mm and uniform
thickness. The modulus for steel is 205 kN/mm2 . Calculate the
maximum allowable value of the load P if the maximum strain
in the span AB is not to exceed the material yield strain,
"0 0:2% with a safety factor of 0:7.
[Ans: Ev 14:1 mm, "AB 8:6 mm, P 209 kN]
Calculate the allowable load in the frame in Fig. 9.17 if the
member DB has the same cross-section dimensions as member
AB and cannot be considered to provide a rigid foundation at
B. The member BD is rigidly connected at B to the horizontal
members.
[Ans: P 243 kN]
An arch is constructed from straight steel members, each of length
l and exural rigidity EI, as shown in Fig. 9.18. The arch is
attached to a rigid foundation at A and G and all the members
are welded to each other so that it should be considered as a
rigid-jointed frame. Calculate, taking account of symmetry as

Matrix analysis of rigid-jointed framed structures

265

P
l

E
D

30

l/2

l/2

30

60

60

Fig. 9.18

appropriate, the vertical deection D at the position of the


applied load P. Calculate the magnitude and position of the maximum moment Mmax in the arch. Compare the moment and deection with the corresponding value Mmax1 , D1 obtained from the
CotterillCastigliano approach, chapter 6, considering the arch as
a continuous curved structure.
[Ans: D 0:034Pl 3 =EI, Mmax 0:21Pl, Mmax1 0:105Pl,
D1 0:105Pl 3 =EI

10. Analysis of thin plates


Introduction

Up to now in this book we have considered structures that are composed of one-dimensional linear elements. That is, the variation of
deections and forces are related only to the axis along the length of
the elements. Particularly, the real three-dimensional nature of a
beam is reduced to the one-dimensional variation by the assumption
of linear strain through the depth of the beam section. There is, of
course, a wider class of elements that are common in practical structures, that is plates and shells. Essentially these elements have two
dimensions much greater than their third, which is called the thickness.
The dierence between plates and shells is that the former has an initially at surface dened by the two main dimensions, lx and ly whereas
shells have a surface that may be singly- or doubly-curved. In this
introductory text, we consider only plates.
Just as the analytical model of a beam is reduced to essentially a onedimensional model by a simplifying assumption so also a plate is
reduced to a two-dimensional body with the aim of facilitating mathematical analysis. The assumption leading to this simplication is
identical to that for a beam; that is, the strain is assumed to vary
linearly through the thickness of the plate. This is called the Kirchho
model of plate deformation, named after the famous natural philosopher. It must be appreciated that this is an approximation to the real
behaviour of plates, but within a restricted class of structural elements
the model provides an adequate level of accuracy. The model will have
reduced validity if, for example, the plate is composed of layers of
material with dierent moduli through the thickness, or the forces
imposed on the plate in, or normal to its plane are very localised and
the analysis is concerned with determining the strain distribution in
the region of these loads. However, when we are concerned with the
calculation of the general deections of plates composed of homogeneous material and with a thickness much less than the other dimensions, then the Kirchho model serves very well and provides practical

268

Energy methods in structural mechanics

information at a decidedly acceptable level of accuracy, when compared with corresponding experimental results.
The presentation in this chapter follows closely the pattern of the
presentation of the Bernoulli beam theory in chapter 4. First, we will
analyse the equilibrium of the plate subject to stretching loads, i.e.
loads whose resultants always act in the mid-plane of the plate. Then
we shall concern ourselves initially with the derivation of the expressions for the strain through the thickness, in terms of the deection
of the mid-plane surface of the plate. The expression for the strain
energy in the plate material is developed for the condition where the
plate mid-surface is deformed laterally in response to loads that generally are applied normal to the surface of the plate. We have two choices
for attempting to determine the actual shape and magnitude of these
deections for specic values and patterns of loading. One approach
is to apply the methods of calculus of variations to the energy functional to derive the dierential equation of equilibrium, and then to
attempt to nd a solution to this equation for specic forms of loading.
This was the approach followed by some 18th and 19th century mathematicians and there is an interesting history of the search for the
solutions that are reported in many textbooks. This approach oers
also the possibility of elucidating the boundary conditions relevant
to plate bending, which are not so obvious as in the case of bending
beams. Inevitably, the solution of a fairly complex dierential equation
is dicult and in practice many of the solutions actually achieved
relate to idealised plate shapes, boundary conditions and forms of
loading.
The practical cases of plate bending in engineering tend not to be so
well conditioned and the solution of the dierential equations of
equilibrium is not feasible and so an alternative, and more widely
applicable approach is to be found through the use of approximate
energy methods, like the RayleighRitz procedure or the nite element
method, which is developed later in this book. Once again, it is emphasised that the exact solutions to plate bending, albeit for idealised
conditions, have a very valuable attribute in that they provide a basis
for the calibration and validation for the more approximate solutions
from the energy methods and the numerically based nite element
models used in engineering practice. Of course, as in the case of
beams, it is possible to obtain approximate solutions to the dierential
equation of equilibrium by using the method of nite dierences. However, experience has shown that this method is not so conducive to automatic generation of the plate modelling and the boundary conditions as
the nite element method so that this latter method is now the primary
analysis tool available to structural engineers for plate analysis. For this

Analysis of thin plates

269

reason the nite dierence method is not exemplied here and the main
emphasis is on the approximate solution of equilibrium using the
method of trial functions, including nite elements, applied to the
energy functional. Nevertheless, for the sake of completeness, the differential equations for equilibrium of plates for in-plane and lateral
loading, i.e. bending, are derived here.

10.1

Kirchho 's model of plate deformation

Let us start by establishing a mathematical model of the plate shown in


Fig. 10.1. As was stated in the introduction, the plate is characterised
by the fact that two of its dimensions, lx and ly , are generally much
greater than its thickness hz . The mid surface of the plate is planar,
and therefore we can think of it as being contained in the xy plane.
Actually, the Kirchho model of plate is rather intuitive and its
correspondence with Bernoulli's model of bending beam is to some
extent straightforward.
In chapter 4 we assumed that irrespective of the deformation of the
Bernoulli beam model, its cross sections must remain plane and normal
to the curve representing the deformed axis of the beam. Now, as we
are dealing with a two-dimensional solid, we can assume that, whatever is the deformation of the plate, the segments normal to its mid
plane must remain straight and normal to the surface representing
the deformed mid-plane of the plate.
Thus, Kirchho 's model of a plate essentially incorporates a kinematic assumption regarding the manner in which the plate will
deform. Of course, it must be emphasised once more that real plates

x
y

Fig. 10.1

270

Energy methods in structural mechanics


ux

y
x

Fig. 10.2

do not necessarily deform exactly in accordance with Kirchho 's


assumptions. However, experience has shown that in most practical
cases for isotropic materials the behaviour of practical plates is not
far from the hypotheses stated above. Thus the results that we are
going to obtain are usually acceptably accurate from a practical
engineering viewpoint.
As we could have anticipated, even if the basic assumptions of the
model are very similar to those adopted for the bending beam, the
fact that we are operating over two dimensions instead of one considerably complicates the development of the basic theory and obtaining
practical solutions.
Let us start from the analysis of the strain eld that occurs in the
Kirchho model of plate deformation. The fact that the edges of segments normal to the mid plane, xy, must remain straight and normal
to the surface representing the deformed mid-plane yields
ux x; y; z zy ux x; y; z0

10:1

uy x; y; z zx uy x; y; z0

10:2

uz x; y; z uz x; y; z0

10:3

where ux x; y; z0 and uy x; y; z0 are the x and y components of


displacement of the points lying on the mid-plane, z0  z 0 and
x and y are the angles of rotation along the x and y axes, respectively. The geometrical meaning of the term zy in equation (10.1) is
depicted in Fig. 10.2.

10.2

Equilibrium of stretched plates

We start our analysis of the equilibrium of thin plates from the case of
simply stretched plates, i.e. under the assumption that the resultants of
the surface and boundary loads are contained in the mid-plane of the
plate, as shown in Fig.10.3.

Analysis of thin plates

qx

qy
q

271

x
y
S

Fx

Fy
F

Fig. 10.3

The deformation of the plate, described by equations (10.1)(10.2),


can be simplied in the following manner
ux x; y; z ux x; y; 0

10:4

uy x; y; z uy x; y; 0

10:5

uz x; y; z 0

10:6

so that the strain eld of the deformed plate can be completely


described by the relationships
qux
qx


1 qux quy

2 qy
qx

"xx

10:7

"xy

10:8

"yy

quy
qy

10:9

where "xx and "yy are the direct strains in the x and y directions, respectively, and "xy is the shear strain in the xy plane.
Let us now consider an innitesimal element of plate, exemplied
in Fig. 10.4. This slice has been obtained by means of four innitesimally close conceptual cuts: two in the zx plane and two in the zy
plane.
Resulting from the general theory of strain for linear elasticity,
stated in chapter 3, the resultants of the internal forces per unit

272

Energy methods in structural mechanics

dx
dy

z
y
x

Fig. 10.4

length are given by the relationships




E
"

"

Nx h
xx
yy
1 2


E
Ny h
"yy "xx
1 2

10:10
10:11

Nxy Nyx h2G"xy

10:12

These are a normal force Nx per unit length acting on the face of normal
x, a normal force Ny per unit length acting on the face of normal y and
nally the forces Nxy per unit length acting on both these faces. The
situation is depicted in Fig. 10.5.
On account of the above relationships the internal energy U stored
in our model of stretched plate has the following expression

1
U
Nx "xx Ny "yy 2Nxy "xy dS
10:13
2
S

Nx
Nxy
Nyx
Ny

Fig. 10.5

Analysis of thin plates

273

where the integral is performed over the surface S of the mid plane. At
this point, in order to study the equilibrium of the plate, we can simply
follow the usual procedure by writing the general condition (equation
(1.28)) stated in chapter 1, that is
U W
To simplify the presentation here, we suppose the loading is a distributed load qx; y, acting in the mid-plane of the plate with components
qx x; y and qy x; y, and a set of forces F per unit length on the boundary of the mid-surface, as shown in Fig. 10.3. Thus, the work done by
the external forces is

W qx ux qy uy dS Fx ux Fy uy d
10:14
S

and the condition of equilibrium (equation (1.28)) can be written as

Nx "xx Ny "yy 2Nxy "xy dS


S

qx ux qy uy dS Fx ux Fy uy d

10:15

On account of the general straindisplacement relationships equations (10.7)(10.9), the left hand side of the above equation can be
written as

Nx "xx Ny "yy 2Nxy "xy dS


S



quy
qux
qux quy
Ny
Nxy

dS
Nx
qx
qy
qx
qy

q
q
N u Nxy uy Ny uy Nxy ux dS
qx x x
qy


 


qNy qNxy
qNx qNxy

ux
uy dS 10:16

qx
qy
qy
qx
S

and recalling the divergence theorem, i.e.





qfx qfy

dS fx x fy y d
qx qy
S

10:17

274

Energy methods in structural mechanics

where x and y are the direction cosines of the normal to the boundary , we can write

Nx "xx Ny "yy 2Nxy "xy dS


S

Nx ux Nxy uy x Ny uy Nxy ux y d


 


qNy qNxy
qNx qNxy

ux

uy dS 10:18

qx
qy
qy
qx
S

so that the equilibrium condition (equation (10.15)) becomes




 

qNy qNxy
qNx qNxy

qx ux

qy uy dS
qx
qy
qy
qx
S

Fx Nx x Nxy y ux

Fy Ny y Nxy y uy d 0

10:19

As the virtual displacements ux and uy are arbitrary by denition, the
vanishing of the above equation requires that
qNx qNxy

qx 0
qx
qy
Nx x Nxy y Fx

and

and

qNy qNxy

qy 0 on S
qy
qx
10:20
Ny y Nxy y Fy

on

10:21

These are the equilibrium equations which relate the internal forces Nx ,
Ny and Nxy to the applied loads qx x; y, qy x; y and Fx , Fy in the case
of simply stretched thin plates.
It is worth noticing that the equations (10.20) and (10.21) are
formally independent from the deformation of the plate.

10.3

Equilibrium of bent plates

As we have already done in the case of axially and laterally loaded


beams, we again keep the stretching and the bending behaviour of
the thin plates distinct. Therefore in the present section we deal with
the pure bending of thin plates and suppose that the mid-plane of the
plate does not undergo any stretching.

Analysis of thin plates

275

If we assume that no stretching of the mid-plane will occur, we must


have
ux x; y; 0 uy x; y; 0 0

10:22

and, remembering the line of reasoning which led to equation (4.53),


we can write
quz
qy
qu
y x; y z
qx
x x; y

10:23
10:24

Equations (10.1) and (10.2) now become


quz
qx
quz
uy x; y; z z
qy
ux x; y; z z

10:25
10:26

Finally, the strain eld of the deformed plate can be completely


described by the relationships
qux
q2 u
z 2z
qx
qx


1 qux quy
q2 uz

2 qy
qx
qx qy

"xx

10:27

"xy

10:28

"yy

quy
q2 u
z 2z
qy
qy

10:29

where "xx and "yy are the direct strains in the x and y directions, respectively, and "xy is the shear strain in the xy plane.
Let us now consider again the innitesimal element of the plate in
Fig. 10.4. Resulting from the general theory of strain for linear elasticity, stated in chapter 3, the resultants of the internal forces per unit
length are given by the relationships


h=2

Mx
h=2


E
"xx "yy z dz
1 2

h=2

h=2

 2

E
q uz
q2 uz
 2 z dz
qy
1  2 qx2

10:30

276

Energy methods in structural mechanics


h=2

My
h=2


E
"

"

yy
xx z dz
1 2


h=2

h=2

 2

E
q uz
q2 uz
z dz


qx2
1  2 qy2

10:31

h=2

Mxy Myx

2G"xy z dz
h=2

h=2


2Gz

h=2


q2 uz
z dz
qx qy

10:32

These are a bending moment Mx acting on the face of normal x direction, a bending moment My acting on the face of normal y and nally
what we will call a twisting moment Mxy on both these faces. The situation is depicted in Fig. 10.6.

My

Myx

Mx

Mxy

Mx
y

x
z

Mxy

Fig. 10.6

Analysis of thin plates

277

After a little mathematical manipulation, thanks to the assumption


that the strain through the thickness of the plate varies linearly, equations (10.30)(10.32) become
 2

q uz
q2 uz
Mx D
 2
10:33
qx2
qy
 2

q uz
q2 uz
My D
 2
10:34
qy2
qx
Mxy D1 

q2 uz
qx qy

10:35

The term D is called the bending stiness of the plate, that is


D

Eh3
121  2

10:36

It is worth noting that in the case of the Kirchho plate model the term
D plays a role similar to that played by the term EIx for the Bernoulli
beam model.
We can now tackle the derivation of the equilibrium of the model.
With reference to Fig. 10.6, it is clear that the internal moments Mx ,
My and Mxy are responsible for the energy stored in the innitesimal
slice of plate on account of the work done by the eects of the
curvatures x , y and xy .
As we know from calculus (see also chapter 4), in the theory of
innitesimal displacements, curvatures can be related to the normal
deection of the mid-plane by
x

qx
1
q2 u
2z
ry
qy
qy


1 qx qy
q2 uz

2 qy
qx
qx qy

y
xy

qy
1
q2 u
2z
rx
qx
qx

10:37
10:38
10:39

As an example, the geometrical meaning of equation (10.38) is depicted


in Fig. 10.7.
On account of equations (10.37)(10.39), the internal energy U stored
in the Kirchho 's model of bent plate has the following expression

1
U
Mx x My y 2Mxy xy dS
10:40
2
S

278

Energy methods in structural mechanics

rx

y
z

Fig. 10.7

where the integral is performed over the surface S of the mid-plane. At


this point, in order to study the equilibrium of the plate, we can simply
follow the usual procedure by writing the general condition stated in
chapter 1, that is
U W
In order to simplify the presentation here, we suppose that the loading is a distributed load qz x; y, which is directed normally to the plane
xy of the plate, and has a set of vertical forces F and bending couples
 per unit length on the boundary of the mid surface, as shown in
M
Fig. 10.8.
The work W done by the external forces is



quz
quz



W qz uz dS
Msn
d
10:41
Fuz Mn
qn
qs
S

 sn are the components of


 n and M
where, with reference to Fig. 10.9, M
the generic bending couple along the normal n and the tangent s to the
boundary , respectively.

Analysis of thin plates

279

qz

M
F

Fig. 10.8

Fig. 10.9

In the present case the general condition of equilibrium, see equation


(1.28), can be therefore written as

Mx x My y 2Mxy xy dS


S

qz uz dS
S


quz
quz



Msn 
d 10:42
Fuz Mn 
qn
qs

Now, in order to illustrate what this condition of equilibrium really


means in terms of the displacement eld uz x; y, some mathematics
is required. Even if the procedure can appear lengthy, it is, however,
worth paying attention to each step, because it constitutes a standard

280

Energy methods in structural mechanics

technique in the theory of deformable bodies and gives us a clear


example of the diculties that can underlie even a simplied model,
like the Kirchho model for plate deformation. In this spirit, let us
now proceed along our way.
First of all, let us substitute the expressions, equation (10.37)
(10.39), for the curvatures into the left hand side of equation (10.42).
Thus we obtain


 2 
 2 
 2 
q uz
q uz
q uz
dS
Mx 
My 
2Mxy 
2
2
qxqy
qx
qy


qz uz dS
S


quz
quz



Msn 
d 10:43
Fuz Mn 
qn
qs

We now replace the variational operators by the dierential ones, and


vice versa, and get



q2 uz
q2 uz
q2 uz
dS

2M
y
xy
qx qy
qx2
qy2



 sn  quz d 10:44
 n  quz M
 zM
qz uz dS
Fu
qn
qs

Mx
S

Let us now consider the following relationships




q2 uz
q
quz
qMx
q2 Mx
M

u
Mx

u

x
z
z
qx
qx
qx
qx2
qx2

10:45



q2 My
q2 uz
q
quz
qMx
M

u

u

y
z
z
qy
qy
qy
qy2
qy2

10:46

My





qMxy
q2 uz
q
quz
q

Mxy

uz
Mxy
qxqy
qx
qy
qy
qx
uz

q2 Mxy
qx qy

10:47

Bearing in mind that in equation (10.47) the x and the y coordinate can
be substituted one for the other, on account of the symmetry, we can
substitute the relationships in equations (10.45)(10.47) in the left

Analysis of thin plates

281

hand side of equation (10.44) and obtain




q2 uz
q2 uz
q2 uz
dS
Mx
My
2Mxy
qxqy
qx2
qy2
S


q2 Mxy q2 My
q2 Mx

uz dS

2
qx qy
qx2
qy2
 

q
qMx qMxy
uz

qx
qx
qy

 

qMxy qMy
q
uz

dS

qy
qx
qy

 
q
quz
quz
Mx
Mxy

qx
qx
qy
S



q
quz
quz
Mxy
My
dS

qy
qx
qy
Recalling the divergence theorem, i.e.



qfx qfy

dS fx x fy y d
qx qy
S

10:48

10:49

where x and y are the direction cosines of the normal to the boundary , we can apply it to the last two integrals of equation (10.48). Thus


q2 uz
q2 uz
q2 uz
Mx
dS

2M
y
xy
qx qy
qx2
qy2
S


q2 Mxy q2 My
q2 Mx

uz dS
2
qx qy
qx2
qy2


uz


 

qMxy qMy
qMx qMxy

y d
qx
qy
qx
qy


Mx


quz
quz
Mxy
x
qx
qy

 

quz
quz
My
y d
Mxy
qx
qy

10:50

282

Energy methods in structural mechanics

This expression can still look quite complicated, however, if we dene






qMxy qMy
qMx qMxy
10:51

y
Sn
qx
qy
qx
qy
and remember from calculus that
quz quz
qu

z
qx
qn x qs y
quz quz
qu

z
qy
qn y qs x
we have


quz
quz
Mxy
x
Mx
qx
qy

 

quz
quz
My
y d
Mxy
qx
qy

quz
quz
Mx 2x 2Mxy x y My 2y

qn
qs


2
2
 My Mx x y Mxy x y d


quz
quz
Msn
d
Mn

qn
qs

10:52
10:53

10:54

On account of the transformation rules that link the moments Mn ,


Msn , Mx , My and Mxy (see Fig. 10.10), i.e.
Mn Mx 2x 2Mxy x y My 2y
Msn Mx My x y Mxy 2x 2y

10:55
10:56

we can nally write the condition of equilibrium, equation (10.42), in


the form

 2
q2 Mxy q2 My
q Mx

uz dS
2
qxqy
qx2
qy2
S


quz
quz
Msn
d
Sn uz Mn

qn
qs


quz
quz



qz uz dS
Msn 
d
10:57
Fuz Mn 
qn
qs
S

Analysis of thin plates

283

Myx

0
Mx

My

Mxy
Msn
Mn
n
y

Fig. 10.10

This expression is more appropriate to our purposes and the nal


step in this development consists of grouping the terms, so that

 2
q2 Mxy q2 My
q Mx

2
qz uz dS
qxqy
qx2
qy2
S

 sn Msn quz d
F Sn uz d M
qs

 n quz d 0
Mn M
qn

10:58

As the virtual displacement eld uz x; y is arbitrary, by denition, the


vanishing of this variational expression implies some clearly dened
conditions. First of all we must have
q2 Mxy q2 My
q2 Mx

2
qz 0
qx qy
qx2
qy2

on S

which, by means of equation (10.33)(10.35), becomes


 4

q uz
q4 uz
q4 uz
D
2 2 2 4 qz
on S
qx4
qx qy
qy

10:59

10:60

The requirement of equilibrium of the plate under the assigned loads


for the displacement eld uz x; ywhich denes the deformed conguration of the Kirchho model means the satisfaction of this dierential

284

Energy methods in structural mechanics

equation . As may be noted, there is a strong resemblance between this


equilibrium condition, equation (10.60), and the corresponding equilibrium condition for the Bernoulli beam model, i.e. Euler's equation of
the deformed beam, equation (4.43). Both are fourth-order dierential
equations (even if, of course, equation (10.60) involves partial derivatives as a consequence of the two-dimensional character of the plate
model) and involve the exural stiness EIx and D, respectively, for
the beam and the plate bending. This is not surprising, since the kinematic hypotheses for the models, i.e. linear variation of strain through
the depth of the plate and beam, are essentially the same in both cases.
However, in contrast to the case of Bernoulli's beam, not all the
boundary conditions associated to the dierential equation (10.60)
are obvious from a mechanical point of view. In fact, the vanishing
of the general condition of equilibrium in the form of equation
(10.58) requires also the vanishing of its last integral. This happens if
quz
quz
 n on
0 or Mn M
n )
10:61
qn
qn
This requires that the rotation quz =qn on the boundary coincides
with the imposed value  or, alternatively, the bending moment Mn
 n . Moreover, the vanishing of the
coincides with the imposed value M
second and third integral in equation (10.58) must be treated in conjunction. In fact the derivative quz =qs along the tangent to the
boundary is linked to the variation uz along the boundary itself.
Integrating the third integral by parts we get

 sn Msn uz d
 sn Msn quz d q M
M
qs
qs


 
qMsn qMsn

uz d

qs
qs

10:62

where the rst integral at the right hand side is obviously zero. In fact,
this term is the line integral along a closed path (i.e. the edge of the
plate) of a perfect dierential. Combining this result with the second
integral of equation (10.58), we get

 sn qMsn 
qM


uz d 0
F Sn
10:63
qs
qs

This statement leads to the boundary conditions


 sn
qMsn
qM
F
on
uz uz or Sn
qs
qs

10:64

Analysis of thin plates


[ Msn +

Msn
s ]

285

ds

Msn ds
M

Msn + ssn
Msn
ds
ds

ds
ds
M

Sn = ssn

ds
ds

Fig. 10.11

The meaning of the denition in equation (10.51) is now evident: the


term Sn represents the shear forces normal to the mid-plane xy. Its
existence on the boundary is strictly related to the twisting
moment Msn , as shown in Fig. 10.11, and this explains the mechanical
condition expressed by the second alternative of equation (10.64). In
other words, if we represent the distribution of twisting moments
Msn along the edge of the plate by means of a continuous set of twisting
couples, we are led to the conclusion that the gradient of the distribution of twisting moments Msn is statically equivalent to a distribution
of shear forces of intensity Sn qMsn =qs.
In closing this presentation it is useful to summarise all the equations
that have been yielded by the general condition of equilibrium,
equation (10.42), that must be fullled by the displacement eld
uz x; y
 4

q uz
q4 uz
q4 uz
Field equation on S
D
2 2 2 4 qz
qx4
qx qy
qy
Boundary condition on

uz uz

or Sn

 sn
qMsn
qM
F
qs
qs

286

Energy methods in structural mechanics

Boundary condition on

quz
n
qn

n
or Mn M

Of course the mechanical boundary conditions can also be expressed


in terms of the displacement eld. Thus, because of equations (10.33)
(10.35), (10.51), (10.55) and (10.56), we can write
 3
 3
 

q uz
q3 uz
q uz
q3 uz
Sn D

y
x
10:65
qx3 qx qy2
qy3 qx2 qy
 2

q uz
q2 uz 2
q2 uz
Mn D

21



x
qx qy x y
qx2
qy2
 
 2
q uz
q2 uz 2
y
10:66


qy2
qx2
 2
  2

q uz
q2 uz
q uz
q2 uz
Msn D
 2
 2 x y
qx2
qy
qy2
qx
D1 

q2 uz 2
2y
qx qy x

10:67

It is now clear that the problem of the equilibrium of the plate bending model is not simple, even for the example of an evenly distributed
load qz x; y. Several special solutions have been developed for specic
problems of plate bending. However, the study of these solutions is not
really the purpose of this book and the interested reader can make
reference to specialised texts. We consider that what has been said is
sucient to elucidate the fundamentals of the problem. In the next
section we will deal with a more practical and general procedure that
is developed by making use of the energy approach with the method
of trial functions.

