Download as pdf or txt
Download as pdf or txt
You are on page 1of 362

Developments in Civil Engineering

Vol. 1 The Dynamics of Explosion and its Use (Henrych)


Vol. 2 The Dynamics of Arches and Frames (Henrych)
Vol. 3 Concrete Strength and Strains (Avrametal.)
Vol. 4 Structural Safety and Reliability (Moan and Shinozuka, Editors)
Vol. 5 Plastics in Material and Structural Engineering (Bares, Editor)
Vol. 6 Autoclaved Aerated Concrete, Moisture and Properties (Wittmann, Editor)
Vol. 7 Fracture Mechanics of Concrete (Wittmann, Editor)
Vol. 8 Manual of Surface Drainage Engineering, Volume II (Kinori and Mevorach)
Vol. 9 Space Structures (Avram and Anastasescu)
Vol. 10 Analysis and Design of Space Frames by the Continuum Method (Kollar and Hegedus)
Vol. 11 Structural Dynamics (Vertes)
Vol. 12 The Selection of Load-Bearing Stuctures for Buildings (Horvath)
Vol. 13 Dynamic Behaviour of Concrete Structures (Tilly, Editor)
Vol. 14 Shells, Membranes and Space Frames (Heki, Editor)
Vol. 15 The Time Factor in Transportation Processes (Tarski)
Vol. 16 Analysis of Dynamic Effects on Engineering Structures (Bata and Plachy)
Vol. 17 Post-Buckling of Elastic Structures (Szabo, Gaspar and Tarnai, Editors)
Vol. 18 Fracture Toughness and Fracture Energy of Concrete (Wittmann, Editor)
Vol.19 Pavement Analysis (Ullidtz)
Vol. 20 Analysis of Skeletal Structural Systems in the Elastic and Elastic-Plastic Range (Borkowski)
Vol. 21 Creep and Shrinkage of Concrete Elements and Structures (Smerda and Kf istek)
Vol. 22 Theory and Calculation of Frame Structures with Stiffening Walls (Pubal)
Vol. 23 Time Effects in Concrete Structures (Gilbert)
Vol. 24 Stresses in Layered Shells of Revolution (Kovafik)
Vol. 25 River Intakes and Diversion Dams (Razvan)
Vol. 26 Analysis of Dimensional Accuracy of Building Structures (Vorlicek and Holicky)
Vol. 27 Reinforced-Concrete Slab-Column Structures (Ajdukiewicz and Starosolski)
Vol. 28 Finite Models and Methods of Dynamics in Structures (Henrych)
Vol. 29 Endurance of Mechanical Structures (Nemec and Drexler)
Vol. 30 Shells of Revolution (Mazurkiewicz and Nagorski)
Vol. 31 Structural Load Modeling and Combination for Performance and Safety Evalution (Wen)
Vol. 32 Advanced Analysis and Design of Plated Structures (Kfistek and Skaloud)
Vol. 33 Regular Lattice Plates and Shells (Sumec)
Vol. 34 Combined Ultrasound Methods of Concrete Testing (Galan)
Vol. 35 Steel-Concrete Structures for Multistorey Buildings (Kozak)
Vol. 36 Analytical Methods in Bin-Load Analysis (Drescher)
Vol. 37 Design of Welded Tubular Connections - Basis and Use of A WS Code Provisions (Marshall)
Vol. 38 Fresh Concrete Properties and Tests (Bartos)
STABILITY, BIFURCATION
AND POSTCRITICAL
BEHAVIOUR OF
ELASTIC STRUCTURES

M. PIGNATARO
N. RIZZI
A. LUONGO
Department of Structural and Geotechnical Engineering,
University of Rome 'La Sapienza\ Italy

ELSEVIER
Amsterdam - London - New York -Tokyo
1991
ELSEVIER SCIENCE PUBLISHERS B.V.
Sara Burgcrhartstraat 25
P.O. Box 211, 1000 AE Amsterdam, The Netherlands

Distributors for the United States and Canada:


ELSEVIER SCIENCE PUBLISHING COMPANY INC.
655 Avenue of the Americas
New York, NY 10010, U.S.A.

Library of Congress C a t a l o g i n g - i n - P u b l i c a t i o n Data

i g n a t a r o , M. ( M a r c e l l o )
S t a b i l i t y , b i f u r c a t i o n , and p o s t c r i t i c a l b e h a v i o u r o f e l a s t i c
s t r u c t u r e s / M. P i g n a t a r o , N. P . i z z i , A . L u o n g o .
p. cm. — ( D e v e l o p m e n t s i n c i v i l e n g i n e e r i n g ; v. 39)
I n c l u d e s b i b l i o g r a p h i c a l r e f e r e n c e s and i n d e x .
ISBN 0 - 4 4 4 - 8 8 1 4 0 - 9
1. E l a s t i c a n a l y s i s ( E n g i n e e r i n g ) 2. B i f u r c a t i o n theory.
I . R i z z i , N. I I . L u o n g o , A. I I I . Title. IV. S e r i e s .
TA653.P54 1991
624. 1 '71—dc20 91-24932
CIP

ISBN 0-444-88140-9

® 1991 Elsevier Science Publishers B.V. All rights reserved.

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, recording or otherwise, without the prior written
permission of the publisher, Elsevier Science Publishers B.V./Academic Publishing Division, P.O. Box
1991, 1000 BZ Amsterdam, The Netherlands.

Special regulations for readers in the USA - This publication has been registered with the Copyright
Clearance Center Inc. (CCC), Salem, Massachusetts. Information can be obtained from the CCC about
conditions under which photocopies of parts of this publication may be made in the USA. All other
copyright questions, including photocopying outside of the USA, should be referred to the publisher.

No responsibility is assumed by the Publisher for any injury and/or damage to persons or property as a
matter of products liability, negligence or otherwise, or from any use or operation of any methods,
products, instructions or ideas contained in the material herein.

Printed in the Netherlands


To Professor Warner T. Koiter
vii

PREFACE

The study of the stability of mechanical systems is a fascinating subject


which has stimulated the interest of mathematicians and engineers since the
time of Euler. Apart from his pioneering groundwork, which was limited to
the particular case of the critical and post-critical equilibrium of compressed
rods, all successive literature up to the 1930s attempted solely to determine
the critical loading of mechanical systems, erroneously claiming this to be an
adequate indicator for the evaluation of structural behaviour.
The analytical results obtained, which were based on linear analysis, agreed
within the errors with experimental results for beams and plates, but not for
shells, particularly those of cylindrical form, where the experimental results
showed a very great dispersion.
The work of Marguerre and Trefftz on the post-critical behaviour of plates,
and of von Karman and Tsien on that of cylinders, based on non-linear anal-
ysis, showed how the behaviour of the two structural types was qualitatively
different. Even though these works describe in detail the phenomena of buck-
ling and post-buckling, revealing the inadequacy of the critical load for eval-
uating structural behaviour, the lack of an adequate theory was reflected in
the impossibility of providing indications of a general character, such as for
post-critical behaviour and collapse loads.
Not until 1945 did Koiter, in his doctoral thesis, present a general theory
on the critical and post-critical analysis of elastic structures. As well as con-
stituting a fundamental source for a deeper understanding of buckling, this
also allows us, through the introduction of the concept of "initial imperfec-
tions", to arrive at a quantitative evaluation of the collapse load in a simple
and rational way. It also helps to explain the reason for the wide spread of
experimental results obtained for cylinders.
The work of Koiter, written in Dutch, remained virtually unknown until
1967, when it was translated into English. Meanwhile, towards the end of
the 1950s, a non-linear theory on the post-critical equilibrium of discrete sys-
tems was developing independently in England with the work of Sewell and
Thompson. All the work published in recent years presents either the analy-
sis of continuous systems, following the theory of Koiter, or that of discrete
systems, following the English school. In the present volume the authors
present a treatment, as far as possible organic, of the non-linear behaviour
of discrete and continuous systems, trying to point out the common method-
ology of investigation based on asymptotic analysis. In the light of Koiter's
theory, in fact, it is evident how such a method of investigation, despite its
"local" nature, is quite adequate for providing not only a qualitative but also
a quantitative estimate of the phenomena studied, in this connection better
viii

than costly finite analysis methods are able to do.


The majority of modern texts focus totally or mainly on the construction of
a bifurcated path, whilst in the present volume the authors attach importance
to stability analysis corresponding to the bifurcation which always precedes
the investigation of post-critical behaviour.
As stability is a dynamic concept the authors thought it right to study
it in the wider context of the theory of the stability of equilibrium, accord-
ing to Liapunov. In such a context, and like Koiter, Sewell and Thompson
limiting the study to elastic systems subjected to conservative forces, the dy-
namic criterion is reduced to the total potential energy criterion by means
of the theorems of Lagrange and Koiter for discrete and continuous systems
respectively.
It is necessary to mention that the present volume lacks a systematic
treatment of initial imperfections and of simultaneous modes which have a
marginal importance in the investigation of some simple structural models.
Furthermore, in the review of structural topologies, the study of arches has
been omitted completely; this is to limit the text, keeping it at an elementary
level. In spite of this, the authors hope that they have been able to provide
an exhaustive view of the problems of buckling.

December 1990

M. Pignataro
N. Rizzi
A. Luongo

A CKNO WLEDGEMENTS
The authors wish to thank the publishers Springer Verlag-Vienna, for au-
thorising the reproduction of Figs. 7.86 to 7.40 and 7.45 from the book 'Post-
Buckling Behaviour of Structures' by M. Esslinger and B. Geier, No. 286 of
the series CISM Courses and Lectures, 1975, and Granada Publishing Ltd for
permitting the reproduction of Figs. 7.41 to 7.48 from the book 'Thin-Walled
Structures' by J. Rhodes and A.C. Walker, London 1980, and Figs. 5.1 to 5.5
and 6.1 to 6.4 from the book 'Design for Structural Stability' by P.A. Kirby
and D.A. Nethercot, London 1979.
viii

than costly finite analysis methods are able to do.


The majority of modern texts focus totally or mainly on the construction of
a bifurcated path, whilst in the present volume the authors attach importance
to stability analysis corresponding to the bifurcation which always precedes
the investigation of post-critical behaviour.
As stability is a dynamic concept the authors thought it right to study
it in the wider context of the theory of the stability of equilibrium, accord-
ing to Liapunov. In such a context, and like Koiter, Sewell and Thompson
limiting the study to elastic systems subjected to conservative forces, the dy-
namic criterion is reduced to the total potential energy criterion by means
of the theorems of Lagrange and Koiter for discrete and continuous systems
respectively.
It is necessary to mention that the present volume lacks a systematic
treatment of initial imperfections and of simultaneous modes which have a
marginal importance in the investigation of some simple structural models.
Furthermore, in the review of structural topologies, the study of arches has
been omitted completely; this is to limit the text, keeping it at an elementary
level. In spite of this, the authors hope that they have been able to provide
an exhaustive view of the problems of buckling.

December 1990

M. Pignataro
N. Rizzi
A. Luongo

A CKNO WLEDGEMENTS
The authors wish to thank the publishers Springer Verlag-Vienna, for au-
thorising the reproduction of Figs. 7.86 to 7.40 and 7.45 from the book 'Post-
Buckling Behaviour of Structures' by M. Esslinger and B. Geier, No. 286 of
the series CISM Courses and Lectures, 1975, and Granada Publishing Ltd for
permitting the reproduction of Figs. 7.41 to 7.48 from the book 'Thin-Walled
Structures' by J. Rhodes and A.C. Walker, London 1980, and Figs. 5.1 to 5.5
and 6.1 to 6.4 from the book 'Design for Structural Stability' by P.A. Kirby
and D.A. Nethercot, London 1979.
ix

INTRODUCTION

This volume is intended for students of structural engineering, those study-


ing for a PhD degree and for young researchers interested in problems of
non-linear mechanics, and requires a knowledge of the basics of analysis and
algebra. In the Appendix the authors give a short account of aspects of the
calculus of variations not usually studied in classical courses of analysis, which
is sufficient for the study of this text.
In Chapter 1 the authors discuss the nature of the equilibrium points of
an autonomous system of differential equations, both linear and non-linear,
followed by the concept of stability according to Liapunov and Chetayev.
Then the conditions under which it is possible to analyse a non-linear problem
are discussed making reference to the corresponding linearised problem.
In Chapter 2, after a brief revision of the Lagrange and Hamilton equations
of motion, the authors apply the Liapunov definition of stability to mechan-
ical systems. This is followed by a demonstration of the Lagrange-Dirichlet
theorem and enunciation of the theorems of Liapunov and Chetayev. A dis-
cussion of the criterion of stability of discrete systems follows, providing the
conditions necessary and sufficient for the total potential energy to be positive
definite. In the second part of the chapter some simple mechanical models of
one or two degrees of freedom are examined in detail, with the aim of pre-
senting in an elementary context basic ideas such as primary and secondary
curves of equilibrium, bifurcation and limit points, the energy criterion for
stability and the influence of initial imperfections and simultaneous modes.
In Chapter 3 a local analysis of the properties of points belonging to an
equilibrium path is carried out, determining the conditions which distinguish
a regular equilibrium point from one of bifurcation and from a limit point.
Successively, the use of perturbation analysis in the asymptotic determina-
tion of equilibrium curves through a point is illustrated. The critical points
along a known equilibrium path are then looked for, and local analysis of the
bifurcated path around the bifurcation point is effected. Finally, a system
of two degrees of freedom characterised by a non-linear fundamental p a t h is
analysed and the bifurcation point and the branched curve are determined,
critically comparing the results thus obtained with those corresponding to a
linearisation of the fundamental path.
Chapter 4 extends the results obtained in Chapters 2 and 3 for discrete
systems to continuous systems. Having enunciated the theorem of Koiter
which provides, by means of the potential energy, the sufficient condition
of stability of continuous systems, the authors go on to discuss firstly the
critical condition of equilibrium and then the conditions under which the total
potential energy is positive definite. Finally, an illustration of the construction
X

of bifurcated equilibrium paths by means of perturbation analysis is given.


The last three chapters describe applications of the general theory to par-
ticular structures such as plane frames, thin-walled beams and shells.
In Chapter 5 two models of plane beams are presented, one characterised
by three internal degrees of freedom, two displacement components and a ro-
tation of the section, the other by pure flexural deformation. For both models
the total potential energy is given for a system of beams, up to terms of the
fourth order. For the internally constrained model the energy is modified
by means of additional terms, which take account of the constraint condi-
tions by using Lagrange multipliers. By using the constrained model, various
problems affecting beams subjected to different boundary conditions and ax-
ial loads are successively resolved. Solutions to the post-critical problem are
given for some of these. The Ritz method is often used for determination of
the critical load. Using the same model, the authors examine several types
of simple frames, for which the complete analytical solution is given. At the
end of the Chapter a series of results for complex frames, obtained through a
compatible finite elements technique which makes use of the first beam model,
is presented.
In Chapter 6 Vlasov's theory on thin-walled open cross-sections, based on
the two fundamental hypotheses of the underformability of the section in its
own plane and of the absence of shear deformation in the middle surface of the
beam, is presented. The total potential energy up to the second-order terms
is then given, as this Chapter is for the most part based only on the study of
critical behaviour. The Eulerian equations of the problem are then deduced
for the most general cases of load and boundary conditions. The equations
are successively specialised to the study of a series of classical problems, such
as those of flexural-torsional instability and lateral instability of beams. The
Ritz method of discretisation (for which an application is given) and of com-
patible finite elements are illustrated. Finally, the post-critical behaviour of
thin-walled members under the hypothesis of simultaneous modes is exam-
ined. The analysis is carried out by using the Vlasov model and writing the
total potential energy up to third-order terms.
Chapter 7 concludes the volume with an examination of two-dimensional
structures. After a brief mention of kinematics in general curvilinear coordi-
nates, the authors come to the expression of the total potential energy based
on the simplified kinematic relations of shallow shells, according to Donnell-
Mushtari-Vlasov. The total potential energy of the plate is consequently
obtained through a limit process. Successively, after discussing the stability
of the post-critical behaviour of plates, the authors pass on to the application
of the bifurcation theory to a series of classical problems regarding various
boundary and load conditions. For some problems, to which no exact solu-
tion is known, Ritz's method is used. Mention is also made of the problem
XI

of stiffened plates. In the second part of the Chapter the critical behaviour
of cylinders subjected to radial, axial or hydrostatic pressure, and of spheres
under hydrostatic pressure, is studied. Indications of a qualitative character
only are given for the post-critical behaviour of shells. These are accompained
by a series of diagrams showing experimental results and of photographs of
the models.
Finally, the Appendix reports some essential notions of the calculus of vari-
ations necessary to the understanding of the analytical development reported
in the text.
While not considering the present volume to be a complete work, the
authors nevertheless hope to have stimulated the reader's interest in the study
and further investigation of the subject.
1

Chapter 1

THE LIAPUNOV THEORY OF


EQUILIBRIUM STABILITY

1.1 INTRODUCTION
In the development of the theory of differential equations, it is possible to
distinguish two quite different approaches. The first is characterised by the
search for a solution in closed form or through a process of approximation.
The second can be distinguished by the fact that information on the solution
is sought without actually solving the problem. This qualitative analysis
was introduced by Poincare around 1880 [l] and developed in the following
decades, especially in Russia.
The central problem in qualitative analysis is to investigate the relation-
ship between the solution and its neighbourhood. A solution is a curve or
trajectory C in a certain space. The question is whether any V trajecto-
ries, which for t — 0 start near C, tend to remain near C or move away
from it. In the first case, the trajectory C is said to be stable; in the second
unstable. Liapunov is credited with creating qualitative analysis, which is
generally called the theory of stability. In 1892 he published the first of a se-
ries of fundamental papers "General Problems on the Stability of Motion" [2],
in which he treated the problem of stability in two different ways. His so-called
first method presupposes explicit knowledge of the solution and is applied
only to a limited but important number of cases; the second method, or
direct method, is altogether general and does not require knowledge of the
solution.

1.2 D I F F E R E N T I A L EQUATIONS
From a historical point of view differential equations were introduced by
Newton through the laws of mechanics which define the motion of a freely
falling body or one subjected to a system of forces. Subsequent developments
in physics have shown how a wide range of problems in completely different
fields are governed by laws which are altogether analogous to those of me-
chanics. Thus it is desirable, as a first step, to describe the types of equations
on which we shall be working and their properties.
The differential equations which are the basis of the problems we are to
2

study are essentially of two types [3,5]. The first is represented by an equation
of n-th order

xW = / ( x , ± , . . . , x ( n - 1 } ; t) (1.2.1)

where t is a variable and generally, but not necessarily, represents time, and
k
x( ) represents the A>th derivative of x with respect to the variable. The
second type is a system of n equations of the first order

Xi = X{ {Xj ; t) (1.2.2)

where, unless otherwise specified, the Latin indices are understood to vary
from 1 to n.
The first type can be reduced to the second if we introduce the new varia-
bles Xi, x2, . . . , xn defined by

Xi = x ^ 1 ) (1.2.3)

In this case equation (1.2.1) is replaced by the system

x{ = xt+i (t = 1, 2 , . . . , n - 1)
(1.2.4)
xn = j [Xj 5 t)

As an example the well known equation of van der Pol

x + k (x2 - 1) x + x = 0 (1.2.5)

can be replaced by the system

ii = x2
(1.2.6)
x2 = - k [x\ - 1) x2 - Xi

System (1.2.2), if we consider components of a vector


x , and X\, X2, . . . , Xn as components of a vector X , can be written in the
compact form

x = X ( x ; t) (1.2.7)

In many problems the variable t does not appear explicitly in (1.2.7). In


this case, the system becomes

x = X(x) (1.2.8)
x2

*1

Fig. 1.1 - Integral curve. Fig. 1.2 - Trajectory.

A system of this type is called autonomous. For example, the system


deduced from the van der Pol equation is autonomous. A system of the type
(1.2.7) is non-autonomous.
Once the solution x\ — fi(t) , x^ — ji{fy , . . . , xn = fn{t) have been deter-
mined a curve, called the integral curve, in the space E^1 can be associated
with them. The projection of this curve in the sub-space E" of the x coordi-
nates is defined as the trajectory or simply the motion, and the space E" is
the space of the phases.
Let us now consider the motion defined by the autonomous system (1.2.8)
to which we refer from now on, and assume that for x — a with a constant,
X ( a ) — 0. If we replace x by a in (1.2.8) we can see t h a t the system is
satisfied, and consequently x = a is a solution to the system. From a physical
point of view this means that if the system is initially in a then it remains in
this position, and therefore a is a configuration of equilibrium. The point a
is defined as the critical point or equilibrium point. By introducing the new
coordinates x* = x — a and rewriting system (1.2.8), again substituting x* by
x , it is seen t h a t the equilibrium point coincides with the origin. Therefore in
the study of the stability of a configuration, we can always refer to the study
of the stability of the origin.

1.3 SIMPLE T Y P E S OF EQUILIBRIUM P O I N T S


Let us consider, as a particular case of system (1.2.8), the system [7] of linear
equations with constant and real coefficients:

X\ — a n ^1 + #12 %2
(1.3.1)
x<i — a>i\ X\ + a22 ^2

which allows an equilibrium point at the origin X\ — x2 = 0. Let us exa-


mine the type of trajectory in the neighbourhood of the origin associated
4

with (1.3.1) by solving the system. By choosing two solutions of the type
xi — ax ext and x2 = a2 eXt we determine A by solving the system

( a n - A) a i + au a2 = 0
(1.3.2)
a a a a
21 l + ( 22 — ^) 2 ~ 0

This allows non-zero solutions if

A2 - ( a n + a 22 ) A + ( a n a 22 - a i 2 a 2 i ) = 0 (1.3.3)

Equation (1.3.3) is called the characteristic equation. The coefficients a i and


a2 are obtained, to within a constant factor, from (1.3.2), substituting in
these equations each of the solutions to (1.3.3). If the matrix of coefficients
has determinant det[a t J ] ^ 0 then the characteristic At (z = 1,2) roots are
different from zero.
The following cases can occur.

(a) The roots of the characteristic equation Ai and A2 are real and distinct.
The general solution to system (1.3.1) has the form

xi = claleXlt + c2/31eX2t
(1.3.4)
Xlt
x2 = Cla2e + c2(32eX2t

where a, and /?, (t' = 1,2) are constants which are determined by equation
(1.3.2) in correspondence with A = Ai and A = A2 respectively, and
c\, c2 are arbitrary constants determined from the initial conditions. It
is necessary to distinguish the following sub-cases.

( a x ) Ax < 0 and A2 < 0.


The equilibrium point is asymptotically stable. In fact, if for t = t0 a
point is found within a neighbourhood e of the origin of coordinates,
for a sufficiently large value of t it then passes into an arbitrarily
close proximity 6 to the origin, and with t —► oo it tends towards
this point. In Fig. 1.3 the trajectories near to the equilibrium point
known as the stable node are indicated. The arrows indicate the
direction of motion on the trajectories when t is increased.
( a 2 ) Ai > 0 and A2 > 0.
This case is transformed into the former if — t is substituted for t. So
the trajectories have the same shape as before, except for a reversal
of the direction of motion, which in this case moves away from the
origin. An equilibrium point of this type is called an unstable node
(Fig. 1.4).
5

Fig. 1.3 - Stable node. Fig. 1.4 - Unstable node.


*2

Fig. 1.5 - Saddle point.

( a 3 ) Ax > 0 and A2 < 0.


The equilibrium point is unstable because the components of the
motion Xi = CiCtieXlt, x 2 = Cia2eXlt tend to move the generic point
situated within a sufficiently close neighbourhood e of the origin
away from it. The trajectories are indicated in Fig. 1.5. An equilib-
rium point of this type is called a saddle point.

(b) The roots of the characteristic equation are complex: Ai)2 = p ± i q.


The general solution to system (1.3.1) can be put in the form
X\ — ept (ci cos qt + c2 sin^f^)
(1.3.5)
x2 — ep (c[ cos q t + c^ sin q t)

where cx and c 2 are arbitrary constants and c\, c 2 are given by linear
combinations of cx and c 2 . The following cases can occur.
( M Ai>2 =p±iq, p < 0.
In equations (1.3.5) the factor ept tends to zero with increase of t
whilst the second factor, which is periodic, remains bounded. The
trajectories are represented in Fig. 1.6 and are spirals which approach
the origin of coordinates asymptotically for t —> oo. The equilibrium
point is asymptotically stable and is called a stable focus. The fo-
cus is different from a node, in t h a t the tangent to the trajectories
does not tend to a determined limit when the equilibrium point is
approached.
( b 2 ) Aij2 =p±iq, p > 0.
This case is transformed into the former by substituting — t for t.
The trajectories are therefore not different from those in the previous
case, except that the motion occurs in the opposite direction with
increase of £, as indicated by the arrows (Fig. 1.7). The equilibrium
point is unstable and is called an unstable focus.
( b 3 ) Ai |2 = ±iq-
Due to the periodicity of solutions (1.3.5), the trajectories are closed
curves containing the equilibrium point known as the centre (Fig.
1.8). The centre is a stable equilibrium point as, once a certain
e > 0 has been fixed, it is possible to find a 6 > 0 such t h a t the closed
trajectories, the points of which belong initially to 6, are contained
in e for any value of t > t0.

The roots are multiple: Ai = A2.

(ci) Ax - A2 < 0.
The general solution to system (1.3.1) has the form
xi = ( c i a i + c j / M ) eAl*
(1.3.6)
x2 = (c1a2 + c2/32t) eXlt
As the factor eXlt rapidly tends to zero with increase of £, thus x x and
x 2 tend to zero when t —> oo. In consequence the equilibrium point
is asymptotically stable and is called degenerate stable node. This
node is an intermediate position between the node a i and the focus
b x because, for small variations in the coefficients an , a 12 , a 2 1 , a 22
of system (1.3.1), the multiple root can change into two real and
distinct roots (stable node ax), or into two complex conjugated roots
(stable focus b i ) . The trajectories are indicated in Fig. 1.9(a). If in
(1.3.6) we have /?i = /32 = 0, then the motion is still asymptotically
stable and the trajectories are those indicated in Fig. 1.9(b).
Fig. 1.6 - Stable focus. Fig. 1.7 - Unstable focus.

Fig. 1.8 - Centre.

a) b)
Fig. 1.9 - Degenerate stable node.
a) b)

Fig. 1.10 - Degenerate unstable node.

(c 2 ) Ax = A2 > 0.
The substitution of t by — t leads us back to the former case. The
trajectories are of the type shown in Fig. 1.9(a) (/?x, f52 ^ 0) or
1.9(b) (/?! = /?2 = 0), but the motion diverges. The equilibrium
point is called the degenerate unstable node (Fig. 1.10(a) and Fig.
1.10(b)).

With this, all possible cases which can occur when det[a tJ ] ^ 0 have now
been examined.
Note. If det[a,-j] = 0, then one or both roots of the characteristic equation
vanish. Let us first suppose that we have Ai = 0 and A2 ^ 0. In this case the
general solution to system (1.3.1) has the form

xi = ci a i + c2 /?i eX2t
(1-3-7)
2
x2 = ci a2 + c2 /32 e
By eliminating t we obtain the family of parallel lines (3\(x2 — C\ a2) =
/32(xi — ci a i ) . If A2 < 0, then when t —> oo the points on each trajectory tend
to the straight line X\jx2 — ai/a2, and the equilibrium point X\ — x2 — 0 is
stable (Fig. 1.11). If A2 > 0 then the trajectories are disposed in the same way,
but the motion of the points is in the opposite direction and the equilibrium
point is unstable.
Let us now suppose that Ax = A2 = 0. The general solution to system
(1.3.1) has the form
Ci+ C2t
(1.3.8)
*2 c\ + c\t
9

♦ x2

il

Fig. 1.11 - Stable equilibrium point.

where c\ and c\ are linear combinations of the arbitrary constants C\ and c 2 .


The equilibrium point x\ — x2 — 0 is unstable and the trajectories are
represented by the straight lines (x x — Ci)/(x2 — c{) = c2jc\.

We can conclude the present examination of equilibrium points by affirm-


ing t h a t if both the roots of the characteristic equation have a real negative
part (cases ( a i ) , ( b i ) and (ci)) then the equilibrium point is asymptotically
stable. If at least one root of the characteristic equation has a real positive
part (cases ( a 2 ) , ( a 3 ) , (b 2 ) and (c 2 )) then the equilibrium point is unsta-
ble. The same conclusions are valid for the homogeneous linear system of n
equations with constant coefficients

Xi = dij Xj (1.3.9)

where, following the Einstein convention, a repeated index indicates summa-


tion with respect to the index.
Example. Let us consider the equation of the free oscillations of a system of
one degree of freedom x + 2bx -f a2x — 0. This equation is equivalent to the
system

Xi = X2
(1.3.10)
i2 = — a2 Xi — 2 b x2

The characteristic equation associated with the system, A2 + 2 6\-\-a 2 — 0,


has the solutions Ai)2 = — b ± \/b2 — a2. We consider the following cases.

(1) 6 = 0, that is, there is no medium resistance. In this case the motion is
periodic and the equilibrium point is a centre (stable equilibrium).
10

(2) b2 — a2 < 0 , b > 0. The equilibrium point is a stable focus. The oscilla-
tions are damped.

(3) b2 — a2 > 0 , b > 0. The equilibrium point is a stable node. All the
solutions are damped and do not oscillate.

(4) 6 < 0 , b2 — a2 < 0 (case of small negative damping). The equilibrium


point is an unstable focus.

(5) b < 0 , b2 — a2 > 0 (case of large negative damping). The equilibrium


point is an unstable node.

1.4 EQUILIBRIUM P O I N T S OF N O N - L I N E A R SYS-


TEMS
Let us consider the autonomous system (1.2.8)

x = X(x) (1.4.1)

in which X ( x ) is a non-linear function of x. Let the origin x = 0 be an


equilibrium point, that is X ( 0 ) = 0. The type of stability of the origin
is determined by the pattern of the trajectories in the phases space in its
neighbourhood. With the same terminology as introduced for the linear case,
we can proceed to the following definition: if for t —► -f-oo all the trajectories
tend to 0, assuming the same tangent at this point, then the origin is a
stable node; if the trajectories tend to 0 for t —> — oo, then the point is an
unstable node. If the curves tend to 0 from different directions, some for
t —* +oo and others for t —► — oo, then the equilibrium point is a saddle point
and the equilibrium is unstable. If the solutions form spirals tending to the
origin for t —» -f-oo or t —> — oo, then this point is a focus, stable or unstable
respectively. Finally, if the trajectories do not tend to a point but form closed
curves around the origin, however near to it, then the origin is a centre and
the equilibrium is stable.
The non-linear analysis of the stability of the equilibrium point is very
much more complex than in the linear case. In fact, as the solutions to
(1.4.1) are no longer of the type ex\ the simple eigenvalues analysis presented
in the previous Section does not apply. Furthermore, as this Section is aimed
at examining the solution in the neighbourhood of the origin, it is possible
to think about replacing system (1.4.1) with a linear system obtained by
expressing the functions X ( x ) in MacLaurin series truncated at the first term.
The operation is justified in a heuristic way by the fact that x remains small
during motion, and is suggested by the need to know the qualitative rather
than the quantitative aspects of the solution.
11

In Section 1.7 the limits of validity for the substitution of the non-linear
system by the linearised system will be discussed, with some theorems indicat-
ing in which cases the results obtained for the linear system can be extended
to the original problem. It will, in fact, be demonstrated t h a t analysis using
a linear system can fail in some important cases, in the sense t h a t it cannot
provide an answer on the stability of the equilibrium point. In such cases it
is necessary to solve the non-linear problem.
The integration of the non-linear system can be very difficult, if not impos-
sible, and so in general it is necessary to try to obtain information on stability
using intermediate integrals, without having to perform the complete integra-
tion of the system. This method is particularly advantageous for the study
of conservative systems, where the law of energy conservation constitutes an
intermediate integral. The following example helps to clarify the method.
Let us consider the second-order differential equation [6]

£ + /(x) = 0 (1.4.2)

which governs the motion of a material point of mass m = 1 subjected to a


positional force F = — f(x). The equation can be equivalently written as a
system of two differential equations of the first order:

&i = x2
(1.4.3)
X2 = -f{xi)

where X\ — x identifies the position of the mass and x2 = x its velocity. By


dividing the two equations member by member we obtain

dx\ x<i
(1.4.4)
dx2 f{xl)

from which, by integrating,

\A +I f{xi)dx1 = H (1.4.5)

where H is a constant which depends on the initial conditions of motion.


Once the relation

P(xi) = I f{x1)dx1 (1.4.6)

has been established, the first integral (1.4.5) can be written as

l
-x\ + P{x1) = H (1.4.7)
12

Fig. 1.12 - Total energy and trajectories in a mechanical system.