10.4

An example of the method of trial function for


the solution of bending plates.

It is apparent that the solution of the derived dierential equation


describing the elastic equilibrium does not constitute a simple task
even for the simplied model of bending plates. From an engineering
point of view it therefore appears very worthwhile to make reference
to all those methods capable of providing approximate solutions.
Among the ways we have already studied is the method of trial
function, which we will now apply to the evaluation of the response
of the example plate shown in Fig. 10.12 to the specied applied
loading.

Analysis of thin plates

287

Fig. 10.12

The gure shows a square plate simply supported along two edges
and rigidly clamped along the other two, subject to an evenly distributed load qz x; y. Its thickness is constant, as well as its bending stiness D. To be in a position to apply the trial function method, rst of
all we must write down the expression for the total potential energy of
the system in terms of the displacement eld, uz x; y. This is simply
achieved by means of equations (10.40) and (10.41), and taking into
account equations (10.33)(10.35) and (10.37)(10.39). We thus have
uz U W
 2 2  2 2
1
q uz
q uz
q2 uz q2 uz
D

2
2
qx2
qy2
qx2 qy2
S

 2 2 
q uz
dS
21 
qx qy



quz
quz



Msn
d
qz uz dS
Fuz Mn
qn
qs
S

10:68

For the example plate being analysed here, the boundary conditions
are essentially geometrical




l
l
10:69
uz ; y 0 and uz ; y 0
2
2




l
l
uz x;
0 and uz x;
0
10:70
2
2
 
 
quz
quz
0 and
0
10:71
qy x; y l=2
qy x; y l=2

288

Energy methods in structural mechanics

Since there are no forces applied on the edges, the last term of equation
(10.68) is zero, so we have
uz U W
 2 2  2 2
1
q uz
q uz
q2 uz q2 uz
D

2
2
qx2
qy2
qx2 qy2
S

 2 2 

q uz
dS qz uz dS
21 
qx qy

10:72

Moreover, the fact that the boundary conditions, equations (10.69)


(10.71) are not dependent on Poisson's ratio  (as they would have
been in the case of expressions (10.66) and (10.67) for the bending
and twisting moments were involved on the edge) leads us to conclude
that the solution cannot be dependent on this ratio, as D is also constant. Therefore, if we rearrange equation (10.72) in the following
manner
uz U W
1
D
2


S



S

q2 uz q2 uz
2
qx2
qy

2

dS D1 

 2 2 

q2 uz q2 uz
q uz
dS

qz uz dS
qx qy
qx2 qy2

10:73

we can establish that the term


 2 2 
 2
q uz q2 uz
q uz
dS

2
2
qx qy
qx qy

10:74

will vanish and we have


1
uz U W D
2


S

q2 uz q2 uz 2

dS

qz uz dS
qx2
qy2

10:75

It can be rigorously demonstrated that this simplication holds for all


the plates rigidly supported on the boundary and for all polygonal
plates whose boundary conditions are either uz 0 or quz =qn 0.
We now assume a shape trial function that satises the geometrical
boundary conditions equations (10.69)(10.71). A general format for a

Analysis of thin plates

289

suitable trial function is


uz a1 f xgy

10:76

with the following expressions for x and gy


f x 1

4x2
l2

10:77

8y2 16y4
gy 1 2 4
l
l
Thus we have



4x2
8y2 16y4

uz a1 1 2
1 2 4
l
l
l

10:78

This function satises the geometrical boundary conditions. We can


now substitute it in the expression for the total potential energy functional, equation (10.75). Thus


q2 uz
8
8y2 32y4
a1 2 1 2 4
10:79
qx2
l
l
l



q2 uz
4x2
16 192y2
1

10:80

1
qy2
l2
l4
l2
and
uz

1
D
2

l=2

l=2

l=2 l=2

64a21
l 12

4
2 2
2 2
2 2
4 2
 3l 8l x 32l y 96x y 16y dx dy
l=2

l=2

l=2 l=2

 
4x2 8y2 32x2 y2
q a1 1 2 2
l
l
l4

16y4 64x2 y4

l4
l6


dx dy

10:81

After evaluation of the integrals equation (10.81) is


1 32768 D
16
a1 ql 2
a1 a21
2
175 l 2
45

10:82

Imposing the stationary condition on this expression of the total

290

Energy methods in structural mechanics

potential energy, to satisfy equilibrium, we nally achieve


q
0
qa1

a1

35 ql 4
18432 D

10:83

The deection of the plate can be obtained by substituting equation


(10.83) in equation (10.78). At the centre point of the plate, we have
4

ql
uz 0; 0 0:00190
10:84
D
This value is quite accurate. In fact the analytical solution1 to the
dierential equation of equilibrium yields
4

ql
uz 0; 0 0:00192
10:85
D
We therefore have a solution from the method of trial function with an
error of only 1:1%.

10.5

More examples of application of simple trial


functions to plate bending

The previous section has shown an example of the application of the


method of trial functions for calculating the magnitude of the deections of plate deformed by loads normal to their plane. The present
section expands on that example and exemplies the simplicity of the
method by application to a number of types of boundary conditions
and loading.
With respect to the general denition of the plate geometry shown in
Fig. 10.13, the energy formulation equation (10.75) is written
D
uz
2

a b 
0 0

q2 uz q2 uz
2
qx2
qy

2

a b
dx dy

qx; yuz dx dy 10:86


0 0

Plates subjected to uniform pressure

Three examples of boundary conditions are shown in Fig. 10.14, varying from the situation where all the edges are simple supported to the
condition where all the edges are encastre.
The applied loading is a uniform pressure, i.e.
qx; y q0

10:87

Timoshenko, S. P. and Woinowsky-Kreiger, S. (1959), Theory of plates and shells,


McGraw-Hill, chapter 6.

Analysis of thin plates

291

y
b

Fig. 10.13

Case (a): All edges simple supported


The boundary conditions for this case are
uz 0

at

x 0;

x a;

y 0;

yb

10:88

From section 7.1, it is reasonable to employ the one-degree-of-freedom


trial function in the form
 
 
x
y

sin
10:89
uz x; y u0 sin
a
b
This function satises the geometry boundary conditions along all
edges. Substituting equation (10.89) into equation (10.86) and performing the mathematical operations we have
u0

4 D 4
4q
a 2a2 b2 b4 u20 20 abu0
3 3

8a b

10:90

Applying the condition of equilibrium, i.e.


qu0
0
qu0

10:91

we have the relationship between the magnitude of the deection and


the applied pressure as


16q0
a4 b4
10:92
u0 6
 D a4 2a2 b2 b4

292

Energy methods in structural mechanics

Fig. 10.14

Case (b): Two edges simple supported and two edges encastre
The boundary conditions for this case, see Fig. 10.14(b), are
uz 0

at x 0;

x a;

quz
0
qy

at y 0;

yb

y 0;

yb
10:94

An appropriate trial function satisfying the boundary conditions for


this case is

 
u
x
2y
1 cos
10:95
uz x; y 0 sin
a
b
2

Analysis of thin plates

293

Substituting equation (10.95) into equation (10.86), carrying out the


mathematical operations and applying the equilibrium condition,
equation (10.91), we have


16q0
a4 b4
u0 3
10:96
 D 16a4 8a2 b2 3b4
Case (c): All edges encastre
The boundary conditions for this case, see Fig. 10.14(c), are
uz 0

at x 0;

x a;

y 0;

quz
0
qx

at x 0;

x a;

quz
0
qy

yb
at y 0;

yb
10:97

An appropriate trial function for this case is





u0
2x
2y

uz x; y
1 cos
1 cos
a
b
4

10:98

Substituting equation (10.98) into equation (10.86), carrying out the


mathematical operations and applying the equilibrium condition,
equation (10.91) we have


q0
a4 b4
u0 4
10:99
 D 3a4 2a2 b2 3b4
The accuracy of these solutions can be assessed by comparison with the
exact values for the deection for a square plate, i.e. a b l. The
general formulation is
u0

q0 l 4
D

10:100

and
Case
a
b
c

exact
0:00406
0:00192
0:00126

trial function
0:00416
0:00191
0:00128

As the table shows the solutions from the one degree of freedom trial
function analysis are within a few percent of the corresponding exact
solutions.

294

Energy methods in structural mechanics

a/2
a

y
b/2
b

Fig. 10.15

Plates subjected to central concentrated load

A particular merit of the method of trial functions is that the analysis


can very easily accommodate a range of types of loading without
requiring extensive re-analysis. This is exemplied here by considering
the condition where the uniform pressure of the previous section is
replaced by a concentrated load P acting at the centre of the plate
normal to its plane. This is shown in Fig. 10.15.
The potential energy functional, equation (10.75), is for this form of
loading
D
uz
2

a b 
0 0

q2 uz q2 uz
2
qx2
qy

2
dx dy Pu0

10:101

where u0 is the deection at the position of application of the load.


Evidently the strain energy component of the potential energy is
unchanged by the type of loading.
Consider now a plate with all edges simply supported (Case a) in
which the trial function is described by equation (10.89) and the
energy functional in terms of the deection u0 is
u0

4 D 4
a 2a2 b2 b4 u20 Pu0
8a3 b3

10:102

Thus applying the equilibrium condition equation (10.91) to equation


(10.102)


4P
a4 b4
u0 4
10:103
 D a4 2a2 b2 b4

Analysis of thin plates

295

The corresponding relationships for the boundary conditions


described in Cases b and c are


16P
a4 b4
Case b
u0 4
 D 16a4 8a2 b2 3b4
10:104


4P
a4 b4
Case c
u0 4
 D 3a4 2a2 b2 3b4
The conditions for a square plate, a b l, are simply obtained from
equations (10.103) and (10.104) and are
u0

Pl 2
D

10:105

where
Case
a
b
c

10.6

exact
0.0116
0.0056


0:0103
0:00608
0:00513

Localised RayleighRitz method applied to


plate bending

The method of localised trial functions was introduced in chapter 7


and the concept of these trial functions was applied to the calculation
of the deformation of beams. It is intuitively obvious that this method
can be extended to the analysis of plate bending by the derivation of
localised trial functions in the two dimensions of the plane of a
plate. The method is described in this section.
The localised trial function that ensures continuity of displacement
and slope at the boundaries of the elements of a beam is described
by equation (7.78) and is dened here as
uny i 1 ui 1 1 3i2 1 2i3 1
i 1 i 1 2i2 1 i3 1
ui 3i2 1 2i3 1 i i2 1 i3 1

10:106

where, as shown in Fig. 10.16, the non-dimensional coordinate is


i 1  zi 1 =ln , ln is the length of the n-th element, uz is the deected
shape normal to the coordinate , and u i 1 and ui are the deections
at the nodes bounding the n-th element. Again, we dene i 1 
i 1 ln and i  i ln .

296

Energy methods in structural mechanics


ln

z
ui1
ui

i1
i
zi1
uy

Fig. 10.16

The one-dimensional function can be extended to provide a set of


two-dimensional functions that can be used as localised trial functions
to analyse plate bending for a variety of conditions of loading and
boundary conditions. The plate is sub-divided into zones, or elements,
each of which can have dierent lengths in the x and y directions, as
shown in Fig. 10.17.
The orthogonal functions describing the deformations of an element
n, shown in Fig. 10.18, are:

Element n
i + 1, j
i + 1, j + 1
lxn
i, j

i, j + 1

y
lyn
uz

Fig. 10.17

Analysis of thin plates


i + 1, j + 1

i + 1, j

297

Element n

i, j

i, j + 1

Fig. 10.18

In the x direction
unz n ui 1 3n2 2n3 i n 2n2 n3
ui1 3n2 2n3 i1 n2 n3

10:107

and in the y direction


unz n uj 1 3n2 2n3 j n 2n2 n3
uj 1 3n2 2n3 j 1 n2 n3

10:108

where
n 

x
lxn

n 

y
lyn

10:109

The displacement function for the plate element is obtained by combining the functions in equations (10.107) and (10.108). This is exemplied for the displacement and slopes at node, i, j. The function
associated with the displacement, ui;j is
F1 n ; n ui; j 1 3n2 2n3 1 3n2 2n3

10:110

and for the slope in the x direction


F2 n ; n xi; j  2n2 n3 1 3n2 2n3

10:111

similarly for the slope in the y direction


F3 n ; n yi; j 1 3n2 2n3 n 2n2 n3

10:112

and for the twist term


F4 n ; n xyi; j  2n2 n3 n 2n2 n3

10:113

The graphical representation of these deformation formulations is shown


in Fig. 10.19. This process of identifying the functions is completed for

298

Energy methods in structural mechanics


x

u
b

b
a

i, j

i, j
d

d
a
z

b
a

(a)

(b)
x

xy

b
a

i, j

i, j
d

d
z

z
y

(c)

(d)

Fig. 10.19

all the 16 degrees of freedom associated with element n, i.e.


ui; j ;
x ;
y ;
xy
i; j

ui 1; j ;
ui; j 1 ;

xi 1; j ;
x ;
i; j 1

ui 1; j 1 ; xi 1; j 1 ;

i; j

yi 1; j ;
y ;
i; j 1

i; j

xyi 1; j
xy

10:114

i; j 1

yi 1; j 1 ; xyi 1; j 1

The complete function can be written succinctly as


unz n ; n

16
X
r1

Dnr Fr n ; n

10:115

where Dnr are the elements of the vector collecting the nodal degrees of
freedom of the n-th element, i.e.
2
3
ui; j
6  7
Dn 4    5
10:116

xyi 1; j 1

Analysis of thin plates

299

The formulation of the stiness matrix is simplied and made more


amenable to computer programming by redening new generalised
coordinates. The deection function, equation (10.115), can be
described as
unz A0 A1 n A2 n A3 n2 A4 n n A5 n2 A6 n3
A7 n2 n A8 n n2 A9 n3 A10 n3 n A11 n2 n2
A12 n n3 A13 n3 n2 A14 n2 n3 A15 n3 n3

10:117

The generalised coordinates are related to the nodal coordinates Dnr ,


see equation (10.115), by
A BDn
with

10:118

3
A0
6  7
6
7
A 6
7
4  5

10:119

A15
The transformation matrix B is
2

1
6 0
6
6
6 0
6
6 3
6
6
6 0
6
6 3
6
6
6 2
6
6 0
6
B 6
6 0
6
6 2
6
6
6 0
6
6 9
6
6
6 0
6
6 6
6
6
4 6
4

0
0
0
0
1
0
0
0
0
1
0
0
2
0
0
3
0
0
1
0
0 2
0
0
1
0
0
2
0
3 2
0
3
0 2
0
0
1
0
0
0
2 1
0
6 6
4 9
2
0
1
0
3 4
4 3
2 2

2 6
2 6
1 4

3
0
0
0
0
0
0
0 0 0 0 0
0
0
0
0
0
0
0 0 0 0 07
7
7
0
0
0
0
0
0
0 0 0 0 07
7
1
0
0
0
0
0
0 0 0 0 07
7
7
0
0
0
0
0
0
0 0 0 0 07
7
0
0
0 3
0 1
0 0 0 0 07
7
7
1
0
0
0
0
0
0 0 0 0 07
7
0 3 1
0
0
0
0 0 0 0 07
7
7
0
0
0
0 3
0 1 0 0 0 0 7
7
0
0
0
2
0
1
0 0 0 0 07
7
7
0 2 1
0
0
0
0 0 0 0 07
7
3
6
2 9
6 3
2 9 3 3 17
7
7
0
0
0
0
2
0
1 0 0 0 07
7
3
4
2 6
3 2
1 6 3 2 17
7
7
2
3
1 6
4 3
2 6 2 3 15
2
2
1 4
2 2
1 4 2 2 1

(10.120)
The formulation for the strain energy of deformation of the element of
plate bending Un is obtained from the rst term in equation (10.73).

300

Energy methods in structural mechanics

In non-dimensional notations
D
Un
2lxn lyn

1 1 
0 0

b q2 unz a q2 unz

a qn2 b qn2

 2 n 2 n  2 n 2 
q uz q uz
q uz
21 
dn dn

qn qn
qn2 qn2

10:121

The strain energy for the complete plate can be obtained by summing
the energy contributions of all the elements, i.e.
X
U
Un
10:122
n

However, for a typical element the substitution of equation (10.115)


into equation (10.121) and carrying out the mathematical operation
results in replacement of the energy functional by a function of the
generalised coordinates Ai dened in equation (10.118). That is a quadratic function, F,
Un ) FAi

10:123

This is equivalent to the strain energy being in the form


D
Un
AT CA
2lxn lyn

D
Dn T BT CDn B
2lxn lyn

10:124

Application of the extremum condition, i.e.


qUn
0
qDnr

10:125

provides the condition


D
BT CBDn 0
lxn lyn

10:126

The stiness matrix for a typical element is


K n BT CB

10:127

 is obtained by adding
and the stiness matrix for the whole plate, K,
the stiness matrices of the all the elements according to equation
(10.122).
The expression for total of potential energy is obtained by addition
of the loss of potential energy of the load to the strain energy equation

Analysis of thin plates

301

(10.122). Take as an example the bending of a plate in response to a


concentrated load P applied normal to the plane of the plate. The
loss of potential energy of the load is
P
u0

10:128

where u0 is the deection at the position of application of the load. In


vector form we can write
P

10:129

where the loading vector, P has zeros except at the position corresponding to u0 . The equilibrium condition for the whole plate is

KD
P 0
10:130
The boundary conditions are incorporated into equation (10.130) by
eliminating appropriate rows and columns. The displacements and
slopes at the nodes of the elements are obtained as
 1 P
D K

10:131

The accuracy of the results from a localised trial function analysis


generally increases as the number of elements included in the model
is increased. This is exemplied by application of the method described
in this section to a square plate with edge length, l, subjected to a concentrated load at the centre of the plate as shown in Fig. 10.20. The
analysis of the plate is run with the mesh geometries shown in Fig.
10.21. Note that symmetry conditions of loading and deformations
allow that only one quarter of the area of the plates needs to be
modelled.

P
l

Fig. 10.20

302

Energy methods in structural mechanics


Symmetry

m=1

m=2

m=4

Fig. 10.21

The formulation relating the deection at the centre of the plate, u0 ,


to the applied load P is
Pl 2
D
The solutions are
u0

10:132

Case a: Centrally loaded plate with all edges simply supported


Mesh size
1
2
4


0:0111
0:0115
0:0116

Case b: Centrally loaded plate with all edges encastre


Mesh size
1
2
4


0:00530
0:00548
0:00557

It is evident that even with only a few elements the accuracy of the
results, see section 10.5, is very good. Again, the great merit of
localised trial function approach is that once the element derivation
is programmed for a computer it is very simple to alter the boundary
conditions and to incorporate a variety of types of loading in the
numerical model.

11. The theory of nite elements


Introduction

Preceding chapters in this book have examined various energy


methods of analysis of elastic solids; particularly bars, beams, trusses,
frames, plates. In the case of plane and space trusses and frames we
have established the direct stiness method, a procedure that is
especially suitable for computer programming and therefore for the
automatic analysis of structures, even involving thousands of members. Moreover, we have seen that the treatment of structures involving
plates, even with the simplications allowed by Kirchho's model,
remains a fairly complicated matter. In this form of structural component approximate methods, and particularly the method of trial functions, can still turn out to be a dicult approach. Indeed, the method
of trial functions requires considerable skill in the choice of suitable
approximating functions and substantial eort in the construction of
the algebraic functions. In fact, the choice of the trial functions
depends on the geometry of the structural component under analysis
and on its boundary conditions and form of loading, therefore the procedure cannot be automated and there are ever-present risks associated
with the possibility of human errors that make embarking on lengthy
calculations somewhat hazardous. This diculty is common to all
those structures that are not composed simply of a nite number of
well-dened components, such as bars and beams, whose behaviour
is unequivocally known.
However, we have seen that most of these diculties can be overcome by the method of localised trial functions. In this chapter we
introduce a generalisation of the method of localised trial functions,
i.e. the nite element method, which is without any doubt the most
powerful tool for the analysis of even very complex structures and
the major analytical development in structural engineering in the last
40 years. The method was created by structural engineers, but its
versatility, easy of use and accuracy have rendered this methodology
a standard tool of analysis even for problems which may seem very

304

Energy methods in structural mechanics

far from the structural ones, like heat diusion, uid dynamics and
magnetic eld analyses. Most of the high technology resources available to us, from supersonic aeroplanes to mobile telephones, owe
their actual present state of sophistication to the nite element
method. Thanks to this method, engineers and physicists can now
model mathematically and obtain solutions to a very wide range of
problems that would indeed be virtually impossible to handle in any
other way. Alternative approaches are the numerical integration of
the dierential equations of equilibrium, using the nite dierence
method and perhaps by very costly testing at full- or small-scale. But
these alternatives require special skills in numerical calculus and
model-making that in most cases are obviated by the nite element
programs now available. These programs can be accessed and operated very simply, often by engineers who, unfortunately, sometimes
have little understanding of the fundamental theory of the method
they are using.
Despite these impressive capabilities, the underlying idea of the nite
element method is very simple. It consists in a sub-division of the structure conceptually into smaller zones or areas. Once this sub-division
has been carried out, trial functions are given within each zone, usually
called an element, and the behaviour of each element can be analysed
separately and independently from all the others. The boundary
conditions between these elements can be imposed quite easily,
provided the choice of the trial functions are simple polynomials
that provide the continuity between the element under consideration
and those contiguous with it. Then, exactly as we have done in the
case of pin-jointed structures and frames, the stiness matrix of each
element can be assembled with the stiness matrices of all the other elements to produce the combined stiness matrix of the problem. The
forces applied to the structure are reduced once again to equivalent
nodal loads and the boundary conditions can be imposed very straightforwardly along the edges of the constrained elements. It may be easily
inferred that the fundamental basis of the nite element is in fact the
localised RayleighRitz method introduced and exemplied in earlier
chapters.
The procedure is very general and requires only limited data to
particularise the analysis to any structural form, i.e. the geometry of
the structure, the shape of the elements, the material properties. The
trial functions are often the same for each element, thus an analysis
can be fully automated on a computer by means of a standard set of
instructions. By following the standardised method even large-scale
nite element problems can be treated in a very straightforward and
almost routine manner.

The theory of nite elements

305

It seems reasonable that a better approximation may generally be


obtained by increasing the number of sub-divisions, which is certainly
more simple than including more and more complex trial functions by
the classical Ritz method. However, despite the fact than the subdivision into simple elements and the equations of compatibility and
equilibrium between the elements have a sound physical meaning,
the problem of the convergence of the nite element solution to the
exact mathematical one may be quite complex and has given origin
to many signicant studies during the last 30 years. Apart from
some brief notes regarding the physical interpretation of the nite
element method, we will not deal with the topic of a formal proof of
convergence, i.e. the application of precise mathematical theorems to
the nite element method or the estimation of accuracy inherent in
the nite element approximation, nor will we give any detail beyond
a very simple outline of the basic procedure. Excellent textbooks
exist on this topic and the interested reader can make reference to
them in order to gain a more mathematical understanding of this
basic tool in current engineering practice.

11.1

Sub-division of a structure into nite elements

We illustrate the basic steps of the nite element method by referring to


the example of a simple structure, i.e. a retaining wall with lateral pressure, shown in Fig. 11.1.
As we have already seen in chapter 3, if the dimension of the structure is very large in the z-direction and the lateral pressure is perpendicular to the longitudinal axis z and does not vary along the length,
we may assume that all the cross sections are in the same state. In
fact, we can suppose that the end sections are conned between
P

Fig. 11.1

306

Energy methods in structural mechanics

Fig. 11.2

smooth rigid planes, which implies that displacement in the axial direction is prevented at the ends and, by symmetry, also at the mid-section.
Consequently, we can assume that displacement in the axial direction
is prevented at any cross section. Thus, the deformation of the wall can
be assumed to be the same at any cross section and can be dened completely by the non-zero components of displacement ux and uy , which
are assumed to be functions of x and y only.
Hence, as already noticed in chapter 3, we have
ux ux x; y

uy uy x; y

quy
qux
"yy x; y
qx
qy


1 qux quy
"xy x; y

2 qy
qx


1 qux quz

0
"xz x; y
2 qz
qx


1 quy quz

0
"yz x; y
2 qz
qy
"xx x; y

uz 0
"zz

quz
0
qz

There are several important engineering problems of this kind, like, for
example, tunnels or cylindrical tubes subject to internal pressure, as
illustrated in Fig. 11.2.
Despite the fact that the problem of elastic equilibrium is reduced to
a two-dimensional model in the xy plane, nonetheless it remains a
quite complex problem and, apart from some cases characterised by
a very simple geometry of the boundary, no closed-form mathematical
solutions are readily available to engineers.
We make resort to the nite element procedure and start by conceptually sub-dividing the structure into zones, i.e. elements. This can be
done, for example, by means of four-noded plane elements as shown
in Fig. 11.3.

The theory of nite elements


14

15

16

13
7

11
6
7

12

10

307

Fig. 11.3

In this way we substitute our structure with an assembly of zones,


whose deformations must be characterised by a nite number of
parameters to permit the replacement of the continuum problem by
a discrete one. It is clear that in the vast majority of structural
problems these parameters may be assumed to be the displacements
of the nodes. If we reduce also the applied loads to equivalent joint
loads, the solution of the whole problem follows exactly the same
rules as those applicable to standard discrete systems, like the trusses
and frames examined in chapters 8 and 9.
Of course, the key to the whole procedure lies in the denition of the
behaviour of each individual element in terms of the displacements of
its nodes. From a kinematic point of view this means that the displacements of the internal points of the element must be functions of the
displacements of the nodes and satisfy basic requirements of compatibility at the boundaries. Thus, for example, the displacement along
contiguous sides of adjacent elements may be required to be identical
for any specied displacement of the nodes and, in certain cases, this
requirement of continuity may be extended to the slope of the displacement. The denition and the employment of such interpolating
functions is the subject of the following sections.

11.2

Displacement functions and shape functions

We have already encountered the concept of displacement functions in


the presentation of the RayleighRitz procedure, where trial functions

308

Energy methods in structural mechanics


y

uy

uy
L

ux

ux

l
uy

uy
I
ux

ux

Fig. 11.4

dependent on a certain number of parameters were employed to


represent the displacement eld of the elastic body under analysis.
The values of these parameters were determined by means of a minimisation of the total potential energy of the system. However, in
that case, trial functions extended over the whole structure, whereas
in the nite element method the displacement functions extend over
a single sub-domain of the region under analysis. This concept was
introduced in section 7.6 under the heading localised RayleighRitz.
The localised nature of the trial functions in the nite element
method is further examined in this section.
If we restrict ourselves to the plane strain problem introduced in the
previous section, we can take into consideration the four-noded rectangle shown in Fig. 11.4 and express the displacement eld ux; y
over this sub-region as a function of the displacements of its nodes.
For example, we can assume the components along x and y of the
displacement eld are described by polynomial expressions of the type
ux x; y a1 a2 x a3 y a4 xy

11:2

uy x; y b1 b2 x b3 y b4 xy

11:3

By inserting the coordinates of the four nodes of the element in


equations (11.2) and (11.3) we obtain a system of linear equations in
the eight unknowns a1 ; . . . ; b4 , that is
ux I a1

ux H a1 a2 l a3 l a4 l 2

uy I b1

uy H b1 b2 l b3 l b4 l 2

ux J a1 a2 l
uy J b1 b2 l

ux L a1 a3 l
ux L b1 b3 l

11:4

The theory of nite elements

309

Solving this system and substituting the results in equations (11.2) and
(11.3), we have





x
y
x
y
ux x; y 1
1 ux I
1 ux J
l
l
l
l


xy
y
x
1
u L
11:5
2 ux H
l
l x
l





x
y
x
y
1 uy I
1 uy J
uy x; y 1
l
l
l
l


xy
y
x
1
u L
11:6
2 uy H
l
l y
l
Therefore the displacement functions, equations (11.5) and (11.6), can
be taken to represent the displacement eld inside the nite element
under consideration with respect to the displacement of its nodes.
At this point it must be stressed once again that the displacement
functions, similar to the situation in the RaleighRitz method, are
not exactly the displacement eld that would be provided by the solution of the elastic equilibrium equation for the particular structure
under consideration, but instead, represent an approximation. However, as the solution of the elastic equilibrium problem cannot be
obtained in closed form for the vast majority of engineering cases,
this approximation is very often the only one available to engineers
and in most cases may indeed be made very accurate.
For the sake of generality we can represent the displacement functions, equations (11.5) and (11.6), in matrix form. This is easily
obtained by writing
u  

0 NL 0
NI 0 NJ 0 NH
x

Di
uy
0 NI 0 NJ
0 NH 0 NL
NDi

11:7

where D is the array containing the displacement of the nodes of the


i-th element, i.e.
Di T ux I uy I ux J uy J ux H uy H ux L uy L
11:8
and Nj x; y are the so-called shape functions.
The shape functions have a crucial importance in the nite
element method. By reference to equation (11.7) it seems clear that
in general they have to be chosen such as to provide appropriate

310

Energy methods in structural mechanics

nodal displacements when the coordinates of the corresponding nodes


are inserted in the shape function. In other words, we must have
NI xI; yI 1

11:9

NI xJ; yJ 0

11:10

In the case at hand the shape functions take the form





x
y
1
NI x; y 1
l
l


x
y
NJ x; y
1
l
l
xy
NH x; y 2
l


y
x
1
NL x; y
l
l

11:11
11:12
11:13
11:14

As both the components of displacement are interpolated in an


identical manner, four shape functions are sucient to represent the
displacement functions, equations (11.5) and (11.6). The graphical
form of the shape functions NI NL over the rectangular sub-domain
under consideration are shown in Fig. 11.5.
It is worth noting that both the components of displacement ux and
uy on the four sides of the rectangle, i.e. the straight lines of equation
x 0, y 0, x l and y l, respectively, depend only on the displacement of the nodes that belong to the sides themselves. For
example, the displacement of the side connecting the nodes J and H
is given by the expressions


y
y
ux x l; y 1 ux J ux H
11:15
l
l


y
y
11:16
uy x l; y 1 uy J uy H
l
l
and similar expressions hold true for the remaining sides. This means
that two adjacent rectangular elements, like those represented, for
example, by the pairs 1 and 4 or 4 and 7, in Fig. 11.3, are characterised
by the same displacement on contiguous sides, provided, of course,
these latter are interpolated by means of equations (11.5) and (11.6)
on both the sub-domains.
Now let us turn our attention to the trapezoid elements no. 2, 5 and
8. A generic element of this set is shown in Fig. 11.6. Following the
same line of reasoning adopted for the rectangular elements, we can

The theory of nite elements

311

NI

NJ
L

NH

NL

Fig. 11.5

l
y
uy

uy
L

ux

ux

l
uy

uy
I
ux

Fig. 11.6

ux

312

Energy methods in structural mechanics

again consider the components of displacement in the form given by


equations (11.2) and (11.3). In order to obtain appropriate nodal
displacements corresponding to the coordinates of the nodes, we are
now led to the following system of equations in the eight unknowns
a1 ; . . . ; b4
ux I a1

ux H a1 a2 l a3 l a4 l 2

uy I b1
ux J a1 a2 l

uy H b1 b2 l b3 l b4 l 2
11:17
ux L a1 a3 l

uy J b1 b2 l

ux L b1 b3 l

The substitution of the solution of the above system in the expressions,


equations (11.2) and (11.3), yields





x
y
x
y
1 ux I
1 ux J
ux x; y 1
l
l
l
l


xy
y
x
1
u L
2 ux H
11:18
l
l x
l





x
y
x
y
1
uy I
1
uy J
uy x; y 1
l
l
l
l


xy
y
x
1
u L
11:19
2 uy H
l
l y
l
which we may take as the expressions of the components of the
displacement eld in the trapezoidal sub-domain, Fig. 11.6. Again,
we can express the displacement function in matrix form, equation
(11.7), provided the shape functions Nj x; y have the following
expressions



x
y
1
11:20
NI x; y 1
l
l


x
y
NJ x; y
1
11:21
l
l
xy
NH x; y 2
11:22
l


y
x
1
11:23
NL x; y
l
l
At this point of the presentation we can anticipate that this way of
proceeding may present serious drawbacks. For example, even if we
assume the same general polynomial expression for the displacement

The theory of nite elements

313

functions (in the present example, equations (11.2) and (11.3)) and the
meshing of the structure is achieved using quadrilateral elements only,
it is evident that the resulting expressions for the displacement function
varies according to the shape of the element and, in general, cannot be
derived as a single function for all elements. Moreover, the conditions
fullled by the displacement functions may vary according to the shape
of the element as well. This factor can be easily explained by reference
to the property of two adjacent rectangular elements to present the
same displacement on contiguous sides.
This property does not hold true for the trapezoidal element,
Fig. 11.6. In fact, the side connecting the nodes J and H is a straight
line of equation
x l y

11:24

and the displacements of its points are given by the expressions




y

ux x l y ; y
1
u I
l
x




y

y
1
1 ux J
1
l

l


2
y y

1 ux H

l l 2



y
1 ux L
1
l
11:25


y

1
u I
uy x l y ; y
l
y




y

y
1
1
uy J
1
l

l



y y2
2
1 uy H

l l




y
1 uy L
1
l
11:26
These expressions involve the components of displacement of all the
four nodes of the element. Therefore, the displacement along contiguous sides of adjacent elements is not governed only by the behaviour
of the nodes associated with the sides themselves and, in general, it
will not be the same for both of elements. The elimination of such a

314

Energy methods in structural mechanics

pathological behaviour would imply a dierent choice of the polynomial displacement functions, equations (11.2) and (11.3). Generally
speaking, this means that an appropriate analysis of the displacement
functions is required, at least for any dierent set of elements.
However, in order to obtain a procedure with the greatest degree of
automation, it is desirable that a small amount of data can be used to
represent the various types of elements that may be required in the
analysis of a real engineering problem.
This can be obtained by making resort to the use of intrinsic coordinates and appropriate transformation rules, as we will see in the
next section.

11.3

Element mapping and intrinsic coordinates


isoparametric elements

Up to this point, we have discussed some diculties that arise in the


direct generation of the displacement functions for the few elements
employed for the simple meshing of the retaining wall, Fig. 11.3. If
we are to deal with more complex practical engineering structures,
which are characterised by quite complicated geometrical boundaries,
it would be valuable to nd a way to construct nite element models
using only a fairly restricted library of elements. This can be done by
mapping a simple element (the so-called parent element), such as a
rectangle in the local coordinate system ; , into a more complex
shape in the general coordinate system xy, as shown in Fig. 11.7.
In the present context, mapping means a unique, one-to-one
relationship between the local system of coordinates ,  and the
general reference system xy. It should be noted that in this case the
relationship between the local system of coordinates and the general
one is somewhat broader than the relationship between the local and
L

H
L

I
I

J
J

Fig. 11.7

The theory of nite elements

L
r2

r1
I

r2

J
r1

315

Fig. 11.8

the general reference frames encountered in the matrix analysis of pinjointed and rigid-jointed structures. In fact, in the previous situation
we dealt with a transformation rule that allowed us to establish a
correspondence between systems of Cartesian orthogonal axes that
diered by simple translations or rotations in the three-dimensional
space. Here we have relationships that allow a continuous distortion
of the element, too. For instance, we can consider the relationship
between cylindrical polar and Cartesian coordinates, i.e.
x r cos 

y r sin 

11:27

that may imply the mapping of a rectangular domain in the local


system r  into a distorted element in the xy space. This situation
is illustrated in Fig. 11.8.
For the sake of convenience the local system  is generally
assumed to be a system of natural or intrinsic coordinates, i.e. a
system of non-dimensional coordinates varying between 1 and 1,
see Fig. 11.9.
In the case of four noded elements, this means that they can be
derived from a parent element which is a square whose sides measure
2. With regards to the transformation rules between the coordinate

Fig. 11.9

316

Energy methods in structural mechanics

systems, a very convenient method of establishing them is to make use


of the same shape functions that we use to represent the displacement
function. This means that the array ci of the coordinates of the nodes
of the element in the general reference frame is used to dene the global
coordinates x; y; z of any point in the element itself. For instance,
we can perform this by deriving a set of shape functions Nj ;  in
the local reference frame  according to the general procedure
introduced in section 11.3 (that is, they must have a unit value at the
reference node and zero at all the others) and by writing for each
element the following transformation rules
x;  NI ; xI NJ ; xJ
NH ; xH NL ; xL

11:28

y;  NI ; yI NJ ; yJ


NH ; yH NL ; yL
In matrix form we can write


x; 
N; ci
y; 

11:29

11:30

where N;  is the shape function matrix and ci is the array containing the global coordinates of the nodes of the generic element, i.e.
ci T xI yI xJ yJ xH yH xL yL
11:31
In this manner we obtain a set which is readily suitable for mapping.
In fact, it is evident that the locations of coordinates I,
I; . . . ; L; L, i.e. the four nodes of the parent element, will automatically assume the required coordinates xI; yI; . . . ; xL; yL in
the mapping. Continuity and uniqueness requirements are generally
satised by such a type of transformation, with the exception of
excessive and pathological distortions. The feature can be checked
by making reference to the Jacobian matrix (after Jacobi1 ) of the transformation, that is
3
2
qx qy
6 q q 7
7
6
11:32
J 6
7
4 qx qy 5
q
1

q

Jacobi, Karl Gustav Jacob (Potsdam, 1804 Berlin, 1851), German mathematician.

The theory of nite elements

317

The Jacobian matrix of the transformation can be trivially calculated


from equation (11.30), as
2
3
qNI
qNL
0
...
0
6 q
7 i
q
7c
J 6
11:33
4
qNI
qNL 5
...
0
0
q
q
From calculus we know that the condition for a one-to-one mapping
is that the sign of the determinant of equation (11.33) remains
unchanged at all the points of the domain under mapping. When
such a transformation rule is adopted, the elements will be called
isoparametric. One of the most signicant advantages of the isoparametric formulation is that if two adjacent elements are generated
from the same parent element, whose shape functions make the
displacement of any side dependent on the behaviour of only the
nodes belonging to the side itself, then the distorted elements will be
always contiguous, before and after deformation has occurred. This
can be demonstrated very easily by making reference to Fig. 11.10.
It is evident that along a common edge, such as the line IJ, elements
A and B have the same tangent coordinate, being A B . Moreover,
as the mapping into the distorted elements is governed along the
common edge by the same functions of , i.e. the shape functions Nj ,
J

B
A

Fig. 11.10

318

Energy methods in structural mechanics

and the sets of global nodal coordinates are the same, it follows that
the inter-element boundary curve IJ will also be the same. The
same line of reasoning applied to the displacement functions, equation
(11.7), demonstrates that displacement is also the same along a
common edge. It goes without saying that the continuity of the displacement functions is preserved in the mapping and that isoparametric elements will be invariant as a consequence of the adoption of
the intrinsic coordinate system .
Finally, it must be underlined that, although we have limited our
development here to two-dimensional structures, nevertheless all the
comments and results derived above are equally valid and applicable
to the analysis of three-dimensional structures and components.

11.4

Convergence criteria for the displacement


functions the patch test

Now it is appropriate to consider some conditions to which displacement functions should conform so that it is likely that the exact mathematical solution of the problem of elastic equilibrium will be
approached by the nite element treatment as more and more elements
are used to mesh a certain structure. From a rigorous mathematical
standpoint this is quite a complex matter and in the past 30 years
many studies have been devoted to underwriting the intrinsically
approximate nite element method with a formal mathematical basis.
A review of such studies and the mathematical developments are outside the purpose of this textbook and we limit the following presentation to self-evident and heuristic physical convergence requirements.
These requirements are
.
.

the displacement eld must be continuous within every subdomain. This is a fairly simple requirement and is met by any
polynomial displacement function.
inter-element compatibility should be respected, i.e. there should
be no gaps or overlaps between elements. Moreover, bending
elements (like, for instance, those associated with the Kircho
plate model or the Bernouilli beam model) should not allow discontinuities in slope between elements. The rst point of this
statement is physically evident and we have seen that the use
of the isoparametric formulation provides a straightforward
way of meeting such a condition. The second point is physically
evident, too. However, it is rather more complicated to meet
such a requirement and there are many element formulations
of common use that violate the requirement in a coarse mesh
arrangement. Nevertheless, these elements may be able to

The theory of nite elements

319

yield a good convergence towards the exact mathematical solution because incompatibilities in slope tend to disappear with
increasing mesh renement. Also, incompatible elements are
very often deliberately adopted because they tend to `soften'
the behaviour of the model and counteract the natural overstiness of the displacement functions. In fact it must be
borne in mind that by substituting the real solid by a nite
element model we seek the solution of the elastic equilibrium
(which for linear problems corresponds to a minimum of the
total potential energy) in a class of functions which is a subset
of the class of all the kinematically admissible functions and
therefore provides displacements that are generally smaller
than the real ones.
the elements should be geometrically invariant or isotropic. This
requirement is once again physically evident and means that
the element behaviour should have no preferred direction. It
implies that the polynomial expression adopted to interpolate
the displacement eld over the element sub-domain should be
symmetric with respect to any coordinate. Of course obtaining
a set of isotropic elements is quite easy in an isoparametric
formulation, provided the assumed displacement functions for
the parent element retain symmetry with respect to all the
intrinsic coordinates.
the elements should be able to represent rigid-body motions
exactly. Of course, a reliable nite element model must be able
to represent correctly this situation, which may occur in the
whole structure or only in a part of it. An obvious implication
of this requirement is the following: when the vector of nodal
displacements d represents a rigid body motion, the element
must exhibit a zero strain eld. From a mathematical standpoint, this implies that the polynomial adopted for the representation of the displacement eld must possess appropriate and
adequate constant terms.
the elements should be able to represent constant strain elds.
From a physical standpoint this last requirement may seem
less obvious than the previous ones, but it is nevertheless a
very signicant requirement. We can imagine any continuous
structure as a system characterised by an innite number of
elements, each one displaying a constant strain, which is the
value of the strain at each point of the whole domain. Thus, it
seems appropriate that a nite element is able to reproduce
constant strain elds so that, as the mesh is further and further
rened and the sub-division of the domain becomes ner and

320

Energy methods in structural mechanics


qz (z)

z
y
l

z (z)

Fig. 11.11

ner, elements will approach the state of constant strain in the


neighbourhood of each point.
In order to give a simple example of this situation, we can make
reference to an axially loaded bar, Fig. 11.11.
The form of the axial strain "z z is displayed below the bar. We may
now sub-divide the bar in a specied number of nite elements, say n,
as shown in Fig. 11.12.
If the elements are able to represent a constant strain eld, the actual
strain eld in Fig. 11.12 can be approximated by the piecewise diagram
shown in Fig. 11.13 and as n ! 1 the real strain eld will be
approached by the nite element solution.
qz (z)

z
l/n

y
l

Fig. 11.12

zz (z)

Fig. 11.13

The theory of nite elements

321

Fig. 11.14

If, however, the elements are not able to represent a constant strain
eld, it may be that even when the dimension l/n of each element
becomes very small, the form of the axial strain resulting from the
nite element modelling will still dier signicantly from the exact
solution, as shown in Fig. 11.14.
We recall that in the linear theory of elasticity the strain eld within
the structure is derived from the displacement eld by means of the
well-known relationships
qu
"xx x
qx
...


11:34
1 quy qux
"xy

2 qx
qy
...
and it follows that a complete linear polynomial is required in order to
represent constant strain states. A complete quadratic polynomial is
required in order to represent constant curvatures in models for structural exure, such as Kircho's plates. As a generality, we can say that
it is desirable that polynomial expressions of displacement elds possess the highest order of a complete polynomial for the assigned
number of parameters to be determined. The whole set of complete
polynomials of any order can be shown graphically by means of the
Pascal2 triangle, Fig. 11.15.
However, it must be underlined that this condition, like each of the
previous requirements, if taken alone, is neither a necessary nor a sufcient condition to guarantee the convergence of the nite element
model to the exact mathematical solution. On the whole, it can be
armed that displacement functions that satisfy the above requirements show a good convergence to an accurate solution and do not
display strange or pathologic behaviours even for a relatively coarse
discretisation of the structure.
A practical test of validity for a generic non-conforming element (i.e.
those elements that do not respect the second of the conditions above,
2

Pascal, Blaise (Clermont-Ferrand, 1623Paris 1662), French mathematician, physicist


and religious philosopher.

322

Energy methods in structural mechanics


1

2 2

Fig. 11.15

which is normally the case for elements used to model plates and shells,
as discussed earlier) is the so-called patch test. This test consists of
assembling an arbitrary patch of elements in such a way that at least
one node is completely surrounded by elements, see Fig. 11.16.
Then, a boundary nodal displacement that corresponds to a state of
constant strain is applied. Once the equations of elastic equilibrium
have been solved for the nite element model, the test is considered
to be passed if resulting strains agree exactly with exact values at
every point in any sub-domain. Of course, in order to be signicant,
the test must be repeated for more than one geometry, mesh arrangement and state of strain. In fact, some elements may pass the test in
some conditions and fail it in others. It can be shown that, under
some precise requirements, the patch test is a necessary and sucient
condition for convergence. In any case and apart from precise

Fig. 11.16

The theory of nite elements

323

mathematical reasons, the patch test is also the most eective and
practical benchmark for nite element programming and checking.