The law of energy conservation can easily be recognised in this, as x\/2 is


the kinetic energy, P(xi) the potential energy and H the total energy of the
system, depending on the initial position and velocity.
Geometrically, the first integral represents a surface in the space (xi, £2, H)
(Fig. 1.12). By choosing a discrete number of values of H, H = Hi
(i — l , 2 , . . . , n ) , we obtain curves which, projected on the plane ( x i , x 2 ) ,
represent the required phases. In the example in the diagram, the convex
form of the energy gives rise to closed phases around the equilibrium point
which corresponds to the minimum of the surface. The origin is therefore a
centre.
An effective representation of motion can alternatively be obtained by the
diagram of the total potential energy alone. Figure 1.13 shows at the top
the case of a function P(x\) presenting three stationary points, one minimum
and two maxima. Each horizontal line Hi is associated with a motion which
corresponds to a given energy level. The difference in ordinates between a
line H — constant and the curve P[x\) represents (for every X\ position) the
kinetic energy which in that motion belongs to the material point. As the
kinetic energy is positive definite, the motion can take place only in those
regions in which P[x\) < H. In particular the points in which we have
P{xi) = H are points of inversion of motion as the kinetic energy vanishes.
Once a discrete set of values of H (Hi, H2, . . . , H6 in the figure) has been
chosen, it is possible to trace the corresponding trajectories shown in the
13

Fig. 1.13 - Total potential energy and trajectories.

lower part of the figure. The three stationary points of P(x\) are seen to
correspond to the equilibrium points which are either saddle points (5) or
centre points (C), depending on whether the potential energy has a maxi-
m u m or a minimum, and therefore are unstable or stable equilibrium points,
respectively.
For levels of total energy equal to a relative maximum (Hz , # 5 , in the
example), it is possible to have trajectories which pass through saddle points.
Such curves are known as separatrices and are marked in the figure. If the
state of equilibrium in a saddle point is perturbed and the representative
point on the plane X\ x2 is at the time t — 0 on a branch of the separatrix
directed towards the equilibrium point, then the system tends to return, in
an infinite time, to the state of equilibrium.
If the perturbed state is situated on a branch of a separatrix which moves
away from the saddle point, the system can still return, after an infinite
time, to the unperturbed position, as happens for the separatrix T 3 . This
phenomenon is not observed in a linearised analysis. If the perturbed state
belongs to an open curve, the representative point tends to move away in-
definitely from the equilibrium point; if it belongs to a closed curve then it
oscillates around the centre, periodically passing close to the saddle point.
This phenomenon also does not emerge in a linear analysis, where the trajec-
tories around the saddle point are represented by open curves (Fig. 1.5).
From the qualitative analysis made, based exclusively on knowledge of the
integral of energy, it is possible to understand how the character of the stabil-
14

ity of an equilibrium point depends only on the form of the potential energy
in the neighbourhood of the point itself. This concept will be clarified in the
following Sections, where systematic use of a technique which leaves aside the
dynamic aspects of the problem will be made. However, a dynamic study can
be equally useful for an understanding of some mechanical phenomena, and
as such will be examined later. An application of linearised analysis will be
shown in Section 2.8, and one of non-linear analysis in Section 2.11.

1.5 STABILITY OF EQUILIBRIUM ACCORDING


TO L I A P U N O V
Let us consider the autonomous system (1.2.8) x = X ( x ) and assume t h a t
the point x = a is an equilibrium point, that is, X ( a ) = 0 . By a change in
coordinates it is always possible to have the equilibrium point coinciding with
the origin, t h a t is, X ( 0 ) = 0. Let us indicate by S(R) the spherical region
defined by ||x|| < R and by H(R) the spherical surface ||x|| = R. The norm
of x , ||x||, can be defined in various ways and will later be given in explicit
form. Let us suppose that in a certain open spherical region Q: ||x|| < D
the conditions required by the theorem of existence and uniqueness of the
solution to the system of equations (1.2.8) are satisfied. We can then say [3]
t h a t the origin represents a configuration of equilibrium which is:

s t a b l e if for every R < D there exists an r < R such t h a t a trajectory £ ,


with its origin in a point Xo of the spherical region 5 ( r ) , remains in the
spherical region S{R) when t increases; that is to say, a trajectory with
origin in S{r) never reaches the boundary H(R) of S(R)]

a s y m p t o t i c a l l y s t a b l e if it is stable and, besides, each trajectory Q with


origin in S(r) tends to the origin for t —» oo; or

u n s t a b l e if for a fixed R < D and for any r, however small, there always
exists a point x in S'(r), such t h a t a Q trajectory which originates in x
reaches the boundary H{R).

Figure 1.14 shows a geometrical representation of the cases discussed.


It is possible to give a second definition of the stability of equilibrium
which is particularly convenient for mechanical systems, as will be seen in
the following. Let us consider an initial perturbation at the instant t = 0
defined by x(0) = x 0 and specify the concept of the norm ||x|| = p as the
distance between the configuration of equilibrium x = 0 and the configuration
at the instant t defined by

P*(x) = £ > ( * , • ) ' (1-5.1)


«= 1
15

Fig. 1.14 - Types of equilibrium.

We then say t h a t
the configuration of equilibrium x = 0 is stable if, and only if, for each positive
e there exists a second positive number 6 which depends on e with the property
that
p(x)<e (1.5.2)
for any t > 0 and for any motion whose initial conditions satisfy
Po = p{*o)<6{e) (1.5.3)
Note t h a t the stability of equilibrium is a dynamic problem, whilst equilibrium
is a static problem.

1.6 T H E O R E M S ON T H E STABILITY OF EQUILIB-


RIUM
It was shown in Section 1.4 (Fig. 1.12) t h a t if, in the vicinity of an equilibrium
configuration of a physical system the energy of the system is increasing, then
the equilibrium is stable. Liapunov's theorems cited in the following represent
a generalisation of this idea, and refer to systems of equations of the type
(1.2.8).
16

A very important role is played by scalar functions V ( x ) , which are said


to be positive definite if they satisfy the following conditions:

(a) V'(x) together with its first partial derivatives is continuous in a certain
open region ft around the origin;

( b ) V{0) = 0 ; and

(c) the function V(x) has an isolated minimum at the origin.

If, in addition, dV/dt is non-positive in ft along the trajectories of motion of


system (1.2.8), that is

V = Vti ii = V}i Xi (xx, x2 , . . . , xn) < 0 (1.6.1)

the function ^ ( x ) is called a Liapunov function. It is assumed that X{(x)


and when necessary X t (x) are continuous. It follows that V is a continuous
function in ft. In equation (1.6.1), a subscript preceded by a comma indicates
differentiation with respect to the corresponding variable. Let us examine the
quadratic form

V(x) = CLij X{ Xj (a t y = CLji) (1.6.2)

The necessary and sufficient condition for V^x) to be positive definite is


t h a t the successive principal minors of the symmetrical matrix A have a
positive determinant (Sylvester).
Generally, the function V(x) can be represented as a series of powers in x
in the neighbourhood of the origin

^(x)=Vp(x)+Vp+1(x) + ... (1.6.3)

where V p (x) is a homogeneous polynomial in x of degree p. A necessary


condition for V(x) to be positive definite is that the lowest degree p of the
series of powers (1.6.3) is an even number. Such a condition, however, is not
sufficient. In fact, for p = 2 the function

Vp{*) = xl-x\ (1.6.4)

is positive definite for Xi — 0 and negative definite for x2 = 0 . If p is an odd


number then ^ ( x ) can never be a Liapunov function. Let us now pass to the
enunciation of some basic theorems.
T h e o r e m o n s t a b i l i t y ( L i a p u n o v ) . / / in a certain neighbourhood ft of the
origin there exists a Liapunov function V^(x), then the origin is stable.
T h e o r e m o n a s y m p t o t i c s t a b i l i t y ( L i a p u n o v ) . / / there exists in ft a
Liapunov function V(x) such that V < 0, then the stability is asymptotic.
17

Fig. 1.15 - Liapunov function. Fig. 1.16 - Plane of the phases.

Let us demonstrate the first theorem using the geometrical interpretation


of a positive definite function V^x), represented in Fig. 1.15 for x = { x i , x 2 } .
In Fig. 1.16 the curves V^x) = k are represented by the solid line whilst the
spheres H(R) and H(r) are indicated by dashes.
Given then an R < D (Fig. 1.14) and H(R), we can find a constant k
such t h a t the curve C defined by V(x) = k is contained in H(R) and an
r > 0 such t h a t H(r) is contained in C. Let us now consider a trajectory
Q with initial point x 0 belonging to S{r), the interior of H(r). In x 0 it is
^ ( X Q ) < k. Furthermore, as V^(x) does not increase along the trajectories,
Q never reaches C and so will never reach H(R). Therefore, each trajectory
with origin in S(r) must remain in S{R) and this implies stability.
The demonstration of the second theorem follows from the previous demon-
stration, as V < 0 implies that trajectory Q which starts in x 0 G S(r) tends
to the origin as t —> oo, and this implies asymptotic stability.

T h e o r e m o n i n s t a b i l i t y ( L i a p u n o v ) . Let V^(x), with V(0) = 0 have


continuous first partials in fi. Let V be positive definite and let V(x) be
able to assume positive values arbitrarily near the origin. Then the origin is
unstable.

The demonstration is omitted here. The condition V > 0 implies that the
trajectory Q which starts from x 0 G S(r) where ^ ( x 0 ) > 0 reaches C and
therefore H(R), and so we have instability.
Liapunov's theorem on instability has the disadvantage of requiring the
existence of a whole 17 region around the origin, where the conditions required
by the function V^(x) are satisfied. The following theorem on instability by
Chetayev is less restrictive in this sense.
18

T h e o r e m o n instability ( C h e t a y e v ) . Let ft be a neighbourhood of the


origin. If a function V(x) and a region ft\ in ft with the following properties
are given:

(a) V(x) and V(x) are positive in fi1;"

(b) V(x) has continuous first partials in ft\;

(c) at the boundary points of fix inside ft, V(x) = 0;

(d) the origin is a boundary point of fti;

then under such conditions the origin is unstable.


It is not difficult to see that any trajectory Q starting from a point situated
in fti must leave fi, since it cannot cross the boundary of ft\ inside fi. As
the origin is situated on the boundary of Hi, we can choose some points
arbitrarily close to the origin from which trajectories Q which leave ft start,
and this implies instability (Fig. 1.17).

V=k
Fig. 1.17 - Plane of phases.

Example 1. Analyse the stability of the trivial solution to the system

x — — y — x3
(1.6.5)
y = x - y*

The function V(x,y) = x2 + y 2 satisfies the conditions of Liapunov's the-


orem on asymptotic stability. In fact

(1) V{x,y)>0, 7(0,0)-0 (1.6.6)

(2) V = 2x{-y - xs) + 2y{x - t/3) = - 2(x 4 + t/4) < 0 (1.6.7)

At a point which is arbitrarily near the origin we have V < 0, and so the
origin is asymptotically stable.
19

Example 2. Analyse the stability of the equilibrium point x = y = 0 of the


system of equations

i = V3 + x5
y= x +y
The function V(x,y) = x 4 — y4 satisfies the conditions of Chetayev's the-
orem

(1) V > 0 for |x| > \y\ (1.6.9)

(2) V = 4x 3 (t/ 3 + x 5 ) - 4y 3 (x 3 + t/5) = 4(x 8 - y 8 ) > 0 for |x| > \y\ (1.6.10)

In the neighbourhood of the origin and for |x| > \y\ we have V > 0 , V > 0;
thus the equilibrium point x = y — 0 is unstable.
Example $. Analyse the stability of the trivial solution xt- = 0 (z = 1 , 2 , . . . , n)
of the system of equations
dxj _ d u ( x i , x 2 , . . . , x n )
eft 3xt-
if the function u(x 1? x 2 , . . . , x n ) has a maximum at the origin of coordinates.
Let us take as a function of Liapunov the difference

Vr(x1,x2,...,xn) = u ( 0 , 0 , . . . , 0 ) - u(xi,x2, . . . , x n ) (1.6.12)

which obviously vanishes for xt- = 0 and has a minimum at the origin of
coordinates. For the derivative with respect to time we have
- _ du dxi du du
dxi dt dxi dx{ ~
In this way the conditions of the second theorem of Liapunov are satisfied,
and therefore the trivial solution is asymptotically stable.

1.7 ANALYSIS OF T H E STABILITY OF EQUILIB-


R I U M B Y LINEAR A P P R O X I M A T I O N
Let us consider the autonomous system (1.2.8), x = X ( x ) with X ( 0 ) = 0. If
the functions X, are derivable in a neighbourhood of the origin of coordinates,
then for the second member of system (1.2.8) can be substituted a series
expansion [3]

ii = a{j Xj + Ri ( x i , . . . , x n ) (1-7.1)

where atJ = ( d X t / d x y ) x = o and ||R|| is small with respect to ||x||, that is to


20

say, ||R||/||x|| tends to zero with ||x||. This fact can be expressed by

||R(x)||=o(||x||) (1.7.2)

Instead of investigating the stability of the equilibrium point x = 0 of


system (1.7.1), the stability of the same point of the linear system

Xi = a,ij Xj (1.7.3)

is analysed. System (1.7.3) is called a system of equations of linear approx-


imation with respect to system (1.7.1). The conditions of stability of this
system were examined by Liapunov and successively generalised by Malkin,
Chetayev and others.
The analysis of stability of the system of equations of linear approximation
is a much simpler problem than the study of the original system. In this regard
there are two very useful practical theorems.
Let us suppose that the characteristic roots A, of the matrix of coefficients
[atJ] are real and distinct, and let us apply to system (1.7.1) the linear trans-
formation of coordinates y — P x with P non-singular. As (d/dt) P x = P x ,
by making use of (1.7.1) we can write

y = Px = PAP_1y + PR (1.7.4)

We now choose the matrix P in such a way that P A P - 1 = diag(Ai, A2 ,


. . . , An) and take P R = R*. System (1.7.4) is rewritten as

y = diag(Ax, A 2 , . . . , An) y + R* (1.7.5)

whence it can easily be shown that

lia*(y)ll=o(||y||) (1.7.6)
The transformation of system (1.7.1) into system (1.7.5) is useful for demon-
stration of the following theorem.

T h e o r e m 1.7.1 A sufficient condition for the origin of the non-linear system


(1.7.5.) to be asymptotically stable is that the characteristic roots are all
negative. If there is a single positive characteristic root, then the origin is
unstable.

Two cases can be distinguished in the demonstration.

(a) The Xh roots are all negative. The following function of Liapunov is
assumed

V = yl + y\ + • • • + y2n (1.7.7)
21

from which

V = 2(AX y\ + A2 y\ + ■ ■ ■ + An y») + r(y) (1.7.8)

where r is small with respect to the terms in parentheses. In a suffi-


ciently small ft region around the origin V and — V are positive definite
functions, and so the origin is asymptotically stable.

(b) Some of the Xh roots, for example Ai, A 2 , . . . , Ap (p < n) are positive
and the rest negative. This time we take

V = yl + --- + y l - y l + 1 -,...,-yl (1.7.9)

from which

V = 2(AX y\ + • • • + Ap y2p - A p+1 y * + 1 , . . . , - A„ y2n) + n ( y ) (1.7.10)

where, as before, the r\ term is small with respect to t h a t in the paren-


theses. At some points which are arbitrarily near to the origin (those for
which t/ p+ i = • • • = yn = 0), V is positive. As for V, since Ai, A2 , . . . ,
Ap > 0, it is positive definite in t h a t fix region in ft where V is positive
definite and therefore, according to the Chetayev theorem, the origin is
unstable.

Let us now suppose that some of the A^ are complex. For example,
let A x , . . . , Ap be real and A p + 1 , A p + 1 , . . . , A p + m , A p + m be complex with
p + 2m = n. If A x , . . . , Ap are negative and A p + ^, \p+h have a real nega-
tive part, then we can choose the following Liapunov function

V = y\ + • • • + y\ + t/p+i y p + 1 + • • ■ + tfp+m J7P+m (1.7.11)

and everything proceeds as in case ( a ) , with the origin asymptotically stable.


If, on the other hand, some of the Ai, . . . , Ap are positive or some of the
Ap+h have a real positive part, then we proceed exactly as in case ( b ) and
we find t h a t the origin is unstable. We can therefore enunciate the following
theorem.

T h e o r e m 1.7.2 A sufficient condition for the origin of the non-linear system


(1.7.5.) to be asymptotically stable is that the characteristic roots all have
negative real parts. If there is a characteristic root with positive real part,
then the origin is unstable.

Note that in virtue of the two theorems presented in this Section, the
asymptotically stable equilibrium points of the linear system (1.3.1) (cases
(ax), (b x )) and the unstable equilibrium points (cases ( a 2 ) , ( a 3 ) , ( b 2 ) )
22

remain as such when we pass from system (1.3.1) to a non-linear system,


obtained from (1.3.1) with the addition of the non-linear terms Ri and R2
which satisfy (1.7.2). If a certain number of characteristic roots vanish or have
a purely imaginary value, analysis of the stability of linear approximation is
generally not possible, as the non-linear terms R^ influence the stability of
the system.
Example 1. Analyse the stability of the equilibrium point x = y = 0 of the
system

x = 2x + 8 siny
(1.7.12)
y = 2 — ex — 3 y — cos y

By expanding sin y, cos y and ex in a Taylor series we can write the system
in the form

x = 2x + 8y + Rx
(1.7.13)
y = - x-3y + R2

where Rx = — 4 y 3 / 3 + • • • and R2 = (y 2 — x 2 )/2 + • • •. As the limitations


(1.7.2) are satisfied, we can analyse the stability of the equilibrium point of
the linear system

x = 2x + 8y
(1.7.14)
y = - x- 3y

The roots of the characteristic equation A 2 +A+2 = 0 are Ai 2 = —l/2±i J 7/4;


therefore the equilibrium point x — y — 0 of systems (1.7.12) and (1.7.14) is
asymptotically stable.
Example 2. Let us consider the system

x = y-x/(x,y)
(1.7.15)
V = -x-yf(x,y)

and suppose that the non-linear terms xf and y f satisfy condition (1.7.2)
and also that / ( 0 , 0 ) = 0. The characteristic roots of the linear system are
A1>2 = ± t and therefore the analysis of the stability of the equilibrium point
x = y = 0 of system (1.7.15) depends on non-linear terms. In fact, let us
choose the Liapunov function V = (x 2 + y 2 ) / 2 , from which

V = -(xi + y2)f(x,y) (1.7.16)


23

Three cases can occur:

/ > 0 in an arbitrarily close vicinity of the origin, the origin is stable;

/ < 0 in an arbitrarily close vicinity of the origin, the origin is unstable;

/ is positive definite within a certain vicinity of the origin, the origin is


asymptotically stable.

Note that the system of equations studied in Example 1, Section 1.6, is


of the same type as for system (1.7.15). In fact, as the characteristic roots
are Ai|2 = ± i , the stability of the equilibrium point has been decided by
non-linear terms.

1.8 C R I T E R I O N OF N E G A T I V E REAL P A R T S OF
ALL T H E R O O T S OF A POLYNOMIAL
In the previous Section the problem of the stability of the trivial solution to
a wide class of systems of differential equations was reduced to an analysis of
the signs of the real parts of the roots of the characteristic equation.
If the characteristic equation is a polynomial of high degree, then its solu-
tion is very difficult, and so the methods which allow us to determine whether
the roots do or do not have real negative parts have great importance.
H u r w i t z ' s t h e o r e m . The necessary and sufficient condition for the real parts
of all the roots of the polynomial

p{z) = zn + ax zn~l + • • • + an_x z + an (1.8.1)

with real coefficients to be negative is that each principal minor of the Hurwitz
matrix

( ax 1 0 0 ... 0 \
o>3 a>2 a>\ 1 ••• 0
a5 a4 a3 a2 ... 0 (1.8.2)

V 0 0 0 0 ... an J

is positive.
Example. Let us consider the polynomial

p(z) — zA + axzz + a2z2 + azz + a 4 (1.8.3)


24

The Hurwitz matrix is


( CLx 1 0 0 ^
03 a-i ai 1
(1.8.4)
0 a4 a3 o2
I0 0 0 04 J
The Hurwitz conditions reduce to

ai > 0 , ai a 2 — a 3 > 0 , (ai a 2 — a 3 ) a 3 — a 4 aj > 0 , a 4 > 0 (1.8.5)


25

REFERENCES
[1] H. Poincare: "Sur I'equilibre d'une masse fiuide animee d'un mouvement
de rotation", Acta Math., 7, 1885, 259.

[2] A.M. Liapunov: Probleme general de la stabilite du mouvement (in Rus-


sian), Karkov, 1892; French translation in Ann. Fac. Sci. Univ. Toulose,
9, 1907; English translation: Stability of motion, Academic Press, New
York, 1966.

[3] J. La Salle, S. Lefschetz: Stability by Liapunov's direct method with ap-


plications, Academic Press, New York, 1961.

[4] W. Hahn: Stability of motion, Springer-Verlag, Berlin, 1967.

[5] L. Pontriaguine: Equations differentielles ordinaires, MIR, Moscow, 1969.

[6] H.H.E. Leipholz: Stability theory, Academic Press, New York, 1970.

[7] L. Elsgoltz: Ecuaciones diferenciales y calculo variacional, MIR, Moscow,


1971,

[8] C.L. Dym: Stability theory and its applications to structural mechanics,
Noordhoff, Leyden, 1974.
27

Chapter 2

THE STABILITY OF EQUILIBRIUM A N D


POST-BUCKLING BEHAVIOUR OF
DISCRETE MECHANICAL SYSTEMS

2.1 INTRODUCTION
The problems presented in this Chapter are in many respects only a particular
case of the same problems illustrated in the previous one. However, it seems
best to treat them separately since in some cases they present a different
feature which is discussed here in detail.

2.2 L A G R A N G E A N D HAMILTON E Q U A T I O N S OF
MOTION
Lagrange has demonstrated [2] t h a t the differential equations of motion of a
system of n degrees of freedom can be written immediately if we know the
kinetic potential or Lagrange function defined by

L = K-P (2.2.1)

where K is the kinetic energy and P is the potential energy of the forces
acting on the system.
Let <7i, q<i, . . . , qn be generalised coordinates in which it is possible to
define the configuration of the discrete system, and suppose t h a t the gt
(t = l , 2 , . . . , n ) are chosen in such a way t h a t in the position of equilib-
rium we have gt = 0. Indicating the position vector by r = r ( q ; £), the kinetic
energy of a system of N particles is expressed by the relation

or

K =
o a«J & *" + a« ^ + a ° (2.2.3)
28

where the coefficients a,, , a, and a0 are functions of t and g,- expressed by
224
-" - l - & £ <--'
225
" - £- £t <--'
and summation convention with respect to repeated indices has been adopted.
From now on, unless otherwise specified, all the Latin indices are understood
to vary from 1 t o n .
The formula (2.2.3) shows that the kinetic energy of a system can be
written in the form
K = K2+K1 + K0 (2.2.7)
where K2 is a positive definite quadratic form in q, Kx is a linear form in
q and K0 is independent of q. In the case of a scleronomous system, which
we shall refer to from now on, the time t does not appear explicitly in the
relationship between r and q and therefore dr^jdt = 0(fc = l , 2 , . . . , N). The
kinetic energy (2.2.7) is reduced to

K = K2 = - a{j (q) g, q {a{j = aj{) (2.2.8)

where [a,y(q)] represents the matrix of inertia.


In linearised theory the coefficients at-y(q) are replaced by a t J ( 0 ) . Unless
otherwise stated, we refer from now on to scleronomous non-linear systems.
Let us now introduce the potential energy P — P ( q ) and allow generalised
forces to act upon the system expressed as
Qi = ~ P,i (2.2.9)
where P t = dP/dgt. Forces of the type (2.2.9) satisfy the law of energy con-
servation, and are therefore called conservative. As well as the forces (2.2.9),
we consider some forces which are not derivable from P. The generalised
forces acting upon the system are therefore
Qi = -P,i + Qi ( q , q ) (2.2.10)
where the non-conservative forces Q, depend, in general, on the position g,
and the velocities q t . The Lagrange equations of motion are finally obtained
in the form

(2 2 n)
JtM-^ = Qi^q) --
29

Equations (2.2.11) are of the type

q = f(q,q) (2.2.12)

which can always be reduced to the form

x = S (x) (2.2.13)

The kinetic potential L from which the equations (2.2.11) have been de-
duced depends on the variables q , q, which are called Lagrange variables.
Hamilton proposed to assume as basic variables the quantities q and p , where
p is the generalised linear momentum defined by

Pi =
M = aij (q) 9i (2 2 14)
''
The quantities q , p are called Hamilton variables. By simple steps it is
possible to express the kinetic energy as a function of q and p , arriving at
the expression

K = \lij(<l)PiPi (2-2.15)

The potential energy P(q) in terms of the new variables remains unchanged,
whilst the generalised forces (2.2.10) are written as

Qi = -P,i+Q,;(p,q) (2.2.16)

By introducing the Hamiltonian

H(V,q) = P(q)+K(p,q) +C (2.2.17)

with C as arbitrary constant, it is easily demonstrated that the following


equations hold:

. _ 3H
dpi
(2.2.18)
Pi = - -^ + Qi ( p , q)

Equations (2.2.18) constitute the Hamilton equations of motion and it is pos-


sible to use them as an alternative to (2.2.11), in order to study the stability
of equilibrium. These equations are of the type (2.2.13).
30

2.3 STABILITY OF EQUILIBRIUM ACCORDING


TO LIAPUNOV
Let us consider a system in the state of equilibrium q = 0 , q = 0 and suppose
that we apply at the instant t = 0 a perturbation characterised by

q(0)=qo, q(0)=q0 (2.3.1)

We now introduce a norm p which measures the distance between the state of
equilibrium and the qurrent state and endow p with the following properties:

(1) p(q,q)>0 for q^0,q^0


(2) p{qLl + q 2 , q x + q 2 ) < p(q 1? q x ) + p(q 2 , q 2 ) (triangle inequality) (2.3.2)
(3) p ( a q , a q ) = | a | p ( q , q) (a real)

D e f i n i t i o n 2 . 3 . 1 ( L i a p u n o v ) . The configuration of equilibrium q — 0 is


stable if, and only if, for every positive number e there exists a second positive
number 6 (depending on e) with the property

p[q(t),q(t)}<e (2.3.3)

for any t > 0 and for any motion with initial conditions which satisfy

Po = />(qo,qo)<<5(£) (2.3.4)

Expressions of p which are suitable for the solution to mechanical problems


l
are, for example

P = yQiQi + ?»?. (2.3.5)

p = max \qi\ + max |</t-| (2.3.6)

Note the analogy between (2.3.5) and (1.5.1).

2.4 LAGRANGE-DIRICHLET THEOREM


In this Section we demonstrate the Lagrange-Dirichlet theorem by following
the presentation furnished in [l]. We notice from (2.2.17) that, along the
motion,

• 8H . dH . ,
31

having made use of (2.2.18) where Q{ = 0 has been assumed. Relation (2.4.1)
shows that during motion the sum of the kinetic energy and the potential
energy remains constant. The theorem is therefore enunciated as follows.

T h e o r e m of L a g r a n g e - D i r i c h l e t . / / the potential energy P(q) of a con-


servative system is positive definite in the neighbourhood fi: ||q|| < D of an
equilibrium configuration, then the configuration of equilibrium is stable.

Let us assume that the origin q = 0 is a configuration of equilibrium and


consider the motion arising from the perturbation (2.3.1) impressed at the
instant t = 0. We choose the constant C, which appears in (2.2.17), so that
f f ( 0 , 0) = 0. Of the two terms P(q) and K(p, q) forming the Hamilto-
nian H(p, q), the kinetic energy is always positive definite. If the potential
energy has an isolated minimum in correspondence with the configuration of
equilibrium q = 0, then H(p, q) is also positive definite, and as H = 0 in
conservative systems the function H is then a Liapunov function. So in accor-
dance with the Liapunov theorem in Section 1.6, the position of equilibrium
is stable.

2.5 T H E O R E M S OF L I A P U N O V A N D CHETAYEV

The Lagrange-Dirichlet theorem on the stability of equilibrium demonstrated


in the previous Section does not give any information on the behaviour of a
mechanical system when the potential energy corresponding to a configuration
of equilibrium does not show a minimum. There are two theorems regarding
this, accredited to Liapunov and Chetayev, respectively, which are enunciated
here without proof [2].

T h e o r e m o n i n s t a b i l i t y ( L i a p u n o v ) . / / the potential energy P(q) of a


conservative system has an isolated maximum corresponding to a configura-
tion of equilibrium, then the configuration of equilibrium is unstable.

T h e o r e m o n i n s t a b i l i t y ( C h e t a y e v ) . If the potential energy P(q) of a


conservative system is a homogeneous function of the coordinates q and if,
corresponding to a configuration of equilibrium, P(q) does not have a mini-
mum, then the configuration of equilibrium is unstable.

Example. Let us suppose that the potential energy of a system is of the type
P(q) = Aqi q2 . . . qn and that q = 0 is a configuration of equilibrium. The
aim is to examine the type of equilibrium of such a configuration. According
to the Chetayev theorem, we can assert that the configuration of equilibrium
q = 0 of the system is unstable.
32

2.6 ANALYSIS OF THE STABILITY OF LINEAR


SYSTEMS
The linearised analysis of stability carried out in Section 1.7 is obviously valid
when the differential equations already discussed represent the equations of
motion of a mechanical system. It is necessary, however, to discuss this again
here in order to reach further conclusions of particular interest which can be
obtained for mechanical systems.
Let us consider the equation of motion of type (1.2.8) and assume that the
origin q = 0 represents a configuration of equilibrium defined by X ( 0 ) = 0.
Let us further suppose that the X(q) are continuous and derivable at the
origin, so that we can substitute the system

* = *ij 9j + Xr {qk) (2.6.1)

for system (1.2.8), having set atJ- = (dXi/dqj)q=o and having indicated by
X* high-order terms of the series expansion which satisfy (1.7.2).
The stability analysis of the equilibrium point q = 0 of the non-linear
system (2.6.1) can be replaced, in some cases, by the analysis of the same
equilibrium point of the linearised system <fc = a^qj. Having assumed a
solution to the linear system of the type q = aeXty the conclusion is (see
Section 1.7) that the origin is a stable equilibrium point of system (2.6.1) if
the roots A of the characteristic equation associated with the linear system
have a real negative part. If the real part of at least one of the characteristic
roots vanishes, then conclusions reached on the stability of the equilibrium
point based on the study of the linear system cannot be extended to the non-
linear system. Indeed, let us consider a mechanical system with one degree
of freedom whose potential energy is given [6] by

P{q) = \cq4 ( c > 0) (2.6.2)

The Lagrange equation of motion (2.2.11), assuming K = - q2 and that


Si

the non-conservative Q forces vanish, may be written as

q+ c q3 = 0 (2.6.3)

The linearised equation associated with (2.6.3), q = 0, admits the solution


q = a + bt which is unstable (see 1.3.8), whilst the non-linear system (2.6.3)
admits a stable equilibrium point at the origin, as the potential energy (2.6.2)
is positive definite (Lagrange-Dirichlet).
33

2.7 CRITERION OF STABILITY OF DISCRETE


SYSTEMS
The Lagrange-Dirichlet, Liapunov and Chetayev theorems shown in Sections
2.4 and 2.5 are important because they allow us to decide on the nature of
the equilibrium of a mechanical system subjected to conservative forces on
the basis of knowledge of the potential energy alone. The kinetic energy,
therefore, plays no part.
Let us now suppose that the potential energy P(<fc) is a sufficiently regular
function around the configuration of equilibrium qi = 0 and write in this
neighbourhood

P(ft) = Pti{0)qi + ^Ptij{0)qiqj + -Ptiik{0)qiqjqk

+ ^ Ptiju [0)qi Qj Qkqi + ...

qi qj qk +
=
2 Cij qi qj
~*~ 6 Cijh 24 Cijid qi qj qkqi
'" (2'7'1)
where [CtJ] is called the tangent stiffness matrix in the configuration of equi-
librium. Note that as the velocity and therefore the kinetic energy vanish at
q = 0, the Lagrange equations of motion (2.2.11) for conservative systems
give Pti(0) — 0 and so the series expansion of P(qi) starts with second-order
terms. For (ft sufficiently small, the potential energy is positive definite if
the stiffness matrix is positive definite [stable equilibrium). If [Cty] is either
indefinite or negative definite, the equilibrium is unstable. If [CtJ] is positive
semidefinite (that is, positive definite with a vanishing value corresponding
to a particular n-tuple of coordinates qi) then the higher order terms of the
series expansion of P(<fc) must be examined. This case is called the critical
case of equilibrium, and must be investigated in detail.
Following the presentation of Thompson and Hunt [7], we introduce a
non-singular transformation of coordinates

Qi = otij Uj (2.7.2)

with inverse transformation

m =faq, (2.7.3)
which reduces the quadratic form CtJ <fc qj to a diagonal form. In terms of the
new coordinates, (2.7.1) becomes a new function which we shall still indicate
by P:

1 n 1 1
P u
i i) = o S dii
( u *) 2 + a diik Ui U Uk +
i OA di^ Ui U Uk Ul
' (2-7-4)
34

where the coefficients d u (the repeated index does not indicate a sum), dijk ,
diju represent the derivatives of P(u t ) with respect to u calculated at the ori-
gin u = 0 which, according to (2.7.3), coincides with q = 0. The quadratic
form Cij qi q^ can be made diagonal in an infinite number of ways without
altering its invariant properties. Supposing that we have chosen one of the
transformations of diagonalisation (2.7.2), the number of positive, null and
negative d„ coefficients is independent of the chosen transformation. Obvi-
ously

det[dtf] = dn d22 • • • dnn (2.7.5)

The new coordinates are called principal coordinates and da coefficients of


stability. If a particular da is positive, we say that the equilibrium in the
configuration u = 0 is stable in relation to the corresponding principal in-
coordinate; if it is negative, we say that the equilibrium is unstable in re-
lation to the corresponding coordinate. If a particular coefficient vanishes,
we speak of critical equilibrium in relation to the associated principal coordi-
nate. The number of negative coefficients of stability is defined as the degree
of instability. Finally, the equilibrium is said to be stable, without further
specification, if all the da coefficients are positive, unstable if at least one of
the coefficients is negative, or critical if one of the coefficients vanishes and
the rest are positive.
We now correlate the definitions given here with those expressed at the
beginning of the Section linked to the nature of the quadratic form CtJ ^ ty,
which is not altered by the change in coordinates (2.7.2). To this end, let us
order the coefficients da in such a way that

dn<d22<...< dnn (2.7.6)

If all the coefficients are negative, or if da < 0 (i = l , 2 , . . . , m < n), then


the quadratic form Y£=i ^«»(ut)2 ls negative definite or indefinite, respectively,
and therefore the configuration of equilibrium u = 0 is unstable. If dn > 0
the quadratic form is positive definite, and therefore the equilibrium is stable.
If dn = 0 then the quadratic form if positive semidefinite and the equilibrium
is critical. In this case, we have from (2.7.5) and (2.7.2)

det[diy] = det[a t i ] 2 • det[C 0 ] = 0 (2.7.7)

from which it follows that, in correspondence with the critical condition of


equilibrium, the matrix of the coefficients of the quadratic form is singular.
The critical case of equilibrium obviously needs to be examined in detail, as
anticipated at the beginning of the Section, to verify whether the equilibrium
is stable or unstable. As the examination of the quadratic terms of the series
expansion of the energy (2.7.4) leads to an indecisive case, it will be necessary
35

to examine the successive terms of the series expansion to verify the nature
of the equilibrium.
Let us suppose that d n = 0 , d22 > 0 and assume

Ui = 0 , i > 2 (2.7.8)

The quadratic form X)"=i d t f(u,) 2 is positive definite for any n-tuple ut- whilst
it vanishes in the direction of displacement (2.7.8), that is, in the direction of
U\. Along Ux, (2.7.4) becomes

P(u{) = - d l u u\ + — dun u\ (2.7.9)


o z4
If dill 7^ 0, then the cubic term of (2.7.9) is dominant and the equilibrium in
the configuration u — 0 is unstable because, for any sign of d m , it is always
possible to choose a value of u\ so that d m u\ < 0. A necessary condition for
stability is therefore given by

dm = 0 (2.7.10)

A further necessary condition for stability is that

dim > 0 (2.7.11)

As we shall demonstrate, this condition is however not sufficient. Let us


consider all the possible curves through the point u = 0 and let us suppose
that they can be represented in parametric form by

Ui = Ui(s) (2.7.12)

where s is a parameter which it is not necessary to specify, and which we


suppose to vanish in correspondence with u = 0. If we allow that (2.7.12)
are sufficiently regular we can write

u{(s) =UiS + - Ui s2 + o(s2) (2.7.13)

where a dot represents differentiation with respect to s and the coefficients


Ui, iii are evaluated at 5 = 0. By performing a series expansion of the total
potential energy along the generic curve (2.7.12) or (2.7.13) starting from
s = 0, we have

P[Ui(s)} = p a + lpa* + l p s ' + ±Ps< (2.7.14)


36

where from (2.7.4) t h e explicit expressions of the coefficients are

P = 0 (2.7.15)

P = £ *,«<«< (2.7.16)
n
P = dijkUiUjiik +z Y , ^ ^ ^ (2.7.17)
n
P = dijkl Ui Uj iik ui + 6 dijk xii u y uk + 4 ] T da u\ u{
n
+ 3 53*ffi,-tt< (2.7.18)
t=i

Equations (2.7.15)-(2.7.18) may be considerably simplified by choosing


s = u i , without prejudicing the generality of the results because t h e param-
eter « is arbitrary. Then we have

U! = l , ui = xii = 0 (2.7.19)

Due to t h e diagonalisation carried out on t h e quadratic form by means of


(2.7.2), and bearing in mind t h a t dn — 0 , d22 > 0 then (2.7.16) is a positive
semidefinite quadratic form which vanishes in t h e direction

ut = l , ut: = 0 (t = 2 , 3 , . . . , n) (2.7.20)

Since for any li ^ lii we have P > 0, it is sufficient to analyse t h e higher


order terms in the direction lii. We have

P = dni (2.7.21)
n
P = rfim + 6 dinfi<+ 3 52da(ui)2 (2.7.22)
i=2

Equation (2.7.14) is therefore reduced to

P{ui)^-dmuz1 + — dim + 6 din tii + 3 ^ di{ (u t ) 2 u\ (2.7.23)


t=2

We exclude the case dm ^ 0, which would immediately lead to the conclusion


t h a t t h e potential energy is negative definite or indefinite, with consequent
unstable equilibrium at u = 0 for any value of the coefficient of u\. If d m = 0,
then analysis of the sign of P{u\) moves on to the term of the fourth order.
By calculating the minimum of the coefficient of u\ with respect to u t we get

u{ = - ^ (i = 2 , 3 , . . . , n ) (2.7.24)
da
(no s u m with respect to i), from which
37

dnn + 6 din u, + 3 £ d,-,- (u,) 2 = dnn - 3 £ ^ - (2.7.25)


t=2 .=2 ""

We can see that the coefficient of u\ is positive if

dim = dnn - 3 £ ^ - > 0 (2.7.26)

and therefore conclude that the sufficient condition for the total potential
energy to be positive definite is that

din = 0 (2.7.27)

dim > 0 (2.7.28)

From (2.7.13), (2.7.19), (2.7.20) and (2.7.24) it follows that the curve along
which (2.7.28) must be satisfied is given by

ui = ui (2.7.29)

tt.-(*i) = - l "T 1 u\ + o(u?) (t = 2 , 3 , . . . , n) (2.7.30)


2 da
In the treatment of continuous systems we shall see that the sufficient con-
dition for stability is given by relations which are altogether analogous to
(2.7.27) and (2.7.28). In particular, the direction along which the counter-
part of (2.7.28) must be satisfied, is obtained by adding to the displacement
which makes the quadratic term of the energy positive semidefinite, a higher
order orthogonal displacement in perfect analogy with (2.7.29), (2.7.30).
Let us now take as an example a system of two degrees of freedom whose
potential energy is given by

P(ui) =u\ + u\u2 + cu\ (2.7.31)

Along the direction u 2 = 0, the quadratic and cubic terms vanish and
P{ui)2 > 0 if c > 0. This condition is necessary only, and we intend to
determine the sufficient condition. Note that (2.7.31) can be rewritten in the
form

P{ui) = ( u 2 + \ u 2 ) 2 + (c - i ) u\ (2.7.32)

Along the curve

u 2 = - \ u\ (2.7.33)
38

Fig. 2.1 - Diagram of function (2.7.31) for c < 1/4.

the energy is positive definite if c > 1/4 and is negative definite if c < 1/4.
Therefore the sufficient condition of stability is given by c > 1/4. In Fig. 2.1,
where the function (2.7.31) has been plotted, we can see that whilst the
function increases starting from the origin in the direction of u x and u 2 , it
decreases along the curve (2.7.33) if c < 1/4. We arrive at the same result if
we apply the condition (2.7.28). In fact, we have
dn = 0 , d22 = 2 , dll2 = 2 , dnn = 24c (2.7.34)
from which, by applying (2.7.26), we get
4
dim = 2 4 c - 3 - > 0 (2.7.35)

that is, c > 1/4. Note that d m = 0. From (2.7.24) we finally have
u2 = - 1 (2.7.36)
from which, using the series expansion (2.7.30), we have

u2 M = - - u\ (2.7.37)
which coincides with (2.7.33).
In the following Sections we study in detail five elementary models, of one
or two degrees of freedom. For each of these the possible configurations of
equilibrium, in an exact or an approximate form, are determined and the
nature of the equilibrium is examined. The investigation is conducted on the
one hand with the aim of exemplifying the material in this Chapter, and on
the other of introducing some concepts which will be developed in Chapter 3.
39

2.8 A SYSTEM OF ONE DEGREE OF FREE-


DOM WITH STABLE SYMMETRICAL POST-
CRITICAL BEHAVIOUR
Let us consider the mechanical system illustrated in Fig. 2.2, which consists
of a rigid rod of length t constrained at one end by a hinge to which an elastic
linear spring of torsional stiffness A; is applied, and which is subjected at the
other end to a vertical force [7].

:>J I*

Fig. 2.2 - Rigid rod subjected to vertical load.

The configuration of the system is identified by the angle 0, measured from


the vertical reference configuration, and therefore the system has one degree
of freedom. In the generic configuration the potential energy is

p{6) = -ke> Ni{l-cos0) (2.8.1)

having assumed that for 0 — 0 the spring is unloaded.


Let us now look for the configurations of equilibrium by equating to zero
the first derivative of P(0), or its first variation. Note that in the case of
a function of n variables the two operations are equivalent. We shall prefer
the second terminology, for unity of language in the treatment of continuous
systems, where the condition of stationarity of a functional is obtained by
making its first variation vanish. From (2.8.1) we therefore have

6P(0) = {kO-Nesm0)6O = O V60 (2.8.2)

from which
0
N=± I sin0
(2.8.3)
40

All those pairs of {Ny0) values which satisfy (2.8.2) or (2.8.3) represent
states of equilibrium. From (2.8.2) we can immediately see that for 0 = 0
the equation is satisfied for any value of iV, and therefore in the plane {N,0)
all the points of the axis 0 = 0 represent states of equilibrium. The straight
line 0 = 0 thus represents a curve or equilibrium path, which we indicate by
the Roman numeral I. A second equilibrium path, which we indicate by the
Roman numeral II (Fig. 2.3), is obtained by plotting the function (2.8.3). As
lim0_>o 0/ s'mO = 1, the curve II intersects the axis 0 = 0 at the point

Nc = k/l (2.8.4)
The name fundamental equilibrium path is given to curve I and bifurcated
equilibrium path to curve II. Point B (Fig. 2.3) where we have two equilibrium
paths is called bifurcation point, and the associated value of N provided by
(2.8.4) is the critical load.
♦N
_ stable eq.
-_ unstable eq.

-jt O n

Fig. 2.3 - Equilibrium paths.

The problem can be studied in an alternative way by using, in place of


the exact expression (2.8.1) of P(0), its series expansion starting from the
reference configuration 0 = 0 (see 2.7.1). Analysis of this type is a useful
procedure which can furnish an approximate solution in those cases in which it
is impossible to have an exact solution. Starting from (2.8.1) and performing
the series expansion of cos 0 in 0 = 0 up to fourth-order terms, we have
t
P{9) = l(k-Nt)tF + ^t (2.8.5)

from which we obtain the equilibrium points by imposing that the first vari-
ation should vanish, that is
NI
{k-N L)6 + — 6 60 = 0 V60 (2.8.6)
41

Equation (2.8.6) gives the two equilibrium paths approximated up to the


second order:

6 = 0 VAT (2.8.7)

Note that the second-order approximation gives an exact representation for


the fundamental path, whilst (2.8.8) can be obtained by performing the series
expansion of the second member of (2.8.3) about 0 = 0. By setting 0 = 0 in
(2.8.8) we obtain the bifurcation point Nc = k/t, which coincides with that
in the exact solution.
Our next aim is to analyse the type of equilibrium of points of the fun-
damental and bifurcated paths, making use of the Lagrange-Dirichlet and
Liapunov theorems presented in Sections 2.4 and 2.5.
To simplify the analysis we shall refer to the approximate solutions to
the problem. In correspondence with the bifurcation point, the coefficient
of 02 in (2.8.5) vanishes. As the stiffness matrix [C,-y] is singular, we have a
critical case of equilibrium (see Section 2.7), and so the type of equilibrium is
decided by the higher order terms. As the third-order term is zero and that
of the fourth order is positive definite in the present case, we can conclude
that for N = iVc, P(0) given by (2.8.5) is positive definite, and so according
to Lagrange-Dirichlet the equilibrium is stable. As far as analysis of the
remaining points of the two equilibrium paths is concerned, it is convenient
to calculate the second derivative of P{0) along the paths and verify their
signs. We have

d2P 1
— = k-NL+-NL02 (2.8.9)

Along the equilibrium path J = 0 w e have

d 2P
= k-Nl (2.8.10)
d02
and so P(0) is positive definite for N < kjl and negative definite for N > kjt.
In the first case the equilibrium is stable, in the second unstable. Along the
bifurcated path, by substituting equation (2.8.8) into (2.8.9), we get

d2P k(n2 0 4\ /«o,^


— - = - [e2 H (2.8.11)
V
d02 3\ 4J '
which is positive and so the equilibrium is stable. In Fig. 2.3 the type of
equilibrium along the I and II paths is indicated. All those structures which
42

Fig. 2.4 - Rigid rod with a mass applied to its end.

present a bifurcated equilibrium path of the type examined are called struc-
tures of stable symmetrical post-critical behaviour.
A second constructive way of solving the same problem is to carry out a
dynamic analysis, giving to the system a displacement and an initial velocity
starting from a configuration of equilibrium and studying the consequent
motion. Let us consider the system in Fig. 2.4, identical to the previous
system (Fig. 2.2) except for the fact that at the free end a point mass m is
applied in place of the force in the constant direction, N. Supposing that
for 6 = 0 the spring is unloaded, the vertical configuration is a configuration
of equilibrium. By measuring 0 from this configuration and disregarding the
mass of the rod, we can write the expressions for the kinetic energy and the
potential energy as

K = -mv2 = -m (tof (2.8.12)

P = -k02 -mg£(l-cos0) (2.8.13)

where g is the acceleration of gravity. By introducing the Lagrangian


L = K — P we have the equation of motion (2.2.11)

mfO + kO — mg I sinO = Q (2.8.14)

which we rewrite in the form

mt?d+{k-mgl)0 + R{0) =0 (2.8.15)

where ||i2(0)|| = o(0). By linearising equation (2.8.15) we arrive at

m £ 2 0 - f {k-mgt)0 = 0 (2.8.16)
43

The conclusions on the stability of equilibrium which we reach by studying


equation (2.8.16) are also valid for the real problem described by the non-
linear equation (2.8.14), provided that the roots of the characteristic equation
are all real and negative, and if there are complex roots, all have real negative
parts (see Sections 2.6 and 1.7). We put the solution to equation (2.8.16) in
the form

0{t) = c ext (2.8.17)

By substituting (2.8.17) in (2.8.16) we have

A = ± \mgl
<2-8-i8>
—k ,

v-b-
and so the general solution is

0(i) = cx ext + c2 e~xt (2.8.19)

The following three cases can occur.

(a) m g t — k > 0, that is, m g > k/L


The A roots are real numbers, one of which is positive. The equilibrium
configuration of the linear system is an unstable configuration, called a
saddle point (see Section 1.3). Furthermore, for the first theorem men-
tioned in Section 1.7, the configuration of equilibrium is also unstable for
the non-linear system. By carrying out a static analysis of the problem,
we can write the approximate potential energy as

P(0) = ±(k-mge)e2 + ^ O i
(2.8.20)

As the coefficient of 02 is negative, P(0) is negative definite and therefore,


according to Liapunov, the configuration of equilibrium is unstable.

(b) m g I - k < 0, that is, m g < k/L


The A roots are pure imaginary numbers which we shall write as ±icj.
Solution (2.8.19) can be expressed in the form

Q
6{t) = 0Q s i n u t + — cos CJ t (2.8.21)
(jj

where 00 and 0O are respectively the initial displacement and velocity.


From an examination of (2.8.21) we conclude that the motion of the lin-
earised system is stable according to Liapunov around the configuration
of equilibrium (centre; see Section 1.3) because to a sufficiently small 0O
44

and 0o there corresponds a bounded 0{i). Such conclusions, valid for the
linear system, may not be extended to the non-linear system because
the conditions forecast in the second theorem in Section 1.7 do not oc-
cur, and therefore a linear analysis is completely useless in this case. By
carrying out a static analysis, we can immediately conclude that the con-
figuration of equilibrium is stable, as in (2.8.20) the second-order term
is positive definite.

(c) m g I — k = 0, that is, m g = k/L


The A root has a zero value which is counted twice, and solution (2.8.19)
becomes

0(t) =0o + 0ot (2.8.22)

The motion of the linear system is therefore unstable because, however


small 0O and 0O are, 0(t) increases indefinitely with t (see Section 1.3).
However, as the conditions of the first theorem in Section 1.7 do not
occur, the conclusions reached are not valid for the non-linear system
and so are without significance. A static analysis carried out by using
the potential energy (2.8.20) immediately leads to the conclusion that
the configuration of equilibrium is stable, in that the fourth-order term
is positive definite.

We conclude this Section by returning to the system in Fig. 2.2 to examine


how the results obtained are modified if imperfections are present. By this we
mean a small initial inclination of the rod 0O corresponding to the unloaded
spring, or a small eccentricity in the force N, or a small horizontal force acting
on a point of the rod. Let us consider the case in which the imperfection is
an initial inclination 0O (Fig. 2.5). By keeping the same convention for the
measurement of the angle 0 used previously, we have

P(0 ;0Q)=:-k(0- 0 O )2 -NI[1-COBO)+NI(1- cos0O) (2.8.23)

from which, by carrying out the series expansion of cos 0O around 0o = 0 up


to the fourth order, we get

P{0 ;0o) = ^k{O2 + 02o-20 0O) -NI{1- cos 0) + Nt IS - ^ ) (2.8.24)

Let us now suppose that 0O is so small that we can disregard, with respect to
0 0O, all terms which contain a power of 0O higher than the first. Under this
hypothesis, indicated by the term small imperfections, (2.8.24) reduces to

P(0 ;0o) = lk{02-200o)-Nl{l- cosfl) (2.8.25)


45

Fig. 2.5 - Imperfect rigid rod subjected to vertical load.

The equilibrium configurations are given by the solution to the problem


[k{0-0o) -NtsinO} 66 = 0 V66 (2.8.26)

from which we obtain the equilibrium relation


k Jo_ (2.8.27)
I sinfl I sin*
We can see that the first term on the right side of (2.8.27) is identical to
the second member of (2.8.3), whilst the second term represents a correction
caused by the presence of the imperfection 0O. Equation (2.8.27) has been
plotted in Fig. 2.6, where four families of curves are observed, two correspond-
ing to 0o > 0 and two to OQ < 0. The curves which intersect the horizontal
axis are called natural equilibrium curves or paths, and the others unnatural

stable eq.
unstable eq.

Fig. 2.6 - Equilibrium paths of the perfect and imperfect structures.


46

equilibrium curves. Both a natural and an unnatural equilibrium curve cor-


respond to each value of 0O. The equilibrium paths of the perfect structure,
towards which the equilibrium curves of the imperfect structure tend when
0O —► 0, are also represented in the figure. The curve S has been obtained by
plotting the function
d 2P
—— = k-Nlcos0 = O (2.8.28)
dO
We can see that for points of the plane situated on the inside of the curve 5 ,
d2P/d02 < 0, while for those points situated outside, d2P/d02 > 0. Therefore,
this curve divides the plane into two parts: in the first the equilibrium is
unstable and in the second, stable.

2.9 A SYSTEM OF ONE DEGREE OF FREEDOM


WITH UNSTABLE SYMMETRICAL POST-
CRITICAL BEHAVIOUR
Let us consider a mechanical system [7] made up of a rigid rod of length £,
hinged at one end and connected to a spring of stiffness k at the other (Fig.
2.7). N is a vertical force applied to the centroid of the section, and 0 the
angle measured from the vertical which specifies the configuration. Assuming
that for 0 = 0 the spring is not loaded, we can write the expression of the
potential energy relative to the generic configuration

P{0) = -hi2 sin20-iV£(l-cos0) (2.9.1)


from which we obtain the equation of equilibrium
I sin0(A;£ c o s 0 - N) = 0 (2.9.2)

k
^ T

Fig. 2.7 - Rigid rod subjected to vertical load.


47

N
stable eq.
unstable eq.

Nc=k/

/ *
-1
■jr/2 TC/2

Fig. 2.8 - Equilibrium paths.

Equation (2.9.2) has the solutions

0 = 0 (2.9.3)

N = k I cos 0 (2.9.4)

represented in Fig. 2.8 and marked with the Roman numerals I and II. To
these curves we again give, as in Section 2.8, the names fundamental path and
bifurcated path. Bifurcation occurs where

Ne = kt (2.9.5)

To decide upon the nature of the equilibrium, we evaluate d2P/d02 along


the paths (2.9.3), (2.9.4). From (2.9.1) we get
d?P
— - = k I2 (2 cos2 0 - l) -NtcosO (2.9.6)

which for 0 = 0 gives


d 2P
l{kt-N) (2.9.7)
d02
Therefore along the fundamental path the equilibrium is stable for N < kt,
and unstable for N > k L Along the bifurcated path (2.9.4) we have
d2P
d02 \ K - *2 ?) (2.9.8)

from which, as N < k £, we can conclude that the potential energy is negative
definite and therefore the equilibrium is unstable. At the bifurcation point,
Nc = k£, (2.9.6) vanishes, and so to decide upon the type of equilibrium it is
necessary to examine the potential energy (2.9.1). By writing 0 = 2<p, this is
48

reduced to

P(<p) = 2kt2 sin2tp (cos2<p - l ) (2.9.9)

which is clearly negative definite, and so at the bifurcation point the equilib-
rium is unstable. All those structures which present a bifurcated equilibrium
path of the type examined are called structures with unstable symmetrical
post-critical behaviour.
Let us now pass to an approximate analysis of the problem based on

'<«>=!«'(«'-IT)-*'(?-£) <2-9-io>
which is obtained from (2.9.1) through a series expansion around 0 = 0. By
imposing that (2.9.10) be stationary with respect to 0 we have the approxi-
mate equilibrium paths up to second order terms

0 = 01 (2- 9.11)

N = kt(l-
?) (2- 9.12)

where the bifurcation point is again furnished by (2.9.5). Furthermore, from


d2P
kt2(l - 2 0 Nlll-
-?)
2
= ) - (2. 9.13)
dO2
we have
d2P
AM
= l{kt--N) (2- 9.14)

along the fundamental path, which is identical to (2.9.7) and to the conclu-
sions already discussed, together with

d2P , A ( 4N\
4N\ N2
N<
< 2 - 9 - 15 >
2
d0 = kt {-3 + Je)-T
2

along the bifurcated path. By writing N = akl with 0 < a < 1, (2.9.15)
becomes

— = k l2 ( - 3 + 4 a - a2) (2.9.16)

which is negative, as is apparent from the diagram of function f(a) = — 3 -f-


4 a —a2 (Fig. 2.9) and so, along the bifurcated path, the equilibrium is unsta-
ble. At the bifurcation point, Ne = kl, a = 1 and therefore, from (2.9.16),
d2P/d02 = 0. By using the expression (2.9.10) of the potential energy we
49

f(a)

Fig. 2.9 - Diagram of function f{a).

deduce

P(0) = --kf264 (2.9.17)

which indicates that the equilibrium at Nc is unstable.


Let us now suppose t h a t the system has an imperfection, represented by
an initial inclination of the rod 0O when the spring is unloaded (Fig. 2.10).
The total potential energy (2.9.1) is modified t o

P ( 0 ; 0Q) = -A;(£sin0-£sin0o)2 - Nt(l-cos0)

+ Nl(l-cos0Q) (2.9.18)

By substituting in (2.9.18) a series expansion in terms of 0 and 0Q around

Fig. 2.10 - Imperfect rigid rod subjected to vertical load.


50

6 = 0 , 0O = 0 we get the approximate expression

l
P{e-,e0) = -ke{e'-^-26e^-Ni(^-^ (2.9.19)
having made the hypothesis that the initial imperfection 0O is very small and
therefore having disregarded, with respect to 0 0O> all terms which contain a
power of 0O higher than the first.
From (2.9.19) we get the equation of equilibrium as

0^_0_o
N = kt 1 (2.9.20)
2 0

where terms of the order O(04) and 0(0 0O) have been disregarded with respect
to 02 and 0o/0, respectively. The function (2.9.20) is represented in Fig. 2.11
where, as in the case discussed in the previous Section, we can see four families
of curves, two corresponding to 0Q > 0 and two to 0Q < 0.
As before the name natural equilibrium curves or paths is given to those
curves which intercept the axis of the abscissae, and to the others the name
unnatural equilibrium curves. To each value of 0O there corresponds a natural
equilibrium curve and one of unnatural equilibrium. When 0O —► 0, the
equilibrium curves of the imperfect structure tend towards those of the perfect
structure represented in the figure.
A basic difference distinguishing the mechanical model studied in this Sec-
tion from that analysed in the previous one is that the natural equilibrium
curves reach a maximum N* and then decrease. To find the dependence of
N* on 0Q, we look for the stationary point of these curves. From (2.9.20) we

stable eq.
— unstable eq.

Fig. 2.11 - Equilibrium curves of the perfect and imperfect structures.


51

kl

Fig. 2.12 - Dependence of the collapse load on the initial imperfection.

have

dN 9_o
= kt\-6 + (2.9.21)
dO 02

from which

el'3 (2.9.22)

By substituting (2.9.22) in (2.9.20) we arrive at

N* = kt(l-\oT) (2.9.23)

which is represented in Fig. 2.12. The name collapse load or limit load is
given to the load JV*, and the corresponding point is called the limit point.
Because the curve (2.9.23) is tangent to the axis of the ordinates OQ — 0,
we can deduce that a small initial imperfection 0O induces a collapse load N*
which is very much below the critical load Nc. From this comes the term
structures sensitive to initial imperfections, which is extended to all those
structures whose behaviour is of the type described.
An alternative way of arriving at equation (2.9.23) is to eliminate 0 from
the equation

d2P
= kt(l-202) -ATM - —) = 0 (2.9.24)
dO2

and from the equation of equilibrium (2.9.20). Equation (2.9.24) is plotted


in Fig. 2.11 (curve S). We can see that at points of the plane situated below
the curve S we have d2P/d02 > 0, and therefore the equilibrium is stable,
whilst at those points situated above, d2P/d02 < 0 (unstable equilibrium).
52

2.10 A S Y S T E M OF ONE D E G R E E OF F R E E D O M
WITH ASYMMETRICAL POST-CRITICAL
BEHAVIOUR
Let us consider a mechanical system [7] consisting of a rigid rod of length £,
hinged at one end and connected to a spring of stiffness k at the other (Fig.
2.13). N is a vertical force applied to the centroid of the section and 0 is
the angle which identifies the configuration, measured from the vertical. We
denote by a* the length of the unloaded spring in the reference configuration,
by <p the angle with respect to the horizontal in this configuration, and by a
the length of the spring in the generic configuration. The potential energy is

P{0) = -k{a-a*y -Nl{l-cos0) (2.10.1)

and as

a = (2.10.2)
sin<p
and

a— 1+ | + 2 sin 0 (2.10.3)
V tan<p \^tan<£>

we have

P(9) = -kl> .iH f h2 sin0


\i t a m p ytan^> sirup
- Ne{l-cos$) (2.10.4)

I*

Fig. 2.13 - Rigid rod subjected to vertical load.


53

Let us assume for simplicity that <p — n/4 and substitute in (2.10.4) the
series expansion around 0 = 0. We then have

From the stationary condition of (2.10.5) we obtain the equations of the


equilibrium curves approximated up to the second order

0 = 0 (2.10.6)

_k_t( 3 „ 02
TV 2 1--0 + — (2.10.7)
~ V 4 8

represented in Fig. 2.14 and marked with the Roman numerals I and II. For
0 = 0, (2.10.7) gives the value of N corresponding to the bifurcation point

(2.10.8)
2

By substituting (2.10.8) in (2.10.5) we obtain P{0) = kl2{-03 + 0 4 / 8 )/ 8 >


which is indefinite as the cubic term is dominant with respect to the quartic
term and therefore the equilibrium corresponding to the bifurcation point is
unstable. To examine the type of equilibrium corresponding to other points
of the fundamental and bifurcated paths, we resort to the sign of the second
derivative of P(0). We have

d2
P 1 , ,2 2 - 3 0 02
Nt\l (2.10.9)
M* ~ 4 2

stable eq.
- unstable eq-

Fig. 2.14 - Equilibrium paths.


54

from which
(21010)
£ = «("-") --
along the fundamental path. We immediately deduce t h a t the equilibrium is
stable for N < kl/2 and unstable for N > kl/2. Along the bifurcated p a t h
(2.10.7) we get

d2P kl2 / ,x ,
— = — (-38 + fi) (2.10.11)

and therefore, as the linear term is dominant over the quadratic term around
0 = 0, the equilibrium is stable for 0 < 0 and unstable for 0 > 0.
The analysis can be simplified in all problems in which, as in the present
case, the potential energy P(0) contains non-zero terms of the third order. In
such cases, it is correct to truncate the series expansion (2.10.5) at the third
order terms, and so the equilibrium curve (2.10.7) is replaced by

N
= T (x - i e ) (2 10 12)
- -
whilst (2.10.6) remains unaltered. Equation (2.10.12) is that of the tangent
to the curve (2.10.7) at the point 0 = 0, and is shown in Fig. 2.14. The
fundamental and bifurcated paths therefore reduce to two straight lines. Bi-
furcation still occurs for the value of N given by (2.10.8) and the potential
energy corresponding to this point has the expression P{0) = —kl203/8,
which is indefinite; therefore, at the bifurcation point the equilibrium is un-
stable. At other points of the fundamental path relation (2.10.10) remains
valid, whilst along the bifurcated path (2.10.11) is substituted by

d2P 3

which is positive for 0 < 0 and negative for 0 > 0. The conclusions for the
problem defined by (2.10.5) can therefore be extended exactly to the problem
in which the series expansion is truncated at the third-order terms.
All those structures which present a bifurcated equilibrium p a t h of the
type examined, with non-horizontal tangent, are called structures with asym-
metrical post-critical behaviour.
Let us now suppose that the rod is inclined at an angle 0O with respect to
the vertical in its initial configuration, and a 0 is the length of the unloaded
spring relative to this configuration (Fig. 2.15).
The potential energy is written as
k
P(0 ; 0O) = - (a - a 0 ) 2 - NI (1 - cos 0) + NI (l - cos 0O) (2.10.14)
55

Fig. 2.15 - Imperfect rigid rod subjected to vertical load.

On the other hand from (2.10.3), written for (p = 7r/4, we have

a = ly/2 \ / l + sin0 (2.10.15)

and

a0 = l\[2 yjl + sin 0O (2.10.16)

By substituting (2.10.15) and (2.10.16) in (2.10.14) and carrying out the series
expansion around 0 — 0 , 0O = 0 we obtain

(2.10.17)

As with the cases discussed in the two previous Sections, we have made
the hypothesis that the initial imperfection 0Q is so small that terms which
contain a power of #o higher than the first can be disregarded with respect to
the term in 0 0O. Besides, as cubic terms in 0 are present, the quartic terms
have been disregarded with respect to them. From the stationarity condition
of (2.10.17) we get the equation of equilibrium

"-VH-*) (2.10.18)

which is represented in Fig. 2.16. Four families of curves are observed, two
corresponding to 0O > 0 and two corresponding to 0O < 0. To each value
of 0o there corresponds a curve of natural equilibrium and one of unnatural
equilibrium. In the figure the equilibrium curves of the perfect structure, to
which those of the imperfect structure (2.10.18) tend when 0O —> 0, are also
represented. The curves of natural equilibrium which correspond to 0O > 0
56

stable eq.
unstable eq.

Fig. 2.16 - Equilibrium curves of perfect and imperfect structures.

reach a maximum N* and then decrease, whilst those corresponding to 0O < 0


always increase.
To find the dependence of the collapse load N* on 0O, we eliminate 0 from
the equation of equilibrium (2.10.18) and the equation

d2P k l ^ _ 3 ^ _ N = 0 (2.10.19)
d$2 ~~2
represented in the figure by dashes. We get

(2.10.20)

which is shown in Fig. 2.17. We can see that the straight line (2.10.19) is the
bisector of the angle between the fundamental and the bifurcated paths. Like

2N*
k/

Fig. 2.17 - Dependence of the collapse load on the initial imperfection.


57

the curve N* — O0 obtained in the last Section, the curve (2.10.20) is tangent
to the vertical axis 0O = 0. In consequence a small initial imperfection causes
a drastic reduction in the collapse load N* with respect to the critical load
iVc, and so those structures which have a behaviour of the type examined here
will again be called structures sensitive to initial imperfections.
It is worth noting t h a t structures with asymmetrical post-critical behaviour
have a greater sensitivity to initial imperfections than structures with sym-
metrical post-critical behaviour, in t h a t the dependence of (2.10.20) on the
imperfection is of the type 0 j / 2 , whilst (2.9.23) depends on 0l,s. We can see
that at points of the plane situated to the left of the straight line (2.10.19)
d2P/d02 > 0, and so the equilibrium is stable there, whilst at points of the
plane to the right of the straight line S we have d2P/d02 < 0, and thus
unstable equilibrium.

2.11 A SYSTEM OF ONE DEGREE OF FREE-


DOM WITH NON-LINEAR PRE-CRITICAL
BEHAVIOUR
Let us consider the mechanical system shown in Fig. 2.18, consisting of two
rods which can take only axial deformation, of equal length ao, and suppose
t h a t the hinge can only move vertically. The system is subjected to the action
of force iV, positive if directed as in the figure [7]. We assume as Lagrange
coordinate the angle 0 which the rods form with the horizontal; 0O is the value
of 0 in the absence of the external load. In the deformed generic configuration
the total potential energy is

P{0) =k[Aa0 [0)\2 -N6{0) (2.11.1)


where k is the axial stiffness of the rods and Aao(0) and 6(0) are the elongation
of the rods and the vertical displacement of the point of application of the

Fig. 2.18 - A system of one degree of freedom.