11.5

Some general families of shape functions

In section 11.2 we started from simple polynomial expressions for


the components of the displacement function in the xy plane, i.e.
equations (11.2) and (11.3). The choice of a four-term polynomial
expression in x and y was obviously suggested by the fact that we
were dealing with a four-noded element and thus we had four
conditions to impose for each component of displacement in order
to make the value of the displacement function at the position of
each node equal to the displacement of the node itself. In general
this approach can be repeated for any parent element in intrinsic
coordinates, regardless of the number of nodes. As an example of
this, let us consider the eight-noded element shown in Fig. 11.17.
The components of displacement at eight points must determine
uniquely the variation of the displacement function exemplied by a
polynomial expression in  and  with eight free coecients, such as
ui a1 a2  a3  a4  a5 2 a6 2 a7 2  a8 3
11:35
The substitution of the coordinates of the nodes I; . . . ; P will provide
a system of eight linear equations in the eight unknowns a1 ; . . . ; a8 , i.e.
2
3 2
32 3
1 I . . . 2 II 3 I
ui I
a1
4 ... 5 4... ... ...
...
. . . 54 . . . 5
a8
ui P
1 P . . . 2 PP 3 P
11:36
O

Fig. 11.17

324

Energy methods in structural mechanics

or, in more symbolic form


d i Ca

11:37

The solution of this system provides the unknown coecients


a1 ; . . . ; a8 , that is
a C1 d i

11:38

Thus, equation (11.35) may be written


ui ;  a C1 d i

11:39

where
  1



2

2

2 

3

11:40

Comparison with the general expression, equation (11.7), allows us to


conclude that in this case the required expression of the shape functions is
N NI

. . . . . . NP C1

11:41

However, there are some serious drawbacks to this procedure. First,


the existence of the inverse of C1 is not always guaranteed. Moreover, even when the inversion of C is possible, it is still a quite complicated and lengthy formal procedure for many element geometries.
Finally, the choice of the lowest possible power for each term of the
polynomial expression does not necessarily lead to expressions that
are symmetric with respect to  and . This is the case in equation
(11.35), which is a polynomial of order 3 that does not include the
terms 3 and 2 , as can be easily observed from the Pascal triangle,
Fig. 11.15.
In this manner the element shown in Fig. 11.17 has preferential
directions, allowing a fully cubic variation of the components of displacement only with respect to the  coordinate and thus violates the
third of the conditions stated in the previous section. Therefore it
appears advantageous to seek to establish a procedure to specify the
shape functions directly and systematically with full compliance with
the required conditions. Actually, there are several ways to full this
goal and each one results in a precise family of shape functions.
Here we will consider two basic and popular `families' of shape functions: the Lagrange family and the serendipity family.
The Lagrange family of shape functions stems directly from the
expression of Lagrange polynomials in one coordinate. It is known
from elementary calculus that these polynomials have the following
remarkable property: given n points on the  axis, it is always possible

The theory of nite elements

325
1

Fig. 11.18

to construct a Lagrange polynomial that has unit value at one point,


say the k-th, and zero value at all the other n 1. The situation is
illustrated in Fig. 11.18.
The general expression of any Lagrange polynomial in one coordinate is given by the relationship
Lnk 

 1  2 . . .  k 1  k 1 . . .  n
k 1 k 2 . . . k k 1 k k 1 . . . k n
11:42

which immediately yields


Lnk k 1

and Lnk  j 0

for j 6 k

11:43

It follows from equation (11.42) that a Lagrange polynomial


through n points is a polynomial of order n 1. If we take into consideration a generic rectangular element with r  s nodes, as shown
in Fig. 11.19, a whole family of shape functions can be written down
directly from the expression of Lagrange polynomials.

s r

r
1

326

Energy methods in structural mechanics

In fact, by labelling the node by its column and row number, say h
and k, we have
Nj ;  Lrh Lsk 

11:44

It is clear from equation (11.44) that the fundamental advantage of


such a family of shape functions lies in the extreme simplicity of its general expression. Additionally, as for a rectangular element both r and s
result  2, it follows that the expression, equation (11.44), will always
include a complete polynomial of order 1, thus allowing the representation of rigid displacements and constant strains. Also, if we assume
r s, as is usually the case for parent elements in intrinsic coordinates,
geometrical isotropy is automatically assured.
However, from Fig. 11.19 it is also clear that the Lagrange family
may imply a large number of internal nodes for each element. Despite
the fact that the behaviour of such nodes can be linked to the displacements of the nodes on the edges by means of a standard substructuring procedure (according to the same technique by which,
for instance, we can make the displacement of the internal node M
of the truss of Fig. 11.20 an explicit function of the displacements of
the end nodes I, J, H, L), nevertheless the presence of such degrees
of freedom can adversely aect the convergence of the nite element
model.
Moreover, if we take into consideration the generic shape function
which is derived from equation (11.44) for a 3  3 nodes element, i.e.
Nj ;  a1 a2  a3  a4 2 a5  a6 2
a7 2  a8 2 a9 2 2

11:45
u (H)

u (L)
L

u (I)
I

Fig. 11.20

u (J)

The theory of nite elements


L

327

Fig. 11.21

we see that it contains some high-order terms and it omits some lower
ones. In fact equation (11.45) contains, together with the complete
second order expansion, the higher order terms 2 , 2 and 2 2 ,
while it omits the terms 3 and 3 (needed for a complete third-order
expansion) and the terms 4 , 3 , 3 and 4 (needed for a complete
fourth-order expansion). This holds true for every element of the
Lagrange family, which therefore does not oset the welcome property
that polynomial expressions of displacement elds possess the highest
order of a complete polynomial for the assigned number of parameters
to be determined.
The serendipity family of shape functions is somewhat slightly less
simple to generate, but it oers substantial advantages, too. The name
serendipity was given to this class of shape functions after the princes
of Serendip, who were reported by Horace Walpole3 to be remarkable
for their chance discovery. The basis for the name stems evidently
from the way in which we derive the serendipity shape functions.
In fact, let us consider the four noded square element in Fig. 11.21.
The equations of the lines JH and LH are, respectively,
1 0

1 0

11:46

Therefore the function


P;  1 1 

11:47

will give zero value for the coordinates of the nodes J, H and L. Thus,
we can write
NI ;  a1 1 

11:48

and impose that NI 1; 1 1, which yields a 1=4. Consequently,


we have
NI ;  14 1 1 
3

Walpole, Horace (London, 17171797), English writer.

11:49

328

Energy methods in structural mechanics

Fig. 11.22

By repeating the same reasoning for each of the four nodes of the
element in Fig. 11.21, we are led to the general formula
Nj ;  14 1 j 1 j 

11:50

It is clear that however simple it seems, this procedure requires some


ingenuity for any element we may wish to consider. In fact, if we shift
our attention to the eight-noded element shown in Fig. 11.17, we have
to identify for each node a certain number of lines that pass, in turn,
through all the other nodes and these may not be simply the sides
which do not contain the node in question. As an example, consider
a corner node, such as the node I, for which the required lines
are JP, HM and OM, as shown in Fig. 11.22. Their respective
equations are
1 0

1 0

1 0

11:51

and the appropriate expression for the shape function NI ;  is


NI ;  a1  1 1 

11:52

Then, the condition NI 1; 1 1 yields


NI ;  14 1  1 1 

11:53

For a side node, such as the node J, the required lines are HM, OM
and IO, as shown in Fig. 11.23.
Their respective equations are
1 0

1 0

1 0

11:54

The theory of nite elements

329

Fig. 11.23

and the expression of the shape function NJ ;  is


NJ ;  a1 1 1 

11:55

Once again, the condition NJ 0; 1 1 yields


NJ ;  12 1 1 1 

11:56

By iterating the same line of reasoning, for the example element we


obtain the following general formulae
corner nodes:

Nj ;  14 j  j  11 j 1 j 
11:57

mid-side nodes: Nj ;  12 1 2 1 j 
Nj ;  12 1 j 1 2

for j 0
for

j 0
11:58

In general the serendipity shape functions require fewer degrees of


freedom for a certain complete polynomial expansion than the Lagrangian family. Moreover, they do not require the presence of internal
nodes a priori, but this can become necessary in cases where the
presence of complete polynomials beyond the cubic for mid-side
nodes is required for particular analyses.

330

Energy methods in structural mechanics

11.6

Strain energy, work of applied loads and


equilibrium equations

Once the kinematic behaviour of any element has been interpolated in


terms of shape and displacement functions, the study of the elastic equilibrium of the nite element model becomes simple and straightforward.
For the sake of simplicity and clarity of presentation, we continue
with the example of the two-dimensional plane strain problem
introduced in section 11.1. From a conceptual point of view the
development of the theory for a generic three-dimensional problem
is identical to that for the two-dimensional case and the extension
should present no real diculties.
The simplest way to study the problem consists once again in writing
down the total potential energy of the nite element model and seeking
the values of the degrees of freedom of the system, i.e. the displacements of its nodes, corresponding to the stationary value of this expression. This is exactly what we have already described in chapters 8 and 9
for the systems of bars and beams and most of the observations made
at that stage remain valid also in the theory for nite elements.
Therefore, let us recall the general expression of the total potential
energy, that is equation (6.2)
UW
Also in the present case the strain energy of the whole structure is given
by the sum of the energy stored in each of its elements, i.e.
X i
U
U
11:59
i

where, according to chapter 3, in the global reference system xy we


can write equation (3.107) in a more compact form

1
Ui
" x; yCijhk "hk x; y dA
11:60
2 ij
Si

S is the domain of the generic element and Cijhk are the constants
which dene the elastic properties of the material. Remember that
we are examining a system of unit depth, so that the integrals can be
performed over the area A of the elements. For the sake of compactness, in the case at hand the components of strain "ij x; y can be
ordered in a three-element array, i.e.
3
2
"xx x; y
7
6
11:61
"x; y 4 "yy x; y 5
2"xy x; y

The theory of nite elements

331

and the elastic constants Cijhk can be set in a 3  3 matrix, that is


2
3
1 

0
6 
E
1 
0 7
6
7
11:62
C
4
1 1 2
1 2 5
0
0
2
Notice that the elastic constants matrix (11.62) can be straightforwardly derived from the relationships (3.93)(3.95) and (3.99) by
setting "zz 0 in equation (3.95) and then substituting the resulting
value of zz in the preceding two equations.
The straindisplacement relationships can also be expressed in
matrix form,
2 q
3
3 6 qx 0 7
2
"xx
6
7" #
q 7 ux
6 " 7 6
7
" 4 yy 5 6
11:63
6 0 qy 7 u
y
6
7
2"xy
4 q
q 5
qy

qx

or, more compact form,


" Su

11:64

where S formally represents a dierential operator.


Thus, the strain energy of the generic element can be written as

1
1
i
T
U
" C" dA
uT ST CSu dA
11:65
2
2
Si

Si

By substituting the equation (11.7) for the displacement function, in


equation (11.65) i.e. ux; y Nx; yDi , we get, nally

1
i
U
Di T NT ST CSNDi dA
2
Si

1
Di T
2

NT ST CSN dADi

11:66

Si

The above formula can be put into a form identical to equations


(8.40)(8.42), obtained in the case of frame of structures, by setting

11:67
K i NT ST CSN dA
Si

332

Energy methods in structural mechanics

so that equation (11.66) can be written as


U i 12 Di T K i Di

11:68

We remember that Di is the array that contains the displacements of


the nodes of the generic four-node element in the global reference
frame. Thus for the present example, see equation (11.8)
Di T ux I uy I ux J uy J ux H uy H ux L uy L
K i is therefore an 8  8 matrix. In order to make the formulae as
general as possible, we take into consideration the vector D, which
contains all the nodal displacements of the nite element model of
Fig. 11.3. We have
DT ux 1 uy 1

...

uy 9 ux 9

...

ux 16 uy 16
11:69

At this point, exactly as we have done in the case of framed structures,


we can identify the joints I, J, H and L of any element in the general
numbering system. For example, for element 7 in Fig. 11.3, we have
I

! 10

! 14

! 13

11:70

On the basis of this identication, we can enlarge the matrix K i from


an 8  8 array to a 32  32 array simply by setting its terms in the
appropriate position and leaving all the others as zero. Thus, we can
write
U i 12 DT K i D

11:71

and the strain energy of the whole structure is


X i X1 T i
D K D
U
U
2
i
i

1 T X i
1

D
K D DT KD
2
2
i

where

K

X
i

K i

11:72

11:73

The theory of nite elements

333

Let us now take into account the body forces b that act on the
domain S of the structure, and the tractions f that act on its boundary,
. If we order their components along the axes of the global reference
frame as two arrays, i.e.




bx x; y
fx x; y
bx; y
f x; y
11:74
by x; y
fy x; y
we can write the work of the applied loads as

W bj uj dA fj uj ds
S

ux; yT bx; y dA ux; yT f x; y ds

11:75

Given the property of the addition operator for integrals, the above
expression can also be written as
X T
X T
u b dA
u f ds
11:76
W
i

Si

where S i is the sub-domain of the generic element and i is its boundary.


By substituting the equation (11.7) for the displacement function
into equation (11.76), that is u NDi , we obtain
X i T T
X i T T
W
D N b dA
D N f ds
i

X
i

Si

9
=
D
N b dA N f ds
: i
;
i
i

8
T<

11:77

Again, in order to obtain a formulation formally identical to the one


presented in chapters 8 and 9 for the framed structures, we can set
8
9

<
=
i
T
T
11:78
P
N b dA N f ds
: i
;
i
S

In the present example the vector P i is given by the integration of the


product of an 2  8 matrix, NT , by a two-element vector, i.e. b and
f , which gives an eight-element column array.

334

Energy methods in structural mechanics

Thus, we have
X iT i
D P
W

11:79

Similarly to what has been done before, identication of the joints I,


J, H and L of any element in the general numbering system allows us to
enlarge the eight-element vectors P i to 32-element arrays. This is
done by setting the terms of P i in the appropriate positions and
leaving all the other elements as zero. In this manner we can write
X i
X T i
W
D P DT
P DT P
11:80
i

where
P

P i

11:81

At this point the total potential energy of the nite element model is

 U W 12 DT KD
DT P

11:82

and, from a formal point of view, appears exactly the same as equation
(8.47) obtained in the case of framed structures, as we set out to show.
The solution of the elastic equilibrium problem can be pursued by
seeking the stationary value of the total potential energy expression
equation (11.82) with respect to the degrees of freedom set in the displacement vector D. This means that we have to impose that
q
0
qDj

11:83

where Dj is the generic component of the displacement vector D.


In the case of the retaining wall, Fig. 11.3, equation (11.83) is a linear
system of 32 equations in 32 unknowns, namely the components of the
vector D dened by equation (11.69). In matrix notation it is

KD
P 0
11:84
 dened by equations (11.73) and
It is evident that the matrix K,
(11.67), represents the combined stiness matrix for the structure,
while the vector P, dened by equations (11.81) and (11.78),
represents the equivalent combined loads vector. As was the case
with framed structures, the system equation (11.84) is singular and
not solvable before the eects of restraints are taken into account.
Imposition of the respect of the kinematic constraints at the nodes 1,
2, 3 and 4 of the nite element model in Fig. 11.3 follows the same
procedure as in section 8.2 and is therefore not repeated here. In the

The theory of nite elements

335

present example introducing the boundary constraints yields a reduced


set of 24 linear equations in 24 unknowns, that are the active degrees of
freedom of the structure, i.e. the components of displacement of the
nodes 516. The system can be written as
KI DI PI
11:85
and admits a unique solution.
The reader can also verify that, once the eect of the constraints has
been imposed, the joint stiness matrix of the problem KI is positive
denite. Restraint displacements, too, can be taken into account
following the same line of reasoning as presented in section 9.6.
Again, as in the case of the framed structures of chapters 8 and 9, the
construction of the equilibrium equations (11.85) can be pursued
directly, without the necessity to repeat every time the underlying
reasoning in terms of energy. This constitutes the direct stiness
method previously introduced. In fact, once the actual structure
has been conceptually sub-divided into nite elements, the only
components required to produce the equilibrium equations (11.84)
are the expressions for the stiness matrices K i , given by equation
(11.67), and of the equivalent nodal loads P i , given by equation
(11.78).
However, both equations (11.67) and (11.78) are expressed in the
global reference frame xy, while we generally establish the expressions of the shape and displacement functions for the generic nite
element in the intrinsic reference frame , as seen in sections 11.3
and 11.5. An appropriate manipulation of data involving the transformation rules is therefore required. Moreover, the computation
of the stiness matrices K i and of the equivalent joint loads P i
involves integration of several functions. In dealing with real largescale complex structures it is clear that these integration operations
must be performed in an automated way by means of suitable
numerical procedures. Both these points will be claried in the next
section.

11.7

Evaluation of stiness matrices and equivalent


nodal loads numerical integration

At the conclusion of the previous section is was noted that we generally


map the elements from the intrinsic system of coordinates  to the
global system xy and therefore we possess the expression of the
shape functions matrix N;  in the intrinsic system , while the
integrations in the expressions equations (11.67) and (11.78), are to
be performed in the global system xy. Some simple mathematical
manipulation is therefore required at this stage. We start from formula

336

Energy methods in structural mechanics

equation (11.67), i.e.

K i Nx; yT ST CSNx; y dA
Si

Basically, in order to perform this integration we require the expression of the matrix
Bx; y SNx; y
2 qN
I
0
6 qx
6
6
qNI
6
6 0
qy
6
6
4 qN qN
I

qy

qx

... ...

qNL
qx

... ...

... ...

qNL
qy

7
7
qNL 7
7
7
qy 7
7
qNL 5
qx

11:86

A way to attain this target is obviously by means of the explicit expression of the shape function matrix N in the global reference frame xy.
However, as said above, in the isoparametric formulation we hold the
expression of the shape functions matrix N;  in the intrinsic
system  and the inversion of the transformation rules, equation
(11.30), may turn out to be quite a dicult task. Fortunately, some
very simple manipulation can provide the matrix Bx; y without
requiring the inversion of equation (11.30). If we consider that the
mapping correspondence equation (11.30) yields x and y as functions
of  and , i.e. x x;  and y y; , the chain rule of derivatives
allows us to write
qNj qNj qx qNj qy

q
qx q
qy q
qNj qNj qx qNj qy

q
qx q
qy q
or, in matrix form,
3
3
2
2
qNj
qNj
6 qx 7
6 q 7
7
7
6
6

J
7
7
6
6
4 qNj 5
4 qNj 5
qy
q

11:87

11:88

where J is the Jacobian matrix of the transformation, given by equation (11.33). Consequently, the terms qNj =qx and qNj =qy required for

The theory of nite elements

337

the expression of the matrix Bx; y are yielded by the formula


2
3
2
3
qNj
qNj
6 q 7
6 qx 7
7
6
7
1 6
11:89
7
6
7 J 6
4 qNj 5
4 qNj 5
qy
q
Of course, the existence of the inverse of the Jacobian matrix of the
transformation is guaranteed under the condition of uniqueness on the
mapping discussed in section 11.3.
At this point, computation of the stiness matrices K i involves the
integration of known functions in the global reference frame xy over
the sub-domain of the nite element, Si . In fact, we have

i

K Bx; yT CBx; y dA
11:90
Si

Also, this integral can be transformed back to the intrinsic system of


coordinates . It is known from calculus that

1 1
f x; y dA

j f~;  d d

11:91

1 1

Si

with
j detJ

11:92

f~;  f x; ; y; 

11:93

and
Thus the integration of equation (11.90) can be performed directly in
the intrinsic system of coordinates of the parent element, given that
the expression of Bx; y is already obtained in terms of  and  by
virtue of equation (11.89). Then, we have
K i

1 1

jB; T CB; d d

11:94

1 1

This integration technique applies also to the computation of the


equivalent nodal loads, equation (11.78)
8
9

<
=
i
T
T
P
Nx; y bx; y dA Nx; y f x; y ds
: i
;
i
S

338

Energy methods in structural mechanics

In this case we have the expression of the shape functions matrix


N;  and we can obtain the expressions bx; ; y;  and
f x; ; y;  of the applied loads directly from substitution of
the transformation rules, equation (11.30). Consequently, we have
81 1
1
<
i
T
P
jN;  b;  d d N; 1T f  ; 1 d
:
1 1



N1;  f 1;  d

where

9
=
N1; T f  1;  d
;

fJ
f s 11
fn J11


fn J12
fs J12


f



N; 1T f  ; 1 d

11:95


fJ
s 21
fn J21

fn J22
fs J22

and fs , fn are the components of the tractions f tangent and normal to


the boundary, respectively.
With regards to the eective evaluation of the integral equations
(11.94) and (11.95) in the intrinsic system of coordinates, it must be
noted that, however possible in a small and mostly academic number
of cases, the exact integration of these expressions is quite dicult
and complicated and generally not feasible in the vast majority of practical applications that may involve complex distorted elements.
Making resort to the techniques of numerical integration is therefore
the standard procedure in almost all nite element analyses. Given
the introductory purposes of this textbook, we will only summarise
here a few elementary principles regarding these numerical integration
operations, referring the interested reader to one of the many valuable
treatises readily available on the subject.
One of the most eective methods of evaluating a denite integral
numerically is the Gauss4 quadrature. The basic idea is well-known:
to approximate a certain one-dimensional integral
1
I

f  d

11:96

1
4

Gauss, Karl Friedrich (Brunswick, 1777 Gottingen, 1855), German mathematician


and physicist, one of the greatest mathematicians of all times.

The theory of nite elements

339

Fig. 11.24
f

Fig. 11.25

over the interval 1; 1 we can take the value f 0 at the mid-point of


the interval and multiply it by the interval length, i.e. 2. The result is
I  2f 0
11:97
This is exact only in the case where f  represents the equation of a
straight line, as shown in Fig. 11.24.
It is clear that a better approximation can be achieved by sub-dividing
the interval in many parts, Fig. 11.25, and by assuming
X
I
f i 
11:98
i

where i are the mid-points of each sub-interval and  their common
length.
However, it may turn out that a better approximation can be
achieved with the same number of sampling points for the function
f  if they are chosen at special locations i . Thus, we can write
X
f i Wi
11:99
I
i

where Wi are appropriate weighting coecients, that may vary


according to the sampling point i . Gauss quadrature, on the basis
of the properties of Legendre5 polynomials, species the position of
the sampling points so that for any number n of them a very high
5

Legendre, Adrien-Marie (Paris, 1752 Auteuil, 1833), French mathematician.

340

Energy methods in structural mechanics


Table 11.1 Sampling points and relative weights for Gauss quadrature
n

i

Wi

0:00000
0:57735

2:00000
1:00000

0:77459
0:00000
0:86113
0:33998

0:55555
0:88888
0:34785
0:65214

2
3
4

accuracy is attained. In particular it allows us to integrate exactly a


polynomial of degree 2n 1 by means of n-point sampling. Sampling
points are placed symmetrically with respect to the centre of the interval and each pair of symmetric points is characterised by the same
weighting coecient Wi .
In Table 11.1 some examples of data for one-dimensional Gauss
quadrature are presented.
Extension to the case of n-dimensional integrals is very straightforward; for instance, in two-dimensional problems we can evaluate
the denite integral
1 1
I

f ;  d d

11:100

1 1

by integrating rst with respect to  and then with respect to , i.e.



1 1
1 X

I
f ;  d d 
f i ; Wi d
1 1

X X
j

f i ; j Wi Wj

XX
i

f i ; j Wi Wj

11:101

Of course, as numerical integration always introduces additional


approximations in the nite element analyses, the following question
arises naturally: what is the minimum number of sampling points
necessary for convergence of the problem towards the exact solution?
The answer is strictly linked to the types of element employed and
appropriate tables have been developed and made available in literature. Generally, we must bear in mind that the computational cost of
numerical integration is quite signicant, but poor treatment and
lack of attention to detail at this phase of the preparation of a nite
element program always leads to inaccurate and poorly reliability
for practical structural analyses.

The theory of nite elements

341

p = 300 kN/m2

15 m

300 m

Fig. 11.26

11.8

Numerical example: analysis of a compressed


foundation

The example concerns the analysis of a very long concrete foundation


subject to a vertical pressure p 300 kN=m2 on the top, as shown in
Fig. 11.26.
On account of its length, we can consider the foundation to be in a
state of plane strain and therefore we will limit ourselves to analysing a
cross section of unit depth in the longitudinal direction, Fig. 11.27.

l
y
x

Fig. 11.27

342

Energy methods in structural mechanics


p
1

y
4

3
l

Fig. 11.28

Moreover, on account of the symmetry about the vertical plane


which divides the cross section in two mirrored parts, we can further
simplify our analysis and take into account the scheme of Fig. 11.28,
where the points coincident with the symmetry axis 24 are allowed
to experience vertical displacements only.
Given the simplicity of the problem, we employ triangular elements
and derive the shape functions directly, without making resort to the
concept of parent elements and intrinsic coordinates. Thus, we
assume that the nite element model for the structure in Fig. 11.28 is
composed of two triangular elements, as shown in Fig. 11.29.
As we have seen in the previous sections, the rst step consists in
deriving the expression of the shape functions Nj x; y which represent
the displacement eld in the element as function of the displacements
of its nodes.

1
l
2
3

Fig. 11.29

The theory of nite elements

343

uy (H)

y
H

ux (H)

uy (J)
uy (I)

ux (J)

ux (I)

Fig. 11.30

With reference to the generic triangular element of Fig. 11.30, we


have
 


ux x; y
NI 0 NJ 0 NH
0

Di NDi
uy x; y
0 NI 0 NJ
0 NH
with

ux I

6 u I 7
7
6 y
7
6
7
6
u
J
x
7
Di 6
6 u J 7
7
6 y
7
6
4 ux H 5
uy H
In this example we use the simplest assumption for the deformed
shape of the element, that is, the components along x and y of the
displacement eld are described by rst-order polynomial expressions
of the type
ux x; y a1 a2 x a3 y
uy x; y b1 b2 x b3 y

344

Energy methods in structural mechanics

Notice that this very simple expression has the advantage that it
reduces the complexity of the algebra to obtain the stiness matrix
for the element but the assumed polynomial function also restricts
the variation of the strain in the plate to be piecewise constant. Thus
for the strain "x , in the x direction,
qux
a2
qx
and, similarly, for the other strain components. This simplicity of
representation, i.e. piecewise, of a continuous strain eld may have
implications for the number of elements that are required to obtain
an accurate solution using the nite element method.
By inserting in the above expressions the coordinates of the three
nodes of the element we have the following systems of linear equations
in the unknowns a1 ; a2 ; a3 and b1 ; b2 ; b3 , respectively
8
>
< ux I a1 a2 xI a3 yI
ux J a1 a2 xJ a3 yJ
>
:
ux H a1 a2 xH a3 yH
8
>
< uy I b1 b2 xI b3 yI
uy J b1 b2 xJ b3 yJ
>
:
uy H b1 b2 xH b3 yH
"xx

From linear algebra, the solution of these systems is readily found in


the form
2
3
ux I xI yI
6
7
a1 det4 ux J xJ yJ 5=2A
ux H xH

yH
3
1 ux I yI
7
6
a2 det4 1 ux J yJ 5=2A
1 ux H yH
3
2
1 xI ux I
7
6
a3 det4 1 xJ ux J 5=2A
1 xH ux H
2
3
uy I xI yI
6
7
b1 det4 uy J xJ yJ 5=2A
uy H xH yH
2

The theory of nite elements

1
61
b2 det4
1
2
1
61
b3 det4
1

345

3
yI
yJ 7
5=2A
uy H yH
3
xI uy I
xJ uy J 7
5=2A
xH uy H
uy I
uy J

where A is the area of the triangular element, and


3
2
1 xI yI
7
6
2A det4 1 xJ yJ 5
1 xH yH
On account of these expressions, with simple calculations we nd that
the shape functions Nj x; y take the form
NI x; y fxJyH xHyJ yJ yHx
xH xJyg=2A
NJ x; y fxHyI xIyH yH yIx
xI xHyg=2A
NH x; y fxIyJ xJyI yI yJx
xJ xIyg=2A
At this point we can evaluate the stiness matrix K i of the generic
element, as given by equation (11.90)

i

K BT CB dA
Si

where B and C are, see equations (11.86) and (11.62),


2 qN

6 qx
6
6
B 6
6 0
6
4 qN

qy
2
6
4

0
qNI
qy
qNI
qx

qNJ
qx
0
qNJ
qy

0
qNJ
qy
qNJ
qx

qNH
qx
0
qNH
qy

7
7
qNH 7
7
qy 7
7
qN 5
H

qx

yJ yH
0

0
xH xJ

yH yI
0

0
xI xH

yI yJ
0

3
0
7
xJ xI 5

xH xJ

yJ yH

xI xH

yH yI

xJ xI

yI yJ

346

Energy methods in structural mechanics

2
C

6
E
6
6
1 1 2 4

1 

1 

1 2
2

3
7
7
7
5

It is evident that for the assumed choice of displacement polynomials the deformation matrix B actually does not depend on the
coordinates x and y of the points inside the domain of the element.
Thus, the integration over the domain of the element is simple and
straightforward, and the required stiness matrix K i can be written as
1
E1 
4A 1 1 2
2 2
c4 c21 c4 c1 c4 c1
6
6
c21 c24
6
6
6
6
6
6
6
6
6
sym
4

K i

c4 c5 c1 c2

c4 c2 c1 c5

c6 c4 c1 c3

c1 c5 c4 c2

c1 c2 c4 c5

c1 c6 c4 c3

c25 c22

c5 c2 c5 c2

c5 c6 c2 c3

c22 c25

c2 c6 c5 c3
c26 c23

c4 c3 c1 c6

7
c1 c3 c4 c6 7
7
7
c3 c5 c2 c6 7
7
7
c2 c3 c5 c6 7
7
7
c6 c3 c6 c3 7
5
c23 c26

with
 


1

 

1 2
21 

and
c1  xH xJ

c2  xI xH

c3  xJ xI

c4  yJ yH

c5  yH yI

c6  yI yJ

For the plane analysis being exemplied here, the stiness matrix of the
element 1 (which is dened by the coordinates I 4, J 2, H 1),
K 1 , is readily obtained from the general expression by setting
c1 l

c2 0

c3 l

c4 0

c5 l

c6 l

while the stiness matrix of the element 2 (which is dened by the


coordinates I 4, J 3, H 2), K 2 , is obtained by setting
c1 0

c2 l

c3 l

c4 l

c5 l

c6 0

The theory of nite elements

347

Before we impose any constraints, the nodal displacement vector D


for the model in Fig. 11. 29 is represented by the following eightelement array
DT ux 1

uy 1

...

. . . ux 4 uy 4

so that the corresponding combined stiness matrix for the structure,


 is an 8  8 array. This matrix can be easily obtained by adding the
K,
contributions of the stiness matrices of the elements 1 and 2, i.e. K 1
and K 2 in the appropriate positions, as formally indicated by equation (11.73). However, the presence of the constraints at the nodes 2,
3 and 4 allows only the horizontal and the vertical displacement of
the node 1 and the vertical displacement of the node 2. Therefore,
the nodal displacement vector D reduces to
3
2
ux 1
7
6
DI 4 uy 1 5
uy 2
and the joint stiness matrix of the structure, KI , becomes a 3  3
array. Therefore the joint stiness matrix is seen to be
2
3
1
1
1
K55
K56
K54
6
7
1
1
7
KI 6
K66
K64
4
5
1
2
sym
K44 K66
1
E1 
4A 1 1 2
2
3
 l 2 l 2
l 2
l 2 l 2 
6
7
l 2 l 2
l 2 7
6
4
5
2
2
sym
l l
3
2
1  
 

E1 
7
6

1 
 5
4
21 1 2
sym
1 

given that A l 2 =2.


The analytical method described above is applied here to a concrete
foundation which is composed of material with the following properties: Young's modulus and Poisson's ratio: E 3:11  107 kN=m2 and
 0:15, respectively. The weight per unit volume is 22 kN=m3 .

348

Energy methods in structural mechanics

Thus, we have
 0:176

 0:412
2 :
3
1 412 0:588
0:412
1:412 0:412 5
KI 1:642  107 4
sym
1:412

The vector of the equivalent nodal loads is given by equations (11.78)


and (11.81), i.e.
8
9

<
=
i
T
T
N b dA N f ds
P
: i
;
S
i
X i
P
P
i

with


b


f

22


on the side 12

300

For element 1 we have


N4 l 2 ly=l 2 l y=l
N2 lx=l 2 x=l
N1 lx ly=l 2 y x=l
and for element 2
N4 l 2 lx=l 2 l x=l
N3 lx ly=l 2 x y=l
N2 ly=l 2 y=l
so that the vector of the nodal forces corresponding to the active
degrees of freedom is seen to be
2
3 2
3
0
0
6
7 6
7
l
l
l
6
7 6
7
6
7
6
yx
lx 7
6
7 6 300
22 dx
dy
dx 7
6
7 6
7
l
l
76
7
PI 6
x
0
0
6
7
6
7
0l
1
6
7
6
7
l
l
l
x
6
7 6
7
6
7
6
7
x
y
x
4 22@ dx
dy dx
dyA 5 4 300
dx 5
l
l
l
0

The theory of nite elements

349

Notice that for the sake of clarity we kept the contribution of the body
forces and that of the applied surface load p separate. With a fairly
simple integration we obtain
3
3 2
2
0
0
6
6
l2 7
l 7
7
7 6
6
6 22 7 6 300 7
PI 6
6 76
27
7
7 6
6
5
5 4
4
l2
l
300
22
2
3
3
3 2
2
3 2
0
0
0
7
7 6
6
7 6
: 7
: 7 6
7 6
6
4 8 25 5 4 225 5 4 233 25 5
241:50
225
16:5
It is straightforward to verify that the sum of the nodal forces
corresponding to the body forces due to the self-weight of the concrete
are equal to half of the weight of the whole section, i.e.
 A 22  2:25 49:5 kN (remember that we are dealing with a
plane strain problem and therefore we are examining a slice of unit
depth). Only half the weight is included in the analysis of the deformation because the remaining part of the total weight would be applied to
the constrained nodes 3 and 4 with no eect on the deformation of the
foundation. Conversely, the resultant of the surface load p, i.e.
p  l 300  1:5 450 kN/m, is applied to the nodes 1 and 2 in the
form of two forces of equal value.
Finally, the equilibrium equations of the problem is
KI DI PI )

32
3
ux 1
1:412 0:588
0:412
76
7
6
1:412 0:412 54 uy 1 5
1:642  107 4
uy 2
sym
1:412
2
3
0
6
:25 7
233
4
5
2

241:50

the solution of which is


2
3
2:14418  106 m
6
7
DI 4 0:150952  104 m 5
0:141951  104 m

350

Energy methods in structural mechanics

It is worth noting that if we neglect the eect of the self-weight, we can


write
3
32
3 2
2
ux 1
0
1:412 0:588
0:412
7
76
7 6
6
1:642  107 4
1:412 0:412 54 uy 1 5 4 225 5
sym
and get

1:412

uy 2

225

3
2:148710  106 m
6
7
DI 4 0:144809  104 m 5
0:133029  104 m

The strain for each element can be evaluated by means of equations


(11.64), (11.7) and (11.86), that is
3
2
"xx
6 " 7
i
i
4 yy 5 SND BD
2"xy
Thus, we have
2
3 2
3
"xx
0:145038  104
6 " 7 6 :
7
4 yy 5 4 0 977461  104 5
2"xy
0:655216  105
3
2
3 2
"xx
0
6 " 7 6 :
4 7
4 yy 5 4 0 897944  10 5
2"xy
0
for elements 1 and 2, respectively.
In order to validate the quality of the solution obtained, we can
refer to the exact value of the strain component "y which is
0:964630  104 . It is found that even with the extremely coarse
mesh adopted, the nite element solution diers from the exact one
by no more than 7:5%.
If a more exact solution were required for the nite element modelling there are two possible ways forward. One is simply to rene the
mesh geometry dening the elements, that is to include more elements
in the analysis. This is a simple and straight forward operation once the
element stiness matrix has been formulated in a general manner.
Indeed this approach is the one that most engineers would follow
when they use commercial programs for nite elements. They would
carry out a sensitivity analysis to determine, by progressively rening

The theory of nite elements

351

the element mesh geometry, the geometry that would give an adequately
accurate answer and ensure that the answer is not sensitive to the change
of mesh geometry.
An alternative approach would be to assume a higher order
polynomial to describe the variation of the deformations ux x; y,
uy x; y, in place of the simple linear variation assumed in this example.
This alternative approach requires the re-formulation of the stiness
matrix for a generic element. The potential advantage for this
approach is that possibly the number of elements required to converge
to an adequately accurate solution may be reduced.

12. Stability of equilibrium and


non-linear deformations of
beamcolumns
Introduction

The great majority of this text is devoted to the presentation of the


linear response of systems of bars and plates to applied loading.
That is, the structural deformations can be modelled by a linear relationship between the loading and the strains induced by the loads. A
basic assumption is that the deformations of the structural elements
are suciently small that the points of application and directions of
the loads can be considered as unchanged during the deformations
and always acting on the undeformed structure.
Despite the increased complexity of incorporating non-linearity
into the analysis even for linearly elastic materials, we feel that some
introduction to these types of phenomena and solution techniques is
a valuable complement to the linear analysis methods presented in
earlier chapters. This chapter therefore is concerned with a development of the analysis of some of the simplest structural elements that
exhibit geometric non-linearity, mainly the beam-column. The deformations of this type of element have a signicant eect on the geometry
of the applied loading. In this presentation we shall restrict the theoretical development to linear elastic materials.

12.1

The concept of stability. Initial buckling of


columns

Most of us are familiar with the concept of a mechanical system


which is capable of exhibiting stable and unstable congurations of
equilibrium. Perhaps the most simple of these systems is that already
introduced in chapter 6, i.e. the ball subject to a gravitational eld in
the landscape of hills and valleys of Fig. 6.1.

354

Energy methods in structural mechanics

P
P

Fig 12.1

In section 6.1 we stated that the total potential energy of the system
coincides with the load potential Phs, which represents the capability
of the gravitational force P to perform work on account of the height
hs of the ball. If no work tends to be performed by the gravitational
force, the principle of conservation of energy assures that no kinetic
energy can be gained by the system and it is in static equilibrium.
However, it is intuitively clear that the state of equilibrium attained
at the top of a hill is qualitatively very dierent from the equilibrium
attained at the bottom of a valley. In fact, let us consider the static
forces, acting as a result of a very small displacement about an equilibrium state, as a means of inferring the likely dynamic response. If the
static forces acting upon the system after such a small displacement
about an equilibrium state (i.e. the gravitational force and the frictionless reaction of the support) are in a direction which tends to restore
the system to its original equilibrium state we consider that the original
position is a stable equilibrium state, but if the static forces are such that
they tend to upset the system then the original position is considered to
be an unstable equilibrium state.
In the case under consideration, if the ball is at the bottom of a valley
then the gravitational force P will always tend to restore the ball to its
original position of equilibrium after any small displacement about
this point, as shown in Fig. 12.1. On the contrary, if the ball is at the
top of a hill then the gravitational force P will always tend to move
the ball away from its original position of equilibrium for any small
displacement about this point, see Fig. 12.2.

P
P

Fig 12.2

Stability of equilibrium and non-linear deformations

355

Fig 12.3

Of course, from a theoretical point of view we may also encounter


cases of neutral equilibrium. With reference to the present example
this may happen when the ball is on a perfectly horizontal plane surface and therefore it may be placed in any displaced conguration
and still be in a state of equilibrium, Fig. 12.3. From a mathematical
standpoint a situation of neutral equilibrium can also be imagined as
a limiting case dividing stable from unstable equilibrium, but from
a practical point of view in the analysis of structures this is not a
signicant case.
The simplest problem of elastic instability in the theory of structures
is that concerning lateral buckling of compressed members. With reference to Fig. 12.4, which shows a slender bar subject to a compressive
axial load N, it is easy to nd experimentally that for a certain range
of values of the axial force, say from 0 to Nc , the axial shortening of
the bar is directly proportional to the applied load.
This happens according to the deformation relationship, equation
(4.5), stated in section 4.2 for an axially loaded bar in the linear
theory of elasticity. However, once the axial load has reached the
N
l

Nc

Fig 12.4

356

Energy methods in structural mechanics


N

u (l/2)

Fig 12.5

critical value, Nc , the bar deects laterally and the value of the deection is found not to be directly proportional to the value of applied
load, even if the material is perfectly linearly elastic. At the onset of
the lateral deection we say that the bar experiences initial buckling.
Of course, as long as we remain in the realm of the linear theory of
elasticity and under the hypotheses formulated in chapter 4, such a
behaviour cannot be mathematically described, given that the deformation relationship, equation (4.5), holds for any value of the axial
load. Still, the reason for the discrepancy between the predictions of
the model introduced in chapter 4 and the observations in tests on slender columns is quite simple. Essentially, if we perturb the equilibrium
of a slender axially loaded bar, for example by slightly displacing its
mid-span, the axial load will cause a bending moment given by the
product of the force and the distance between its direction and the
deformed axis of the beam, Fig. 12.5.
As long as the elastic stiness of the beam is capable of restoring
the original conguration of equilibrium, the equilibrium is stable
and the only equilibrium conguration is the straight one. However,
when the magnitude of the axial force is such that the bending
moment yielded by the perturbation exceeds the restoring capacity
of the elastic stiness of the beam, other displaced congurations of
equilibrium become possible and the column will bend laterally. In
this case the straight conguration of equilibrium is classied as
unstable, because, as in the case of the ball at the crest of the hill of

Stability of equilibrium and non-linear deformations

357

Fig. 12.2, the column always departs from the crest as a result of any
small displacement of its position. From a practical point of view it
must also be said that it is experimentally very dicult to keep the
column in a straight conguration of equilibrium once the critical
value of the axial load Nc has been attained because this implies the
absence of any imperfection in the specimen or external perturbation
of the column.
It is now clear that the modelling of axially loaded bar illustrated in
chapter 4 ignores the additional moment induced by perturbations or
imperfections of the axis of the column and is therefore incapable of
describing the buckling phenomena. Intuitively it is anticipated that
initial buckling aects only slender bars, being dependent on the
length and on the bending stiness of the element. In fact, for a
squat structural member it is found that yielding and plastic deformation of the material occurs before any lateral initial buckling takes
place. However, there is a large range of practical compressed elements
that can be analysed using the theory of linear elasticity.
The mathematical problem of the lateral instability of compressed
members was solved in the 18th century by the Swiss mathematician
Leonhard Euler, who also formulated the equation of the equilibrium
for laterally loaded beams presented in chapter 4. In the realm of large
displacements this theory leads to the formulation of what is generally
known as the Euler's equation of the elastica. We will present the Euler
theory of compressed bars, i.e. struts, in section 12.4, limiting our
treatment to relatively small deections. First, however, in the next
section, we present an energy criterion for stability and give some
elementary examples.

12.2

Energy criterion of stability

According to the simple example of the ball discussed previously, the


understanding of the dynamic response of mechanical systems to disturbances is the basis of stability studies. However, from our point of
view, that is the analysis of structures, this route is probably not the
most convenient one to follow and we shall now introduce a very
useful and intuitive alternative approach based on the potential
energy functional.
We will assume that if the potential energy is a local minimum the
structure is stable with respect to small disturbances, and if the potential energy is a relative maximum the structure is correspondingly
unstable.
In order to attempt a rigorous proof of the link between dynamical
response to disturbances and the potential energy, a great deal of work,
which is outside the scope of this introductory textbook, would be

358

Energy methods in structural mechanics

required. It must also be noted that in recent times some scientic


objections, related to very special physical systems, have been directed
towards this classical energy criterion. However, these special systems
do not impinge on the structural models considered here. Therefore,
given its intuitive basis, we will assume the criterion above to be a statement which is conrmed by correspondence between experimental
results and analytical predictions made using this assumption.
Alternatively, the following simple line of reasoning can be considered. In chapter 6 we stated the principle of stationary value of the
total potential energy, that is the relationship  0 holds in the equilibrium state of conservative systems (which in statics of structures is
always the specic state for which the stability is investigated). Let
us assume that the potential functional  can be made dependent on
a system parameter that species the position, i.e.  q. As a
consequence, q0 0 holds for the value q0 which corresponds to
equilibrium. For a small variation of q, we have


q
1 q2 
q0 q q0 q
q2   
qq q0
2 qq2 q0
q0 q0 12 2 q0   
q0 12 2 q0   

12:1

Thus, in the vicinity of the equilibrium position dened by q0 the


potential energy  is a relative maximum if 2  < 0. This means
that potential energy is released and kinetic energy gained when q is
varied from q0 and the system will tend to move away from the
equilibrium position that is therefore unstable. On the contrary, if
2  > 0, potential energy is gained when q is varied from q0 and the
system cannot therefore leave the position of equilibrium by itself,
that is the equilibrium is stable. The stability boundary is given by
2  0, which corresponds to neutral equilibrium. As an historical
remark, it should be noticed that these results make the energy criterion formally identical to Dirichlet's theorem for a point system.

12.3

Some illustrative examples of stability conditions.


Branching points and limit points.