58

force, respectively, both depending on the angle 0. The kinematic study of


the system gives

K }
(2.11.2)
2 Vcos0 o cos0/

6 (0) = - (tan 0O - tan 0) (2.11.3)

from which the total potential energy, in an explicit form, is

ki2 / 1 1 \2
P{0) = + 2A(tan0-tan0o) (2.11.4)
V cos 0O cos0/

having for convenience set


N
A= (2.11.5)
kl
By imposing the stationarity condition on P(0) we obtain the following
equation of equilibrium

A = sin0 f —J (2.11.6)
\ COS 0 O COS0/

which describes the equilibrium path A = A(0;0 O ). The diagram of the


function is shown in Fig. 2.19 for — n/2 < 0 < n/2. The function is anti-
symmetric, has two asymptotes at 0 = ±7r/2 and vanishes for 0 = ±0 O and
0 = 0. Therefore three positions of equilibrium are possible in the absence of
load: the reference configuration, the symmetric configuration with respect

-jr/2

Fig. 2.19 - Equilibrium curve.


59

to the straight line joining the fixed hinges and the configuration in which the
three hinges are aligned. In the first two the rods are stress-free, and in the
third they are compressed. By increasing the load from zero, the three con-
figurations of equilibrium change, two of them occurring for positive values of
0 and the third for negative values. For A —* Ac, the first two configurations
tend to the same limit 0 = 0C; correspondingly, the equilibrium path shows a
maximum. By imposing the condition dX/d0 = 0, we find

cos 3 0C = cos 0O (2.11.7)

and by substitution in (2.11.6)

Ar = tan 3 0r (2.11.8)

Equations (2.11.7) and (2.11.8) define the coordinates 0C, Ac of the stationary
points of the load path which will be examined in detail in the next Chapter.
To determine the stability of the equilibrium points, it is necessary to
calculate the second derivative of P(0) at these points and check its sign. We
have

d 2P kl2
( l + 3 tan 2 0)
d02 2 Lcos20
tan0
( l + 2 tan 2 O) + 2 A (2.11.9)
cos 0 cos 0O V~ ' ~ ~ " J cos 2 0j

By eliminating the parameter 0O between (2.11.6) and (2.11.9) we get

d2P , m9 1
kl2 (tan30-A) (2.11.10)
d02

from which, for A < tan 3 0 we have d2P/d02 > 0 and the equilibrium is stable;
for A > tan 3 0, d2P/d02 < 0 and the equilibrium is unstable; for A = tan 3 0,
d2P/d02 = 0 and the equilibrium is critical. On the plane A, 0 the curve
A = tan 3 0 separates the region of stable configuration from the unstable
region (Fig. 2.20). It intersects the path (2.11.6) at the limit points (see
(2.11.8)) which are therefore also critical equilibrium points.
To decide upon the nature of the equilibrium at the critical points, it is
necessary to calculate the third derivative of P(0). We have

d 3P sin# COS0
2 ( l + 3 tan 2 0) ( l 4- 2 tan 2 0)
d03 cos 3 0 COS 0o

2 (
COS0 \COS0
3
2
COS0n
) 1 o\

+ — T z(l + 2s\n2e) (2.11.11)


60

stable eq.
unstable eq.

Fig. 2.20 - Equilibrium curve and characterisation of the equilibrium.

By substituting (2.11.7) and (2.11.8) in (2.11.11) we obtain

* P3 = 3 4* ^ 0 for o < ee < - (2.11.12)


d0 cos 0C
from which it follows that, as (2.7.28) is violated, the critical equilibrium is
unstable.
We now have all the elements necessary to describe the mechanical be-
haviour of the model. Let us consider the system in the natural state O
(Fig. 2.21) and increase the load, starting from zero. The representative
equilibrium point moves along the curve O C describing the non-linear load
path, characterised by configurations of stable equilibrium. If the system is
unloaded, starting from any point B belonging to the piece OC, then the

F!

Ac c

B
/ \
V* J
°\ / °\
•- ± \ A
D
~K

Fig. 2.21 - Snap-through phenomenon.


61

representative point moves backwards along BO. However, if with an in-


crease of A the representative point reaches the maximum C, the position
of equilibrium becomes unstable and the system passes to E where it finds,
under the same load, a position of stable equilibrium.
The phenomenon called snap-through corresponds physically to the instan-
taneous change of configuration from t h a t in which the loaded hinge is above
the straight line which joins the fixed hinges, to the configuration in which
it is below the line. If the load is increased further the representative point
moves along the curve E F. If, along this curve, we proceed to a complete
unloading, the system moves to 0\ symmetrical to O. If A decreases further,
that is if it assumes negative values, another j u m p occurs in D and the system
moves to A, which belongs to the original branch of the curve. All the config-
urations of equilibrium in the piece C D are unstable and cannot be reached
from the natural configuration. If the system is in one of these configurations
a perturbation, however small, tends to take the system into one of the two
positions of stable equilibrium corresponding to the given load.
The analysis conducted, based solely on examination of the total poten-
tial energy and quite apart from the dynamic aspects of the system, allows
us to determine all the configurations of equilibrium of the system and the
corresponding types of equilibrium, thus exhausting the investigation which
was proposed. However, a dynamic description of the phenomenon which
also takes into account the effect of the forces of inertia, although not offering
new results with respect to those already acquired, can be useful for a better
understanding of the mechanical behaviour of the system.
To this end we determine the Lagrange equation of motion of the system
in Fig. 2.22, obtained from that in Fig. 2.18, substituting for the force N a
mass ?n = N/g where g is the acceleration of gravity.
The kinetic energy K(0, 0) resulting from (2.11.3) is

K(0 , 0) = - m 62 = — -^— (2.11.13)


v
' } 2 8 cos 4 0 V }

where a dot indicates differentiation with respect to time. The kinetic energy,
like the total potential energy, is therefore a non-linear function of 0 and

Fig. 2.22 - Dynamic system with one degree of freedom.


62

the equation of motion is also non-linear, due to the choice of the Lagrange
coordinate. With the aim of linearising the problem we assume for P(0) and
K{0, 0) a quadratic approximation, by expanding in series the two functions
starting from a generic configuration of equilibrium, which belongs to the path
(2.11.6). Let 0e be the configuration of equilibrium and

<p = 0-0e (2.11.14)

the angle which specifies the generic configuration assumed by the system
during motion. We have

^d^H^-ify^"2' <2iii5>
% , p ) = ! » l ' — r r *>' + <>(*>') (2.11.16)
8 cos* ue
having made use of (2.11.10) and (2.11.5). The linearised equation of motion
is
k cos 4 J g / « mo\ ,
<p + 4-^—^(t&n3ee--f)<p = 0 2.11.17)
m sin 20 e \ kt /
from which, if the solution is assumed in the form (p(t) = eat, we have

(2.11.18)
V m sin20 c V kl
If mg/kt > tan 3 0 c , then the configuration of equilibrium <p = 0 is a
saddle point in the plane <p, <p and the equilibrium is unstable (Section
1.3, item (a 3 )); linearised analysis then shows a motion which diverges. If
mg/kt = tan 3 0e the eigenvalues a vanish, and the origin of the plane of the
phases is a non-simple point. Linearised analysis yields a uniform motion
which therefore diverges. This case, by virtue of (2.11.8), corresponds to
the snapping point. If mg/kt < tan 3 0e then the origin is a centre (Section
1.3, item (b 3 )) and therefore stable. Linearised analysis yields a sinusoidal
motion, the trajectories of which are ellipses with their centres at the origin.
As seen in Section 1.7, a saddle point for the linearised system is still a
saddle point for the non-linear system, that is, the equilibrium point is un-
stable for both systems. This is not the case for a centre, which is stable
for the linear system but can be either stable or unstable for the non-linear
problem. Therefore linearised analysis cannot decide the nature of the equi-
librium. This problem presents itself every time a conservative system is
studied. In fact, according to the law of energy conservation, the real part
of the characteristic roots can be either positive or zero, but never negative,
which excludes asymptotic stability. The theorems in Section 1.7 relating
63

a) k,<Xc b) A!<A2<AC c) A = AC
Fig. 2.23 - Total potential energy and trajectories for different load values.

to linearised analysis are therefore applicable only in predicting instability,


never stability.
In the case under examination, an investigation of the type presented in
Section 1.4 can be carried out, based on the construction of the trajectories
of the system starting from a first integral of motion, such as that which ex-
presses the law of energy conservation. Figure 2.23, taken from [4], shows the
diagrams of the total potential energy P(0) and the corresponding trajectories
on the plane 0, 6 for three different values of A = mg/kL
For \ = \i < Xc (Fig. 2.23(a)) the total potential energy shows three sta-
tionary points, two minima and one maximum. Corresponding to the first
two we have two centres A and B , and to the third a saddle point C. If the
configurations A and B are perturbed, the system oscillates according to a
periodic law (which in general is not simply sinusoidal) around the configu-
rations of equilibrium, and so the configurations are stable, as determined by
linearised analysis. If the configuration C is perturbed several possibilities
can occur: (i) the system oscillates around A or around B; (ii) the system
describes a periodic motion around the three configurations of equilibrium;
(Hi) the system tends to return to C after an infinite time. Point C is un-
stable, as determined by linear analysis, but the motion is not exponentially
divergent and cannot be studied by local investigation.
For A = A2 with Xx < A2 < Ac (case (b)) the configurations of equilibrium A
and C tend to move towards each other, as already observed in examination
of the load path in Fig. 2.21. The configuration A is still stable, but the
maximum range of perturbation which allows the motion to remain of the
oscillatory type around A decreases.
64

For A = Ac (case (c)) the configurations of equilibrium A and C coincide


and P(0) has a horizontal inflexion at that point. If the system is perturbed,
starting from C, then we have the following possibilities: (i) the motion is
oscillatory around B\ (ii) the motion is periodic around the two configurations
of equilibrium A and B; (Hi) the motion tends to return the system to the
original configuration C. This type of behaviour is not accounted for by
linearised analysis, which predicts a uniform motion in the neighbourhood of
the critical point, thus mistakenly indicating a neutral equilibrium.

2.12 A S Y S T E M OF T W O D E G R E E S OF F R E E D O M
The system examined [5] consists of a rigid rod of length £ constrained by a
pivot at one end, to which are applied two linear elastic springs of stiffnesses
ki and k2. At the other end a vertical force iV, acting on the centroid of the
section, is applied (Fig. 2.24(a)). Unlike the systems studied in the last four
Sections, this mechanical system has two degrees of freedom. Denoting by
<Pi > <P2 > V?3 the angles between the rod and the coordinate axes (Fig. 2.24(b)),
it is obvious that cos 2 <pi-fcos2 <p2+cos2 (P3 = 1, and so it is sufficient to specify
two of the three <£>,- angles to identify the configuration of the system.
This mechanical system is known as the Augusti model. By introduc-
ing the angles 0\ — TT/2 — (p\ , 02 — ^r/2 — £>25 bearing in mind that
c o s ^ i = s i n # i , cos<£>2 = sin# 2 , and making use of the relation between the

a) b)
Fig. 2.24 - Rigid rod subjected to vertical force:
(a) in the reference configuration,
(b) in the generic configuration.
65

cosines of the angles £>,-, we obtain the potential energy expression

P(0ly 02) = - M J + - M 2 -Nl(l-yjl- sin 2 0X - sin 2 02 ) (2.12.1)

in which we substitute the series expansion up to terms of the fourth order


around 0X = 0, 02 = 0

In writing (2.12.1) it has been implicitly assumed that in the reference config-
uration (Fig. 2.24(a)) the springs are not loaded. By imposing the stationarity
condition of (2.12.2) we obtain the equations of equilibrium

0, V 6 2JJ
(2.12.3)

and
42 /)2
02 l = 0 (2.12.4)
^ 2 6

where N t/ki = p and k2/ki = c has been assumed. The system has the
solutions

01 = 0 , = 0 (2.12.5)

*1
02 = 0 , p= l + (2.12.6)

0 1 = O, (2.12.7)

0i 0j (l+f-f)=0 (2.12.8)
1-p 1
6 2 = 0,
The functions (2.12.5) to (2.12.7) are represented in Fig. 2.25 and the cor-
responding curves are indicated respectively by Pi, P2 , P3. The two branches
marked P4 and P$ correspond to the equations (2.12.8). The system therefore
presents a fundamental equilibrium path Pi and four bifurcated equilibrium
paths P2 , P3 , P4 , P5, of which the first two intersect Px at the points p = 1 and
p = c corresponding to two different values of the critical load (in the figure
c > 1 has been assumed) whilst the curves P4 and P$ intersect the curve P2 in
points which are symmetrical relative to the plane 0i = 0. These points are
called secondary bifurcation points. For a more precise idea of curves P4 and
66

— stable eq.
- unstable eq.

Fig. 2.25 - Equilibrium paths.

P5 it is convenient to project them on the horizontal plane. By eliminating p


from (2.12.8) we get

6 (c - 1) - (c + 3) 6\ + (3 c + 1) 0\ = 0 (2.12.9)

Equation (2.12.9) represents two curves symmetrical with respect to the


axes $i and 0 2 , which intersect the axis 02 in imaginary points and the axis
0i in real points placed at a distance d from the origin:

d=± 6(c-l) (2.12.10)


3+ c

These curves are represented in Fig. 2.26(c). By projecting the curves P4 and
Ps on the plane p, 62 we obtain, by means of a series expansion up to terms
of the second order,

H
(2.12.11)
4 V 3

which is represented in Fig. 2.26(b). It can immediately be verified t h a t the


point of coordinates (d, (3 + c)/4) satisfies equation (2.12.6), and therefore
the curves P4 and P$ each have a point in common with P2> respectively E
67

%
\ y
\ /

1 \
\ stable eq.
C) unstable eq.
/F E
/ \
\
/ \

Fig. 2.26 - Projection of equilibrium curves on the coordinate planes.

and F in Fig. 2.25. Finally, by projecting the curves P4 and P5 on the plane
p, 0i we obtain
(2 i2i2)
'-*&k -
Curve (2.12.12) is represented in Fig. 2.26(a). Discussion of the nature of
equilibrium, in this case, is obviously less simple than in that of the models
treated in the previous Sections, as the potential energy is a function of two
variables.
It is known from analysis that for P($i, 02) to be positive definite at a
point, the first derivatives must vanish and also the matrix of the second
derivatives must be positive definite, that is, according to the Sylvester cri-
terion (see Section 1.6)
d2P «i *IM
= *! \ (
d0\
d2P d2P
x (l-p)(c-p)-p(l-c)^ (2.12.13)
~dej ~d$J \d01d02J
1 2 ^1 2 u
2 j *? 1
+ P(l - e) p >o
7
68

must be satisfied at the point. In our case, if we write the series expansion of
P{0\, 02) starting from a configuration of equilibrium of coordinates (0{, flj),
we have

+ (9 s:)(9 + (2i2i4)
(S,,,. '- '-« -
as the first derivatives in (Q{, 6\) vanish according to the equilibrium, and we
have set P(0{, Q*2) = 0. Then (2.12.14) is positive definite if the conditions
(2.12.13) are satisfied. If the terms of second order vanish, the test on the
stability of the equilibrium needs to be transferred to terms of a higher order
(see Section 2.7).
Along the fundamental path / \ (2.12.13) leads to

*i (1 - p) > 0

kx(c- p) > 0 (2.12.15)

kl{l-p){c-p)>0
Under the hypothesis c > 1 the three inequalities are satisfied for p < 1 and
violated for p > 1, and so the equilibrium is stable up to the first bifurcation
point and unstable above it. At the bifurcation point in particular it is
immediately clear that the potential energy (2.12.2) is positive definite. In
fact, by observing that there are no cubic terms and applying the criterion
(2.7.29) with dn = 0 , d22 = c — 1, dn2 — 0 and d i m — 1> we get dim = 1?
and so the equilibrium is stable.
Along the bifurcated path P2 we have

2 kx (p - 1) > 0

kx (3 + c - 4 p) > 0 (2.12.16)

2k\(p- l)(3 + c - 4 p ) >0


from which we deduce t h a t the three inequalities are satisfied in the interval
1 < p < (3 + c)/4 and are violated outside the interval, and therefore the
equilibrium is stable in that piece of the path between E and F and unstable
in the rest. Along the path P$ the conditions

kx (1 - 4 p + 3c) > 0

2kx {p-c) > 0 (2.12.17)

2 f c J ( p - c ) ( l - 4 p + 3c) > 0
69

hold, which are satisfied for p < (1 + 3 c ) / 4 and p > c. As p cannot satisfy
both inequalities it follows that relations (2.12.17) are violated and therefore
the equilibrium along the path ?z is unstable.
In correspondence with the curves P4 and ,P5, relations (2.12.13) reduce to

*1 {\ + lc-r)>°
e
kl(\+ --P)>o (2.12.18)

k\ - 8 p 2 + 8p(l + c ) - - c 2
- 5 c - - >0

We can see that whilst the first two of equations (2.12.18) are satisfied for
p < (3 4-c)/4, the third is always violated with these values of p and therefore
the equilibrium along the curves P4 and P$ is unstable. The results discussed
are shown in Figs. 2.25 and 2.26, where the branches of the equilibrium curves
with stable equilibrium are indicated by continuous lines, and those branches
corresponding to unstable equilibrium by dashes.
Let us now examine the case of the model with initial imperfections <pio
and <£>20 corresponding to the unloaded springs. By introducing the angles
#io — TT/2 — <P\o and 02o = TT/2 — ^20 and again indicating the angles relative
to the generic configuration by 0X and 02> w e have the following modified
expression of the potential energy

P{0\,02; Oio,02o) — An -\01—29i 0ioj + - \02 ~~ ^0 2 #20)


L2
_p(e_l +
e
_l_?L-k + W\ (2.12.19)
y K
\2 2 24 24 4 J\ '
under the hypothesis that the initial imperfections are small, in the sense
discussed in the previous Sections. The equations of equilibrium

0,-0 10 = 0
V 6 2
(2.12.20)
c(*2-M-p(*2-f + ^ l = 0
correspond to (2.12.19). The solutions to these equations provide natural
equilibrium curves which intercept the horizontal plane and unnatural equi-
librium curves for pre-determined values of the initial imperfections 01O and
020- In Fig. 2.27 we have a single curve of natural equilibrium together with
the equilibrium curves of the perfect structure. As seen in the models in
Sections 2.9 and 2.10, we observe that such a curve reaches a maximum p*
and then decreases. The value of p* depends both on the values of 0i O , #20
70

Fig. 2.27 - Equilibrium curves of the perfect and imperfect structures.

and on those of the coordinates of the points of secondary bifurcation E and


F , which are a function of c.
A detailed analysis from such parameters of the dependence of the limit
load p* is quite complicated, and outside the scope of the present treatment.
We believe it is sufficient here to give the qualitative behaviour of the phe-
nomenon, which shows how the Augusti model represents a type of structure
sensitive to initial imperfections. It is finally demonstrated that the equi-
librium is stable along the increasing piece of the equilibrium curve of the
imperfect structure, and unstable along the decreasing piece.
We complete this Section by examining how the preceding results are mod-
ified under the hypothesis that the two springs have the same stiffness, that
is c = 1. The bifurcation points along the fundamental path coincide and
therefore the curves P2 and Pz intersect at p = 1. Furthermore, the points
of secondary bifurcation also coincide with the bifurcation point on the path
Pi, since the abscissa of points E and F reduces to zero and the ordinate to
1. The projections of curves >P4 and P5 on the horizontal plane reduce to two
straight lines passing through the origin 0i/O2 = ± 1 , whilst the projections
on planes p — 0\ and p — 62 are two identical parabolae. Figures 2.28 and
2.29 illustrate the new situation. The conditions of stability (2.12.15) on the
fundamental path are again satisfied for p < 1 and violated for p > 1, whilst
the stable piece between points E and F of the curve P2 disappears. By
applying (2.12.16), (2.12.17) and (2.12.18) we finally verify t h a t along the
curves P2, P3 , P4 and P5 the equilibrium is unstable. The results are plotted
71

A / perfect stable
/ % perfect unstable
t
K
/ / imperfect stable
, \\
i
/
y
,
1
X
V imperfect unstable
1 \

! *A
l/ /M
i u
/ : 1D W\
/ i '
! 3
A
/1
'
'
!
1 1/ ^
/» -" r
\
\ / ■ '' * i ^
"/ N
I
V* / \'

'*,

Fig. 2.28 - Equilibrium paths of the perfect and imperfect structures


for c = 1.

I'P
v-\\3\ n

^i i3; ^ *

\\ % /
\\ perfect stable
\\ / //
\ perfect unstable
\) imperfect stable
/ /\ imperfect unstable
/ /
//

r*2

Fig. 2.29 - Projections of equilibrium curves on coordinate planes for c = 1.


72

in Figs. 2.28 and 2.29, where a natural equilibrium curve of the imperfect
structure has also been traced. Obviously, unlike the case c > 1, the value of
p* depends only on the amplitude of the initial imperfections.
73

REFERENCES
[l] J. La Salle, S. Lefschetz: Stability by Liapunov's direct method with ap-
plications, Academic Press, New York, 1961.

[2] F . Gantmacher: Lectures in analytical mechanics, MIR, Moscow, 1970.

[3] L. Meirovitch: Methods of analytical dynamics, McGraw-Hill, New York,


1970.

[4] H.H.E. Leipholz: Stability theory, Academic Press, New York, 1970.

[5] J.G.A. Croll, A.C. Walker: Elements of structural stability, Macmillan,


London, 1972.

[6] W.T. Koiter: Theory of elastic stability, lecture notes, University of


Technology, Delft, academic year 1972-73.

[7] J.M.T. Thompson, G.W. Hunt: A general theory of elastic stability,


J. Wiley & Sons, London, 1973.

[8] W.J. Supple: Aspects of postbuckling behaviour, lectures delivered at


the Istituto di Scienza delle Costruzioni of the Faculty of Engineering,
University of Rome "La Sapienza", March 1980.
75

Chapter 3

ANALYSIS OF BIFURCATION FOR


DISCRETE SYSTEMS.
CHARACTERISATION OF THE POINTS
OF A N EQUILIBRIUM PATH FROM
EXAMINATION OF LOCAL PROPERTIES

3.1 INTRODUCTION
Let us examine a conservative system of n degrees of freedom subjected to
external forces which depend only on one parameter A and with holonomous,
perfect constraints. Let

P(gi;X)=W(qi)-V(qi;X) (i = 1 , 2 , . . . , n ) (3.1.1)

be the total potential energy function which describes the system under
examination, where the arguments of P are the n Lagrangian coordinates
qi and the parameter A. In the following, unless otherwise specified, it is
supposed t h a t the subscripts vary from 1 to n. As we know, the conditions
of equilibrium of the system correspond to those of stationarity of P with
respect to the Lagrangian coordinates, t h a t is

Pti{qj ; A) = Wti(gj) - Vti(qj ; A) = 0 (3.1.2)

where (•),,- = d(-)/dft.


All the qi, A solutions to equations (3.1.2) which give rise to regular curves
in the space of n + 1 dimensions <fc , A can be thought of as descriptions of the
evolution of the system under the action of external forces, and in this sense
reference is made to them as equilibrium paths.
Only in very particular cases is it possible to determine explicitly all the
solutions to problem (3.1.2), so in general, in relation to the problems under
examination, we must have recourse to approximate solutions or to local
investigations. Furthermore, as equations (3.1.2) are generally non-linear, it
is possible t h a t under certain conditions the solution ceases to be unique and,
consequently, through a point in the space g t , A two or more equilibrium
paths can pass (bifurcation points).
Phenomena of this type are assigned great importance in the study of the
response of mechanical systems, as the loss of uniqueness of the solution allows
76

the system to pass from one path to another with, sometimes, substantial
changes of behaviour. An analysis which tries on the one hand to investigate
the problem of the determination of the bifurcation points, and on the other
to study the behaviour of the system at least in their proximity, is therefore
necessary.
Here we limit ourselves to considering the case of simple bifurcation, in
which we have two equilibrium paths which pass through the same point.
In this context, we observe that the behaviour of the simple discrete models
illustrated at the end of the previous Chapter is of this type.
With the aim of outlining the phenomenon of simple bifurcation, it is useful
to give an analytical characterisation of some of the properties of the points
of an equilibrium path. Using the elements emerging from such an analysis,
we shall then pass to the study of the problem of the search for possible
bifurcation points in a known equilibrium path, and to the construction of
the bifurcated path in asymptotic terms. We then go on to illustrate the
calculation procedures by means of a study of a simple model of two degrees of
freedom and we shall discuss critically the limitation, apparently very serious,
imposed by the fact that all that is said is applicable only to the case in which
an equilibrium path is known.

3.2 LOCAL ANALYSIS OF THE PROPERTIES OF


P O I N T S BELONGING TO A N EQUILIBRIUM
PATH
Let us suppose that the mechanical system under examination is described
by (3.1.1) and that Q is a known equilibrium point. Let us suppose that Q is
not isolated, in the sense that at least one equilibrium p a t h C passes through
it. A basic condition for the successive developments is that C is regular, in
that it can be expressed in the form

Qi = qi[t)
(3.2.1)
A = X(t)
where <fc(£) , X(t) are continuous functions of an auxiliary parameter t with
continuous derivatives up to the order desired.
Because of the regularity of (3.2.1) it is possible to write MacLaurin series
expansions with initial point Q

Qi = Qio + fto t + - qi0 t2 + o(t2)


(3.2.2)
2 2
A = \o + \0t + -i0t + o{t )
77

assuming, without losing generality, that Q has coordinates

too = tt (0)
(3.2.3)
A0 = A(0)

From now on, until it is otherwise explicitly established, we shall indicate


by (") differentiation with respect to t and with subscript ' 0 ' all functions
evaluated at t = 0.
The equations (3.2.2) show that for an approximate expression of (3.2.1) it
is necessary to know the n+1 groups of quantities <7t0 , A0 , qi0 , A0 , . . . which
describe the local properties of C in Q. The more of these quantities known,
the better the approximation with which C is known.
It is possible to develop a method which allows easy calculation of the
quantities mentioned, and we shall deal with this in more detail in later
Sections.
Instead, we discuss here some properties of the points of an equilibrium
path connected to the phenomenon of bifurcation and, as will be seen in the
following, to the stability of the equilibrium configuration.
It should be observed that equations (3.2.1) describe, by definition, an
equilibrium path, and therefore by substituting these relations in (3.1.2) we
obtain a function of t which is identically zero

Pii(qj(t);\(t)) = 0 Vt (3.2.4)

By now differentiating, up to the order desired, we have

^ {P.-(«>(*); A(*))} = 0 VA,Vt (3.2.5)

and in particular

^ {P,-(g,-(0;A(0)}« = o = 0 Vfc (3.2.6)

The relation (3.2.6) corresponds, for each value of k, to a linear algebraic


system of n equations in n -f 1 unknowns. Equation (3.2.6), when we set
k — 1, gives the linear homogeneous algebraic system

P0)ij qj0 + Po,iX A0 = 0 (3.2.7)

where [Po,ij] — [P,ij{t)]t=o ls a matrix of dimensions n x n and


[■Po.tA] — [-P,tA(£)]*=o is a vector of dimensions n x 1. Using

L0EE(P0,,> Po.ix) v T = (</,- A) (3.2.8)


78

we can write (3.2.7) in the form

L0v0 = 0 (3.2.9)

Denoting the rank of L 0 by p(L 0 ), the following situations may occur

p(Lo) = n (3.2.10)

p(L 0 ) < n (3.2.11)

Before going on to discuss the two cases (3.2.10) and (3.2.11), we remind
the reader of two theorems of linear algebra on the existence and number of
independent solutions to a system of linear equations. In the form shown in
[10] they are
I . R o u c h e - C a p e l l i T h e o r e m . The system A x = y is compatible if, and
only if, the matrices A and A x have the same rank fA x is the matrix obtained
by adding the column y to A).
I I . The solutions to a homogeneous linear system of n equations in n un-
knowns constitute a vector space of dimension n — p(A) (that is, they are
OQn-p(A)jm

As far as the search for solutions is concerned, we must remember t h a t


they are found by limiting ourselves to considering p equations associated
with a non-vanishing determinant.
Let us analyse the case in which (3.2.10) occurs together with p([Po,ti]) = n»
For theorem I, system (3.2.7) is compatible whilst theorem II assures that its
solutions are oo 1 due to the indeterminacy of A0. By bearing in mind (3.2.2),
we can see that all the solutions describe the curve

<7t = fto + <7t01


(3.2.12)
A = AQ + AQ t

that is, the tangent to C in Q. The indeterminacy of the numerical values


depends on the arbitrary choices allowed for the parameter t. So we can
conclude t h a t
a sufficient and necessary condition for a point Q to belong to one and only
one regular equilibrium path is that p(L 0 ) = n.
A case of particular interest is that in which p(L 0 ) = n together with
p([-Po.ti]) = rc — I- It is easy to see from (3.2.7) that this implies A0 = 0. This
signifies that the projection on the axis A of the tangent to the equilibrium
curve through Q degenerates at one point, t h a t is the tangent plane to C in
Q is orthogonal to A. In Fig. 3.1, the situation in a two-dimensional space
q, A is shown.
79

Fig. 3.1 - Limit point.

Points of this type denote stationarity for the load and for this reason are
generally called limit points. It is therefore possible to state t h a t
a sufficient and necessary condition for a point Q belonging to a regular equi-
librium path to be a limit point is that p(L 0 ) = n and p([Potij]) — n — 1.

Finally, we observe t h a t in the case P0,tA = 0, the only solution to (3.2.7)


is qio = 0 , A0 7^ 0, when Q belongs to an equilibrium p a t h whose tangent has
null projection onto the sub-space </,, that is, it is parallel to the axis A. In
Fig. 3.2 this situation is shown in a two-dimensional space X-q.
If the p a t h C is linear, then it coincides with its tangent in Q, and the
system behaves as if it were rigid. An equilibrium p a t h of this type has been
called fundamental in the examples given at the end of the previous Chapter.
We must now consider the solution qi0 = 0 , A0 = 0 associated with the
case P0,«A — 0- Anticipating the results, we can see t h a t the perturbation
equations succeeding (3.2.7), obtained through (3.2.6) by setting successively
k — 2 , 3 , . . . give in this case qi0 = 0 , A0 = 0 , qi0 = 0 , A0 — 0 etc., and
therefore from (3.2.2) g, = <7t0 , A = A0 and the point Q is isolated. This
contradicts the hypothesis that an equilibrium curve C passes through <3,
and for this reason the solution &0 = 0 , A0 = 0 is discarded.

Fig. 3.2 - Limit point for the displacement.


80

Let us now examine what happens when (3.2.11) occurs, making reference
however only to the case p(Lo) = n — 1. By again using theorems I and
II, we can state that system (3.2.9) is compatible and admit oo 2 solutions.
Bearing in mind the arbitrary choice of the parameter, we should conclude
that there are oo 1 curves of the type (3.2.12) passing through Q. In reality
this is not so, and we can only say t h a t in this case relations (3.2.7) or (3.2.9)
are no longer able to give all the information on the coefficients of (3.2.12).
To convince ourselves of this, it is sufficient to write (3.2.6) for k > 1 and
note t h a t the coefficients of (3.2.12) appear in all the relations together with
the other coefficients of (3.2.2).
Leaving out further details, it is sufficient to know here t h a t although in the
case previously examined (p(L 0 ) = n) the solution to the algebraic problems
which we get from (3.2.6) for each value of k depends only on the solution to
all the problems of the order i < A:, in the case p(L 0 ) = n — 1 it depends also
on the problem of the order i = k + 1.
To clarify the solutions to (3.2.9), it is therefore necessary to write (3.2.6)
with k — 2. This reads

Po,ij QJO + Po,iX Ao + Po,ijk Qjo Qko + 2 Po,ij\ <7yo A0 + Poti\\ A0 = 0 (3.2.13)

which can also be written in the form

L0v0 = b0 (3.2.14)

having introduced the vector b 0 = M 0 v 0 v 0 with M 0 a matrix of the third


order.
Still bearing theorem I in mind, it is clear t h a t when L 0 is not singular,
(3.2.14) admits solutions V b 0 , while if Lo is singular, solutions exist provided
that

p(L 0 ) - p(S 0 ) (3.2.15)

where S 0 = [L 0 : b 0 ] .
We now show how the condition of compatibility (3.2.15) leads, in our
case, to an algebraic equation of second degree in the components of v 0 . In
fact, let us consider the homogeneous problem

L£ x = 0 (3.2.16)

where L j denotes the transpose of L 0 . For each of the x solutions to (3.2.16)


the relation

x T L0 v0 = x T b 0 = 0 Vx (3.2.17)
81

must hold. This is generally called a bilinear identity or condition of orthog-


onality and is a different way of expressing (3.2.15), that is theorem I, as
known from algebra. L j is a matrix of n + 1 rows and n columns of rank
/?(LQ ) = p(L 0 ) = n — 1. Having chosen a minor of maximum order, the oo 1
solutions to system (3.2.16) are written as

%i = a, xa Via (t — 1 , . . . , n) (3.2.18)

with aa = 1.
Similarly, the oo 2 solutions to system (3.2.9) are

v,-o = c{ vp0 + di v^ V vp0 , v^o (t = 1 , . . . , n)


(3.2.19)
A = c\ vp0 + dx VTO

where c t , d,-, c^, dA are coefficients with cp = 1, d^ = 0 and with c 7 = 0 ,


d1 = 1. Equation (3.2.17), showing the dependence on xa , vpo, t57o> becomes

x T (x a ) b 0 (fy 0 , vl0) = 0 (3.2.20)

Bearing in mind (3.2.13), this can be written in the explicit form

\Potijk [Oi Cj ck v2p0 + (at- Cj dk + at- cjt dy) v^o v^o + «t- dy djk t)20J

+ 2 Po.ijA [fli Cj cx Vp0 + (at- Cj dx + a{ dj cA) vpo V70 + a i dj dx ^ \

+ Po.ixx (di c\ v}0 + a{ cx dx vpo v^ + at- d\ v^j } xa = 0 Vxa


(3.2.21)

which implies the vanishing of the expression inside the braces. Excluding
the case vp§ — 0 , t)7o — 0 corresponding to an isolated equilibrium point, it is
possible to state that all the vectors v 0 , and only they, which at the same time
satisfy (3.2.9) and (3.2.21), give rise when substituted in (3.2.12) to straight
lines which are tangent to the equilibrium paths through Q.
As (3.2.21) is a quadratic equation, it has two solutions v^o/v^. If we
consider the case in which these are both real, we can say that
a necessary and sufficient condition for two regular equilibrium paths to pass
through the point Q is that p(L 0 ) = n — 1.
Note that this condition does not exclude, obviously, that some (not all)
of the components of v 0 are zero in one or both of the solutions. Then
the tangent to that path will have a null projection on the axis or axes in
correspondence with which the components of v 0 vanish. In particular, if we
have A0 = 0 along a path, then Q is a bifurcation point and a limit point at the
same time. Some of these cases are illustrated in Fig. 3.3 in two dimensions.
The occurrence of the condition 0 < p(L 0 ) < n — 1 expresses the fact
that Q belongs to more than two equilibrium paths. We do not take this
82

q ' q

Fig. 3.3 - Cases of simple bifurcation.

circumstance into consideration here except to mention the fact that, even
if the number of paths is related to the value of p(Lo), this relation is not
one-to-one, as happens in the case of p(L 0 ) = n — 1. We can see this if we
realise that in this case the solution to (3.2.16) is of the type

X{ = at xa + bi £p + (3.2.22)

By substituting (3.2.22) into (3.2.21), this gives rise to a system of non-


linear algebraic equations for which, ELS we know, the number of real solutions
is not determined.
The analysis of problems of this type has particular relevance in the sphere
of the "Theory of Catastrophies" [8].

3.3 PERTURBATION METHOD IN THE ASYMP-


TOTIC DETERMINATION OF REGULAR
EQUILIBRIUM PATHS THROUGH A POINT Q
Let us consider, as in the previous Section, a mechanical system described by
(3.1.1), where Q is a known equilibrium point belonging to a curve C which
is a regular equilibrium path in the sense pointed out previously.
We have seen how it is possible to think of an asymptotic representation
in terms of an auxiliary parameter t of the curve C, through the expressions
(3.2.2). It is evident from what was said in the previous Section that the
solution to the algebraic system discussed there requires the determination
of the coefficients of (3.2.12), that is, the construction of an approximate
expression for an equilibrium path through Q.
Whereas we have so far developed a procedure for a classification, though
only brief and limited, of the Q points, here we would like to re-examine the
procedure with the aim of showing how it is possible to institute a rigorous
method useful in all those cases where it is impossible or inconvenient to
search for solutions in the closed form (3.1.2) and we must be content with
83

an approximate solution of at least one equilibrium path, starting from the


knowledge (which is always available) of one of its points.
In this Section, therefore, we again take up the material discussed previ-
ously, with the intention of incorporating the relevant facts more precisely. A
correct interpretation of these facts will be the starting point for understand-
ing the consistent formulation of a procedure suitable for use in the study of
the bifurcation from a known path.
The method we want to discuss is known as the method of static pertur-
bation, which was introduced in rigorous form in the analysis of problems of
this type by M.J. Sewell [l],[3].
Each of the relations which come from (3.2.5), assigning a value to A:, is
called a perturbation equation of the order k. These equations are written
here up to the third order

Po,ij qj0 + Po.ix A0 - 0 (3.3.1)

Potij QJO + Po,i\ Ao + -Po.iy* Qjo Qko + 2 Po,ij\ fyo Ao + Po,ixx A0 = 0 (3.3.2)

Po,ij QJO + Po,ix A0 + 3 Po}ijk Qjo Qko + 3 Potijx ( Qjo Ao + £yo Ao )


+ 3 Potixx Ao A0 + Potiju <7yo Qko Qio + 3 Po,ijkX Qjo Qko Ao
+ 3 Pojjxx Qjo A0 + Po,ixxx A0 = 0 (3.3.3)

We note, first of all, that the matrices P0,iy > -RVA, appear in each of the
equations (3.3.l)-(3.3.3) independently of their order fc.
The problem of the construction of an approximate path through Q there-
fore leads to that of the solution to a succession of linear algebraic problems, of
which the first is homogeneous and those following are all non-homogeneous.
The essential feature of the perturbation method then becomes apparent, in
that the problem of the solution to a non-linear system of equations (3.1.2)
is transformed into that of the solution to a sequence of systems of linear
equations in which the fundamental role is given to the matrix L 0 obtained
by the linearisation of (3.1.2).
This last consideration helps us to understand the whole generality of
the method, as what is said in this field, or in that of algebraic systems,
is transferred exactly, as will be seen in later Chapters, to the differential
systems which we meet in the analysis of continuous systems. The name
tangent matrix is often given to Lo.
We can now understand the importance of the study of Lo and in particular
of its rank, as this obviously conditions the behaviour of equations (3.3.1)-
(3.3.3). Supposing that (3.2.10) is satisfied, each perturbation equation allows
oo 1 solutions in the increments v , v etc. of the vector v of components qt;, A.
84

Fig. 3.4 - Straight line x = 2y.

It may seem strange at this point that all of these oo 1 solutions describe
a single curve, which is the equilibrium path C through Q. This apparent
contradiction depends on the fact that in (3.2.2) the auxiliary parameter t
has not been specified. For instance, the straight line in Fig. 3.4, can be given
a representation in the form

(3.3.4)
\ y = bt
where a/b = 2.
The choice of different values of the pair (a, b) means only a scaling for t.
In fact, if we indicate by tA , tB the values which t assumes corresponding to
points A and B of d = r (Fig. 3.4), we have

for (2,1): tA = l tB = 3
(3.3.5)
for (4,2): tA = 0.5 tB = 1.5

The indeterminacy of the solution to the perturbation equations can be


removed by normalising it, thus prescribing the choice of the parameter t.
Since the choice of t is not immaterial, for reasons which will be explained
later, it is in general useful to proceed by deriving the condition of normal-
isation from this choice, rather than vice versa. Operating directly on the
parameter, on the other hand, at times allows its identification with quanti-
ties which clarify the physical interpretation of the results.
It is generally appropriate, at least in the sphere of mechanical problems, to
choose t from among the class of linear combinations with constant coefficients
of the increments of v starting from Q. This is equivalent to writing

t = /.-(*-*o)+f(A-A0) (3.3.6)

Bearing in mind (3.2.1), we have

t = U (ft(0 - fto) + / (A(0 - A0) (3.3.7)


85

By writing (3.3.7) in incremental form, up to the third order, we have

1 = *.•*(*)+**(<) (3-3-8)

0 = liSi{t)+li{t) (3.3.9)

0 = lili{t)+li{t) (3.3.10)

which calculated at Q, that is, for t = 0, give

1 = /,' fto + / A0 (3.3.11)

0 = /,• h0 + / Xo (3.3.12)

0 = /,- L + / *o (3.3.13)

Equation (3.3.7) is the condition of normalisation, which is to be associated


with (3.1.2) to obtain (3.2.1) unequivocally. As (3.1.2) has been replaced in
the perturbation procedure by (3.3.1)-(3.3.3), it is with these that equations
(3.3.11) to (3.3.13) must be associated to give (3.2.2) unequivocally.
Note that the non-homogeneous equation (3.3.11) in Vo is associated with
(3.3.1) which is homogeneous, whilst (3.3.12) and (3.3.13), which are homo-
geneous, are associated with (3.3.2) and (3.3.3) respectively, which are non-
homogeneous. In fact, in a perturbation equation of order fc, there appear the
&-th increments of v as well as all its increments up to the order k — 1, which
are all known since they follow from the solution to the previous perturbation
equations.
This explains what was said in the previous Section, that each perturbation
equation depends only on the solution to the preceding problems or, in other
words, a problem of higher order does not influence those of lower order.
We again notice that since (3.3.11) is non-homogeneous, the possibility that
all the components of v 0 vanish simultaneously and that Q is therefore an
isolated equilibrium point, is excluded.
A common choice, in problems of mechanics, is to identify the parameter
t with the increment of A, that is [6]

t = A - A0 (3.3.14)

This implies the conditions of normalisation

1 = A0 , 0 = X0 , 0 = X0 (3.3.15)
86

It is therefore necessary that, for the problem under examination, we have


A0 7^ 0. Thus we must exclude this choice of the parameter in the case in
which Q is a limit point.
Obviously this is also valid for any other component of v 0 , and in general
we can say that it is not possible to identify the parameter with a component
va of v , if the tangent to C in Q has a null projection upon it, that is, if the
regular curve C is not locally invertible in Q in terms of va [9].
At this point, we have an asymptotic expression for C with initial point
Q in terms of the chosen parameter £, and the higher the order of the last
perturbation problem solved, the better the approximation of (3.2.2) in the
neighbourhood of Q. Such an approximation nevertheless depends greatly
on the choice of the parameter £, as will be shown in a numerical example
reported in Section 3.6.
What must be underlined here is the possibility of obtaining more expres-
sions of the type (3.2.2) for the same C, starting from a known expression,
each in terms of a different parameter, for all or some of the components of
v. Indeed, let

r] = bi{qi - qiQ) + 6(A - A0) (3.3.16)

where the coefficients 6t and b respectively are not all equal to the coefficients
/, and / which appear in (3.3.6). Considering a generic component g, of v , we
want to obtain the expression

* f a ) = *o + <7,'o n + - <fio 1* + o(r/2) (3.3.17)

where the dash denotes differentiation with respect to rj and the suffix 'o'
identifies the values of the functions calculated at the point rj = 0, which for
(3.3.16) corresponds to t = 0.
According to (3.2.1) and (3.3.16) we can also write

V{t) = bi {qi{t) - qi0) +b{X{t) - A0) (3.3.18)

where r\ [t) is a regular function of t since it is a linear combination of regular


functions of t. We can therefore give an asymptotic expression for it with the
initial point t = 0

2
^W = */o* + ^ o ' + o(*2) (3.3.19)

By substituting (3.3.19) in (3.3.17) and neglecting terms which are higher


than the second order, we have

* (*) = fto + <lio *<>t + \ (<& ^o + rfo tfo) *2 + °(' 2 ) (3-3-20)


87

which, from comparison with the first of relations (3.2.2) gives

9,!o = fa/to (3.3.21)

9,"o = {ho ~ 9.o hoi no) / vl (3.3.22)


Now differentiating (3.3.16), we get at t = 0

Vo = bi 9.o + b A0 (3.3.23)

ho = bi ho + bi0 (3.3.24)

from which, by substituting in (3.3.21) and (3.3.22), we have

9,'o = W {bj Qjo + b A0) (3.3.25)

z <lio{bjqjo + b\0)
9,"o = /(bjqjo + bXo)2 (3.3.26)
i Oj qjo + o A0 .

which are simple algebraic expressions providing the unknown quantities


present in (3.3.17) as functions of the given quantities 6,, b and of the quan-
tities v 0 , v 0 furnished by the previous analysis.
We can immediately see from (3.3.21) and (3.3.22) that, if the function q{
is not locally invertible in Q in terms of rj (that is, if fj0 = 0), then (3.3.25)
and (3.3.26) lose their significance.
Let us now consider the case in which p(Lo) = n — 1, which as we have
already stated, implies simple bifurcation. In this case (3.3.1)-(3.3.3) do not
give a single solution. In fact, the algebraic system consisting of
• the perturbation equation of the first order (3.3.1)
• the condition of normalisation (3.3.11) written twice, and
• the condition of orthogonality (3.2.21)
provides two solutions for v 0 , which we shall call VQ and VQ . Note that the
sequence of the operations indicated corresponds to the identification of the
two curves and to choosing the normalisation on these separately.
Having determined the first-order quantities, we can then write the pertur-
bation equation of the second order (3.3.2) which, according to the solutions
found, splits into

POM 9,o + Po.iX Ao = - (Po,,;* 9 $ $ + 2 Pom $ W + Po.ixx W)


(3.3.27)

Po,a 9 i 0 + ^0,-A *o = - {Po,ijk «}? &S + 2 Po.ax 9 ^ W + Po.ixx W)


88

that is, bearing in mind (3.2.14),

L0 v 0 = b ^
(3.3.28)
T * 1.(2)
L 0 v 0 = b^ '
As VQ and VQ ' have been obtained in such a way as to satisfy (3.2.21), both
systems (3.3.28), according to theorem I, are compatible and, by theorem
II, allow oo 2 solutions. It is clear in this case, as in the previous one, that
because v 0 appears in the perturbation equation of the third order, it must
satisfy a condition of orthogonality.
In particular, in order that the perturbation equation of the third order
also allows solutions it must be, analogously to (3.2.17),

xrLo4o=*r(c?, + 4 1 ,
)=° V
*T
(3.3.29)
xrLov0=xr(c(2) + d(2))=0 V F

where x is the solution to system (3.2.16), c£ lj = T 0 v £ 1 } v ^ v ^ ,


<£> = T o v f ' v ^ v ? 1 , d<1} - M o v ^ v o , 4 2 ) = M 0 v < 2 ) v o , and T 0 is a ma-
trix of the fourth order. Thus, Co contains cubic terms in the components of
v 0 whilst d 0 contains bilinear terms in the components of v 0 and v 0 . From
(3.3.3), bearing in mind (3.3.18), we can write (3.3.29) in explicit form

13 \Potijk a>i <$ Qko + po,ijX ai U$ *o + §yo ^oj + Po,iX\ ai ffl *o

+ P n nW \{1)2 + P n nW n{l) A ( 1 ) l + P a n{1) nW nW


+ -n>,ijAA ai QjO ^0 + "0tijk\ &i QjO QkO A0 + "o,ijkl a% QjO QkO <7/0

+ Po.ixxx a,i\y >xa = 0 V xa


(3.3.30)

13 \P0tijk a>i qfo Qko + Po,ij\ «• Ufo A0 + lj0 ^o]) + Pojxx a. ^>2) A0

+ Potijxx a, $ A<2)* + P0tijkx ^ $ gg J A<2)] + P0tijkl a, $ $ $

+ Potixxx a>i A^2) I xa = 0 Via

The solution to the perturbation problem of the second order is obtained


by solving the two algebraic systems which consist of

(a) first system: the first of equations (3.3.27), (3.3.12) and the first of
(3.3.30);
89

(b) second system: the second of equations (3.3.27), (3.3.12) and the second
of (3.3.30).

Both of these systems are linear and therefore each of them has a solution,
which we shall call v 0 , v 0 . When these solutions are associated with VQ '
and VQ respectively, two groups of coefficients are obtained which, when
inserted into (3.2.2), give rise to the approximate expressions of the second
order of the two paths through Q.
Should we wish to pass to a third order solution to the perturbation prob-
lem, we would find a solution with analogous structure to that described for
the problem of second order, that is to say, we would arrive at two systems
of linear equations, for which we believe it superfluous to go into detail. We
can therefore conclude that in the case in which L 0 is once singular, that is,
p(L 0 ) = n — 1, (3.3.l)-(3.3.3) give rise to a sequence of algebraic systems
made up as follows

(a) to the first order: a system of n linear equations plus a quadratic equation
of orthogonality;

(b) to all successive orders: two systems of linear equations.

With respect to the previous case in which p(L 0 ) = n, we now find ourselves
in the presence of the following substantial differences:

(a) instead of a sequence of linear algebraic systems there is a double se-


quence of such systems with the addition, to the first order, of a non-
linear equation;

(b) the solution to the perturbation problem of the r-th order depends on a
condition which must be imposed on the perturbation equation of order
r + 1.

It is useful here to repeat that it is possible to express the two equilibrium


paths in terms of two distinct parameters. To do this, it is sufficient to
write the two different expressions (3.3.6), form two systems (3.3.11)-(3.3.13)
and use the quantities of the first system to normalise the increments of
VQ and those of the second to normalise the increments of v^ \ We must
again note that, even if to solve the perturbation problem of order r it is
necessary to resort to that of the order r + 1 , the condition of orthogonality
contains increments of v of the order r only [6]. This is of fundamental
importance, because making the problems for various orders self-contained
permits practical use of the method.
90

3.4 SEARCH FOR CRITICAL P O I N T S ALONG A


K N O W N EQUILIBRIUM PATH A N D LOCAL
ANALYSIS OF T H E B I F U R C A T E D PATH
I N T H E V I C I N I T Y OF T H E B I F U R C A T I O N
POINT
Having assumed a knowledge of a non-isolated equilibrium point Q for the
mechanical system described by (3.1.1), criteria for characterising the point
were given first in the preceding Sections, examining in particular the case
of the simple bifurcation of the equilibrium path through it. A procedure
aiming to furnish asymptotic expressions of the equilibrium paths for Q was
described, always limited to cases in which their maximum number was two.
We now suppose we know an equilibrium path Q for system (3.1.1), or a
finite part of it; let

Qi = 9i{t)
(3.4.1)
A - g(t)

be its expression in terms of a parameter t.


In this Section we shall deal first with the problem of the search for the
possible bifurcation points along the known path and then, in the case of
the existence of such points, of the construction of an asymptotic expression
for the only bifurcated path about the point of bifurcation. To identify the
possible bifurcation points it is sufficient to write a relation analogous to
(3.2.9) or to (3.3.1), deriving it, however, from (3.2.5). In this case we get

L(t) v = 0 Vt (3.4.2)

that is, in explicit form

P.H (ft(«)! 9(t)) 9j{t) + P,iX (9i{t); g{t)) g{t) = 0 V* (3.4.3)


As the condition of bifurcation implies p(L 0 ) = n — 1, (3.4.2) is the ex-
pression of the algebraic problem of the search for the values of £, which we
shall indicate by ta, corresponding to which the matrix L(£) is singular. We
can see that L is linear in v but not, generally, in t. The values of ta thus
determined identify the set {<?&} of all the bifurcation points belonging to the
known path. We can therefore say that
the problem of the search for bifurcation points along a known equilibrium
path is reduced to the algebraic problem of the search for the t values of the
parameter which make the linear operator L singular.
Now that the bifurcation points are known, since for each of them what was
said in the previous Section is valid, the second problem which was posed, that
91

of the asymptotic construction of bifurcated paths starting from a bifurcation


point, is solved.
We can say at this point that the target which was set has practically
been reached, even though some considerations on the operative level of the
instruments used for the investigation, and on the opportunity of introducing
some modifications which increase their efficiency, impose themselves. The
procedure presented in the previous Section leads to the determination, in
asymptotic terms, of both the equilibrium paths passing through the bifurca-
tion point, and it has also been observed how this is closely connected with
the existence of equation (3.2.21), which is the only non-linear relation to
appear in the whole perturbation process. The latter brings about two kinds
of complication. The first lies in the non-linear nature of (3.2.21), which may
cause difficulties, above all in applications of automatic calculation. The sec-
ond complication is the occurrence of pairs of linear algebraic equations at
each successive level of perturbation analysis.
Both of the difficulties mentioned have, as a counterpart, the fact that
they furnish the asymptotic expressions of both equilibrium paths through the
bifurcation point; but this becomes irrelevant when one of the two equilibrium
paths is known.
We can ask ourselves at this point if it is possible to modify the per-
turbation procedure in such a way as to clear the procedure of both the
inconveniences mentioned.
This is, in fact, possible, and to obtain what we want it is sufficient to
express (3.3.l)-(3.3.3) in terms of new Lagrangian coordinates. Such a trans-
formation, in general, is expressed by the relation

<li = /t K ' , t)
(3.4.4)
A - /(uy,t)

where Uj, t are the new coordinates. The transformation (3.4.4) is required
to satisfy the following conditions

/i(o,t) = gi{t)
(3.4.5)
/(o,*) = g{t)

where the terms on the right of (3.4.5) are the same as those in (3.4.1), that
is, the functions which describe the known equilibrium path. In the new
coordinates the old path is described by the relations

u{ = 0 Vt (3.4.6)
92

Fig. 3.5 - Incremental coordinate u.

and therefore it coincides with the coordinate line t. Of all the possible forms
t h a t (3.4.5) can assume, we consider the simple linear transformation

Qi = 9i (*) + Ui
(3.4.7)
A = g{t)
In a two-dimensional space q, A, the meaning of the symbols is illustrated
in Fig. 3.5.
Denoting by u the vector
Ui
u (3.4.8)
t
we can construct the new total potential energy
P{u) = P (u, ;t) = P {{gi{t) + u t ) ; g{t)) (3.4.9)
which describes the same system as described by (3.1.1) in terms of u . From
now on we indicate by the symbol ( )>t differentiation with respect to the
Lagrangian coordinates which appear as arguments of the function being
considered (that is, P t — dP/dqi whilst P t = dP /du{). The conditions of
equilibrium of our system, expressed through (3.4.9), are

Pti (uy ; t) = 0 (3.4.10)


All the Uj , t solutions to (3.4.10) which give rise to a regular curve in the
(n + l)-dimensional space to which u belongs are equilibrium paths of the
system, and coincide with those obtained by (3.1.2).
We now choose a point Qc G {Qb} of coordinates

qu = 9i{tc)
(3.4.11)
K = g{tc)
93

Through this, by virtue of what has been said, two equilibrium paths pass,
for which we can give a representation in terms of v and in terms of u. In this
second case, by adopting parametric expressions, the path (3.4.1) is written
as
Ui = 0
(3.4.12)
t = t(e)

whilst the bifurcated path is


u,- = Ui (e)
(3.4.13)
* = t(e)

where it is intended, without any limitation of generality, that corresponding


to the value e = 0 of the parameter, we have
t(0) = tc (3.4.14)
for which u,(0) = 0 follows from the first of relations (3.4.13). Therefore,
from (3.4.7), (3.4.11) and (3.4.14) we get
g,(*(0)) = qic
(3.4.15)
A(t(0)) = Ac
By virtue of the regularity of (3.4.13) we can give for these equations a
MacLaurin series expansion in terms of e

Me) = uice + -uice2 + o(e2)


i (3.4.16)
2 2
t{e) = tc + ice+-tce + o{e )

where the symbol (') denotes differentiation with respect to e and the suffix
c that the values of the function to which it refers are calculated at e = 0,
that is, at the bifurcation point.
Now bearing in mind (3.4.13), we can think of (3.4.10) along the bifurcated
path as a function only of the variable e and we can write
Pii(uj(e);t(e))=0 Ve (3.4.17)
which is similar to (3.2.4). By differentiating (3.4.17) with respect to e we
get a perturbation process equal to that in Section 3.2

v k v e
^ {P,i («,•(«); *(*))> = o » ( 3 - 4 - 18 )
94

and

^ to M O ;*(«))>«=„ = 0 V* (3.4.19)
which are analogous respectively to (3.2.5) and (3.2.6).
From (3.4.18) we derive the perturbation equations, which are in explicit
terms up to the second order
i
hi "i + h = ° (3.4.20)

fi
hi i + h * + hit "i Uk + 2 Ptij tiy i + Pti i2 = 0 (3.4.21)

An important simplification of (3.4.20) and (3.4.21) can be made imme-


diately by observing that the known equilibrium path (3.4.12) must be the
solution to (3.4.20) and (3.4.21). In this case, for the first of (3.4.12), there
must exist the asymptotic solution

in (e) = Ui (e) = u{ ( € ) = . . . = 0 Vc (3.4.22)

for any £(e), that is, for any i(e) , i(e) , . . . and this obviously implies

pi(e) = Pti{e)= ... = 0 (3.4.23)

Equations (3.4.20) and (3.4.21) are therefore written as

Ptij Uj = 0 (3.4.24)

fi
hi i + hit iy *k + 2 Ptij u, i = 0 (3.4.25)

To solve (3.4.24) and (3.4.25) it is necessary to construct the function


P(ui, t) first and this, even though it is a conceptually simple operation,
introduces notable computational difficulties in the procedure. It is therefore
appropriate to rewrite (3.4.24) and (3.4.25) having recourse, as far as possible,
to the known function P((fc, A) and to its derivatives. With this in mind,
we can see that for (3.4.9) and using (3.4.7), remembering that we indicate
differentiation of a function with respect to its own arguments by the comma,
we have

P,i^l,t) = -^{P{qi{m{t),u,);X{t))}

= P-(*(»(*),«!); A(«)) §*J


= P<fai(ft(t),«i);A(0) (3.4.26)

/>,,-, ( u , , t ) = P, t f («(ffj(*),«j);M')) {ZA.27)


95

and so on, having made use of the relation dq8/dui = 6{8 where 6i8 is the
Kronecker symbol. As far as terms in which the derivatives of P with respect
to t appear, we have

£.-(««. 0 = J^,-(«(ft(*).«i);A(t))}
= P,H (ft (gi (0. «i); MO) ^ + P,ix (ft (a (*). «i); MO) A
(3.4.28)

4(«<>0 = ^{^o-(«(ftW.««);M0)>
= P,m (ft (ft (<).««); MO) & + ^OA (ft (a (0. «0; MO) A
(3.4.29)
>,.•; («J , 0 = {P,m {q. (s.(0.«.); MO)»
+ J,«*A(?.(ff.(0.«.);M0)*ftk}
+ J,«*fa.(ft>(0.«.);M0)&
+ {*>«** («.(ff.(0>«.); MO) &
+ P,ijx\{q.{g.{t),u.);Ht))X}\

+ PiiiX {q, {g,{t),u,); A(0) A (3.4.30)

and so on. Note that (3.4.28) calculated at any point of the known path is
equivalent to (3.2.7), and this gives additional significance to (3.4.23).
Bearing in mind (3.4.27) and (3.4.7), we write (3.4.24) as

p,ij ((ft(0 + ««); s ( 0 ) «y = ° (3-4-31)

This is valid for any equilibrium path and therefore, in particular, for the
known path along which uj = 0. This allows us to write (3.4.31) in the form

^i(ft(0;»(0)«i = 0 (3.4.32)

Equations (3.4.32) are the formulation of the algebraic problem (generally


non-linear) for the search for eigenvalues t1 of the matrix [P,tj(01" They
identify the set of critical points {Qc} of the equilibrium path which is made
up of the union of {Qb} and the set of limit points. However, as Q is known,
{Qb} is immediately obtained by {Qc}, through the condition g(t) ^ 0. It is
important to note that expression (3.4.32), relative to (3.4.2), is characterised
by the important fact that the matrix [Pt%j] which takes the place of L is
symmetrical.
Returning now to (3.4.19), we can write the perturbation equations which
96

come from it. Bearing in mind (3.4.24)-(3.4.30) we can write in terms of


P(tt,A)

PcijUjc = 0 (3.4.33)

PcijUjc + Pc,ijkUjcUkc + 2(Pctik9kc + Pcjj\k)ujcic=0 (3.4.34)

S
PcM J« + 3 PCtijk Ujc Ukc + 3 (Petijh 9kc + PctijX A c ) (uje ie + Uje L)

+ Pcjjkl Ujc ^kc Ulc + 3 yPcijU 9lc + Pc,ijk\ ACJ Ujc Ukc te

+ 3 [^P C | fj« 9lc + PcjjkX K) 9kc + Pc,ijk 9kc

+ (Pc.ijkx he + Pc,ijx\ Ac) Ac + PCtijX AJ itjc t\ = 0 (3.4.35)

Relations (3.4.33)-(3.4.35) are the representation of (3.3.1)-(3.3.3) in terms


of the components of the vector u .
Although in appearance the new perturbation equations do not show any
particular differences with respect to (3.3.1)-(3.3.3), we shall show t h a t they
have a completely different structure which allows us to eliminate the prob-
lems connected with the analysis of bifurcation conducted through (3.3.1)-
(3.3.3). As the functions in (3.4.33)-(3.4.35) are calculated at a bifurcation
point, obviously p([-Pc,ij]) = *i — 1, and this implies the necessity of imposing a
condition of orthogonality on the known term of the perturbation equation of
the order r + 1 to solve the problem of order r. Such a condition is obtained
by requiring the known term of each (r + l ) - t h equation to be orthogonal
to the solutions to the problem [P 0 ,«y] r x = 0 (see eq.(3.2.16)). In this case
however, as the linear operator which appears in the problem is self-adjoint,
we can conclude t h a t
the condition of orthogonality for each perturbation problem of order r is ob-
tained by equating to zero the product of the known term of the perturbation
equation of order r + 1 with the solution to the perturbation problem of the
first order uic.
This is the first simplification which avoids the necessity of solving the
problem (3.2.16). Furthermore, we can see t h a t whereas, in the perturbation
equations (3.3.1)-(3.3.3), to each order r there appear all of the derivatives
of v up to the order r inclusive, in (3.4.33)-(3.4.35) the derivative of the t
component of u appears up to the order r — 1.
We pass now to an examination of the operational details of the solution
to (3.4.33)-(3.4.35). First of all, we write the explicit expressions of the
conditions of orthogonality on the perturbation equations of the second and
97

third order respectively, which are analogous to (3.2.21) and (3.3.30). These
are

{Pe,ijk V>ic Ujc Ukc + 2 (Pe,ijk Uic Ujc 9kc + PC,ij\ U>ic Ujc K) *C} U>ac = 0
V u a / 0 (3.4.36)

{ 3 [Pc,ijk Uic Ujc Ukc + \Petijk 9kc + Pc,ij\ K) (v>jc U + Uje *c) Uic

+ \Pc,ijkl 9lc + Pc,ijk\ K) U{e Uje Ukc tc + [{Pe,ijkl 9lc + PC,ijk\ K) 9kc

+ Pctijk 9kc + (Pc,ijk\ 9kc + Pe,ijXX K) K + Pc,ij\ K J Uie Uje tcJ

+ Pc,ijki u>ic ujc iikc uu] uac = 0 V uac ^ 0 (3.4.37)

We now denote by

Uic = *i uac V uac (3.4.38)

the solution to system (3.4.33), which is equivalent to system (3.2.16), rewrit-


ten in the new coordinates.
Equation (3.4.36) is a quadratic equation in ii c , which by using (3.4.38)
can be written as

[Pc,ijk ti tj ek uac + 2 (PCtijk t{ ej gkc + PCtijX e, e, Ac) *c] u2ac = 0 (3.4.39)

One solution to (3.4.39) is uac = 0, from which using (3.4.38) it follows t h a t


uic = 0. This solution represents the first-order part of the known equilibrium
p a t h . The second solution is obtained from the linear equation

Pc,ijk ti Cj ek uac + 2 [PCiijk ti tj 9kc + PCiij\ *i *j K) ic = 0 (3.4.40)


We now have all the elements necessary to proceed to the construction in
asymptotic terms of the only bifurcated p a t h . T h e perturbation problem of
the first order will consist of
• perturbation equation of the first order (3.4.33)
• condition of orthogonality (3.4.40)
• condition of normalisation (3.3.11) written, for example, in terms of the
increments of u with respect to e.

This is a linear algebraic system of n + 2 equations in the n + 1 unknowns


Uic j *o non-homogeneous by virtue of (3.3.11), where the matrix of coefficients
has rank n + 1. This has a unique solution ii c .
The perturbation equation of the second order will consist of
• the perturbation equation of the second order (3.4.34)
• the condition of orthogonality (3.4.37) which is linear in u c .
98

• the condition of normalisation (3.3.12) written in terms of u.

This also is a linear algebraic system of the same type as that obtained from
the perturbation problem of the first order, and has only one uc solution. We
proceed in the same way for all the problems of higher order, and thus we
again obtain the sequence of linear problems which allows us to treat the case
of simple bifurcation in the same way as that in which the solution is unique.
We notice that the structure of the perturbation problems, to any order,
does not change and is characterised by the presence of
(a) one of the equations from (3.4.33)-(3.4.35),
(b) a condition of normalisation,
(c) a condition of orthogonality.