To illustrate the application of the energy criterion, let us start by


considering the elementary model in Fig. 12.6. This is a rigid cantilever
elastically hinged at the bottom end and subject to a vertical force P at
the upper end. The rotational spring is an idealised device that is
capable of resisting rotations by a reactive moment. A familiar
example might be a common door spring that restores the door to

Stability of equilibrium and non-linear deformations

359

l cos

M = k

Fig 12.6

its closed state. The relationship between rotation and the reactive
moment M is
M k

12:2

 being the rotation about the hinge axis and k the rotational stiness
of the spring.
By virtue of Clapeyron's theorem, the strain energy stored in the
spring upon rotation is
U 12 k2

12:3

In order to compute the work done by the vertical load P, we must


remove the hypothesis of small displacements. The total potential
energy of the system becomes, consequently
 12 k2 P 12 k2 Pl1 cos 

12:4

and the equation of equilibrium is


 0

q
k Pl sin  0
q

12:5

This equilibrium equation is always satised by the trivial solution


 0, which corresponds to the perfectly vertical position of the
bar. If we wish to examine the stability of such a conguration, we

360

Energy methods in structural mechanics


P

Pc

Fig 12.7

must examine the term



1 2
1 q2  2
 0

2
2 q2 0
Since 2 is always positive, the above term reduces to

q2 
k Pl cos0
q2 0

12:6

12:7

This means that for P < k=l the conguration of equilibrium  0 is


stable, while for P > k=l the equilibrium  0 is unstable. The value
of the applied load P k=l constitutes the boundary between stable
and unstable equilibrium and that load is called the critical load, Pc .
Apart from the trivial solution, equation (12.5) is also satised by
the values of P and  satisfying the relationship


k

P
12:8
l sin 
In the space P  this equilibrium path has the form shown in Fig. 2.7
and intersects the trivial or fundamental path,  0 coincident with the
load axis at the point Pc k=l.
Actually, this point, apart from being the boundary between stable
and unstable congurations on the fundamental path, constitutes a
branching point in the equilibrium path. Thus the critical load Pc
makes a bifurcation in the deformation state.
The rising post-buckling path, i.e. increasing deformation with
increasing load, given by the equation (12.8) is easily found to be

Stability of equilibrium and non-linear deformations

everywhere stable, given that



q2 
k Pl cos  > 0
q2 

for

 6 0

and P >

k
l

361

12:9

In other words, if we perturb the equilibrium of the cantilever in its


fundamental position of equilibrium, it will tend to return to it as
long as P < k=l. On the contrary, if P > k=l any small disturbance
will cause the cantilever to move away from the fundamental path
 0 and it will tend to become on the rising equilibrium branch,
which is always stable.
Of course a perfectly straight cantilever loaded by a perfectly axial
force is a pure idealisation and in reality some imperfections will
always aect this simple model. In order to examine such a situation,
let us consider the cantilever in Fig. 12.8, where the vertical load P is
slightly oset from the vertical axis of the bar.
In such a case the total potential energy of the system becomes
 12 k2 P
12 k2 Pl1 cos  Pe sin 

12:10

P
e sin

l cos

M = k

Fig 12.8

362

Energy methods in structural mechanics

on account of the presence of the small eccentricity e. Thus, the


condition of equilibrium is
 0

q
k Pl sin  Pe cos  0
q

12:11

By using the ratio


e
"
l

12:12

as a measure of the loading imperfection, the equation of equilibrium


(12.11) becomes


k

P
12:13
l sin  " cos 
Plotting the equilibrium paths yielded by equation (12.13), as was done
in Fig. 12.9, we notice that there is no longer a bifurcation condition.
This is due to the fact that, on account of the loading eccentricity,
the trivial equilibrium path has disappeared. Nevertheless, the value
of the critical load Pc k=l for the corresponding perfect system
retains all its importance because it represents the magnitude of the
load at which the system starts to present three possible and dierent
congurations of equilibrium. Moreover, if we proceed to examine the
stability of the equilibrium paths by studying the sign of the parameter
q2 
k Pl cos  Pe sin 
q2

12:14

Fig 12.9

Stability of equilibrium and non-linear deformations

363

P
l
cos

Fig 12.10

we discover that the equilibrium paths 1 and 3 are stable, while the
equilibrium path 2 is unstable.
To summarise the results, we may say that the imperfect system of
Fig. 12.8, loaded from its unloaded state, will follow a constantly
rising stable path so that no instability is encountered. The deections
will simply grow more rapidly as the critical load of the perfect system
is passed. In addition to this natural equilibrium path, the system will
also exhibit a complementary equilibrium path in the other quadrant
of the P  space, which is partly stable and partly unstable and
cannot be normally encountered in a natural loading process from
P 0.
To complete our presentation of simple examples, we include a
dierent type of instability, which may occur also in perfect systems
without the presence of bifurcation of the equilibrium path. To this
purpose, we can consider the shallow arch of Fig. 12.10.
The arch comprises two axially deformable bars pinned to each
other and to rigid foundations. It is evident from the simple
equilibrium of rigid bodies that for any position, dened by the
rotation angle , the bars are always subject to axial loading only.
The deformation law (4.5) allows us to write
l

N
l
N

EA cos k

12:15

where k is the axial stiness of the bar. By dening the rotation angle 
as the parameter describing the conguration of the system, we can
write the following expression for the strain energy of the system
 
2 
1
l
l
U2 k

12:16
2 cos cos 

364

Energy methods in structural mechanics

The corresponding work of the applied load P is


W Pl tan l tan 

12:17

and the total potential energy of the system is therefore


2 
 
1
l
l

 2 k
Pl tan l tan 
2 cos cos 
12:18
The present example is simplied if we restrict it to a shallow arch,
that is we can take both and  to be small and write
cos  1
tan  ;

2
;
2

cos   1

2
;
2

12:19

tan   

so that equation (12.18) becomes


 14 kl 2 4 2 2 4 3 4 Pl

12:20

Once again, the condition of equilibrium is


 0

q
kl 2 2 2  3 2 3 Pl 0
q

12:21

and yields
P kl2 2 3  2

12:22

By plotting this relationship in the P  space, we can visualise the


equilibrium path, see Fig. 12.11.
P

Fig 12.11

Stability of equilibrium and non-linear deformations

365

First of all we notice that there are three equilibrium positions for
P 0, i.e.  0,  and  2 . In order to investigate the
stability of the equilibrium path, we again make reference to the sign
of the second derivative
q2 
kl 2 2 2 6  32
q2

12:23

In this manner we nd that the two unloaded equilibrium congurations at  0 and  2 are both stable, while the unloaded
equilibrium conguration at  is unstable. Moreover, the path
0A is stable, as well as the rising path CE, while the path ABC
is unstable. In other words, starting from the unloaded origin of the
co-ordinates, the arch equilibrium will be described by the stable
rising path 0A. At this point the equilibrium would require a diminishing load along the path AB until zero magnitude is attained at B.
From B to C the magnitude of the load would be required to increase
again but with opposite direction from the initial one, until point C is
reached. In reality, on attaining the point A under the prescribed value
of the load P, the arch will move dynamically to the state D, this is
termed snap buckling. In a practical case the energy released in the
rapid transition from the geometry at A to that at D will eventually
be dissipated by vibrations due to the presence of some damping and
the system will come to rest in this position. Subsequent increments
of the load will cause the inverted arch equilibrium states to follow
the stable rising path DE.
In conclusion, in this example there are no branching points and the
quality of the equilibrium changes from stable to unstable at some
points, namely A and C, which we will call limit points. In other
words, this is a form of progressive instability in the sense that
the structure tends to lose its stiness gradually as an eect of the
change of conguration due to increasing values of loading.
At the end of the presentation of these simple examples, it seems
clear that when we have to investigate the stability of equilibrium we
must start by removing one of the fundamental assumptions of the
linear theory of elasticity, that is on account of the smallness of the
displacement eld the forces acting on the body can be always referred
to the undeformed conguration. In the next section we will see what
this implies for Euler's equation of beam deformation.

12.4

Euler's equation of beam deformation in the


presence of axial loads

We have already encountered Euler's equation of beam deformation


when dealing with laterally loaded beams in chapter 4. Euler was

366

Energy methods in structural mechanics


P
B

z
A

Fig 12.12

also the rst mathematician to provide a complete and correct solution


to the problem of an axially loaded elastic bar. Thus, at this point of
the presentation we wish to illustrate how it is possible to write the
total potential energy functional for a beam subject to axial compressive loading and establish, exactly in the same manner as we did in
chapter 4 for the laterally loaded beams, the dierential equation of
the problem.
We start by considering the simply supported beamcolumn of
Fig. 12.12. However, the procedure is absolutely general and can
account, as we will see, for any kind of boundary conditions.
For the sake of simplicity, we consider the beamcolumn to have
uniform bending stiness EIx . If we neglect the axial extensibility,
the only component of the strain energy is that due to bending, is
once more given by equation (4.33)
l
1
U
EIx u002
y dz
2
0

Since the load P is axially applied, its potential is related to the vertical
displacement of the sliding support of the structure . Given that we
have neglected any axial extensibility, this displacement is dependent
only on the lateral deection of the beamcolumn. However, in

Stability of equilibrium and non-linear deformations

367

chapter 4 and in the realm of the theory of very small displacements,


we assumed that
u02
y 1
and therefore this term was discarded in order to obtain the simplied
expression of the curvature, equation (4.25), from the exact relationship, equation (4.23). This also implied that the axis of the deected
cantilever of Fig. 4.8 had, according to equations (4.27) and (4.28),
the same length of its projection on the z-axis so that the horizontal
displacement of its right end could be considered zero under bending.
We now need to remove this hypothesis in order to evaluate the displacement of the sliding support. However, it must be stressed that
we will not remove the assumption u02
y  1 with reference to the simplication of the curvature relationship equation (4.23) into equation
(4.25). In other words, we assume to have small strains but, with
respect to the applied loading, not negligibly small displacements.
This fact may appear incongruous from a mathematical point of
view, but it must be borne in mind that we are always working with
idealised models of real systems and that in order to describe a
phenomenon a model must be simplied enough to allow a solvable
mathematical treatment. Formulating a problem is, in fact, asking
questions whereas the solution is answering those questions. As long
as we obtain a solution which gives acceptable correspondence with
experimental observations, we may be satised that we have grasped
the essence of the phenomenon under investigation. Of course, every
simplication introduces a degree of approximation and we are
going to see how the present treatment will allow only the evaluation
of the initial buckling load, inasmuch as it is not able to provide any
information about the post-buckling equilibrium states.
Thus, let us consider a generic element of the beamcolumn in its
displaced conguration, as shown in Fig. 12.13. From the theorem
of Pythagoras1 we see that
dz duz 2 dz2 du2y

12:24

so that we have
q
dz duz dz2 du2y

1
s
 2
duy A
dz
duz @1 1
dz
12:25

Pythagoras (Samos, 571479

BC),

Greek philosopher and mathematician.

368

Energy methods in structural mechanics


duy

duz

dz
dz

Fig 12.13

Also, from calculus we know that, by means of a truncated series


expansion, we can write
p
x
1x1
12:26
2
thus equation (12.25) becomes
duz 12 u02
y dz

12:27

As a result, the vertical displacement of the sliding roller of the beam


column of Fig. 12.12 can be computed as
l
1 02
 duz
u dz
2 y
l
0

12:28

and the total potential energy functional of the system can be written
as
l
l
1
1 02
002
EIx uy dz P
u dz
uy U P
2
2 y
0

12:29

Stability of equilibrium and non-linear deformations

369

Once again, the stationary value of the functional uy gives the equilibrium states of the system. Thus
l


EIx u00y u00y

dz P u0y u0y dz 0

12:30

By repeatedly integrating by parts, and taking into account the same


boundary conditions exemplied in section 4.4 (and in a more general
way, in section 4.5), we get
l


EIx u0000
y uy

dz P u00y uy dz 0

12:31

which, since the virtual displacement eld uy is arbitrary, is equivalent


to the fourth-order dierential equation
00
EIx u0000
y Puy 0

12:32

This equation is called Euler's equation of beam buckling, and its solution represents the deformed conguration of the elastic beamcolumn
under the specied axial load P.
By dividing both terms by the bending stiness EIx and setting
s
P

12:33
EIx
we can write
2 00
u0000
y  uy 0

12:34

The general integral of such equation is


uy z A sin z B cos z Cz D

12:35

where A, B, C and D are constants which need to be determined by


means of the geometrical and mechanical conditions at the supports
A and B. These constraints require that
uz 0

for

z0

Mz EIx u00y 0

and z l
for

z0

12:36
and z l

12:37

370

Energy methods in structural mechanics

Therefore we have an homogeneous linear system of four equations in


the four unknown constants A, B, C and D. The system is
8
uy 0 B D 0
>
>
>
>
>
< EIx u00y 0 EIx 2 B 0
12:38
>
uy l A sin l B cos l Cl D 0
>
>
>
>
:
EIx u00y l EIx A2 sin l EIx B2 cos l 0
or, in matrix form
2
0
0
6
0
2
6
6
4 sin l
cos l
2
2
 sin l  cos l

1
0
l
0

2 3
0
6
7
6
7
0 76 B 7 6 0 7
7
76 7 6 7
1 54 C 5 4 0 5
1

32

12:39

A homogeneous linear system of equations always admits the trivial


solution A B C D 0, which corresponds to the fundamental
equilibrium of the beamcolumn, i.e. its straight conguration.
From linear algebra we also know that other solutions are possible
if, and only if,
3
2
0
0
1 1
6
0
2
0 07
7
6
det6
12:40
70
4 sin l
cos l
l 15
2 sin l

2 cos l

Therefore we are led to a transcendental equation (it involves trigonometric functions) in the parameter . The solution of equation (12.40)
is found to be
PE n2

2 EIx
l2

12:41

where n is a generic integer. For n 1 we get the minimum magnitude


of the axial load for which a deected conguration of equilibrium
is possible. This is usually called the Euler buckling load, PE , or the
critical load.
It must be noted that in correspondence to this buckling load the
matrix equation (12.39) will admit a solution which is dierent from
the trivial one, but the solution is modied by an arbitrary constant.
Thus the buckled conguration represented by the equation (12.35)

Stability of equilibrium and non-linear deformations

371

can be obtained in shape, but not in magnitude. In Fig. 12.14 the


buckled shapes for n 1, n 2 and n 3 are represented.
As we anticipated before, the fact that we are left only with the
critical buckling load and not with a relationship between the load
and the corresponding deection is to be ascribed to neglecting the
eect of higher-order terms in the deectioncurvature relationship,
equation (4.23). A fully non-linear equilibrium equation enabling the
study of post-buckling equilibrium paths can be obtained quite
easily by formulating the exact expression of the curvature, but it is
not integrable in an elementary manner and is therefore outside the
scope of this introductory book.2
It is however interesting to notice that a similar behaviour to the
Euler initial buckling formulation might be found also in the previously
examined example of the elastically hinged cantilever of Fig. 12.6.
In fact, if we expand the total potential energy expression (equation
(12.4)) with reference to the term cos  and retain only up to the second
2
Nevertheless, the analysis of the post-buckling behaviour of the simply supported
beam of Fig. 12.12 can be tackled in an approximate way by means of the trial function
method. In fact, with a line of reasoning similar to equation (4.23), we can write the
curvature of the deected column as

u00y
1 d

q
r ds
1 u02
y

where, however, this time the derivatives are taken with respect to the curvilinear coordinate of the undeformed axis of the beam, s. Moreover, we can write
 2
u002
1
y
02

 u002
y 1 uy
r
1 u02
y
Thus, the total potential energy is


l
l
1
P 02
02
EIx u002
1

ds

u ds
y
y
2
2 y
0

By assuming a trial function of the type


 
ns
uy A sin
l
and performing the substitution in the total potential energy expression, setting
d=dA 0 gives
P

n2 2 8EIx l 2 n 4A2 EIx n3 3 4EIx l 2 sin2n A2 EIx n2 2 sin4n


4l 4 2n sin2n

which constitutes the loaddisplacement relationship in post-buckling.

372

Energy methods in structural mechanics

a2

a3

a1

PE =

EI
l2

PE = 4 2 EI
2

PE = 9 2 EI
2

Fig 12.14

order terms, we have


1
1
2
 k2 Pl1 cos  k2 Pl
2
2
2

12:42

and the equation of equilibrium results


 0

q
k Pl 0
q

12:43

Equation (12.43) gives the value of the critical load Pc k=l as the
value of the load at which a non-trivial conguration of equilibrium
becomes possible, but it fails to give information about the bifurcated
equilibrium paths.
However, from a practical point of view engineers are mostly
interested in designing structures with a degree of safety such that
the critical load is not attained and therefore the knowledge only of
the magnitude of such a load is generally sucient for design analysis
purposes together with an understanding of the eects of imperfections
in the structure loading or geometry.

12.5

Analytical treatment of some examples of


columns

The analysis in this section will apply quite systematically the method
introduced above for obtaining the load at which the onset of lateral
deformations are initiated. In the previous section we have seen that
the load at which the perfect column becomes unstable and develops
lateral deformations is called the Euler, or critical load. The form of

Stability of equilibrium and non-linear deformations

373

buckling considered in that section is thus called Euler buckling or


sometimes bar buckling to distinguish the phenomenon from the
other forms of buckling exhibited by plates and shells.
It has also been noted that a perfect column is an idealisation of
reality and implies that
.
.
.

the column is perfectly straight before the application of the


perfectly axial loading
the material is perfectly linearly elastic, homogeneous and
isotropic
the column cross section is perfectly uniform along its length.

Therefore, in the following we will take into account other very


common examples of beamcolumns, these are the rotational
restrained column, the column with a lateral load and the column
with initial out-of-straightness. The analytical treatment in this section
reduces the degree of idealisation and the eect of lateral loads on the
column behaviour eliminates the condition of the unique critical load
behaviour. Nevertheless, the later analysis emphasises the value of the
Euler load as an important parameter characterising the behaviour of a
column subjected to increasing load levels

Column with various degrees of rotational restraint

A column very seldom exists as a separate entity, it is usually a component in a frame and as such it has a exural restraint at its ends from
the contiguous members of the frame. The eect of this restraint, which
is a function of the relative exural stiness of the column and the
members attached to it are considered here.
Figure 12.15(a) shows the deformed shape of an initially perfect
column subjected to a load P applied along the initial axis. The
moments m at the end of the column are applied restraints from
other members attached to the column and in the present example
the restraints are equal at both ends. Note that these restraining
moments have zero value until the column begins to deect and that
they will increase in magnitude as the magnitude of the column lateral
deection increases.
We refer to Euler's equation of initial buckling (12.32)
00
EIx u0000
y Puy 0

and again, by setting


s
P

EIx

374

Energy methods in structural mechanics


P

P
m
Q
z

uy (z)

Q
m

P
(a)

(b)

Fig 12.15

we obtain
2 00
u0000
y  uy 0

for which the general solution is given by equation (12.35)


uy z A sin z B cos z Cz D
where A, B, C and D are constants determined from the boundary
conditions as follows
uz 0
M

EIx u00y

for
m

z0

and z l

for z 0 and

12:44
zl

Thus, we can write the following system


8
uy 0 B D 0
>
>
>
>
< EI u00 0 EI B2 m
x y
x
>
u
l

A
sin
l

B cos l Cl D 0
y
>
>
>
:
EIx u00y l EIx A2 sin l EIx B2 cos l m

12:45

12:46

Stability of equilibrium and non-linear deformations

375

which, contrary to the linear system equation (12.38) found in the


previous section, is not homogeneous in the present case.
Its solution yields


m 1 cos l
m
l
A

tan
EIx
sin l
EIx
2
12:47
m
m
B
C

0
D

EIx 2
EIx 2
so that
uy z

m
l
m
m
tan sin z
cos z
2
EIx
2
EIx 
EIx 2

Now, let us dene the coecient of restraint as


m



12:48

12:49

then, since we have


u0y 0 0

m
l
tan
EIx
2

12:50

we get
tan

l
2EIx

2
l

12:51

l
2 l

2
2

12:52

or also
tan

with  l=EIx .
The degree of exural restraint applied to the column from a contiguous member provides the specic value of in equation (12.49) and
hence in equation (12.52) which can then be solved to provide the
appropriate value of the Euler load.
The Euler load obtained from the solution of equation (12.52) can be
written as
2 EI
12:53
l2
where is the particular coecient related to the value of . An alternative approach and convenient to describing the Euler load is to
dene an eective length le such that
PE

PE

2 EIx
le2

12:54

376

Energy methods in structural mechanics


10

P
09
le

le /l

08
P

(b )

P
m

07

le
06

m
P

(a )

05
0

10

15
= l/EI

20

25

30

Fig 12.16

where

l
le  p

12:55

Curve (a) in Fig. 12.16 shows the variation of the eective length
with increasing values of . Curve (b) show the corresponding variation for columns in which one end is restrained and the other is
simply supported. The free body diagram for this case is shown in
Fig. 12.15(b) in which the asymmetry of the constraint moments
requires that a lateral load Q be applied at the ends of the column.
This load, as in the case of the restraint moments m is zero until
the column experiences lateral deections at the Euler load.
The physical interpretation of the eective length le is made more
clear in Fig. 12.17 where the deected shape of the columns at the
respective Euler loads are shown.
The analysis shows that for a column simply supported at each end,
the eective length is le l, and for a column encastre at both ends,
le l=2. The corresponding condition for a column encastre at one
end and simply supported at the other has le 0:71l. These are
shown graphically in Fig. 12.17(a) to (c) respectively. The eective
length is in fact the distance between the points of inection, i.e. zero
moment, in the deected shape of the columns. The case shown in

Stability of equilibrium and non-linear deformations


P

377

l
l e = 071l
l e = 05l

le = l

(a)

l e = 2l

(b)

(c)

(d)

Fig 12.17

Fig. 12.17(d) is for a column encastre at one end and totally free from
constraint at the other. The analysis shows that le 2l, that is the
Euler load for this column is 25% of that for a column simply
supported at both ends. This form of buckling mode is called sway
buckling and is of particular importance to the design of frames since
the buckling load is so low.

Column with a lateral load

In practice a column very seldom is in a condition where it experiences


only a load aligned with the axis. More often, the loading is a combination of axial and lateral loading, as shown in Fig. 12.18.
The following analysis considers the deections imposed on a column
due to the loading. This is for simplicity of presentation, but the method
of analysis and, more importantly, the results can easily be applied to the
other forms of boundary constraint considered up to now.
The lateral load Q is applied at mid-length of the beam, whose
behaviour is therefore symmetrical with respect to this cross section.
Given the presence of the lateral load Q, the total potential energy
of the system is given by the equation (12.29) plus the addition of
the potential relative to this lateral load, that is
uy U P1 Q2
l
l
1
1 02
002

EIx uy dz P
u dz Q2
2
2 y
0

12:56

378

Energy methods in structural mechanics


P
1

2
z

Fig 12.18

On account of the symmetry of the deection with respect to the midspan section, we can express the lateral deection at the mid-span 2 as
follows
l=2

2
0

u0y

l
1 0
dz
u dz
2 y

12:57

so that
l
l
l
1
1 02
1 0
002

EIx uy dz P
u dz Q
u dz
2
2 y
2 y
0

12:58

As usual, the equilibrium is given by


l


EIx u00y u00y

l
dz P

u0y u0y

l
1
dzQ
u0y dz 0
2

12:59

and by integrating by parts and taking into account the boundary


conditions, we get
l

0
EIx u000
y uy

l
dz P
0

u0y u0y

l
1
dzQ
u0y dz 0
2
0

12:60

Stability of equilibrium and non-linear deformations

379

which, since the virtual eld of rotations u0y is arbitrary, is equivalent


to the third-order dierential equation
0
EIx u000
y Puy

Q
2

12:61

By integrating both the left and the right hand terms, we obtain
EIx u00y Puy

Q
z const:
2

12:62

and by evaluating the above equality at z 0, it is immediatly possible


to see that the constant of integration is zero.
Thus, using the denition in equation (12.33) once again, we have
the following equilibrium equation
u00y 2 uy

Q
z
2EIx

12:63

Note that also in this case the equilibrium equation is no longer a


homogeneous equation, as it was the case for the perfect column,
and the general solution of equation (12.63) is
uy z A sin z B cos z

Q
z
2EIx 2

12:64

where A and B are constants evaluated using the boundary conditions,


that are, on account of the symmetry with respect to the mid-span
uy 0 0

and

u0y l=2 0

12:65

Finally, the deected shape of the column is


0
1
Q
sin z
uy
zA
@
2EIx 3 cos z
2

12:66

and the deection at mid-length of the column is




Q
l l

uy l=2
tan
2
2
2EIx 3

12:67

Note that as
P ! PE

l

!
2
2

uy l=2 ! 1

12:68

that is, as the axial load increases close to the magnitude of the Euler
load for the column, the deection becomes very large. The magnitude

380

Energy methods in structural mechanics

of the deections of the column are claried by using the approximation


   2
 3
 l

l

1
l 4 2
2
12
tan
12:69
 2  2
2

l

2
2
This approximation enables equation (12.67) to be written as
0
1
Ql 3
1
12:70
uy l=2
PA
48EIx @
1
PE
Thus it can be seen that when the axial load, P, is zero the initial deection at the mid-length of the beam corresponds to that due to the
lateral load Q only. This is
uy l=2

Ql 3
48EIx

12:71

As the axial load is increased in magnitude, the deection increases


in a non-linear manner until, as stated above, when the axial load
approaches the magnitude of the Euler load the deections become
very large. If the initial deection uy l=2 is dened as the deection
at zero value of P, then the variation of deection with load is
uy l=2

uy
1

P
PE

12:72

It is evident from equation (12.72) that the Euler load is the governing parameter in determining the increase in the deection as the axial
load is applied. The eect of the amplication factor is shown in
Fig. 12.19.

Column with initial out-of-straightness

The previous example has shown how a column with a lateral load Q
initially will have a lateral deection and that deection will grow nonlinearly as the axial load P is imposed. It was assumed in section 12.4,
where the perfect column was analysed, that the initial form of the
column was perfectly straight. That is an idealisation of the actual
initial condition of a practical column which due to the manufacturing
process will have some degree of out-of-straightness along its length.
The present example deals with the evaluation of deections from

Stability of equilibrium and non-linear deformations

381

PE

uy (l/2)

uy

Fig 12.19

that initial condition as the axial load P is imposed on such an imperfect column.
The initial deformations of the column are dened as u0y z, as
shown in Fig. 12.20(a). According to equation (4.25), they dene the
curvature of the axis as
1
u000y
r0

12:73
P

u0y (z)

uy (z)

P
(a)

Fig 12.20

(b)

382

Energy methods in structural mechanics

so that the total potential energy of the system is


l
l
1
1
00
00 2
02
EIx uy u0y dz P
u02
uy
y u0y dz
2
2
0

12:74

The equilibrium equation is therefore


l


EIx u00y

u000y u00y

dz P u0y u0y dz 0

12:75

and by integrating by parts and taking into account the boundary


conditions, we get
l

EIx u00y

u000y u00y

dz P uy u00y dz 0

12:76

which, since the virtual eld of curvatures u00y is arbitrary, is equivalent


to the second order dierential equation
u00y u000y uy 0

12:77

with  given by equation (12.33).