If we use a condition of normalisation of the type

£ = /,. (ut- - uic) (3.4.41)

that is, a condition without the term in the increment of parameter £, we


can immediately see that the system constituted by (3.4.33) and by the in-
cremental equation to the first order, derived from (3.4.41), defines u tc in an
unequivocal way and so the condition of orthogonality determines tc. The
same occurs for all successive perturbation problems.
In the following, we give the explicit expressions of tc and ic

te = - - PcMkUicUicUkc ^ (3 4 4 2 )
2
Pctijk Uic Ujc ()kc + Pctij\ Uic Ujc Xc

tc = — j Pc,ijkl ^ic Ujc Ukc Uic + 3 Pc,ijk Uic Ujc Ukc

,ijkl U^ Ujc Ukc 9lc + Pc,ijkX Uic Ujc Ukc K

+ Pctijk U^ Ujc gkc + Pc,ijX U^ Ujc K)

,ijki uu Ujc gkc die + 2 PCtijkx u^ Ujc gkc K

+ Pc,ij\\ Uic Ujc A c + Pc,ijk Uic Ujc gkc

+ Pc,ijX Uic Ujc A c ) j JZ (PCyijk Uic Ujc he + Pc.ijX Uic Ujc K) (3.4.43)

In many mechanical problems it is possible to have an explicit expression


of the known equilibrium path in terms of the load parameter A. Then to
write the equations of the perturbation process in terms of A, it is sufficient
to put A = t and repeat the previous steps. In this case (3.4.1) and (3.4.7)
99

become

ft = U (A) (3.4.44)

ft = fi (A) + u,- (3.4.45)

respectively, where (3.4.45) describe the function v i-> w , being

w
= ( "' ) ( 3 - 4 - 46 )

Equation (3.4.9) in terms of w will be

£(w) = £ (u,; A) = P ((/,(A) + ut) ; A) (3.4.47)

and the point Qc G {Q&}, in whose vicinity the analysis of bifurcation is


conducted, will from (3.4.11) have the coordinates

Qu = fi (Ac) (3.4.48)

The parametric expressions of the fundamental and bifurcated paths in w


will, from (3.4.12) and (3.4.13), be respectively

Ui = 0
(3.4.49)
A = A(e)

Ui = Ui (e)
(3.4.50)
A = A(e)

The asymptotic expression of the bifurcated p a t h , from (3.4.16), is

Ui (e) = uie e + - uic e2 + o(e 2 )


(3.4.51)
1 - 2 2
A (e) = Ac + Ac6 + - A c 6 + o(6 )
z
The equations of the perturbation process which replace (3.4.18) and (3.4.19)
are

^ {£,,. («,-(£); A(e))} = 0 Vfc.Ve (3.4.52)

^ {f,-(«iW;A(€))}€=0 = 0 VA: (3.4.53)


100

Finally, the eigenvalue problem for the search for the bifurcation points is
rewritten as

Pty(/,(A);A)uy = 0 (3.4.54)

If we denote differentiation with respect to A, for successive relations, by


the symbol ("), we can write equations analogous to (3.4.33)-(3.4.35), (3.4.42)
and (3.4.43) using the following substitutions

tr. — A-
^C A* — 1
(3.4.55)
te = \c xc = o
We obtain

Pe,ij uje = 0 (3.4.56)

Pe,ij uje + PCiijk ujc iikc + 2 Pe,ijk 9ke uje \c + 2 Pe>ij u]e Xc = 0 (3.4.57)

Pc,ij Wjc + 3 Pc,ijk Ujc Ukc + 3 [Pe,ijk 9kc + Pe.ij ) Ujc A c

+ 3 [PCtijk 9kc + Pc,ij) Ujc Ac + PCtijkl Ujc Ukc Ulc

+ 3 (Pc,ijkl 9lc + Pctijk) Ujc Ukc Ac + 3 [(Pc,ijkl 9lc + Pc,ijk) 9kc

+ Pc.ijk he + Pdik 9kc + Pc,ij ] Ujc Ac2 = 0 (3.4.58)

In this case the conditions of orthogonality give

Xc = - 1 Pc^cUicUkc ( 34 5 Q )

2 Pc,ijk Uic Ujc <jkc + Pc,ij Uic Ujc

Ac = — { Pc,ijkl U^ Ujc Ukc Uic + 3 PCtijk U^ Ujc Ukc

+ 3 ( P C ,ijkl Uic Ujc Ukc 9ic + Pc.ijk Uic Ujc 9kc

+ Pcjjk Uic Ujc Ukc + PClij Uic Ujc) Ac

+ 3 (Pc,ijki Uic Ujc 9kc 9ic + 2 Pc,ijk Uic Ujc 9kc + Pc,ijk Uic Ujc gkc

+ PCtij Uic Ujc ) \2C I / 3 (Petijk Uic Ujc 9kc + Pctij Uic Ujc) (3.4.60)

As a final remark it is worth noting that, in the case of simple bifurcation,


the equilibrium curves behave as shown in Fig. 3.6 for the two-dimensional
case.
101

a) Kc^o b)Xc=0,lc>0 c)i. c = 0 , A c < 0

Fig. 3.6 - Bifurcation in the load-incremental displacement plane.

3.5 ANALYSIS OF BIFURCATION FOR A SYSTEM


OF TWO DEGREES OF FREEDOM
In this Section we show an application of the theory of bifurcation of discrete
systems described in previous Sections. Some considerations on the problem
of the determination of the fundamental path, a knowledge of which is always
assumed a priori, will follow.
Let us consider the system shown in Fig. 3.7 proposed by Kerr and Brant-
man [7].

i i ,
1
I
I
+
# 0 +a
0) (A
<W oo O
-o
1 o
-J
L C\J
^ -M i
O Q.
+u

t/>.
r
o '
'
+a
o
Lsin (# G )
Lsin@0+|8)
h ■ H

Fig. 3.7 - Pin-jointed system of two degrees of freedom.


102

In Fig. 3.7 the symbols adopted are indicated. It is easy to see that

A = L [sin(0o + P)- sin(0 o + a)] (3.5.1)

h = L [2 cos 0O - cos(0 o + a) - cos(0 o + P)\ (3.5.2)

The constraint devices are characterised by the following constitutive relations

m = sa (3.5.3)

F = Jfci A + k2 A 2 (3.5.4)

The total potential energy is

P(a,/?;A) = 5 (a +/?) 2 + ^ kx L2 [sm{00 + P) - 8m[0o + a)}2

+ -k2L3 [sin(0o + P) ~ sin(<?0 + a)}3


o
- X L [2 cos 0O - cos(0 o + a) - cos(0 o + /?)] (3.5.5)

The equations of equilibrium (3.1.2), in the case under examination, are writ-
ten as

Pta = 2 5 (a + /?) - iki L2 [sin(0o + /?) - sin(0 o + a)] cos(0 o + a)


- k2 L 3 [sin(0o + P)~ sin(0 o + a)] 2 cos(0 o + a)
-ALsin(0o + a)=O (3 5 6)

P, = 2 5 ( a + /?)+A: 1 L 2 [sin(0o + ^ ) - s i n ( 0 o + a)]cos(0o + /?)


+ k2 L 3 [sin(0o + 0) - sin(0 o + a)] 2 cos(0 o + P)
-XL sin(0 o + P) = 0

It is simple to verify that the curve represented by the following parametric


relations is an equilibrium path, in the sense that it satisfies the equation of
equilibrium for every value of t

a = t = gx(t)

P t = a
= *M (3.5.7)
_ 23 2t _
A
" Tsm(0o + t)-9(t)
103

When the equilibrium path (3.5.7) is known, we can go on to the search


for the bifurcation points upon it. To do this, it is necessary to write (3.4.32)
explicitly and so calculate the derivatives of (3.5.5). These are

P,aa = 2s + kiL2 [sin(0o + P) - sin(0o + ot)} sin(6>0 + a)


+ kiL2 cos2(0o + a)
+ k2 Ls [sin(0o + 0)- sin(0o + a)] 2 sin(0o + a)
+ 2k2L3 [sin(0o + 0) - sin(0o + <*)] cos2(0o + a)
-XL cos(0o + a) (3.5.8)

P.00 = 2s-k1L2[sm{9o + 0)-sm{0o + a)}sm{9o + 0)


2 2
+ kt L cos {6o + 0)
- k2 L3 [sin(0o + 0)~ sin(0o + a)}2 sin(0o + 0)
+ 2k2Lz [sin(0o + 0) - sin(0o + a)] cos2(0o + 0)
-XL cos(0o + 0) (3.5.9)

P.cfi = 2 s - kt L2 cos(0o + a) cos(0o + 0)


-2k2L3 [sin(0o + 0) - sin(0o + a)] cos(0o + a) cos(0o + 0)
(3.5.10)

The expressions (3.5.8)-(3.5.10), calculated along the equilibrium path


(3.5.7), give

P,aa (ft(0 5 g{t)) = 2s + klL2 cos2(0o +1) - ?(*) L cos 6


( o + t) (3.5.11)

PfiP (ft(«) ; g{t)) =2s + kiL2 cos2(0o + 0 - g{t) L cos(0o + t) (3.5.12)

P,a0 (ft(«); </(*)) = 2 * - *! L J cos2(0o + *) (3.5.13)

whilst the perturbation equations of the first order (3.4.32) give rise, in this
case, to the algebraic system
[1 + k{ cos2(0o + t)-X* cos(0o + t)\ tii + [1 - K cos2(0o + «)] «2 = 0
(3.5.14)
2 2
[1 - *; cos (0o + 0] «i + [1 + *i cos (0o + t)-X* cos(0o + *)] "2 = 0

where we have used

v = « i , *; = * ^ , *; = ^ p.5.15)
x 2 v
2s ' 2s ' 2s '
104

The homogeneous algebraic system has solutions different from the trivial
(til = ti2 = 0) only if the determinant of the coefficients is equal to zero. This
is expressed by the condition

det [Ptij (gi(t); g(t))] = [l + *J cos 2 (0 o + *) - A* cos(0 o + *)]*

- [l - JfeJ cos 2 (0 o + t)]2 = 0 (3.5.16)

which is called the characteristic equation of the problem. This equation can
be simply factorised in the form

[2 - A* cos(0 o + *)] [2 *i cos


(^o + 0 - A*] cos(0 o + *) = ° (3.5.17)

The values of t which satisfy this equation are the eigenvalues, whilst the
solutions uic, u 2c corresponding to each of them are the eigenvectors. We
note that in the case under examination the eigenvalue problem is non-linear
in t. We are obviously interested in the search for the lowest eigenvalue and
for the corresponding eigenvector. By taking

jfcj = 1 0Q = 10° (3.5.18)

(3.5.17) yields

tc = 27.73° = 0.4840 rad (3.5.19)

To the value tc given by (3.5.19) there corresponds a point on the equilibrium


path (3.5.7) with coordinates

etc = 9i {te) = 27.73°

Pc = 92 {tc) = 27.73° (3.5.20)


Ac = g{tc) = 2 y 1.58178

It is easy to see that this is not a stationary point for the load because Ac =fi 0
[see (3.5.45)], so we can conclude that it is certainly a bifurcation point. We
note that, in correspondence with the eigenvalue (3.5.19), the term contained
in the second parentheses of the characteristic equation (3.5.17) vanishes and
so we can write

2 cos(0 o + tc) - A* = 0 (3.5.21)

This implies

1 + cos 2 (0O + tc) - A* cos {0O + te) = l - cos 2 ($0 + tc) (3.5.22)
105

for which the system (3.5.14), corresponding to (3.4.33), is written as

uic + u2c = 0
(3.5.23)
uu + ti2c = 0
with the solution

uu = ~ u2e (3.5.24)

Let us now identify the parameter e introduced in (3.4.13) with the coor-
dinate ui. In this case the equation of normalisation (3.3.6), written in terms
of e, assumes the form

e = ui (e) (3.5.25)

From this come the incremental conditions analogous to (3.3.1l)-(3.3.13)

ule = 1 (3.5.26)

uu = 0 (3.5.27)

5lc= 0 (3.5.28)

In this case equations (3.5.23) and (3.5.26) have the solution

uu = 1
(3.5.29)
u 2c = -1

We can now pass on to the determination of tc. To do this, it is necessary


to calculate all the third derivatives of the total potential energy function
(3.5.5). These are

Ptaaa = *i L2 [sin(0o + P) - sin(0 o + a)] cos(0 o + <*)


-3kxL2 sin (0o + a) cos(0 o + a)
+ k2 L3 [sin(0o + P) ~ sin(0 o + a ) ] 2 cos(0 o + ot)
-6k2Lz [sin(0O + P) - sin(0 o + a)] cos(0 o + ot) sin(0 o + OL)
-2k2Lz cos 3 (0 O + a) + A L sin(0 o + a) (3.5.30)

P,aaP = kx L2 cos(0 o + P) sin(0 o + a)


+ 2k2Lz [sin(0o + P) - sin(0 o + a)] cos(0 o + P) sin(0 o + a)
+ 2 k2 L cos(0 o + P) cos 2 (0 o + a)
z
(3.5.31)
106

PjW = -fc 1 L 2 [sin(d 0 + /3)-sin(do + a)]cos(d 0 + /?)


- 3fcjL 2 cos(d0 + /?) sin(d0 + 0)
- k2 L3 [sin(d0 + 0)- sin(d0 + a)] 2 cos(d0 + 0)
- 6 k2 L3 [sin(d0 + 0)- sin(d0 + a)] cos(d0 + 0) sin(d0 + /?)
3 3
+ 2k2L cos (d0 + /?) + A L sin(d0 + 0) (3.5.32)

P*M = *i L2 cos(d0 + a) sin(d0 + 0)


+ 2fc2L 3 [sin(d0 + /?) - sin(d0 + a)] cos(d0 + a) sin(d0 + 0)
-2k2L3 cos(d0 + a) cos2(d0 + 0) (3.5.33)

P,«aA = - L cos(d0 + a) (3.5.34)

P^px = 0 (3.5.35)

Pjilix = - ^ cos(d0 + 0) (3.5.36)


Furthermore, the derivative of A with respect to t is
£ = i f sin (d0 + a) - f cos (d0 + a)
L sin2(d0 + a) l
' * '
For the bifurcation point, that is, for the value tc = 0.4840 rad, putting
k2 = 0.5 and remembering that k\ = 1 and dp = 10°, the expressions (3.5.30)-
(3.5.37) assume the values

Pc,aaa = 2 s [ - 3 k{ sin(d0 + te) cos(d0 + te) - 2 k*2 cos3(d0 + te)

+ X*c sin(d0 + te)] = - 1.9574 s (3.5.38)

Pc,a*f> = 2s[k{ cos(d0 + tc) sin(d0 + te) + 2 A;2* cos3(d0 + te)]


= 1.9574 s (3.5.39)

^C/J/5* = 2 5 [fcj cos(d0 + ie) sin(d0 + te) - 2fc2*cos3(d0 + te)]


= -0.0215 s (3.5.40)

PcfiW = 2 a [ - 3 *; cos(d0 + t e ) sin(d0 + te) + 2 A:2* cos3(d0 + t e )


+ A* sin(d0 + te)] = 0.0215 « (3.5.41)
107

-Pc.aaA = -L cos(0 o + *«) = - 0.7909 L (3.5.42)

PeM\ = -L cos(0 o + te) = -0.7909L (3.5.43)

Pc^ix = 0 (3.5.44)

. = 4 , s i n ^ + g-fcos(go + g = 2 4 4 7 ,
2
L sin (^0 + ^) ^
Remembering the first of equations (3.5.15), we can also write

A* = Ac — = 1.2239 (3.5.46)
2s
We have, at this point, all the elements for calculating the value of tc. It
is sufficient, for this purpose, to replace in each of the terms which appear
in (3.4.42) the values of the expressions (3.5.38)-(3.5.46), remembering that
ui c = 1 , u2c = - 1 , ,&c = 1 , Pc = 1. We have

Pc.ijk «ie Uje Ukc = Pc.aaa Ulc + 3 Pc,aa0 « i c "2c + 3 Pc,ppa «2c " l c

+ Pcjw t»L = - 7.9156 5 (3.5.47)

Pc.ijk w,c Ujc 9kc = Pc,aaa u2u occ + Pc>aap (u2lc (5c + 2 uu u2c &c)

+ PCDPCL («L OCc + 2 tile U2c & ) + Pc,000 u\c J3e


= -3.8719 s (3.5.48)

Pc,ijX Uic Ujc Xc = A c (^P c ,aaA ^ l c + 2 Pc,a/3X ^lc ^ 2 c + -Pc.flSA ^ 2 c J

= -3.8719 5 (3.5.49)

By substituting (3.5.47)-(3.5.49) in (3.4.42), we obtain

tc = - 0 . 5 1 1 1 (3.5.50)

Bearing in mind (3.4.7), the relations which identify the bifurcated path are
written as
ab = a + Ui
b
(3.5.51)
P = P + u2
A being the same as defined on the fundamental path. Equations (3.5.51), by
virtue of the expression of the known path (3.5.7), of (3.4.13) and of the first
108

of equations (3.5.15), can be written in the parametric form

ab{e) = t{e)+u1{e)

(3>(e) = t(e) + u2(e) ( 3 5 5 2 )

2t{€)
Vie) =

By deriving (3.5.52) with respect to e and calculating the values of the deriva-
tives at e = 0, and using (3.5.50), (3.5.46) and (3.5.29), we obtain

«5 = ic + uu = 0.4889
Pbc = ic + u2c = -1.5111 (3.5.53)

A* = A* ic = - 0.6255

The asymptotic expression of the path (3.5.52), approximated to the first


order, is therefore written as

ab(e) = 0.4840 + 0.4889e


(3b{e) = 0.4840-1.5111e (3.5.54)
A* (6) = 1.5817-0.6255 6

These results obviously coincide with those found by Kerr and Brantman
[7]. It is interesting to note that in [7] the authors use the method of in-
vestigation presented in Section 3.3 of this Chapter, for which they arrive
at a non-linear condition of orthogonality of the type (3.2.21), in place of
(3.4.40). A comparison of the two methods of investigation is instructive, as
it demonstrates the greater efficiency of the procedure illustrated in Section
4.4 and used here.
The projections on the plane a, /3 of paths (3.5.7) and (3.5.54) are repre-
sented in the following figure.
Note that the point A, in which the straight line B in Fig. 3.8 intersects
the axis a, has the abscissa a = 0.6406 rad and corresponds to the value
6A = 0.3203 of the parameter. For point B , the intersection of the bifurcated
p a t h with axis /?, we have

<*B = 0 , 13B = 1.980 rad , eB = - 0.9900

It is appropriate to call attention to the fact t h a t e = Ui represents the


difference between the values that the angle a assumes along the bifurcated
p a t h and along the fundamental path. Positive values of e therefore describe
the piece of the straight line B in Fig. 3.8 to the right of a c , while we have
negative values of e on the piece of the straight line to the left of a c .
109

<$ : fundamental path


^ : bifurcated path

Fig. 3.8 - Projection of the linearised bifurcated path on the plane a, /?.

We now have an asymptotic expression of the bifurcated path, approxi-


mated to the first order. Naturally, such an approximation is quite rough, so
it is of interest to determine at least the second-order terms of the equations
(3.4.16). The coefficients which appear in the first of these equations are ob-
tained by solving the system (3.4.34). To do this, it is necessary to find the
values that (3.5.11)-(3.5.13) assume at the bifurcation point. We have

Pctaa = Pc,a!3 = Pc,fifi = 2 8 [l + k{ COS 2 (0 O + te) - Ac* COs(0 O + tej\

= 0.74894 6 (3.5.55)

Equation (3.4.34) is equivalent to a non-homogeneous linear algebraic sys-


tem of two equations in two unknowns, which are obtained by putting i = a
and i = ft respectively. We have the following explicit expressions of the
known terms

Pc,ijk Ujc Ukc — Pc,iaa Uu + 2 PCtiafi UU U2c + Pc,ipp ^2c (3.5.56)

2 ie PCtijk ujc gkc = 2 ic ( P C | t a a ulc ac + Pc%iafi ulc fic

+ Pcti/3a V>2c OLc + Pcti00 U2c $c) (3.5.57)

2 tc Xc Pc,ij\ Ujc — 2tcXc (Pc,iaX Uu + PCtij3\ U>2c) (3.5.58)


110

From (3.5.38)-(3.5.45), (3.5.50) and (3.5.29), and remembering that


&c — Pc — 1? we obtain

Pcajk Ujc Ukc = ~ 5.8937 6 (3.5.59)

Pcfijk u>jc Ukc = 2.0219 s (3.5.60)

2 ic Pc,ajk ujc gkc = 1.9789 5 (3.5.61)

2 ic Pc%fijk iijc gkc = - 1.9789 s (3.5.62)

2*e Ac Pc>ajX ujc = 1.97895 (3.5.63)

2tc k PCipj\ ujc = - 1.97895 (3.5.64)

Therefore the equations of the secondary critical mode are written as

0.74894 uu + 0.74894 u2c = 1.9359


(3.5.65)
0.74894 ulc + 0.74894 u 2c = 1.9359
One can immediately verify that the two equations (3.5.65) are linearly
dependent and in fact, in this specific case, they are the same. System (3.5.65)
can therefore be solved and this, as we know, depends on having imposed the
condition of orthogonality (3.4.40).
Bearing in mind the equation of normalisation (3.5.27), the solution to
system (3.5.65) is written

ule = 0
(3.5.66)
u2c = 2.5848
To complete the perturbation analysis of the bifurcated paht, up to the
second order, it is again necessary to calculate the value of the parameter
tc. From the relation (3.4.43) we can see that to calculate the terms which
appear in it, it is necessary to write all the fourth-order derivatives of the
total potential energy (3.5.5). These are

P,aaaa = - *i L2 [sin(0o + P) - sin(0 o + OL)] sin(0 o + a)


-4kxL2 cos 2 (0 o + a) + 3 Jki L 2 sin 2 (0 o + «)
-2k2Lz [sin(0o + P) - sin(0 o + a)] cos 2 (0 o + <*)
- k2 L3 [sin(0O + P) - sin(0 o + a)}2 sin(0 o + a)
+ 12 k2 Lz cos 2 (0 o + a) sin(0 o + a)
-6k2Ls [sin(0o + P) - sin(0 o + « ) ] [ - sin 2 (0 o + a)
+ cos 2 (0 O + a)] + A I cos(0 o + a) (3.5.67)
Ill

PjpfiP = k
i L<t
l s i n (0o + 0) - sin(0 o + a)] sin(0 o + 0)
- 4 kt L2 cos 2 (0 o + P) + 3k1L2 sin 2 (0 o + 0)
-2k2L3 [sin(0o + 0) - sin(0 o + a)] cos 2 (0 o + /?)
+ A;2 L3 [sin(0o + 0) - sin(0 o + a ) ] 2 sin(0 o + /?)
- 12 A;2 L 3 cos 2 (0 o + 0) sin(0 o + /?)
- 6k2 L3 [sin(0o + 0) - sin(0 o + a)] [ - sin 2 (0 o + /?)
+ cos 2 (0 o + 0)] + XL cos{90 + 0) (3.5.68)

P.aaap = &1 £ 2 COs(0o + /?) COs(0o + « )

- 6 fc2 L 3 cos (0O + a) sin (0O + a) cos(0 o + 0)


+ 2k2L3 [sin(0O + /?) - sin(0 o + a)] cos(0 o + 0) cos(0 o + a)
(3.5.69)

Pfitapp = -hL2 sin(0 o + 0) sin(0 o + a)


+ 2k2L 3
cos 2 (0 o + /?) sin(0 o + a)
-2k2L3 [sin(0o + 0) - sin(0 o + a)] sin(0 o + 0) sin(0 o + a)
-2k2L 3
sin(0 o + /?) cos 2 (0 o + a) (3.5.70)

P,a000 = &1 ^ COS (0O + a ) COS (0O + 0)


+ 6k2L3 cos(0 o + a) cos(0 o + 0) sin(0 o + /?)
3
+ 2 fc2 L [sin(0o + 0)- sin(0 o + a)] cos(0 o + a) cos(0 o + 0)

(3.5.71)

P,aaa\ = L sin(0 o + a) (3.5.72)

i%«/»A = P.afi(>X=0 (3.5.73)


J'jW* = L s i n ( 0 o + y9) (3.5.74)

The values assumed by (3.5.67)-(3.5.74) at the bifurcation point are

Pc,aaaa = 2 5 [ - 4 *; cos 2 (0 O + t e ) + 3 fcj sin 2 (0 o + te)

+ 12 k2 cos 2 (0 o + t„) sin(0 o + te) + A* cos(0 o + *e)]


= 4.3381s (3.5.75)
112

,0000 = 2 s [ - 4 k{ cos2(0o + te) + 3 k{ sin2(0o + *c)

- 12 k*2 cos2(0o + <c) sin(0o + U) + X*e cos(0o + tcj\


= -4.84875 s (3.5.76)

Pe,aaap = 2 8 [k[ COS2(0O + U) - 6 fcj sin(0 O + t«) COS2(0O + t„)]

= -1.04566 s (3.5.77)

Pc,aapp = 2s[-k\ sin2(0o + te)] = -0.74894s (3.5.78)

Pe*000 = 2s[k{ cos2(0o + te) + 6fc2*cos2(0o + te) sin(0o + *«)]

= 3.54777 s (3.5.79)

Po,aaaA = 0.61194 L (3.5.80)

Pc,aa0X = Pc,a00\ = 0 (3.5.81)

i'ejj/j/JA = 0.61194 L (3.5.82)


Equations (3.5.75)-(3.5.82) allow calculation of the values of all the terms
which appear in (3.4.43). Their expressions and the corresponding numerical
values in the case under examination are given in the following.

+ 4 Pctappp file U2c + Pe,000fi U2c


= -15.01275 (3.5.83)

3 Pc,ijk Uic Ujc Ukc — 3 yPC)aaa Uu file + PC,aa0 V>U fi2c

+ 2 Pc,aPa file fi2c file + 2 Pc,af3p file ^2c ^2c

+ Pc,0j3a U2c file + -Pc,/9^ U2c fi2cJ

= 15.6786 5 (3.5.84)

3 *c -Pc>«jib fitc fijc 9kc — 3 ^Pc,aaa file file &c + -PC,aa^ file file A:

+ -Pc,a/9a ^10 fi2c «c + -Pefa/00 file fi2c Pc

+ Pc£*« U2cfilc&c + Pcjap fi2c filc &


+ -P Cf fl8« fi2c fi2c OLc + ^ c , / W fi2c fi2c Pc) U

= -7.6725 5 (3.5.85)
113

3 tc Xc PCtij\ Uic Ujc — 3 \PC,OLQLX Uu Uu + Pe,a0X Uu U2c

+ Pct0aX U^c Uic + Pc,00X U2c U2e) tc K

= -7.6725 5 (3.5.86)

o tc "cjjkl Uic Ujc Ukc 9lc =


Uu OLe + PCi*a«p Uu Pc

+ 3 Pe,a«0ct UU U2c &c + 3 Pc,aa00 Uu U2c Pc

&
+ 3 Pc,a00a *>U u\c c + 3 Pc,a000 UU \i\c (3c

+ Pcjfifta ulc &c + Pc,0000 ^2e Pc)


= -28.172 5 (3.5.87)

3 tc Xc PC}ijkX Uic Ujc Ukc = 3 te Xc {Pc,aaaX Ulc + 3 Pc,aa0X UU U2c

+ 3 PCta00X Uu u\c + PC,000X U2c)

= 0 (3.5.88)

3
i\ Pc,ijkl Uic Ujc Qkc gic = 3 t\ [Pctaaaa u\c &l + 2 Pc,aaa0 u\e Ctc $e

U
+ PctaaM Uu (5C +2 PCtafiaa U U2c &c
U
+ 4Pc,a0afi UU U2c &c J3C + 2PCfa000 U U2e fi\

+ Petfi0aa u\c &\ + 2 Pc,00a0 u\e &C 0c

+ Pc,MW u\c pi)


= 0.773665 (3.5.89)

6 t\ Xc Pc>ijkX Uic Ujc Qkc = 6 t\ Xc [Pc,aaaX u\c 6cc + PC,aa0X u\c (5c

+ 2 Pc,a0aX Uu U2c &C + 2 Pc,a00X UU U2c fie

+ Pet00aX u\c ac + Pc,000X u\c 0e)

= 4.6954 5 (3.5.90)

3 i\ Xc PCtijx uic Ujc = 3 i\ \c (Pc,aax u\c + 2 PCi0tpX uu u2c

+ PcMxu\c) -3.9219 5 (3.5.91)


114

We can see in (3.5.91) that use has been made of the value A,., which the
second derivative of A = g(t) with respect to t takes at the bifurcation point.
By making use of (3.5.45) and bearing in mind (3.5.18) and (3.5.19), we obtain

Ad = { - 2 sin(0 o + te) cos(0 o + tc) [sin(0o + tc) - te cos(0 o + te)}

+ tc sin 3 (0 o + *«)} 4 y sin~ 4 (0 o + *„) = - 3.1638 y (3.5.92)


' Li Li
Also, from the first of equations (3.5.15), we get

XI = - 1.5819 (3.5.93)

Finally, substituting the values of (3.5.84)-(3.5.92) in (3.4.43) and remember-


ing (3.5.49) and (3.5.50), the result is

ic = - 1.4403 (3.5.94)

At this point we can determine the expression of the bifurcated path, in


asymptotic terms, correct up to the second order. To write this expression,
however, it is necessary to evaluate the coefficients abc, /3b, A* of the series
expansion.
By differentiating (3.5.52) twice with respect to e and calculating the values
of the derivatives at the bifurcation point, we obtain

a"e = te + uu = -1.4403

Pi = ic + u2c = 1.1445 (3.5.95)

A = A t + A tc = -2.1760
having made use of (3.5.46), (3.5.50), (3.5.66), (3.5.93) and (3.5.94). Finally,
we can write
a 6 (e) = 0.4840 + 0 . 4 8 8 9 e - 0 . 7 2 0 1 5 e2
0b{e) = 0 . 4 8 4 0 - 1 . 5 1 1 1 6 + 0 . 5 7 2 2 e2 (3.5.96)
A* (e) = 1.58178 - 0.6255 e - 1.0880 e2
Due to identification of the parameter e with Ui, (see (3.5.25)), the relations
(3.5.96) express the variables ab, (3b, A* as functions of the difference (a 6 —a).
It is legitimate, however, to object that taking the results of the investigation
from (3.5.96) does not give an immediate interpretation of the mechanical
phenomenon under examination, whereas an identification of the type e = A*
would produce more effective relations.
We remember at this point that in Section 3.3 of this Chapter, simple
expressions which allow a change in parameter were given. Using these rela-
tions, we want to find the asymptotic expression of the bifurcated path as a
115

function of (A* — A*). To do this we write (3.3.18) putting &■


, = 0 , b = 1 and
t = e:

V(e)=y(e)-K (3.5.97)

We can verify immediately that such a choice of parameter is legitimate,


in the sense that the curve of the bifurcated path is locally invertible in t] in
the bifurcation point, as i]c = A* ^ 0. Comparing (3.5.97) with (3.3.18) we
realise that in our present case e has the same role as had the parameter t.
The relations (3.3.21), (3.3.22) written for a and /?, respectively, now give

of = al/ie = -0.7816
(3.5.98)
ri = Pc/Vc = 2.4158

a? = (abe-ahcric/r,c)/ec = -8.0283
(3.5.99)
0? = {0bc-0bctic/tc)/ec = 16.361

where use has been made of (3.5.53) and (3.5.95) and of the relations rjc =
K 5 Vc = K derived immediately from (3.5.97).
The expressions for the bifurcated path for which we are looking are

a (77) = 0.4840 - 0 . 7 8 1 6 (A* - A * ) - 4.0141 (A* - A * ) 2


(3.5.100)
jSfa) = 0.4840+ 2.4158 ( A * - A * ) + 8.1806 ( A * - A * ) 2

Figure 3.9 shows the graphs of the projections on the plane a,/? of the
functions (3.5.7), (3.5.96) and (3.5.100).
Note that, even if (3.5.96) and (3.5.100) represent an asymptotic expression
to the same order of the same function, with coincident initial points, their
graphs are very different. This shows the importance of the choice of parame-
ter when we are interested, in addition to a local analysis, in an approximate
expression of the function under examination within a finite proximity of the
bifurcation point.
Obviously at the bifurcation point the curves Si and S2 have the same
tangent and the same curvature. Figure 3.10 shows the fundamental and
bifurcated paths giving, for the second, the results of a numerical solution to
the non-linear equations of equilibrium (3.5.6).
It is evident, when making a comparison with Fig. 3.9, how although curve
81 can be considered an acceptable approximation of the bifurcated path in
the whole range of e examined, the same cannot be said for curve B2.
116

:
# fundamental path

i ' bifurcated path d e s c r i b e d by p a r a m e t e r r\


[
3$2 bifurcated path described by p a r a m e t e r £

Fig. 3.9 - Effects of the change of parameter on the representation of the


bifurcated path.
117

Fig. 3.10 - Exact solutions to equations of equilibrium.

3.6 OUTLINES OF THE ANALYSIS OF BIFURCA-


TION STARTING FROM AN APPROXIMATE
EQUILIBRIUM PATH
The basic assumption of the procedure for the construction of the bifurcated
path is that of knowledge of an equilibrium path of the structural system
under examination. All that has been said so far would therefore not be
of practical use if we did not know, case by case, at least one solution to
the equations of equilibrium (3.1.2). This at first sight may seem extremely
restrictive, as the search for even one solution to problem (3.1.2) is often very
complex. The problem is greatly simplified, however, if we look only for an
approximate solution. In this case we ask ourselves whether, by using such a
solution as the known equilibrium path, it would not be possible to proceed
to an analysis of bifurcation adopting the procedure shown previously, and
what validity the results found would have.
It is not easy to give a general answer to this problem, especially as an
estimate of errors in this type of procedure is difficult to make; after all,
approximate results are introduced in an approximate procedure (the pertur-
bation procedure). However, looking at the results which many researchers
have obtained in problems of structural engineering, both for continuous and
118

discrete models, using even rough approximations of the known equilibrium


path, we can confirm that in most cases results have been obtained which
agree not only qualitatively but also quantitatively with the experimental re-
sults. Here we shall show an example in which an approximation of a known
path is constructed using the perturbation method, that is to say, the path
will be extrapolated from a known equilibrium point. A procedure of this
type is easy to use as it is always possible to know an equilibrium point
and furthermore, as it utilises the same ingredients as bifurcation analysis,
it is particularly suitable for use in automatic calculation. Let the point of
extrapolation be
qi0 = 0
3.6.1
A0 = 0
and the extrapolated path be expressed in the form

ft(0 = qiot + ^hof + htio**+ <>{?)