The initial deected shape of a column will be dependant on the
manufacturing process and the treatment the column experiences
during fabrication. The analysis is simplied if the initial deection,
u0y z, is represented as a trigonometric series
u0y z b1 sin

z
2z
3z
b2 sin
b3 sin

l
l
l

12:78

The components, b1 , etc., could be evaluated if an actual column


was being analysed and its initial deected shape had been measured.
However, for the moment it suces that the trigonometric series
representation is feasible. Substitution of equation (12.78) into
equation (12.77) gives
 2
 2

z
2
2z
u00y uy

b1 sin
b2 sin
l
l
l
l

12:79

Stability of equilibrium and non-linear deformations

The general solution of equation (12.79) is


 2

z
l
b1 sin
uy z A sin z B cos z  2
l

2
l
 2

4
2z
l

 2
b2 sin
l

2
4

l

383

12:80

where the constants A and B are determined by the following boundary


conditions
u0

for

z0

and z l

12:81

Hence
B0

and

A sin l 0

12:82

If sin kl 0, then at distinct values of the load, that is the eigenvalues l n, the magnitude of A can have any value. But at a
value of loading dierent from the eigenvalues, sin kl 6 0 and therefore A 0. Thus, provided the consideration is restricted to load
levels less than the Euler load, it is permissible to substitute A 0 in
equation (10.80). Consequently
uy

b1

P
1
PE

sin

z

b2

P
1
4PE

sin

2z

b3

P
1
9PE

sin

3z

l
12:83

It is clear from equation (10.83) that as the axial load P increases,


the denominators in the right hand terms will decrease non-linearly.
Thus, provided the magnitudes of the coecients b1 , b2 , etc. are
fairly similar, the contribution to the magnitude of the deection of
the column will increase and the rate of increase with the load will
be greater for the term with b1 compared to the others. Actually,
generally it is b3 > b3 > b3 > . . ., so that as P ! PE the contribution
of the other terms becomes negligible and without introducing much
error, equation (12.83) can be approximated as
uy

b1
z
sin
l
1 P=PE

12:84

384

Energy methods in structural mechanics

and the deection at mid-length of the column is


uy l=2

b1
1

12:85

P
PE

Note the similarity between equation (12.85) and equation (12.72).


In the latter the initial deection is specied by the actions of the lateral
load Q, and in the former the initial deections occur due to manufacturing eects. In this second case the practical levels of initial deection
are determined by measurements made on a wide range of practical
column geometries. In both cases, the initial deections are amplied
by the action of the axial load P and it is evident that the Euler load
PE , although derived on the basis of idealised column geometry and
loading, is a very important parameter in determining the practical
behaviour of columns.

12.6

Method of trial functions for the initial buckling


of columns

The method for applying trial functions to the evaluation of the Euler
loads for columns is very similar to that considered in chapter 7 for
beams. The method is exemplied rst by application to the perfect
columns analysed above, with uniform cross-section geometry and
with simple support and encastre boundary conditions, and then by
showing the power of the method by considering a perfect column in
which one half has a exural rigidity dierent from the other half.
For a perfect column with uniform cross-section geometry, the total
potential energy of the system, equation (12.29), is
l
l
1
1 02
002
uy U P
EIx uy dz P
u dz
2
2 y
0

As usual, the application of the method of trial functions is simplied if


the energy formulation and the trial functions are non-dimensionalised. Thus we will assume
EI
 3x
2l

1  2
1  2 2
d uy
duy
P
d
d
2
2l
d
d
0

Simply supported column

with  

z
l

12:86

The rst example concerns the derivation of the Euler load for the
column simply supported at both ends and subjected to an axial

Stability of equilibrium and non-linear deformations

385

load P, shown in Fig. 12.12. The boundary conditions, in non-dimensional form, are
uy 0 uy 1 0

12:87

The simplest function satisfying these conditions is


uy uy sin 

12:88

Substituting equation (12.88) into equation (12.86) and performing the


integration gives


EIx 4 2 P2 2
uy
u
4l y
4l 3

12:89

Thus the functional relationship, equation (12.86), has been replaced


by an approximate but very simple algebraic expression. The equilibrium condition is
q
0
q
uy

12:90

Hence


EIx 2
P uy 0
l2

12:91

Two solutions exist for equation (12.91), either uy 0 or


PE

2 EIx
l2

12:92

Thus we have obtained exactly the same results for the Euler load
which were obtained from the exact analytical approach presented in
section 12.4. However, in this case it is evident that the choice of
trial function coincided with the exact mode of deection of the
column at the Euler load.

Encastre column

The boundary conditions in non-dimensional form for this column


(Fig. 12.17(b)) are
uy 0 and

duy
0
d

at  0

and  1

12:93

A simple function that satises these conditions is


uy 12 uy 1 cos 2

12:94

386

Energy methods in structural mechanics

Substitution of this into the energy functional, equation (12.86), and


performing the integration gives


EIx 4 2 P2 2
u
uy
4l y
l3

12:95

from which by means of the equilibrium condition the Euler load is


obtained as
PE

42 EIx
l2

12:96

which again corresponds exactly to the expression we would have


found by the analytical approach.

Column with step-change in exural rigidity

The previous two examples of the application of the method of trial


functions to the evaluation of the Euler load have shown how simple
the method is, but have not provided information beyond that available using the exact analytical approach. However, we know that the
merit in the method of trial functions is that it can provide practically
useful approximate solutions in cases where the exact analysis is quite
complex and sometimes not feasible. This advantage is exemplied
here by applying the trial functions to the evaluation of the Euler
load for a column with a step change of exural rigidity.
Figure 12.21 shows a perfect column that has half its length with a
exural rigidity of I1 and the other half with I2 . The total potential
energy functional for this condition is
2
3
1=2
1  2
1  2 2
 2 2
d uy
d uy
duy
EI1 6
7 P
dI
d5
d
 34
2l
d
2l
d 2
d 2
0

where
I

I2
I1

1=2

(12.97)

12:98

We start the analysis by means of a one-term trial function. The


column is simply supported at both ends therefore a suitable trial function is
uy uy sin 

12:99

Stability of equilibrium and non-linear deformations

387

EI1

l/2

EI2
l/2

Fig 12.21

which substituted in equation (12.97) gives, after integration




EI1 2 4
4 2 P2 2
I
u
u
12:100

4
4 y
4l y
2l 3
Applying the equilibrium conditions, the Euler load is obtained as
1 I 2 EIx
12:101
2
l2
Evidently when I 1, i.e. the column has a uniform exural rigidity
along its length, the Euler load corresponds to the exact value.
The accuracy of the one-term trial function can be tested using more
extensive trial functions. A two-term function can be, for example,
PE

uy u1 sin  u2 sin 2

12:102

Substituting this function in the total potential energy functional, performing the integrations and applying the equilibrium conditions, i.e.
q
q
0;
0
12:103
q
u1
q
u2

388

Energy methods in structural mechanics

gives the equations


1 I
u1

32
1 I
u2 p
u1 0
3

12:104

32
1 I
u1 161 I
u2 4p
u2 0
3
where
p

2Pl 2
2 EI1

12:105

The value of the critical loads are obtained from the zero value of the
determinant of equation (12.104), that is


32


1 I p
1 I


3
12:106
0


32


1 I
161 I 4p
3
Expanding the determinant gives the quadratic equation
2 


32
2
2
4p 20p1 I 161 I
1 I
0 12:107
3
Figure 12.22 shows a plot of the Euler load obtained from the solution of equation (12.107) for varying values of I, compared to the
corresponding values from equation (12.101).
It may be seen that for I ! 1 the two solutions converge and give
the Euler load that agrees with the exact analytical value. However,
as the degree of step-change of the exural rigidity along the column
25
p1 one degree of freedom

p1, p2

20

p2 two degrees of freedom

15

10

05
02

04

06

08
I

Fig 12.22

10

12

Stability of equilibrium and non-linear deformations

389

increases the one-term solution diverges from the supposedly more


accurate two-term solution.

Column with step-change in exural rigidity and a lateral load

We now apply the method of trial functions to the case where the
column with a step-change in exural rigidity has also an initial lateral
load Q. This load is applied at mid-length, as shown in Fig. 12.23.
The non-dimensional expression of the total potential energy functional for this condition is
1  2
1  2 2
d uy
duy
EIx
P
 3
d
d Q2
12:108
2
2l
d
2l
d
0

With reference to Fig. 12.23, the last term in this functional is relative
to the potential energy of the lateral load during the increase in the
deection 2 at the point of application of the load.
The two-term trial functions described in equation (12.102) are applied
to the energy functional, equation (12.116) and after the appropriate
P

EI1

l/2

EI2
l/2

Fig 12.23

390

Energy methods in structural mechanics

integration operations, the equilibrium conditions equation (12.103)


q
q
0;
0
q
u1
q
u2
are applied. Note that in this example, and for the trial functions
chosen, the deection at the point of application of the lateral load
is specied by 2  u1 , so that the potential energy term associated
with the lateral load is Q2  Q
u1 .
Thus, we get
32
u2 q
1 I p
u1 1 I
3
12:109
32
1 I
u1 161 I 4p
u2 0
3
where
p

2Pl 2
2 EI1

and q 

2Ql 3
4 EI1

12:110

Once more, these equations can be put in matrix format


3
2
32
   
1 I
1 I p
q
7 u1
6
3

12:111
5
4
32
u2
0
1 I
161 I 4p
3
The solution of equations (12.111) to evaluate the development of
the lateral deections are exemplied in Fig. 12.24 for the conditions
I 0:2, q 0:1.
Figure 12.22 shows that the non-dimensional Euler load is found to be
p 0:745 by means of the same two-degrees of freedom trial function,
20

u1
u2

u1 (p), u2 (p)

15

10

0
0

Fig 12.24

01

02

03

04
p

05

06

07

08

Stability of equilibrium and non-linear deformations

391

25
20

uy ()

15
10
05
0
05

02

04

06

08

10

12

Fig 12.25

equation (12.102). Figure 10.24 shows that as the value of p increases to


close to the Euler load the deections of the column increase in a nondimensional manner. The deected shape is a combination of the two
terms in the trial function. The shape is exemplied for p 0:7, where
Fig. 12.25 shows the deection shape is asymmetrical about the midlength of the column due to the step-change of the exural rigidity.

12.7

Localised trial functions applied to the initial


buckling of struts

Section 7.6 introduced the concept of localised trial functions that in


eect are an automated way of choosing appropriate functions. This
approach actually can be considered the basis for the nite element
method and chapter 11 has presented in some detail the more formalised aspect of the nite element application in practice.
The treatment in this section starts once again by considering the socalled perfect column, that is the condition in which the column is
initially perfectly straight and the load P is applied perfectly co-axial
with the column, then it takes into account the column with a lateral
load and the column with a step-change of exural rigidity. The
development follows closely that of section 7.6.

Perfect column

The total potential energy of the system is given by equation (12.29), i.e.
l
l
1
1 02
002
EIx uy dz P
u dz
uy U P
2
2 y
0

The practical application of the method is exemplied with respect to a

392

Energy methods in structural mechanics


P

Node 1

Element 1

Node 2

Element 2

Node 3
P

Fig 12.26

column built-in at both ends, Fig. 12.26. For the sake of simplicity we
treat the column as having two elements.
The strain energy can be particularised for an element, that is
a length of the column ln between nodes, as a function of the nondimensional length . Thus we can write
1  2 n 2
d uy
EI
Un 3
d
12:112
d 2
2ln
0

Exactly in the same manner we can evaluate the shortening of the


vertical projection of each element as
1  n 2
duy
1
d
12:113
n
2ln
d
0

so that the total potential energy of the system can be written as


2
3
2 1
3
1  2 n 2
 n 2
n
n
X
X
d uy
duy
EI
1
4
4
d 5P
d 5 12:114
n
3
2
2l
d
d
2l
n
n
i1
i1
0

In the particular case considered here, n 2 and ln l=2.

Stability of equilibrium and non-linear deformations

393

The development in section 7.6 demonstrated that for laterally


loaded beams an eective trial function, yet maintaining simplicity, is
uny  ui 1 1 3 2 2 3 i 1  2 2  3
ui 3 2 2 3 i  2  3

12:113

where ui 1 , ui , are the displacements at nodes i 1 and i, respectively.


As before, for the slopes we set n  n ln .
Substitution of equation (12.113) into equation (12.114) and performance of the stationary conditions for equilibrium, i.e.
qn
0;
q
ui 1

qn
0;
qi 1

qn
0;
q
ui

qn
0
qi

12:114

yields, in the present case


2
32 3
u1
24
12 24
12
0
0
6 12
7
6
8 12
4
0
0 76 1 7
6
7
6
76 7
24 12
48
0 24
12 76 u2 7
4EI 6
6
76 7
6 7
l3 6
4
0
16 12
47
6 12
76 2 7
6
76 7
4 0
0 24 12
24 12 54 u3 5
0
0
12
4 12
8
3
32 3
2
u1
36
3 36
3
0
0
7
6 3
6
4 3 1
0
0 76 1 7
6
7
76 7
6
36 3
72
0 36
3 76 u2 7
P 6
76 7 0
6

6 7
15l 6
0
8 3 1 7
76 2 7
6 3 1
76 7
6
12:115
4 0
0 36 3
36 3 54 u3 5
0
0
3 1 3
4
3
The boundary conditions correspond to
u1 1 u3 3 0
thus, the system equation (12.115) reduces to

 

 
4EI 48 0 u2
P 72 0 u2

0
15l 0 8 2
l3
0 16 2

12:116

12:117

Dening a non-dimensional load as


p

Pl 2
60EI

12:118

394

Energy methods in structural mechanics

equation (12.117) can be written as



 

 
72 0 u2
48 0 u2
p
0 )
0 8 2
0 16 2

 
u2
48 72p
0
0
0
82 p 2

12:119

As usual in the case of perfect columns, the system equation (12.119)


admits the trivial solution, i.e. u2 2 0, for any value of p. Values
of p that allows the system equation (12.119) to have a solution dierent from the trivial one correspond to the zero value of the determinant
on the matrix. That is


48 72p

0

0
12:120

0
82 p
Expanding this determinant, we have
0:776 p2 p 0

12:121

Evidently the roots of this equation are p 0:667 or 2. Intuitively the


lesser of these values is the one that is practically relevant since the
lowest buckling load of a column will govern the onset of the buckling
deformations. Thus, in this case p 0:667 is the relevant value and
gives the value of the load at which instability of the column occurs
as PE 40EI=l 2 . This compares closely with the analytical value
PE 42 EI=l 2 even if, of course, the choice of the trial function,
equation (12.113), cannot represent the trigonometric shape yielded
by the Euler dierential equation.
It is interesting to notice that in the development shown above it is
possible, and trivial, to expand the determinant since there are only
two elements and therefore with the boundary conditions this reduces
the problem to a 2  2 matrix. When the number of elements included
in the modelling is increased the order of the matrix increases also and
the eort to expand the determinant becomes prohibitive. Apart from
this fact, algebraic equations of an order greater than four cannot be
solved algebraically. The approach that is used to obtain the lowest
buckling load in these circumstances is therefore dierent.
If we put equation (12.119) in a totally symbolic form, we can write
Au pBu A pBu B1 A pIu 0
12:122

Stability of equilibrium and non-linear deformations

395

where u is the vector of the degrees of freedom of the system and I is


the identity matrix.
Thus, the required values of the critical loads p are provided by the
eigenvalues of the matrix B1 A, which is a standard algebraic eigenvalue problem which can be solved by means of a classical Jacobi
procedure. Of course, the least of these values is the relevant value
for the onset of large deformations.
The method of localised Rayleigh functions, or nite elements, is
further exemplied here by considering the analysis of a column
simply supported at its ends. In such a case the boundary conditions
for the two-element model are
u1 u3 0

12:123

so that eliminating the corresponding rows and columns from equation


(12.115) we can arrange the matrix equation in the form (12.122)
Au pBu 0
where

Pl 2
;
p
60EI
2

6
6 3
6
B 6
6 1
4
0

6
6 12
6
A 6
6 4
4
3

72

12

7
0 12 7
7
7;
0 16 4 7
5

48

0
12 4 8
3
2 3
0
1
7
6 7
6 u2 7
37
7
6 7
;
u

7
6 7
7
6 2 7
1 5
4 5
4

12:124

3

The result from the calculations are that the eigenvalues of the matrix
B1 A are 0:166, 0:894, 2:145 and 3:36. The lowest of these values is
the one of practical signicance and gives the buckling load as
PE 10EI=l 2 , which as in the previous example, compares very
well with the analytical value PE 2 EI=l 2 .

Column with lateral load

In this section the formatting of the localised trial functions method is


considered for the case of a column with a lateral load. The method
is exemplied by application to a column with uniform exural rigidity
and simple support conditions at both ends. This is shown in
Fig. 12.27(a).

396

Energy methods in structural mechanics


P
Node 1

Element 1

Node 2
Q

Element 2

Node 3

(a)

(b)

Fig 12.27

The energy formulation, in non-dimensional terms and for n


elements, is
2
3
1  2 n 2
n
X
d uy
4 EI
n
d 5
3
d 2
2l
n
i1
0

n
X

41
P
2ln
i1

3
1  n 2
duy
d 5 Q2
d

12:125

where 2 is the lateral displacement of the node at which Q is applied.


By dividing the column in two elements, Fig. 12.27(b), we have 2  u2 .
With the employment of the trial function equation (12.113), the
formulation of the stiness matrix A and of the axial loading
matrix B follows as in the development presented for the perfect
column. Thus the equilibrium equations, after imposition of the
boundary conditions, can be written in the form
Au pBu qE

12:126

Stability of equilibrium and non-linear deformations

397

10
08

P/PE

06
04
02
0
0

18

2/02

27

36

45

Fig 12.28

where
p

Pl 2
;
60EI

q

Ql 3
;
48EI

2 3
0
617
7
E 6
405
0

12:128

The matrices A, B and the vector u are as dened in equation


(12.124). The system equation (12.126) is no longer, as may be expected
given the presence of the lateral load, an homogeneous one. The displacement vector can be obtained for a specic value of the applied
axial load P and of the initial lateral load Q from
u A pB1 qE

12:129

Figure 12.28 shows the variation of the deection at mid-length of


the column, i.e. where the lateral load is applied, with the increase in
the axial load.
The deection is plotted as the ratio of the total deection 2 to the
initial deection caused by the lateral load alone 02 . The variation of
the axial load is plotted as the ratio of the load to the Euler load PE .
Not surprisingly, the localised trial function method replicates the
non-linear increase in the deection that is predicted by the exact
analytical method.

Column with step-change of exural rigidity

In this example the localised trial function method is applied to a


column that has a step-change of exural rigidity form EI1 to EI2 at
mid-length, see section 12.6. The matrix that is aected by this
change of rigidity is A and using the denition I  I2 =I1 , the matrix

398

Energy methods in structural mechanics

for the complete beam for a two-element model can be written as


2
3
24
12
24
12
0
0
6 12
8
12
4
0
0 7
6
7
6 24 12 24I 1 12I 1 24I
7
12I
6
7
A 6
4 12I 1
8I 1 12I
4I 7
6 12
7
4 0
0
24I
12I
24I 12I 5
0
0
12I
4I
12I
8I
12:130
Applying the boundary conditions relevant to simple support conditions,
u1 u3 0

12:131

and eliminating the appropriate rows and columns in the matrix above,
the resulting system is again in the form shown in equation (12.122)
Au pBu 0
with

p

Pl 2
;
60EI1
2

4
6 3
B 6
4 1
0

8
12
4
6 12 24I 1 12I 1
A 6
4 4 12I 1 8I 1
0
12I
4I
2 3
3
1
3 1
0
6
7
72
0
37
6 u2 7
7;
u

6 7
0
8 1 5
4 2 5
3 1
4
3

3
0
12I 7
7
4I 5
8I
12:132

This equation can be solved for specic values of I and the results
compared to the corresponding values from section 12.6. The following typical results are obtained
PE

2 EI1
l2

Trial function
One-term
Two-term
Localised (two elements)

I 1
1
1
0.99


I 0:5
0:75
0:67
0:64

I 0:2
0:60
0:37
0:30

Stability of equilibrium and non-linear deformations

399

Thus the localised trial function agrees fairly well with the two-term
trial functions described earlier with the greater advantage that the
localised method can be automated and the numbers of elements
increases very easily.

12.8

Method of trial functions for the initial buckling


of frames

The previous sections, namely section 12.6 and section 12.7, have
presented the application of the method of trial functions to the
calculation of the critical load of columns with various end geometric
constraints. A column is, of course, usually a component in a frame
structure and it has the important attribute that when it attains its
critical load the frame, or perhaps a region of the frame, will experience
the onset of buckling.
The ultimate load analysis of frames is the subject of many papers
and textbooks and it is not the purpose of the presentation in this
section to emulate the scope of these publications. Instead this section
introduces, by means of a simple example portal frame, the application
of the method of trial functions to the calculation of the critical loads
for frame structures.
The presentation starts by considering the analysis of the portal
frame shown in Fig. 12.29, when subjected to the loads P=2 aligned
co-axially with the vertical members of the frame.
Both the vertical members have the same exural stiness EI1 and
the horizontal member BC has a exural stiness EI2 .
As in all applications of the trial function method, we must rst
consider the likely modes in which the frame will deform at the critical
loads. Figure 12.30 shows two possible modes: (a) sway and (b)
symmetric.
There may be other modes of deformation, and each will have its
corresponding critical load. However, observations of simple tests
show that the two modes in Fig. 12.30 are related to the two lowest
critical loads. The trial functions we choose to represent these modes
are developed and incorporate the compatibility condition that the
frame is exurally rigid at joints B and C. Thus the basic assumption
is that the changes of slope of the vertical and horizontal members
have a common value at these joints.

Sway mode

With reference to Fig. 12.31, and using the non-dimensional variables


1 

z1
l

2 

z2
l

12:133

400

Energy methods in structural mechanics


l

P/2

P/2
B

C
EI2

EI1

EI1

Fig 12.29

The appropriate mode shapes are


Member AB
uAB
y

a1

1
1 cos
2


12:134

This function satises the geometric constraints of zero slope and zero

P/2

P/2

(a)

Fig 12.30

P/2

P/2

(b)

Stability of equilibrium and non-linear deformations

401

B
P/2
B

a 2 = a 1/4

B
B

C
u (z 2)

a1
z2
u (z 1)

(b)
z1

A
(a)

Fig 12.31

deection at 1 0. The displacement and slope at 1 1 are


AB
1 duy

AB
AB
uy B a1
and
 B
a1

l d
2l
1

Member BC
uBC
y a2 sin22

1 1

(12.135)
12:136

The slope of member BC at the joint with the vertical member, i.e. at
2 0, is
BC
1 duy
2
a2
12:137
BC B
l d2 2 0
l
Thus from the condition of compatibility of slope at the joint,
a
12:138
a2 1
4
Hence
a1
uBC
sin22
12:139
y
4
It is evident that in this manner the compatibility condition at the
joint C with the column CD is automatically satised, provided we
assume for the column CD the same shape function of the member
AB, equation (12.134).

402

Energy methods in structural mechanics

The expression of the total potential energy for the frame results
  1  2 AB 2
1  2 BC 2
d uy
d uy
EI1
EI2
2
d1 3
d2
2l 3
d12
2l
d22
0

  1  AB 2
duy
P

d1
2l
d1

12:140

Substituting the chosen trial functions, we have




EI1 4 2 EI2 4 2
2 2
a
a
a

P
1
1
32 1
l 3 32
l3 2

12:141

The equilibrium condition is obtained from the stationary value of the


total potential energy, thus

 4
d
 EI1 4 EI2 2 P
a 0

12:142
da1
8l 1
16l 3
2l 3
and, consequently, we have
PE

2 EI1
1 8I
2l 2

12:143

where I  I2 =I1 .
When the member BC has a very small value of exural stiness
compared to the vertical members, AB and CD, i.e. I ! 0, it is
evident that the critical load for the frame approaches the critical
load for initial buckling of a simple cantilever encastre at the bottom
end, i.e. PE 2 EI1 =4l 2 , by virtue of the fact that the load P is
applied equally to the two vertical members.