\ (3.6.2)
2 3 3
MO = A0* + ^ i 0 * + ^ i * + o(* )
As it is useful in what follows, we write an asymptotic expression of the
total potential energy function with the initial point coinciding with the ex-
trapolation point:

PfeiA) = Po,AA + - P o , , ; ^ g i + P o , , A g . A + - P 0 , A A A 2

+ g Potijk Qi Qj Qk + ~ Po,\jk A qj qk + - Po.AA* A2 qk

+ g PO.AAA A3 + — Po,iju q% qj qkqi + - PO,\JU A ?,- qk q\

+ T Po,\\ki A qkqi + - PO.AAAJ A qi + — io.AAAA A


4 6 24
+ ... (3.6.3)
where the suffix "0" indicates that the differentiations are evaluated at the
extrapolation point. The bifurcation point on the approximated path, if it
exists, is found by using (3.4.32)

Pert {qi{t); A(0) uj = {Potij + Po,ijk qt{t) + POMJ A(t)

+ - Po,ijki qt{t) qi{t) + Pot\ijk <7*(*) \{t)

+ Po,AA,yA2(0 + . . . } Uj = 0 (3.6.4)
By substituting for q(t), \(t) the asymptotic expressions, we obtain a sys-
tem of polynomials in t which are equal to zero. In our example we assume a
119

linear approximation of the equilibrium path (3.6.2): <fc = &o (t), A = Ao (t).
Then (3.6.4) becomes

{Po,ij + (Potijk Qko + PQMJ Ao) t} Uj = 0 (3.6.5)

which is an eigenvalue problem, linear in t.


To the model of two degrees of freedom studied previously we now apply
the present considerations, adopting a linear extrapolation of the fundamental
path starting from the origin. By solving the simple algebraic system of
two equations in two unknowns which is obtained by writing (3.2.6) with
k — 1, t = 0, and remembering that P , are the functions (3.5.6), we obtain

a(t) = t
P(t) = t (3.6.6)

sin #o
We can immediately see that relations (3.6.6) are the first-order part of
the asymptotic expression with initial point t = 0 of (3.5.7). By using the
expressions (3.5.8)-(3.5.10) and (3.5.30)-(3.5.36), calculated at t = 0, we get
the coefficients of (3.6.5):

1 + k{ cos 2 0O 1 - £1 cos 2 0O
Potii = 25 (3.6.7)
1 - k{ cos 2 0O 1 + K cos 2 0O

— 2 k\ sin 0O cos 0O 2 k\ sin 0o cos 0O


Po,ijk %k — 2S (3.6.8)
2 k\ sin 0O cos 0O — 2 k{ sin 0O cos 0O J

[ — 2 sin 0O/ cos 0O 0


Po,\ij A0 = 25 (3.6.9)
0 — 2 sin 0Q/ cos 0O

In explicit form (3.6.5) may be written as

[1 + k\ cos 2 0O - [k{ sin 2 0O + 2 / tan 0O) t] tii


+ [l-k{ cos 2 0O + {k{ sin 2 0O) t] u 2 = 0
(3.6.10)
[1 - k\ cos 2 0O + [k[ sin 2 0O) t] ux
+ [1 + k\ cos 2 0O - [k{ sin 2 0O + 2 / tan 0O) t] u2 = 0
120

The characteristic equation of the eigenvalue problem (3.6.10) is

-r ( 7T + K sin 2 0 t2
tan 0o Vtan B0 J
- \k{ sin 20 o + —^-r- ( l + *I cos 2 $o)] t (3.6.11)
L tan do 1
+ 4k{ cos 2 0O = 0

from which the following results are obtained:

te = 0.1763

A* = 2.0308 (3.6.12)
UU = U2c

tc = 0.16128

K = 1.8575 (3.6.13)

uu = -<l"ic

Omitting lengthy algebraic manipulations, we have the value of tc and the


first order approximation of the bifurcated p a t h as

ic = -0.1457 (3.6.14)

ab(e) = 0.1613-|-0.8543e
b
(3 (e) = 0 . 1 6 1 3 - 1.1457 e (3.6.15)
A*(e) = 1.8575-1.67816

As can be seen from Figs. 3.11, 3.12 the approximation of the bifurcated
path, despite of the favourable value obtained for the critical load, is very
poor. This is because the fundamental equilibrium curve of the system if
highly non-linear and the bifurcation point is "a long way" from the extrap-
olation point. It is legitimate to ask, therefore, how this can be reconciled
with what has previously been stated regarding the possibility of obtaining
adequate results, for structural problems, by means of this type of analysis.
The explanation of this apparent contradiction lies in the fact t h a t the range
of application of the method shown here is defined by the meaning t h a t is
given to the words "a long way", the exact determination of which presents
enormous difficulties. To try here to give some indications, however simply,
seems inopportune, and for this reason it is necessary to interpret the pre-
vious example only as a test which suggests a cautious use of the type of
approximation discussed here.
i

A* s£ \ fundamental path
8% : bifurcated path

2.0

1.5

1.0 Nv ^ ^ f

\f ^\
0.5

nn . 1 1 1 _^k -
20° 40° 60°

Fig. 3.11 - Projection on the plane a , A* of the tangent to the bifurcated


path at the bifurcation point.
,i
A* S&/

2.0

Jt^^
S\5

1.0

0.5

00 I | I -I ►
20° 40° 60° 80° $

Fig. 3.12 - Projection on the plane /3, A* of the tangent to the bifurcated
path at the bifurcation point.
122

REFERENCES
[l] M.J. Sewell: "The static perturbation technique in buckling problems",
J. Mech. Phys. Solids, 13 (1965), 247.
A.D. Kerr, M.T. Soifer: "The linearisation of the pre-buckling state and
its effect on the determined instability loads", J. Appl. Mech. (1969),
775.
M.J. Sewell: "On the branching of equilibrium p a t h s " , Proc. R. S o c ,
London, Ser. A 315 (1970), 499.
J.G.A. Croll, A.C. Walker: Elements of structural stability, Macmillan,
London, 1972.
W.J. Supple (Editor): Structural instability, IPC Science and Technology
Press, Guildford, 1973.
J.M.T. Thompson, G.W. Hunt: A general theory of elastic stability,
J. Wiley & Sons, London, 1973.
A.D. Kerr, R. Brantman: "On non-existence of adjacent equilibrium
states at bifurcation points", Acta Mech., 23 (1975), 29.
T. Poston, L. Steward: Catastrophe theory and its applications, Pitman,
London, 1978.
G. Iooss, D.D. Joseph: Elementary stability and bifuraction theory,
Springer-Verlag, New York, 1980.
W. Ledermann: Handbook of applicable mathematics, VoL I: Algebra,
J. Wiley & Sons, 1980.
123

Chapter 4

STABILITY OF EQUILIBRIUM A N D
POST-CRITICAL BEHAVIOUR OF
CONTINUOUS SYSTEMS

4.1 INTRODUCTION
The theory presented in this Chapter represents the extension to continuous
systems of the theory of discrete systems developed in Chapters 2 and 3. For-
mulated by Koiter just after the war [1,2,3,5,6], chronologically it preceded
the theory of discrete systems developed mainly in Great Britain by Thomp-
son and Sewell ([3,6] Chapter 3). The theory is based on the hypothesis that
the material of the structure is elastic and that the external loads acting are
conservative. Following the order of the two previous chapters, we shall first
enunciate the theorems on stability and instability of equilibrium, follow with
a discussion on the criterion of stability, and finally expound the procedure
for the construction of equilibrium paths, with particular reference to the
bifurcated path.

4.2 T H E O R E M S OF STABILITY A N D INSTABIL-


ITY
Let us consider an elastic body subjected to conservative forces and suppose
that there exists an equilibrium configuration which we shall denote by the
Roman numeral I and call fundamental, of which we wish to study the sta-
bility. Let u and li be respectively the vector displacement measured from
configuration I and the velocity at the instant t, consequent on a perturbation
applied at the instant t = 0. Let T be the absolute temperature at the instant
t and let us introduce the functional of the available energy defined by [6]

D ( u ( x , 0 , * ( x , t ) , r ( x , t ) ) = P(u)+K(u) + ^Jv„^.(TI- TfdV

(4.2.1)

where

K{u) = - f fiu-udV (4.2.2)


124

is the kinetic energy, P ( u ) is the potential energy, \x represents the mass


per unit volume V, Tj denotes the temperature relative to configuration I
and C* represents the specific heat at constant deformation measured at the
temperature T*, with Tj < T* < T. After cumbersome manipulations which
make use of the Clausius-Duhem inequality, the inequality of Fourier for heat
conduction and the first law of thermodynamics, it is shown [6] t h a t during
the evolution of the system the available energy never increases, t h a t is,

t> < 0 (4.2.3)

Such a property is used to prove the theorem of stability. However, it is


convenient first to make some observations and provide some definitions which
are useful for this purpose.
In discrete mechanical systems stability with respect to a norm implies
stability with respect to any other norm, because of the equivalence of norms
in a finite-dimensional space. In continuous systems, on the contrary, it makes
sense to speak of stability only with respect to a particular norm, since norms
are not equivalent in infinite-dimensional spaces. In the definition of stability,
therefore, it is necessary to choose a norm which, far from being of academic
interest, leads to realistic results which involve, possibly, global properties
of the system. The condition, for example, t h a t the displacements ||ut-||, the
displacements gradient ||u t J || and the velocities ||ut-|| remain bounded in the
motion caused by an initial perturbation, is sometimes too restrictive. It
has, in fact, been demonstrated [4] that by applying a radial perturbation,
however small, to an elastic sphere, the gradients of the displacements at the
centre of the sphere grow to infinity because of the singularity of the solution,
and therefore the fundamental configuration of equilibrium is unstable even
in the absence of load. A less restrictive measure of the distances suitable for
providing realistic results is represented by the norm L 2 of the displacements
and the velocities defined by

HI? = jjfv im{x,t).u{x,t)dV (4.2.4)

IN? = ^ Jv Mu(x,t) • ii(x,0 dV (4.2.5)

where M is the total body mass. We now introduce the following definition.
The equilibrium in the fundamental configuration is called stable if, and
only if, for every pair of positive constants a\ and CLI there exist three pos-
itive constants fa, fa, fa with the property that all motions due to initial
perturbations which satisfy the inequalities

IM|o < fa , H o < fa , Do < fa ,


125

also satisfy the inequalities

whatever the time t.

Making use of the inequality (4.2.3), the following two basic theorems have
been proved [6]

T h e o r e m o n s t a b i l i t y ( K o i t e r ) . The equilibrium in the fundamental con-


figuration I is stable if the functional of the potential energy is positive defi-
nite.

T h e o r e m o n i n s t a b i l i t y ( K o i t e r ) . The equilibrium in the fundamental


configuration I is unstable if the potential energy is indefinite or negative
definite.

It is clear t h a t the theorem on stability is the extension to continuous


systems of the theorem of Lagrange-Dirichlet on discrete systems, whilst the
theorem on instability is the counterpart of the Liapunov theorem, mentioned
in Sections 2.4, 2.5 respectively.

4.3 CRITICAL CONDITION OF EQUILIBRIUM


Let us denote by II the generic configuration of the body near to the funda-
mental configuration I [1,7,8] and let

lii = 2 (Ui>> + Uj
"'*' + U f c Uh j
' >) (4.3.1)

be the components of the associated strain tensor. All the indices which
appear in (4.3.1) and in what follows, unless otherwise specified, are intended
as varying from 1 to 3. Furthermore, an index preceded by a comma indicates
differentiation with respect to the corresponding variable. Let us assume
t h a t the density of elastic energy is a regular function and make the series
expansion starting from configuration I

having assumed w(7 t J )l — 0- Let us now introduce the tensors


126

where 5ty is the stress tensor in the fundamental configuration I and Eijhk ls
the elastic tensor. By making use of (4.3.1), (4.3.2), (4.3.3) and of the fact
that Sij = Sji, we can write an expression for the elastic energy of the system

W(u) = j I Si:i ujti + - S{j uhti uhtj + - Eijhk(uitj + ujti)(uhtk + ukth)

+ - Eijhk [(uitj + ujti) ul>h ultk + ( u M + ukth) umti umJ]

+ ~ Eijhk utii uij umth umA dV (4.3.4)

Let us denote by V(u; A) the potential energy of the external loads and
suppose that it is possible to write the series expansion in terms of u starting
from configuration I

V(u; A) = V'u + \ V > 2 + \ X'^ + YA X""A + •■■ (4-3.5)

where "VQ = V(0;A) and so on. In (4.3.5) a dash denotes a Frechet differen-
tiation with respect to u and A is the load parameter, which is kept constant
in the expansion. The total potential energy of the system is therefore [1,2,5]

3 4
P(u)=P> + ^ V + ip;u +i C u +- (4.3.6)

having put

P> = j SijUjjdV-%VL (4.3.7)

i PS u 2 - i Jv {Sij ttkiI- uKi + Eijhk e,y ehk) dV - i X ^ (4-3.8)

- P^ u 3 = i j
l
y Eijhk ei5 ulth u,,4 dV - i X « 3 (4-3.9)

^ PS" ** = \Jv EijHk ulti uu umM um,k dV-± X" « 4 (4-3.10)

with tij — (uij + Ujti)/2, PQ = P'(0; A), and so on. Also, use has been made
of the symmetry of the elastic tensor Eiju — E^j. If the potential of the
loads is bilinear in A and u is VQ U ^ 0 whilst all the other terms of the series
expansion VQ U 2 , V™ u 3 , . . . are equal to zero. From (4.3.7), by splitting the
term "VQ U into the contribution due to the forces per unit volume X{ and to
the forces per unit surface Q,, we have

Pf0 u = j {S^ ujti - Xi Ui) dV - f Qi Ui dS (4.3.11)


127

from which on integrating by parts we have

P<5 u = f (Su Hi - QA Uj dS - f {Sijti + Xj) Uj dV (4.3.12)


Js Jv
The two terms on the right side of (4.3.12) vanish, since the first integrand
in parentheses represents the mechanical boundary conditions and the second
integrand the equation of equilibrium relative to the configuration I; therefore
PQU = 0. Equation (4.3.6) then becomes

P ( u ) = i PS u 2 + i P»' u 3 + 1 PS" u* + ■ ■ ■ (4.3.13)

If the u displacements are sufficiently small, a necessary condition for P{u)


to be positive definite (stable equilibrium) is that

PoU2>0 VU (4.3.14)

the equality sign being valid only for u = 0. Condition (4.3.14) is not suffi-
cient. Let us consider, in fact, an expression of the potential energy in the
form of the integral [2]

P(u) = J1 (x2 u'2 - t / ) dx (4.3.15)

where u(x) is the only non-zero component of the displacement and a dash
indicates differentiation with respect to x. It is assumed that the geometrical
boundary conditions are u(0) = u(l) = 0. The first derivative PQU of (4.3.15)
is identically zero and the second

- P£u2= [lx2u'2dx (4.3.16)


2 Jo
is positive definite. Let us now consider a one parameter family of the dis-
placement fields u(x;e)

u = e (x - 2 x2le + xs/e2) 0< x < e< 1


}
' ' " " (4.3.17)
u=0 e<x<1

Both the displacements are continuous with their derivatives and tend to
zero for e —► 0. The increase of energy for (4.3.15) is then
(4318)
'<«(*«»-!£'"-£''
where the first term is due to the second derivative P{f u2/2. Equation (4.3.18)
is obviously negative for any value of e in the interval 0 < e < 1 and so
the functional P(u) does not have a minimum at u = 0. For Po( u ) to be
positive definite, it must be that PQU2/TQU2 > d [d positive), T"u2 being
128

Fig. 4.1 - Schematic representation of functionals PQ U 2 and TQ'U 2 .

an auxiliary homogeneous quadratic functional whose typical integrand is a


positive definite quadratic function of all arguments which appear explicitely
in the energy increase. This situation is shown schematically in Fig. 4.1 for
the case of one variable u. The rigorous proof of the sufficiency is given by
Koiter [2,5].
With the aim of simplifying the analysis we exclude the occurrence of
cases of the type illustrated in the previous example, which rarely occur in
practice, and assume that (4.3.14) is a necessary and sufficient condition for
P ( u ) to be positive definite. This assumption is currently made in classical
treatments of mechanics.
A particularly interesting case occurs when P£u2 = P"(0; A)u 2 is positive
semidefinite, that is

Pc"u2=° (4.3.19)

Pc" u 2 > 0 for u ^ ux, u ^ 0 (4.3.20)

where Pc" = P"(0; A) identifies the particular case under examination. As we


are, from this point on, exclusively interested in the study of this problem,
we shall read all the previous formulae with an index "c" in place of "o".
Introducing a field of virtual displacements 6u ^ a u x (a real), we have
for u = iii + <$u

Pc" [m + 6u}2 = Pc" u 2 + 2 Pc" U l 6u + Pc" [6u]2 > 0 (4.3.21)

As the first term on the right side vanishes by virtue of (4.3.19) and the
third term is positive definite by virtue of (4.3.20), for (4.3.21) to be satisfied
for £u ^ a u i , we must have

P^lu16u = 0 (4.3.22)
129

Fig. 4.2 - Beam subjected to axial load.

for any 6u.


Equation (4.3.22) is an eigenvalue problem, the eigenvectors u x of which
are defined as critical modes. The corresponding eigenvalues Ac are called
critical loads. To decide upon the type of equilibrium, it is necessary to
examine the higher order terms of the series expansion (4.3.13). The case
in which P" u 2 is positive semidefinite corresponds, as in the treatment of
discrete systems, to the case of critical equilibrium. The essential difference
between the two problems lies in the fact that in discrete systems the operator
[Cij] or [da] is an algebraic operator, whilst in continuous systems the singular
operator P" is a differential operator. From this we can immediately realise
that the difficulties connected with the solution of continuous problems are
greater than those which are met when solving a discrete system. Let us now
examine a specific problem to see the explicit form of the terms of the series
expansion (4.3.13) and of the variational problem (4.3.22) associated with the
condition of critical equilibrium.
Let us consider a beam which in the fundamental state I is acted upon by an
axial force A applied to the centroid of the end section (fig. 4.2). By assuming
that the beam is inextensible, a point of the axis is identified by the same
value of the curvilinear abscissa in any configuration. We limit ourselves to
the case in which v{x) is the lateral displacement in the direction of a principal
axis of inertia of the cross-section and the beam is shear undeformable. The
elastic energy is then zero in configuration I and is equal to EIx2/2 Per
unit length of the beam axis in configuration II, where we have indicated the
curvature by x and the flexural rigidity by El. The work of the axial force A
is equal to the product of the force and the displacements of the end of the
beam when passing from the initial to the final configuration. The increase
in energy P ( u ) can therefore be written in the form

PW-jf 1 „, »"*
+ x(Vi 1)
dx (4.3.23)
130

having indicated differentiation with respect to x by a dash. By making a


series expansion of the terms 1/(1 — v' ) and v 1 — v*2 and re-grouping terms
of the same order, we have

P'cv =0 (4.3.24)

P'c' v2 = f (El V"2 - A t ; ' 2 ) dx (4.3.25)

P"'vs = 0 (4.3.26)

Pc"" v* = 3 f (4 El v'2 v"2 - A t/ 4 ) dx (4.3.27)

The variational problem (4.3.22) associated with the critical condition of


equilibrium is

p c " v6v= f {El v" 6v" -Xv' 6v') dx = 0 (4.3.28)


./o
from which, by integrating by parts

/ ' {EIv"" + X v") 6v dx - \{EIv'" + A v1) 6v][ + [EIv" 6v']l0 = 0 (4.3.29)


./o

From (4.3.29) the differential equation

EIv"" + Xv" = 0 (4.3.30)

is obtained, with the boundary conditions

v (0) - v(l) = 0 , Elv"(0) = EIv"(l) = 0 (4.3.31)

We can immediately verify t h a t the function

v(x) = ak sin k — (k = 1,2,...) (4.3.32)

where ak are arbitrary constants, satisfies the geometrical and mechanical


boundary conditions and the differential equation if the external load A is

A:27r2 El
Xc = ^ ^ ± (k - 1 , 2 , . . . ) (4.3.33)

Equation (4.3.33) gives the Eulerian critical loads and (4.3.32) the asso-
ciated buckling modes. The constants a* can be determined by choosing a
suitable normalisation, as we shall see in the next Section.
131

4.4 CRITERION OF STABILITY OF CONTINUOUS


SYSTEMS
In a similar way to the treatment for discrete systems, we furnish in this
Section the necessary and sufficient conditions for stability in the case of
critical equilibrium [1,2]. It can immediately be verified that in the direction
of the critical mode Ui the series expansion (4.3.13) is reduced to

p{*i) = lrr*i+jipr*i M.I)


from which it follows that the necessary condition of stability is given by

P?u* = 0 (4.4.2)

Pc"X>0 (4.4.3)

To find the sufficient condition, we introduce a generic displacement field

u = aiii + w (4.4.4)

where a is an arbitrary constant which we suppose to be "small" and w is


a displacement field which we assume to be "small" with respect to a u i . In
order to obtain results which are comparable with those obtained for discrete
systems and directly usable in the next Section, we suppose that u is a regular
function of a parameter e such that u(0) = 0, corresponding to the critical
point. By performing a series expansion starting from e = 0 [9] we obtain

u(e) = u e + - u e2 + o(e 2 ) (4.4.5)

From comparison with (4.4.4) we see that aui = i i e , w = u e 2 / 2 and


e = a,xi — Ui. Within the frame of the new notations, equations (4.3.19),
(4.3.20), (4.3.22), (4.4.2) and (4.4.3) respectively read

P c "u 2 = 0 (4.4.6)

P"u2 > 0 for u^u.BjiO (4-4.7)

P,."ui5u = 0 (4.4.8)

Pc"'u3 = 0 (4.4.9)

Pc""ii4>0 (4.4.10)
132

Let us now write the series expansion (4.3.13) of the total potential energy
corresponding to the displacement (4.4.5)
l2

p {*(*)) = \p? ue + - ue + - P'"


c ue + - ue
2 ^6
n4

24 c lie + - ue
+
= - e2 P" ii 2 + - e3 P"' ii 3 + - e3 P" u u + — e4 P'"' ii 4
2 6 2 24

+ - e4 P f ii 2 u + - e4 P," u 2 + o(e 4 ) (4.4.11)


4 8
Using (4.4.6), (4.4.8) and (4.4.9), (4.4.11) is simplified into

P(u(6)) = l ^ p ; ' ' u 4 + i 6 4 P ; ' u 2 u + i 6 4 P ; u 2 + o(6 4 ) (4.4.12)

Our aim is to determine the function ii and the parameter e which minimise
the functional P(u(e)). To do this, we analyse the functional

J(ii) = - e 4 P ; u 2 + - e 4 P c m u 2 u (4.4.13)

for a fixed arbitrary value of the parameter e and look for the function u
which minimises it. By indicating with ii* the solution to the problem, we
introduce a field of virtual displacements 6u. Obviously J(ii* +6u) > J(ii*),
from which

I e4 Pc" [ii* + 6 u] 2 + \ e4 P? u 2 [u* + 6 u]


8 4

8 4 4 4

+ - e4 P" [6 u] 2 > - e4 P» ii*2 + - e4 P'" ii 2 ii* (4.4.14)


c l J c c
8 8 4
By carrying out simplifications, we obtain

P?ii* 6u + P'c"u26u + - P? [6uY >0 (4.4.15)

As the last term of (4.4.15) is positive definite by virtue of (4.4.7), for the
inequality to be always valid for any 6u it must always be that
P"ii*<5u+P'"u 2 <$u = 0 (4.4.16)
Here (4.4.16) is a non-homogeneous variational equation, the homogeneous
part being coincident with (4.4.8). As the operator P" is singular, equation
(4.4.16) has a solution if and only if the second term satisfies the orthogonality
condition (Fredholm) with respect to the solution to the homogeneous adjoint
133

problem, which coincides with problem (4.4.8) as the operator P" is self-
adjoint. Such a condition is written as

Pc"'ii3=0 (4.4.17)

and is obviously satisfied because of (4.4.9). The general solution to the


variational problem (4.4.16) can then be put in the form

u* = a i i + ii p (4.4.18)

where mi (a an arbitrary constant) is the integral of the associated homoge-


neous equations and up is a particular integral.
This solution defines an absolute minimum of the functional (4.4.13) and so
is necessarily unique. This can be proved by assuming that equation (4.4.16)
has two distinct solutions, u{ and ii*,. Their difference must satisfy the equa-
tion

P c "[ii;-ii;]<5u = 0 (4.4.19)

from which, by putting 6u = u j — ii*,,

P'c' [ul - u;] 2 = 0 (4.4.20)

Equation (4.4.20) is obviously in contradiction with equation (4.4.7) unless


ii* = ii*.
The minimum value of the functional (4.4.13) can be assessed even before
the solution to equation (4.4.16), by an appropriate use of this equation. By
writing 6u = ii* in (4.4.16) we have

Pc"ii*2 = - P c ' " u 2 i i * (4.4.21)

from which

Min - e4 (- Pc" ii 2 + P'» ii 2 ii) - - - e4 Pc" ii*2 (4.4.22)

Equation (4.4.22) shows that the minimum of functional (4.4.13) is always


negative. By substituting for ii* the solution to equation (4.4.16), we can
calculate the value of the minimum.
Let us now turn to the final discussion of (4.4.12). From (4.4.22) we have

P (u(e)) > ^ £4 (Pr u 4 - 3 Pc" u' 2 ) (4.4.23)

from which, by setting


pm. tf = pm tf _ 3 p,, ^ (4A.24)
134

we obtain the sufficient condition for stability

Pc'"ii3 = 0 (4.4.25)

P c "''ii 4 >0 (4.4.26)

Equations (4.4.25) and (4.4.26) are analogous to equations (2.7.28) and


(2.7.29) obtained for discrete systems. Returning now to (4.4.18), we point
out that to determine a it is necessary to normalise the solution. This pro-
cedure leads us to the identification of the curve along which P ( u ) reaches
its minimum value. By introducing an arbitrary positive definite quadratic
functional T"u 2 , we require that

T"u2 = e2 (4.4.27)

from which, by virtue of (4.4.5),

e2 T" u2 + 63 T" li ii + o(€3) = €2 (4.4.28)

As (4.4.28) must be valid for any e we must have

T"ii 2 = 1 (4.4.29)

T"iiu = 0 (4.4.30)

The condition of normalisation (4.4.29) is to be associated with the eigen-


value problem (4.4.8), and equation (4.4.30) with the variational problem
(4.4.16). By introducing (4.4.18) into (4.4.30) we obtain

aT"u2 + T"uup = 0 (4.4.31)

from which, on account of (4.4.29), we have

a = -T"iiup (4.4.32)

In conclusion, as in the case of discrete systems (see (2.7.30), (2.7.31)), the


curve along which (4.4.26) is satisfied is composed of a first-order term in the
direction of the critical mode and a second-order term which is orthogonal to
it, in the sense that it satisfies (4.4.30), that is

u(c) = u e + - [ ( - T" u u p ) li + iip] e2 + o(e 2 ) (4.4.33)

Because it is arbitrary, we can, for example, choose for the functional T" u 2
the second-order terms of the elastic energy.
As an example, let us apply the criterion of stability to the problem of the
Euler beam introduced in the previous Section. A critical case of equilibrium
135

manifests itself when the axial load A is equal to the Euler load

A = —£- (4.4.34)

The corresponding buckling mode (4.3.32) can be normalised by introduc-


ing a slightly modified form of (4.4.27), T"u2 = Eln4 e 2 /2£ 3 , where for T"u2
we choose the elastic energy / 0 EIv" dx. Equation (4.4.29) is then replaced
by

„,„ . , Eln4
T (4 4 35)
* ~2iF --
whilst (4.4.30) remains unaltered. By making use of (4.4.35) we have
7T X
v = sin — (4.4.36)

As P"' v3 = 0 from (4.3.26), the second term of the variational equation


(4.4.16) vanishes and the equation reduces to (4.3.28), from which equation
(4.4.30) is obtained with the boundary conditions (4.3.31). From the vari-
ational problem (4.4.16) it follows that up = 0 and therefore a = 0 from
(4.4.32). Hence ii* = 0 from (4.4.18).
The term Pc""ii4 defined by (4.4.24) reduces to Pc"" t;4 which, by virtue of
(4.3.27), (4.4.34) and (4.4.36) yields

PT»> = - p E l l (4 .,n" - cos' - - cos* - ) dr

= 3~ ^ (4-4*7)
8 £5 v }

As both (4.4.25) and (4.4.26) are satisfied, we conclude that the equilibrium
of the beam in the fundamental configuration subjected to the Euler load
(4.4.34) is stable.

4.5 C O N S T R U C T I O N OF EQUILIBRIUM PATHS


B Y M E A N S OF P E R T U R B A T I O N A N A L Y S I S
The analysis of equilibrium configuration stability, presented in the previous
Section, has without doubt a fundamental importance from the practical
viewpoint, but in order to have a complete picture of the behaviour of the
system, we need information about all (or some) of the possible equilibrium
paths. The aim in this Section is to determine the natural equilibrium path,
that is, the equilibrium curve passing through the origin, to identify the
bifurcation points and then to trace the bifurcated equilibrium curve passing
through the first of them [1].
136

Let
p ( w ; A) =W(w) - V ( w ; A) (4.5.1)
be the total potential energy of the elastic system and let
P ' ( w ; A)<5w = W ( w ) < 5 w - T ; ' ( w ; \)6V/ = 0 (4.5.2)
be the condition of stationarity with which the equilibrium of the body is
associated.
The solution to system (4.5.2) provides all the equilibrium curves of the
body. For the problems encountered in practice their determination, however,
is in general very difficult or impossible, especially if we look for solutions in
a closed form. There are, however, numerous problems of great practical
interest in which the natural equilibrium path with which we are concerned
is linear or even trivial, and can therefore be determined exactly. If this is
not so, then it is possible to determine the equilibrium path approximately
by means of numerical procedures such as perturbation analysis, working in
exactly the same way as indicated in Section 3.6 for discrete systems. The
numerical implications in the determination of the bifurcation point and of the
bifurcated equilibrium curve lie outside the interests of the present treatment,
and will therefore not be examined.
Let us now suppose that we know the natural or fundamental equilibrium
path, the solution to system (4.5.2), in the explicit form
v = v(A) (4.5.3)
By writing w = v(A) + u, where u represents an additional displacement
measured from the fundamental path (Fig. 4.3), we make a series expansion
of the total potential energy in terms of u starting from a generic point on
the path. From (4.5.3) we have

P(v(A)+u;A) = ip"(v(A);A)u2 + ip'"(v(A);A)u3


Z o

+ ^ " " ( v ( A ) ; A ) u 4 + ... (4.5.4)

where the first term of the series expansion does not appear because of (4.5.2).
As the points of critical equilibrium are characterised by the quadratic func-
tional P"(v(A); A) u 2 being positive semidefinite, they can be determined by
solving the eigenvalue problem (4.3.22)
P"(v(A); A)u<5u = 0 (4.5.5)
which is linear or non-linear in A according to (4.5.3). The critical points
identified can be either limit points or bifurcation points. Let us suppose
that the first critical point is a bifurcation point. We can now construct the
bifurcated path by using the asymptotic method.
137

A
fundamental
p a t n
\ S

v=v(A) / yu

^bifurcated
X yfc path

v,u

Fig. 4.3 - Fundamental and bifurcated paths.