Symmetric mode

We assume now that the frame experiences the form of buckling shown
in Fig. 12.30(b). The deformation conditions for the members of the
frame are shown in Fig. 12.32.
Members AB and DC
As by the nature of the method we can assume any trial function which
satises the appropriate geometric constraints, we propose a function
in the form
2
5 3
2 4
uAB
y 1 b1 1 3 1 3 1

12:144

Stability of equilibrium and non-linear deformations


B

403

z2

P/2

b2 = b1 /3

B
B

B
u (z 2)
u (z1)

(b)

z1

A
(a)

Fig 12.32

which evidently satises the following boundary conditions for the


symmetrical mode
duAB
y
0
d1

uAB
y 0

and

uAB
y

at 1 1

at 1 0

12:145

The slope of this member at the joint with BC, that is at 1 1


AB
1 duy
b
AB
12:146
1
 B

l d1 1 1
3l
Member BC
A suitable deformation trial function is
uBC
y 2 b2 sin2
The slope of this member at joint B is
BC
1 duy
b
2
BC B
l d2 2 0
l

12:147

12:148

Hence from compatibility of the slope at joint B


b2

b1
3

12:149

404

Energy methods in structural mechanics

Substituting the trial functions in equation (12.140) gives the equilibrium condition as


d
EI1 8 EI2 2 P 4
3
b 0

12:150
db1
l 105 1
l3 5
l 9
That provides the critical load for the symmetric mode of frame buckling as
PE

4:262 EI1
1 0:343I
l2

12:151

where, as in the sway mode, we have here I  I2 =I1 . In this case therefore the exural stiness of the member BC has a much less eect in
increasing the critical load of the frame than is the case for the sway
buckling mode. Notice as I becomes very small the critical load
approaches the value for the strut pinned at one end and encastre at
the other end, taking into account that the load P is applied equally
to the two vertical members of the frame.
The method of analysis presented above required a separate identication of the two signicant modes of initial buckling associated with
the critical loads of the portal frame. However, it goes without saying
that the choice of deformation mode does not have to be made at the
outset of the analysis. If sucient degrees of freedom are included in
the trial functions, then the calculation of stationary values of the
potential energy will identify the buckling modes automatically.
This is simply attained by incorporating the two buckling modes, i.e.
the sway and symmetric, as one trial function with two degrees of freedom. Thus, the trial function for the members AB and CD becomes




1
5 3 2 4
AB
2
uy 1 a1 1 cos
b 1 1 1 1
12:152
3
3
2
and by reason of compatibility of rotation at the joint B, the trial function for member BC is
uBC
y 2

12.9

a1
b
sin 22 1 sin 2
4
3

Analysis of a frame loaded by vertical and


lateral loads

12:153

In this section, we shall consider the situation in which a load P is


applied to the frame at mid-length along the horizontal member as
shown in Fig. 12.33.
An additional horizontal load Q is applied at B. This implies that
there will be some initial bending in the frame and this modies the

Stability of equilibrium and non-linear deformations

405

l/2

P
B

Q
EI2

EI1

EI1
l

Fig 12.33

response to the applied load from that of the frame analysed in the
previous section in the same manner as for a column, which was
shown in section 12.5.
Let us assume the combined trial functions given by equations
(12.152) and (12.153) where, by means of equation (12.149), we express
b1 in terms of b2 . The total potential energy functional is

 1  2 AB 2
1  2 BC 2
d uy
d uy
EI1
EI2
2
d1 3
d2
2l 3
d12
2l
d22
0

  1  AB 2
duy
P

d1 Pb2 Qa1


2l
d1

12:154

Substituting equations (12.152) and (12.153) in the above expression and


performing the indicated derivatives and integrals we substitute the
functional equation (12.154) with a simple function a1 ; b2 of the parameters a1 and b2 . Thus, the equilibrium conditions are simply given by
q
0;
qa1

q
0
qb2

12:155

Finally, we have
1 8I 4pa1 0:0611 0:2443pb2 0:405q
0:0611 0:2443pa1
0:2631 0:343I 0:1235pb2 0:405p

12:156

406

Energy methods in structural mechanics

where
a1 

a1
;
l

b2 

b2
;
l

I

I2
;
I1

p

P
;
PES

q

Q
PES

12:157

PES 22 EI1 =l 2 is the critical load for a column built-in at one end
and simply supported at the other.
Putting equation (12.156) in matrix notation, we get
Au pBu q
where

"

A
"
B

1 8I
0:0611
4:00

12:158
0:0611

0:2631 0:343I
#
0:2443

0:2443 0:1235
" #
" #
q
a1
:405
u
q

0
p
b2

12:159

Therefore, the system equation (12.158) is not a homogenous one and


this happens not only on account of the transverse load Q but also if we
set Q 0, which implies q 0, as the vector q does not become zero.
This happens on account of the mid-span position of the vertical load
P which always produces a deformation of the member BC that is
transmitted to the columns AB and CD by means of the slope at
the junctions B and C. We come therefore to the conclusion that the
trivial solution u 0 is not admissible for any value of the loading.
In order to quantify this eect, we consider rst the case Q 0.
Figure 12.34(a) shows the deection amplitudes a1 and b2 for I 0
plotted against a range of values of p.
Note that the deections remain quite small until the load p
approaches the symmetric critical value, p 17:67. Thus the eect of
the positioning of the load P is to introduce some deections in the
frame at loads less than the critical value and as the load is increased
to the critical value the deections grow very rapidly. This is a very
similar behaviour to that shown by the individual struts considered
in section 12.5. Notice also that even if the bending stiness of the
horizontal member is set to zero, the corresponding deection b2
does not tend to innity because the member is described by means
of the trial shape function equation (12.153) and b1 3b2 on account
of equation (12.149). Therefore the deection b2 is related to the deection of the columns by means of the equation (12.152).

Stability of equilibrium and non-linear deformations

407

1000

p
q

800

a1*, b*2

600

400

a*1
b*2

200

20

15

10
p

(a)
20

a*1
16

a1*

12

20

15

10
p

05

(b)

Fig 12.34

However, the more important result from this analysis is that the
frame responded to the application of the load P as if there was no
sway buckling mode. This mode has an eigenvalue much less that
the symmetric mode and therefore it would have been anticipated
that deections in the sway mode would have developed at quite
lower values of p. However, if a small lateral load Q is applied, then

408

Energy methods in structural mechanics

the picture changes quite dramatically. Considering the condition in


which the lateral load has a value of 1% of the vertical load P so
that q 0:01p, it may be seen from Fig. 12.34(b) that the magnitude
of the deections in the sway mode, i.e. a1 gradually increases with
increasing values of p, until at the eigenvalue for the sway mode, i.e.
p 2, the deections grow very rapidly.
This simple example illustrates one of the most dicult aspects of
the non-linear buckling analysis of structures which is that unless all
the potential buckling modes are activated by the form of loading
there is a danger that the lowest buckling mode of the structure will
be missed and an engineer designing a structure can easily overestimate
the load carrying capacity of the structure.

12.10

Localised trial function analysis of frame


buckling

The method of localised trial functions, the progenitor of the nite element method, has been applied to the analysis of a frame with a linear
response to the applied loading. The method has also been shown
earlier in this chapter being applied to the analysis of columns with
various boundary conditions and step changes of section geometry.
In the present analysis the method is applied to the evaluation of the
initial buckling load for the simple frame previously considered.
The geometry of the frame is shown in Fig. 12.35. Once again the
analysis is carried out under the hypothesis that the axial deformation

l
z2

y2

y3
3

EI2
EI1

z1

Fig 12.35

y1

z3

EI1

Stability of equilibrium and non-linear deformations

409

of the elements, due to their axial stiness, is negligible with respect to


the eects of the lateral deection.
The potential of the vertical load P is related both to the bending
deection of the horizontal beam and to the vertical displacements
of the nodes 3 and 5. Thus, the potential energy of the system  in
terms of the deection of its members is
uyn U P
EI
1
2

l 
0

EI
1
2

q2 uy1
qz21

l 
0

2

EI
dz1 2
2

q2 uy3
qz23

2

l 
0

q2 uy2
qz22

2

dz2

dz3

8 2l
39
2
2

l 
<
=
quy3
1
1 4 quy1
P
dz1
dz3 5
;
2 :2
qz1
qz3
0

Puy2 l=2

12:160

The members are each composed of two elements with the nodes
numbered 1 to 7 as shown in Fig. 12.36. The interaction between the
elements, i.e. their boundary conditions, is claried in Fig. 12.36
where the nodes at the corners of the frame are separated into the
3"

Fig 12.36

5'

3'

5"

410

Energy methods in structural mechanics

nodes 30 and 500 , associated with the vertical members of the frame, and
300 and 50 , associated with the horizontal members of the frame. The
generic displacement function for each element is obtained from equation (12.113). The boundary conditions at corner nodes are
u300 0;

30 300 ;

u50 0;

50 500

12:161

The boundary conditions at the encastre base of the vertical members


are
1 0;

u1 0;

7 0

u7 0;

12:162

Taking these conditions into account, the equilibrium equations for


the frame, which come from the stationary value of the total potential
energy functional, equation (12.160), are

KD

Pl 2
Pl 3
L1 D
L2 0
120
4

12:163

where the matrices are


2

48

0 24

12

6 0
16 12
4
0
6
6
6 24 12
24
12
0
6
6
4 12 8 8I 12I
6 12
6
6 0
0
0
12I
48I
 EI1 6
K
6
6 0
0
0
4I
0
6
6 0
0
0
0
0
6
6
6 0
0
0
0
12I
6
6
4 0
0
0
0
0
0
2

72

4I

12I

16I

4I

24

12

24

4I

12 8 8I

24

12

0 36

3 0 0

1 0 0

36

3 0 0

6 0
8
6
6
6 36 3
6
6
6 3 1
6
6 0
0
6
L1 6
6 0
0
6
6 0
0
6
6
6 0
0
6
6
4 0
0
0

12
4
0

0 0 0

36 3

72

07
7
7
07
7
7
07
7
07
7
7
07
7
37
7
7
1 7
7
7
05

0 0 0

3 1

4 0 0

0 0 0

0 0 0

0 0 0

36

0 0 0

3 36
4

12
48

07
7
7
07
7
7
07
7
07
7
7
07
7
12 7
7
7
47
7
7
05

0 16

Stability of equilibrium and non-linear deformations

411

2 3
0
607
6 7
6 7
607
6 7
607
6 7
6 7
617
7
L2 6
607
6 7
6 7
607
6 7
607
6 7
6 7
405
0
As usual, D is the vector containing the nodal degrees of freedom,
i.e.
3
2
u2
6  7
6 27
7
6
6 u30 7
7
6
6  0 7
6 3 7
7
6
6 u4 7
6
D 6  7
7
6 4 7
7
6
6 u500 7
7
6
6  00 7
6 5 7
7
6
4 u6 5
6
and I  I2 =I1 .
Notice that the two rows and columns in L1 associated with u4 and
4 are completely null. This is because the vertical load P plays its role
through two distinct terms: the rst one is related to the shortening of
the vertical members on account of their lateral deection, the second
one is related to the vertical displacement of the mid-span of the
horizontal beam on account of its bending. However, these rows and
columns are required to maintain the correct order of matrix L1 to
be compatible with the vector D.
It is evident that the rst two terms in equation (12.163) determine
the eigenvalues and corresponding modes of the initial buckling of
the frame while the third term reects the bending deection of the
horizontal member in the frame. Eliminating for the moment the

412

Energy methods in structural mechanics


P/2

P/2

EI2

EI1

EI1

Fig 12.37

third term, equation (12.163) becomes


 L D 0
K
1

12:164

where


Pl 2
;
120EI1

 
K

1 
K
EI1

The formulation in equation (12.164) is equivalent to divide the load


P such that it is aligned with the two vertical members of the frame as
shown in Fig. 12.37.
 gives ten eigenvalues, whose number is equal to
Setting I 0 in K
the number of degrees of freedom in the nite element model, of which
the lowest is
c1 0:0425

i:e:

PE1

0:5172 EI1
l2

12:165a

with the eigenmode deected shape shown in Fig. 12.38(a).


The initial buckling load PE1 is evidently associated with a sway
form of buckling. The next lowest eigenvalue is
c2 0:348

i:e:

PE2

4:232 EI1
l2

12:165b

with the corresponding deected shape shown in Fig. 12.38(b), i.e. a


symmetric deected buckling mode. These initial buckling loads and
modes correspond to those obtained in section 12.5 for columns
encastre at one end and with zero exural restraint, i.e., simply
supported, at the other, as indeed would be the case for I 0.

Stability of equilibrium and non-linear deformations

= 00425
PE1
2

P/2

P/2

P/2

P/2

= 0348

I=0

0258 2 EI1
l2

PE2
2

(a)

= 01772
PE1
2

212 2 EI1
l2

(b)

P/2

P/2

P/2

P/2

413

= 06667

I=

108 2 EI1
l2

PE2
2

(c)

405 2 EI1
l2

(d)

Fig 12.38

 the lowest two eigenvalues from equation


Applying I 1 in K,
(12.164) are
c1 0:1772

i:e:

c2 0:6667

i:e:

Pc1
Pc2

2:152 EI1
l2
8:112 EI1

12:166

l2
with the corresponding deformed shapes shown in Figs 12.38(c) and (d).
These two initial buckling loads also correspond to sway and symmetric
modes of buckling but now for the condition where the horizontal
member of the frame has an innite exural stiness.
The value of the lowest eigenvalue is dependent on the value of I,
that is, the ratio of the exural stiness of the horizontal and vertical
members of the frame. This was shown also in section 12.8 for the
application of the method of trial functions for the initial buckling

414

Energy methods in structural mechanics

Coefficient of initial
buckling load

4
3
Variation of the initial buckling load
with I (2 term trial function)

2
1
0

05

10

15

Value of I
(a)
Coefficient of initial
buckling load

12
10
08

Variation of the initial buckling load


with I (finite element)

06
04
02
0
0

05

10

15

Value of I
(b)

Fig 12.39

of frames. The analysis in section 12.8 involved two degrees of freedom


whereas the present analysis using the localised trial function method
has ten degrees of freedom. Figure 12.39(a) shows the variation of
the lowest initial buckling loads, as I is increased. Figure 12.39(b)
shows the corresponding variation of the initial buckling loads
obtained from the two degree of freedom analysis in section 12.8.
It is evident that in the case at hand incorporating a larger number of
degrees of freedom predicts lower buckling loads.

Example analysis for I 2:68

Section 6.5 has shown the linear analysis of a plane frame with a
vertical load at the mid-span of the horizontal member. The ratio of
the exural stiness of the horizontal and vertical member is
I 2:68. The same frame geometry is considered here from the
viewpoint of the eects of the vertical load to cause initial buckling
of the frame.
 and obtaining the lowest eigenvalue from
Putting I 2:68 in K,
equation (12.164) gives
c1 0:1264
and equation (12.163) can be written
 L D0 30L
K
12:167
1

Stability of equilibrium and non-linear deformations

415

where
1
D0  D;
l


K

1 
K
EI1

thus
 L 1 L
D0 30K
1
2

12:168

Substituting various values of  into equation (12.168) the


corresponding values of the deections of the frame at the nodes can
be evaluated. It is evident from the symmetrical form of the frame
that if the vertical load is applied to the mid-span of the horizontal
member then the sway mode of deection may not be detected. This
was shown in section 12.8 where it was clear that in order to ensure
that the lowest buckling mode was initiated in the analysis, a lateral
load has to be applied to the frame. Similarly, in the present analysis,
a small non-dimensional lateral load is required to ensure that the
deformations calculated using equation (12.168) correspond to the
lowest value of the initial buckling load, i.e. the sway mode of buckling. Figure 12.40(a) shows the arrangement in which the lateral
load, i.e. 0:01 is classied as an initial imperfection.
This loading is incorporated in equation (12.168) by modifying the
vector L2 to be
3
2
0
60 7
7
6
6 : 7
6 0 01 7
7
6
60 7
7
6
7
6
61 7
7
L2 6
60 7
7
6
7
6
60 7
7
6
60 7
7
6
7
6
40 5
0
Figure 12.40(b) shows the variation of the horizontal deection at
node 3 with increasing values of . The applied load , has been
normalised by dividing by the lowest eigenvalue c1 .
It is relatively simple in the present example to ensure that the initial
imperfection, i.e. the lateral load, will indeed induce deformation that
correspond to the lowest initial buckling load. In other cases of nite

416

Energy methods in structural mechanics


30
03

Normalized vertical load

(a)
Variation of lateral deflection with increasing vertical load
(1% lateral load)
10
08
06
04
02
0

05
10
15
Non-dimensional lateral deflection

20

(b)

Fig 12.40

element modelling, where there may be thousands of degrees of freedom, the application of an initial loading system may be much more
complex. In such cases the eigenvalues can be used to introduce a
system of initial geometric imperfections that will indeed ensure that
the lowest initial buckling deformations are not overlooked in the
analysis.
The procedure is straightforward. Equation (12.164) is used to
calculate the eigenvalues of the system and the corresponding eigen~n . These eigenmodes are then introduced into equation
modes, D
(12.167) as follows
 L D0 30L a D
~ 0 a D
~0   
K
12:168
1

Stability of equilibrium and non-linear deformations

417

Normalized
vertical load

Variation of lateral deflection with increasing


vertical load (initial geometric imperfection)
10
08
06
04
02
0
0

1
Non-dimensional lateral deflection

Fig 12.41

~1 is the eigenmode for c1 , etc., and a1 ; a2 ; . . . are constants


where D
that are generally chosen such that the form of the initial deection
of the structure is as general as possible and the magnitude of the initial
deection corresponds to measurements on practical structures.
~0i  D
~i =l.
Again, D
Using equation (12.168) the deformations at the nodes of the
model can be calculated for increasing values of . Figure 12.41
shows the results from such a procedure in which a1 0:1 and
a2 a3    0. Comparison between Figs 12.41 and 12.40(b)
show that the procedure of using the eigenmodes is very similar, in
the present example, to applying a small articial lateral load.

Suggested exercise problems


1.

The column shown in Fig. 12.42 is simply supported at both


ends and has an initial deection at mid-length of 50 mm. The
cross-section is a cylindrical tube with 300 mm outside diameter
and 10 mm wall thickness. Calculate the magnitude of the applied
P

oc = 50 mm
15 m

Fig 12.42

418

Energy methods in structural mechanics


P

oc =
l

l
500

axial load, P, such that the maximum strain does not exceed the yield
strain, "0 0:0015. The modulus of the material is 205 kN/mm2 .
[Ans: P 673 kN]
The column shown in Fig. 12.43 has a length l, and is built-in at
both ends. The column has an initial deection of 0c l=500.
The column is subjected concurrently to a lateral uniformly distributed load, q, and an axial load P. Derive an expression for
the total deection at mid-length, c , of the column in terms of
P, l, q and the exural stiness, EI.
"
#


l
1:3ql 3
Pl 2
1
c
1 2
Ans:
500
EI
4 EI

2.

3.

A column, pinned at its base, is attached to a cross-beam as shown


in Fig. 12.44. The column has an initial deection at mid-length of
0c l=300 and both the column and the cross-beam have the
P

Mode a
l

l
oc =

l
300

Mode b

Fig 12.44

Stability of equilibrium and non-linear deformations

419

same exural rigidity, EI. Show, using the method of localised trial
functions, whether the rate of increase in the total strain, "t at midlength of the column is greater in mode a or mode b.
[Ans: mode b has an eigenvalue 70% of the value of the eigenvalue
of mode a, therefore the rate of growth of bending strain in mode b
is greater than in mode a at the same applied load.]

Index
amplication factor, 380
anisotropic material properties, 63
bar buckling, 373
Beltrami, Eugenio, 81
Beltrami's equations, 81
Beltrami's yield criterion, 88
bending moment, 37
bending moment diagram, 39
Bernoulli, Daniel, 106
Bernoulli, Jacob, 94
Bernoulli's model of beam bending, 94
Betti's theorem, 148
bifurcation condition, 360
branching point, 360
buckling sway, 377
buckling, bifurcation, 360
buckling, snap, 365
bulk modulus, 90
calculus of variations, 108
Castigliano, Carlo, 155
Clapeyron, Benoit-Paul-Emile, 82
Clapeyron's theorem, 82
closed system, 5
combined joint load vector, 254
compatibility equation, 214
concatenated displacements, 28
conservative systems, 154
coordinate functions, 170, 188
Cotteril, James, 155
critical load, 360, 370
Croll, James, 106
degrees of freedom, 6, 174, 211

Descartes, Rene, 12
direct stiness method, 221
Dirichlet's theorem, 358
displacement work, 147
ductility, 54
eigenvalue problem, 395
eigenwork, 148
elastic, 54
elastic energy, 3
element, 191
elements, non-conforming, 321
energy, 1
energy conservation, 5
engineering shear strain, 70
equilibrium, neutral, 355
equilibrium, unstable, 354
equilibrium, stable, 354
equivalent joint loads, 254
Euler buckling, 373
Euler buckling load, 370
Euler, Leonhard, 20
Euler's equation of beam buckling, 369
Euler's equation of beam deformation,
109
extensometer, 51
external constraints, 21
rst law of thermodynamics, 5
exural stiness, 110
free body, 36
fundamental path, 360
Gauss, Karl, 338
generalised coordinates, 6

422

Index

generalised coordinates, 174


generalised displacement, 149
generalised mass, 9
Green, George, 66
Green's strains, 66

Newton, Isaac, 1,12


node, 234
nominal strain, 52
nominal stress, 52
non-conforming elements, 321

heat, 3
Hooke, Robert, 57
Hooke's law, 57

Pascal, Blaise, 321


patch test, 322
perfect column, 391
perfect system, 362
plane strain, 79
plate bending stiness, 277
plate deformation, 267
Poisson, Simeon-Denis, 61
Poisson's ratio, 61
principal components of strain, 77
principle of conservation of energy, 5
principle of stationary work, 12
principle of substitution of constraints,
33
principle of superpositon, 143
pure bending, 99
Pythagorus, Samar, 367

imperfections, 357
ineective constraint, 25
innitesimal displacements, 102
initial buckling, 356
initial imperfection, 415
initial out-of-straightness, 380
internal constraints, 21
intrinsic coordinates, 315
isoparametric, 317
isotropic material, 63
Jacobi, Karl, 316
Jacobian matrix, 316
kinetic energy, 3, 8
Kircho, Gustav, 144
Kirkho model of plate deformation,
267
Lagrange family of shape functions,
324
Lagrange, Joseph-Louis, 12
Lagrangian coordinates, 12
lateral load, 377
Legendre, Adrien-Marie, 339
limit points, 365
linear elastic, 56
localised Rayleigh-Ritz method, 189
Luder's plateau, 54
material modulus, 57
Mises, Richard von, 90
mode, sway, 399
mode, symmetric, 402
modulus, shear, 85
moment of inertia, 101
neutral equilibrium, 355

Rayleigh, Lord, 174


relative centre of rotation, 27
rigid joints, 234
Ritz, Walther, 174
rotational restraint, 373
Saint-Venant, Adhemar-Jean-Claude,
89
Saint-Venant's yield criterion, 89
second moment of area, 101
serendipity family of shape functions,
324
shape function, 170
shape functions, 309
shape functions, Lagrange family, 324
shape functions, serendipity family,
324
shear force, 37
shear force diagram, 39
shear modulus, 85
shear strain, 69
Shute, Nevil, iv
snap buckling, 365

Index
static equilibrium, 9
statical determinacy, 35
statical indeterminacy, 35
stiness matrix, 215
strain compatibility, 80
strain energy, 9
strain energy density, 83
strain invariants, 76
superposition, 143
sway buckling, 377
sway mode, 399
symmetric mode, 402
theory of innitesimal displacements,
102
thermodynamics, 5
Timoshenko, Stephen, 290

423

trial displacements, 10
true strain, 52
true stress, 52
virtual displacements, 10
Von Mises yield criterion, 90
Walker, Alastair, 106
Walpole, Horace, 327
Woinowsky-Krieger, Stephen, 290
work, 1
yield plateau, 54
yield strain, 57
Young, Thomas, 57
Young's modulus, 57

Anda mungkin juga menyukai