Let
P ' ( v + u ; A)<5u = 0 (4.5.6)
be the condition of stationarity of the total potential energy. The solution to
(4.5.6) in the space u,A gives the bifurcated curve u = u(A). In this space
the fundamental path is represented by the equation u = 0. Let us suppose
that the bifurcated curve is represented in the parametric form
u = u(e)
(4.5.7)
A = A(e)
Assuming e = 0 in correspondence with the bifurcation, we then have
u ( x ; 0 ) = 0,A(0) = AC (Fig. 4.3). Let us suppose t h a t relations (4.5.7) are
regular and write the series expansions with initial point e = 0

u(e) = i i c e + - ii c e 2 + o(e 2 )
(4.5.8)
A(e) = Ac + Ac 6 + - Ac e2 + o(e2)

where a dot indicates differentiation with respect to the parameter e. Along


the bifurcated equilibrium curve we obviously have

5?{P'(v(c)+u(6);A(€))*u} = 0 Ve (4.5.9)

for k = 1 , 2 , . . . where v = v(A(e)) = v(e). By writing (4.5.9) for k — 1


corresponding to e — 0, we have
P" K ; Ac) [vc + ii c ] 6 u + Ac P' ( v c ; Ac) 6 u = 0 (4.5.10)
where the symbol (") indicates differentiation with respect to A, v c = v(0)
and v c = v(0). Equation (4.5.10) is greatly simplified if we note t h a t , start-
ing from (4.5.2) and evaluating the derivative with respect to e along the
138

fundamental p a t h written in the parametric form v = v(e) ,A = A(e), we


obtain

P " (v(e) ; A(e)) v 6 u + A P ' (v(c); A(e)) j u - 0 (4.5.11)

valid for any e. Corresponding with 6 = 0, (4.5.11) is written

P " ( v c ; Ac) v c 6 u + Xc P' ( v c ; Ac) 6 u = 0 (4.5.12)

Consequently (4.5.10) is simplified t o

P " ( v c ; A c )ii c <5u = 0 (4.5.13)

The variational equation (4.5.13) represents the same eigenvalue problem


as equation (4.5.5), and therefore yields the same eigenvalues Ac and eigen-
vectors ii c . Using the logical procedure followed so far it is therefore possible
to think of passing from the calculation of the fundamental p a t h t o t h a t of
the bifurcated path, and eliminating the intermediate step of identification
of the bifurcation point, which is given directly by the first equation in the
perturbation process set up for the determination of the bifurcated curve.
By taking k — 2 and k — 3 in (4.5.9), evaluating the derivatives corre-
sponding t o 6 = 0 and simplifying the terms relative t o the perturbation
equations along the fundamental path, as before, we arrive at

P'c'uc6u = -P;'[iic2 + 2Acvciic](5u-2AcP;iic(!)u (4.5.14)

P;ucSu = -P™[3\lYlue + 3\eveiil + ul\6u

- 3 P'" [Ac VC ii c + Ac v c ii c + A* v c ii c + ii c u c ] 6 u

- 3 Ac Pc" ii c 6 u - 3 Ac Pc"' [iic2 + 2 Ac v c u c ] 6 u

-3\2cP'luc6u-3\cP'c'uc6u (4.5.15)

where P" — P"(v c ; A c ), etc., and we have made use of

vc - Acvc (4.5.16)

vc = A c v c + Ac2^c (4.5.17)

obtained by differentiating v = v(A(e)) at e = 0. It is worthwhile to note the


analogy between equations (4.5.13), (4.5.14), (4.5.15) and (3.4.56), (3.4.57),
(3.4.58) respectively, valid for discrete systems.
In many problems which are encountered in practice the potential energy
of the loads "V(u; A) is a bilinear form in u and A. In this cause all the mixed
derivatives of the total potential energy which are of an order higher than
139

one in either of the arguments u and A vanish, and equations (4.5.14) and
(4.5.15) are greatly simplified.
As the operator P" is singular, equation (4.5.14) has a solution if, and only
if, the right side satisfies the Fredholm condition of orthogonality

Pc'" [iic2 + 2 Ac v c ii c ] ii c + 2 Ac Pc" iic2 - 0 (4.5.18)

from which we obtain


1 P ' " ii 3
Ac = - - ^ - ^ (4.5.19)
2 P c '''v c ii? + Pc»uc2
by analogy with (3.4.59) for discrete systems. Knowing Ac we can solve equa-
tion (4.5.14) in terms of ii c , obtaining a solution in the form u c = auc + upc
with a an arbitrary constant which is determined from the normalisation
condition and ii p c a particular integral. By applying the Fredholm condition
of orthogonality to the right side of (4.5.15), we obtain the expression for Ac:

+ P : UC UC + p ' : u»] + \\ [PT ♦; u2c+p'c" h ui

+ 2 P'c" v c uc2 + P"c ul ] } / [ i f v c u? + P ; ul] (4.5.20)

analogous to (3.4.60). If Ac = 0 then (4.5.20) is simplified into


4 2
1 P
P""ii
c uc - "* P"i'i
6Pcuc
3 P-v.u^ + P " ^
having used (4.5.14), written for 6u = uc.
Once the coefficients of the series expansion (4.5.8) have been calculated, it
is possible to trace the bifurcated equilibrium curve A = A(u), approximated
up to the second order. If, for example on the basis of the normalisation
chosen in solving the problems (4.5.13) and (4.5.14), we have e = u, where u
is the non-dimensional displacement of a point or a rotation, then from the
second of equations (4.5.8) we have

^ = 1 + A1u + A2u2 + o ( u 2 ) (4.5.22)


Ac
with Ax = A c /A c and A2 = A c /2 Ac, which is the equation of the curve A = A(u).
Figures 4.4 (a), (b), (c) illustrate all the possible cases connected with the
value and the sign of the coefficients of equation (4.5.22). It is evident t h a t
we have a repetition of the situation already encountered in systems with a
finite degree of freedom illustrated by the models in Sections 2.8-2.10. In
140

A/A, stable eq.


unstable eq.
A2<o-

*N A2<0

a) b) c)

Fig. 4.4 - Possible types of post-critical behaviour of a structure near a


bifurcation point.

the figure we also have the equilibrium curves of the imperfect structures, for
which no details are given as the subject of imperfections is not treated in
the present volume.
It should be noted that in general, in all cases in which Ai ^ 0, the cal-
culation of A2 is omitted. This is equivalent to approximating the behaviour
of a structure with the tangent to the effective equilibrium curve at the bi-
furcation point, with results which are wholly acceptable in practice. The
analysis is obviously greatly simplified in that it is not necessary to solve the
variational equations (4.5.14) to determine u c and hence Ac. In fact, it is
exactly the solution to equations (4.5.14) which constitutes the most difficult
part of the whole analysis.
It is simple to verify that along the fundamental path the equilibrium
is stable or unstable, depending on whether the external load is smaller or
larger t h a n the critical load. In fact, as the second variation of the energy
P ( v + u ; A)

^P"(v(A);A)u2 \ W" (v(A)) u 2 - \ V (v(A) ; A) u 2 (4.5.23)

is positive semidefinite for A = Ac, it is obviously positive definite for A < Ac


in t h a t the elastic energy prevails over the energy of the loads, and negative
definite for A > Ac for the opposite reason. Consequently, the equilibrium is
stable for A < Ac and unstable for A > Ac. The demonstration of the type of
equilibrium along the bifurcated paths in Fig. 4.4 is even more complicated.
The reader who is interested in this can refer to [1].
Before closing this Chapter it is useful to make some considerations of a
general character, even though for simplicity we refer to a particular case.
Let us demonstrate first of all that the term P" ii 2 is always negative. We
141

consider the term P"(v(A); A)ii* and make the series expansion in terms of A
starting from A = Ac:

P" (v(A) ; A) ul = P? u2e + (X- Ac) Pc" uc2 + • • • (4.5.24)

By virtue of (4.4.6), equation (4.5.24) is reduced to

P " (v(A) ; A) ul = (A - Ac) P," u* + ■ ■ ■ (4.5.25)

Having shown that along the fundamental p a t h the left side is positive
definite for A < Ac and negative definite for A > Ac, it follows t h a t P"ul is
negative.
Let us now consider a type of structure for which v = 0 and Ac = 0. A
structure of this type could be, for example, the Euler rod with inextensible
axis previously examined. This hypothesis implies t h a t the term P"'vcul
vanishes identically. From (4.5.21) we therefore have the sign of Ac coinciding
with the sign of the numerator. As the numerator coincides with P""u*
(see (4.4.24)), it follows t h a t if at the bifurcation point the equilibrium is
stable (Pc""ii* > 0) the concavity of the curve is upwards (Fig. 4.4(b)), and
if unstable (Pc""ii* < 0) the curve has a downwards concavity (Fig. 4.4(c)).
Finally it can immediately be verified that if at the bifurcation point the
equilibrium is unstable because of violation of (4.4.25), t h a t is, Pc'"u;? ^ 0,
then Ac ^ 0 follows from (4.5.19). This means that the slope of the bifurcated
equilibrium curve at the critical point is different from zero (Fig. 4.4(a)).
142

REFERENCES
W.T. Koiter: On the stability of elastic equilibrium (in Dutch), Disserta-
tion, Delft, 1945; English translation published as NASA T T F-10, 833,
1967 and A F F D L Report T R 70-25, 1970.

W.T. Koiter: "Stability of equilibrium of continuous bodies", Techn.


Report No. 79, Brown University (1962).

W.T. Koiter: "On the concept of stability of equilibrium of continuous


bodies", Proc. K. Ned. Akad. Wet. Ser. B, 66 (1963), 173.

R.T. Shield, A.E. Green: "On certain methods in the stability theory of
continuous systems", Arch. Ration. Mech. Anal., 12 (1963), 354.

W.T. Koiter: "The energy criterion of stability for continuous elastic


bodies", Parts I and II, Proc. K. Ned. Akad. Wet. Ser. B, 68 (1965), 178.

W.T. Koiter: "On the thermodynamic background of elastic stability


theory", in Problems of hydrodynamics and continuum mechanics, SIAM,
Philadelphia, 1969.

R. Baldacci: Strength of materials (in Italian), Vol. I, U.T.E.T., Torino,


1980.

C.L. Dym, I.H. Shames: Solid mechanics: a variational approach,


McGraw-Hill Kogakusha Ltd., Tokyo, 1973.

B. Budiansky: "Theory of buckling and post-buckling behaviour of elas-


tic structures", in Chia-Shun Yih (editor): Advances in applied mechan-
ics, vol. 14, Academic Press, New York, 1974.
143

Chapter 5

ANALYSIS OF BEAMS A N D P L A N E
FRAMES

5.1 INTRODUCTION
In this Chapter we shall deal with the analysis of bifurcation in structural
systems consisting of one-dimensional continua (beams) in a two-dimensional
space. In this context the simplest system is without doubt t h a t of a single
rectilinear element, and Fig. 5.1, together with the four following figures taken
from [23], illustrate for a particular condition of constraint and load its well
known behaviour at bifurcation.

IK

Fig. 5.1 - Single b e a m

If in general such a simple system has mainly exemplifying and historical


interest, in some particular cases it can be of help in the study of more complex
systems. Figure 5.2 shows the behaviour at bifurcation of a system consisting
of two rectilinear rods hinged to each other and to the body of reference.
From examination of the figure we can understand how, in this case, the
144

Fig. 5.2 - Frames with hinged bars.

behaviour of the system is identified with that of one of its components and
therefore the analysis of problems of this type can be reduced to that of the
single beam.
It is possible to illustrate further examples in which the information looked
for is substantially contained in the case of the single beam. Examples of this
type, however, are in general uncommon since most of the time the behaviour
of systems of beams depends precisely on the collaboration between various
elements [3-5,8,11,14,22]. To get an idea of the importance of this fact, we
consider the previous system in which the rods are connected to each other
by a device which does not permit relative rotation of the bars at the ends.
Figure 5.3 shows how the mutated constraint conditions substantially change
the behaviour of the system at bifurcation and how the collaboration of both
the rods considerably increases the value of the critical load.
Figure 5.4 shows another classical example of the bifurcation of a plane
frame and illustrates the behaviour of the system as a whole. This structure
will be examined in detail in future Sections, where we shall show how not
only the geometrical parameters of the single rod, but also the span/height
ratio can also greatly influence the qualitative behaviour of the frame.
In figure 5.5 two multi-storey frames are illustrated, unbraced and braced,
respectively. In the first case bifurcation is associated with a field of asym-
metrical displacements, whilst in the second case it is associated with a field
of symmetrical displacement. The influence of bracings on the value of the
critical load is very great.
In this Chapter, before examining more complex cases, we shall present
145

Fig. 5.3 - Frames with rigid node.

the analysis of some sample structures which have a double importance: on


the one hand they constitute, to a certain extent, a collection of the principal
types of behaviour of plane frames, and on the other they give the possibility
of comparison of the theoretical results obtained by the application of a non-
linear analysis with experimental results [7,15].
Before passing to the analytical treatment of the various cases, it is neces-
sary to call attention to a question of great theoretical and technical impor-
tance: the structure of the mathematical model of beams to be used. Tradi-
tionally, in this type of analysis, a model of beams characterised by inextensi-
bility and shear undeformability is used; this type of simplification is perfectly
consistent with the behaviour of the rods used in the applications. The adop-
tion of this type of model of beams, however, far from being a simplification,
constitutes rather a complication for the occurrence of non-linear kinematic
boundary conditions at the nodes [26]. This operational difficulty can be
overcome in various ways. In particular, it is possible to pass directly to a
model of three degrees of freedom [18,20,27] or introduce auxiliary constraint
conditions through Lagrange multipliers [26]. Both of these possibilities have
advantages and disadvantages, and we shall present both procedures here.
It remains to show how the adoption of ad hoc methods of investigation in
buckling phenomena, widely used in the past and already presenting great
problems at the level of the traditional linear analysis of elastic stability, be-
come impracticable for non-linear systems. On the other hand it will emerge
from the following t h a t , despite appearances, an instrument such as the non-
linear theory of elastic stability is convenient and suitable for synthesis of
more complex cases.
146

P«0 F«0

vt&rn w^r

i *

r&gfim ^rtygjpm

Fig. 5.4 - Hinged portal frame.


147

Fig. 5.5 - Multi-storey frame with and without bracings.


148

5.2 BEAM MODELS

5.2.1 Model in which the configuration is identified by


three functions
Let us consider the beam as a one-dimensional continuum whose configura-
tions are represented by a regular curve in the plane which we call the beam
axis, and let us associate with each point of the curve a cross-section. We
shall assume t h a t in the configuration of reference the beam axis is rectilin-
ear and the cross-sections are orthogonal to it. Further, we shall denote the
abscissa on the reference beam axis by x £ [0,£].
Let us introduce the displacement components

<p(x) , w ( x ) = u ( x ) i + v(x)j (5.2.1)

where the vector w ( x ) represents the displacement of the points belonging


to the beam axis and the scalar function <p(x) represents the rotation of
the cross-sections. The knowledge of the displacement function allows us to
identify the generic configuration of the beam.
Let us define the position vector r(x) and the unit vectors a(x) , b ( x ) (Fig.
5.6) as orthogonal and tangent, respectively, to the beam cross-sections. In
the undeformed configuration the unit vectors a and b do not depend on x
and are all parallel to i and j respectively.

Fig. 5.6 - Current and reference configurations of beams.


149

It is easy to verify the relations


r(x) = x i + w(x)
a(x) = cos <p(x) i + sin <p(x)j (5.2.2)
b(x) = — sin (p(x) i + cos <p(x) j
Let us introduce the deformation measures e , 7 , x which are defined by
the expressions [20]
r' = (l + e)a + 7 b
(5.2.3)
X = <P
where a dash indicates differentiation with respect to x. By making use of
(5.2.1) and (5.2.3) we arrive at the relations
e = (1 + u') cos (p + v' sin (p — 1
7 = v' cos <p — (1 -f u') sin <p (5.2.4)

X = <p'
We can see t h a t relations (5.2.4) are exact in t h a t they vanish in correspon-
dence with a rigid displacement. Indeed, if we consider a generic rigid dis-
placement u r (x) = u0-hx(cos<£>0 — l)j vr[x) = t>0-|-xsin<£>05 ^ r ( ^ ) — V^o? where
u
o ? vo ? <Po a r ^ arbitrary constants, and substitute these expressions in (5.2.4)
we obtain e = 7 = x = 0-
Let us introduce the scalar fields N, T, M so t h a t the internal force and
bending moment are represented by the vectors
t = iVa + T b
(5.2.5)
m = Ma x b
where the symbol ( x ) denotes the vector product. The constitutive relations
are taken as linear and are written in the form
N = EAe
T = GA7 (5.2.6)

M - EIX
being E A, G A, El respectively the axial, shear and flexural rigidities.
The construction of the Total Potential Energy (TPE) functional (4.5.1)
for the single beam element is now easy
Pe(w;A) = W e (w) - TUw ; A)

= - f (EAe2 + GAi2 + EIX2)dx


2 Jo ^ '
— X L(w) + boundary terms linear in u , v , <p (5.2.7)
150

where we denote by L(w) a linear functional in w so t h a t the potential energy


of the external loads is bilinear in w and A. Note t h a t from now on w collects
all displacement components listed in (5.2.1). Substituting (5.2.4) in (5.2.7)
we have

Pc(w;A) = - / {J5A[(1 + u') cos <p + v'sirup- l]2


2 Jo ^
+ GA [v' cos <p - (1 + u') sin <p]2 + EI(p,2}dx

— A L(w) + boundary terms linear in u , v , <p (5.2.8)


By applying the procedure described in Section 4.5, it is possible to write
explicit expressions of the differential equations and the related natural bound-
ary conditions which follow from the perturbation process. In particular, from
(4.5.13) and putting Nc = EAec, Tc = GA^C, Mc = Elxc, the first-order
perturbation equations are written as
EAu':-{Tc<pc)' = 0
{Ne<pe)' + GA{v'c-<pc)' = 0 (5.2.9)
(Nc -GA) (v'c - <pe) -Tcu'c-EI<p'! = 0
with the boundary conditions
{{EAu'c-Tc<pc)6u + [Nc<pc-GA{<pc-v'c)}6v

+EI<p'c6<p}l0 = 0 (5.2.10)
The second-order perturbation equations, from which it is possible to find
the secondary mode, are written from (4.5.14) as
EAu't-(Tc<pc)' = {(EA-2GA + Ne)(ipl+2\e(pe<pe)

-2{EA-GA) [v'c<Pc + KiKvc + Kvc)]}'

{Nc<pey + GA(v'c-<pc)' = {-2{EA-GA)[u'e<pe

+ Xc (u'c<pc + u'c<pc)}'

+ Te(<pl + 2\e<pe<pe)}' (5.2.11)

(Nc - GA) (v'e - <pc) - Tcu'c -EI<p» =

2{EA-2GA + Ne)[u'e<pe + Xc {u'c <pc + u'c <pc)\

-2{EA-GA)[u'ev'e + K (u'c v'e + u'c v'cj\

+ Tc [2 ipcv'c-(p2c + 2 \c (<pc v'e + (pcv'c- <Pc <Pcj\


151

with the boundary conditions

{[-Nc(<p2c+2\c<pc<pc)

+ EA(2i/e<pe - <p2c + u'c + 2 Ac {v'e<pe + v'c<pc - <pc<Pc))

+ 2GA (<p2c -<pcv'c + K {2<pc<pc - <pc v'c - <pcv'c)) - Tc<pc] 6u

+ [Nc<pc + 2EA(<pcu'c + \c {u'e<pe + u'c<pc))

+ G A (- 2u'c<pc - <pc + v'c - 2 \c (ti'c <pc + u'c<pc))

-Tc (<pl+2\c<pc<pc)]6v + EI<p'c6<pYQ = 0 (5.2.12)

Note t h a t the preceding expressions have been obtained assuming t h a t the


displacements along the fundamental path are zero. Note also t h a t from now
on, differentiations with respect to A refer to the fundamental p a t h v = v(A).
Besides, as in Section 4.5, the symbol (") denotes differentiation with respect
to A along the fundamental path.
It is appropriate at this point to make some observations which can furnish
further details on the mechanical interpretation of equations (5.2.9)-(5.2.12).
They have been obtained by performing a series expansion (with initial point
w = 0) of the expression of the virtual work principle P'Svs = 0 and therefore
are, in our beam model, the incremental equations of equilibrium, written in
the generic configuration. To make clear these facts, which have already
been mentioned in Chapter 3, we first write the equations of equilibrium in
a direct form and then perform an asymptotic expansion with initial point
w = 0, showing how the expressions obtained in this way coincide with
(5.2.9)-(5.2.12).
Let us consider, with this in mind, a beam element of length dx (Fig.
5.7). The following two figures, the symbols of which are obvious, show the
reference configuration, an arbitrary configuration and the forces acting on
the beam element.
The equations of equilibrium in scalar form in the deformed configuration
are written as

- (N cos tp)' + (T sin tp)' = 0


- (N sin (p)' - (T cos <p)' = 0
(5.2.13)
TV [v1 cos <p — (1 + u') sin (p\
-T [v1 sin (p + (1 + u') cos <p] - M' = 0
152

N+dN

M+dM

Fig. 5.7 - Components of displacement and of internal forces.

The third of equations (5.2.13) can, bearing in mind (5.2.4), be written in


the more expressive form

-M' + N1-Te-T = 0 (5.2.14)

The three differential equations of equilibrium can finally be written as

(— N cos <p + T sin (p) — 0

(— N sin <p — T cos <p) = 0 (5.2.15)

-M' + Ni-Te-T = 0

From these the first- and second-order incremental equations are obtained

(— N cos <p + N (p sin <p + T sin (p + T <p cos <p) = 0

(— N sin (p — N (p cos <p — T cos <p + T <p sin (p) = 0 (5.2.16)

-M' + N-l + Ni-fe-Te-f = 0


153

f — N cos <p + 2 N <p sin <p + N <p sin <p + N <p2 cos <p

+ T s i n <p + 2T <p cos (p + T <p cos <p — T (p sin <pJ = 0

f — TV sin £> — 2 JV <p cos <p — N (p cos <p + TV <p2 sin <p (5.2.17)

— T cos (p + 2T<p sin <p -\- T<p sin <p + T<p cos £>J = 0

- M ' 4 - i V 7 + 27V7 + i V 7 - f ' 6 - 2 r e - r e - f =0

These equations, evaluated at the bifurcation point for which u = v = <p =


0 , e = 7 = x = 0? become

(-AT c + T c ^ c ) ' = 0

(-fc-Nc<pc)' =0 (5-2.18)

-M'c + Ncic-Tcec-fc = 0

( - TV, + Tc <pc + Nc <p\ + 2 fc <pc)' = 0

( - f c - JVcV3c + T c # - 2 7V C ^ C )' - o (5.2.19)

-M[ + iV c 7 c - Tc ec - fc + 2iV c 7 c - 2TC ec = 0

Equations (5.2.4), in incremental form, are written as

*e = U'e €c = u'e-(p2e+2 v'c <pc

ic = v'c--<Pc 7 C = v'c - <pc - 2 u'c <pc (5.2.20)

Xc = <P'C Xc = <P'C

By now writing (5.2.18), (5.2.19) and (5.2.20) along the two equilibrium
paths and by subtracting, bearing in mind (5.2.6) and the relation v c = Ac v c ,
we obtain the expressions (5.2.9) and (5.2.11) given previously. T h e algebraic
steps, as well as the expressions of Ac and Ac, are omitted for brevity.
At this point we have all we need to perform the critical and post-critical
154

analysis of the system of beams. In fact, the T P E for these will be of the
type

P(w;A) = £ > e ( w ; A ) (5.2.21)


e=l

From the perturbation process, n systems of differential equations (5.2.9),


(5.2.10) and (5.2.11), (5.2.12) will originate and will be solved, case by case,
taking into account the boundary conditions. Note that, having formulated
the problem by adopting u, v, <p as independent variables, the boundary condi-
tions are easily expressed and, except for particular cases of constraint which
are not considered here, they are linear.

5.2.2 Internally constrained model


If the continuum is internally constrained, then there are fewer degrees of
freedom and non-linear boundary conditions arise. In particular, it is simple
to derive a model constrained axially and in shear. From (5.2.4), setting
e = 7 = 0 and taking <p as independent variable, we obtain

u' = cos <p — 1

v* = sin (p (5.2.22)

X = <P*

and relations (5.2.6) reduce to

M = EIX (5.2.23)

Equation (5.2.7) therefore becomes

A(w;A) = I f''EI<p'2dx
L JO

+ energy of external loads


+ boundary terms in <p (5.2.24)

As we see, the constraints expressed by (5.2.22) lead to a functional T P E


in <p(x) only and therefore to a variational problem in <p. This represents,
indeed, a great simplification from an operational point of view, but there
remains the difficulty of expressing the boundary conditions in u and v in
terms of (p. This leads, as is evident from (5.2.22), to non-linear relations.
155

Obviously, this fact does not appear in linear analysis because then (5.2.22)
becomes
= 0
v1 — <p (5.2.25)

X = <P

and therefore, by expressing <p in terms of r, we arrive at a problem with


linear boundary conditions. However, when we are interested in the post-
critical analysis, the linearisation of (5.2.22) is insufficient and we must take
into account terms of higher order than the first in the series expansion, con-
sistent with the order of the perturbation problem under consideration. What
has been said corresponds to searching for the stationarity of the functional
(5.2.24) in the sub-space of regular functions <p which satisfy (5.2.22). The ad
hoc solutions to a problem of this type are not practicable and are extremely
laborious [10]. It is therefore worthwhile to solve the problem of constrained
stationarity by constructing a modified TPE (MTPE) functional, obtained
by introducing the auxiliary constraint conditions into (5.2.24) by means of
Lagrange multipliers.
Let us consider the generic beam element in Fig. 5.8 in the straight con-
figuration, where the positive directions of the kinematic quantities and end
forces are indicated [26]. From the first two relations of (5.2.22) we get

u(£) - u(0) - / (cos <p - 1) dx R = 0 V R


Jo
(5.2.26)
v(t) — v(0) — / sin <pdx S = 0 V5
Jo
The two terms which appear on the left of (5.2.26), if R and S are interpreted
as forces, have the dimensions of energy; in particular, they represent the
energy associated with that part of the nodal displacements u and v which

v(x)
<Kx).
R x=Q
^ u(xj x=L
A

Fig. 5.8 - Components of displacement and of end forces.


156

violates the kinematic compatibility. Because of this, it is necessary to impose


t h a t these terms be zero.
Equation (5.2.24), modified by the addition of the terms (5.2.26), becomes
under the hypothesis of external loads applied to the nodes

ne(w;A) = i llEI<p'2dx
I JO

+R u(£) - u(0) - / (cos (p - 1) dx


Jo

+ S v(l) — v{0) — / sin <pdx AL(w)


Jo
+ linear boundary terms in <p (5.2.27)

Note t h a t in (5.2.27) w has as components the rotation <p, the nodal


displacements and the Lagrange parameters, whilst the only components of
w are the nodal displacements. It is easy to see by applying (4.5.13), under
the hypothesis <pc — 0, that the critical mode is described by the system of
equations

-EI& + Rc<pc-Sc = 0

uc{t)-uc{0) =0 (5.2.28)

Jo
with the boundary conditions

[Rc6 U + Sc6 v + EI <p'c6 tp}1 = 0 (5.2.29)

and the secondary critical mode is described by the system obtained from
(4.5.14)

- EI <p"c + Rc <pc - Sc = - 2 Rc <pc + Sc <p\

- 2 Ac (RC <pc + Rc <pc + Sc <pc tp^j

[i -ft (5.2.30)
uc(l) - uc(0) = - <pcdx-2\c (pc<pcdx
Jo Jo

vc{£) - v c (0) = / <pe dx


Jo
with the boundary conditions

[Re6 u + Se6 v + EI <p'e6 <p\£ = 0 (5.2.31)


157

The explicit expressions of the terms which appear in the relations which
give Ac and Ac are

n>c3 = f (3<p2cRe + Sc<pl)dx (5.2.32)

n'c"vcuc2 = j (2<pc<pcRc + Rc<p2e + Sc<pe<pl)dx (5.2.33)

n'c'V = £(-Rc<p4c+4Sc<p*)dx (5.2.34)

rC'v c u c 3 = j (-Rc<pc<p3c + 3<pcSc<p2c + Sc<pl)dx (5.2.35)

K'*Wc = jT (-Re<p\<p\+2Setpeip\ + 2<p\Sc<pe)dx (5.2.36)

n™vii c u e = 1 (<Pe <Pc Re + <Pc Re <Pc + Re <Pe <Pe

+ Sc (pc (pc <pc) dx (5.2.37)


Relations (5.2.28)-(5.2.31) correspond to (5.2.9)-(5.2.12) written for the
beam model without internal constraints, and can be used only in problems
relative to a single beam. The generalisation to the case of systems of beams
does not present any particular difficulty, as is illustrated in the following
[26].
Let us consider a system of n beams and m nodes. Let e be a generic
element of the system and i and j the nodes, t h a t is, the points at which
e is connected to the other elements of the system (Fig. 5.9). To describe
the displacements of the element points (and the rotation of the sections

v(x e ) / j

u(x e )

-R
<p. "e

Fig. 5.9 - Beam element.


158

connected to them) we adopt a local system of coordinates where the xe axis


coincides with the beam axis.
Let us now introduce a global system of coordinates (x, y). Denoting the
angle between x c and x by 0e, the kinematic relations

u(0) = Ui cos 0e + V{ sin 0e


v(0) — — U{ sin 0e + Vi cos 0e
(5.2.38)
u(te) = Uj cos 0e + Vj sin 0t
v(te) — — Uj sin 0e + Vj cos 0e

hold true.
Expressing the equations (5.2.26) in terms of node displacements by means
of (5.2.38) we obtain, with <pe = <p(xc),

Re (UJ — U{) cos 0e + (VJ — Vi) sin 0e

- / (cos <pe - 1)dxf 0 VJL


o
(5.2.39)
— (UJ — u^ cos 0t + (VJ — Vi) sin 0t

— / sm <pe dxe = 0 V5,


Jo

The M T P E of the element e, in terms of the node displacements % and j ,


is obtained from (5.2.27), making use of (5.2.39). We have

1 /■«.€
nc(wc;A) = - f EIt<p'l dxe
L JO

+ Re | (UJ - Ui) cos 0C + (VJ - Vi) sin 0e

/ (cos <pe - 1) dxe


Jo

+ Se | - (UJ - u^ sin 0e + (VJ - Vi) cos 0e

- / sin <pe dxe


Jo
— X L(v7e) + boundary terms linear in <p (5.2.40)

where w c = {^>c, J? c , 5 c , u t , v t , u ; , vy}. The M T P E of the whole system will


159

therefore be

n(w;A) = ]Tne(we;A) (5.2.41)


e=l

By using (5.2.40), the first-order perturbation equations (4.5.13), corre-


sponding to (5.2.28) and (5.2.29), are written as

-EIe<pl + Rec<pec-Sec = 0

{UJC ~ uic) cos 0e + [ijc - vic) sin 0t = 0 (5.2.42)

-{ujc - uic) sin 0e + (vjc - vic) cos 0e = / <pec dxe


Jo

Rec cos 0e 6(UJ — uij + Rec sin 0e 6(VJ — V{)

— Sec sin 0e6(uj — Ui) + See cos 0C 5(vy — vt) (5.2.43)

+ EIe<p'ec{Q 6<pe{Q - EIe<p'«{0) 6<pe{0) = 0

The second-order perturbation equations (4.5.14), corresponding to (5.2.30)


and (5.2.31), are

-EIe<p"tc + Rec<pec - Sec = -2Rec<pec + Sec<p]c

- 2 Ac [Rtc <pec + Rec <pec + Sec <pcc <pec)

(5.2.44)
fie
(ujc - uic) cos 0e + (vjc - vic) sin 0t = - / <pec dxe
Jo
. fie
- 2 Ac / <ptc <pec dxe
Jo

-{tijc - Uic) sin 0e + (vjc - vic) cos 0t—\ <pec dxe


Jo

Rec cos 0e 6{uj - Ui) + Rec sin 0e 6(VJ - v{)

- Sec sin 0e 6(UJ - Ui) + Sec cos 0t 6(VJ - v,-) (5.2.45)

+ EIe <p'ec(Q 6<pe{Q- EIe $tc(0) 6<pe(0) - 0

To obtain the differential and algebraic equations and related boundary


conditions, suitable for solving problems of systems of rods, it is necessary
to use the perturbation process starting with the functional (5.2.41) instead
of (5.2.40). However, as this is obtained from the sum of n terms of type
160

(5.2.40), its variations can be obtained by first finding the variations of the
single terms and then adding them together. In particular

ntteue6u = f2 n';c uec s u (5.2.46)


e=l

leads to a system of n differential equations of the same type as the first of


(5.2.42) and to 2n algebraic equations of the same type as the second and
third of (5.2.42). Equation (5.2.43) on the other hand is written
n

Y^ [Rec COS 0e 6(UJ - Ui) + Rec sin 0e 6(VJ - v t )

- Sec sin 0e 6(UJ - u^ + Sec cos 0e S(VJ - V{) (5.2.47)

+ EIe<p'ec(Q6<pe(Q - £ J . # c ( 0 ) * < p . ( 0 ) ] =0
which provides the boundary conditions of the system of differential-algebraic
equations mentioned above. The same reasoning is valid for (5.2.44) and
(5.2.45).
The values of Ac and Ac are calculated by using the expressions (4.5.19)
and (4.5.20) respectively. It is necessary only to observe that the various
terms which appear in these expressions are now obtained by adding the
contributions of the various beam elements. For example

IC *J = E C ui = ± I'' (3 <pl Rec + Sec tfc) dxe (5.2.48)


e=l e=lJ0

and so on, for each of the expressions (5.2.32)-(5.2.37).


Lastly, we would like to point out the difference between R and S defined
for the constrained model and the generalised forces N and T defined in the
previous Section. First of all, whilst N and T are defined for each point of
the beam, R and S are defined only at its ends. Furthermore, the values
t h a t TV and T assume at the ends of the beam do not generally coincide with
those of R and S. This is due to the fact that N and T are the components
of the resulting force expressed in a reference frame fixed to the cross-section
of the beam and so dependent on its configuration, whilst R and S are the
components of the same resulting force expressed in a frame independent of
the section configuration. It is easy to make the relationship between the
forces mentioned above explicit (Fig. 5.10)

Rej = Nej cos <pej - Tej sin tpej


(5.2.49)
Sej = Nej sin (pej + Tej cos <pej

where j is the generic end of the beam. The incremental expressions of


161

-R.

Fig. 5.10 - Components of end forces in different reference frames.

(5.2.49) with reference to the perturbation analysis described so far are, up


to the second order,

Rcc = Nec - Tec (pec


(5.2.50)
Sec = Nec (pec + fec

R,r — Nec - Ttc <pec - Nec <p2ec - 2 fec <pec

- 2 \ c (Nec <pcc <pec + fec <pec + fec <Pcc)


(5.2.51)
&pr — fec + Nec <pec - Tec <pl + 2 Nec <pec

- 2 Ac (Tec <pec <pec + Ncc <pcc + Nec <pe^

5.3 CRITICAL LOAD A N D P O S T - C R I T I C A L B E -


HAVIOUR OF B E A M S LOADED AXIALLY AT
ONE E N D

5.3.1 The Euler beam


Let us consider a simply supported beam under axial load, as introduced in
Section 4.3, and divide it into the beam element and two nodes, as shown in
Fig. 5.11.
The explicit expression of (5.2.41) is

n ( w ; A) = [ \- EI<p'3 - Rlcos tp - 1) - S sin ip] dx


Jo L2 J
+ RuB + \uB (5.3.1)
162

B X

x=0

Si v(x) st
R
A T <K
A Kx) * U(x) B

Fig. 5.11 - Euler's beam.

where w = {<p,R,S,uA, vA,uB,vB} and we have used the boundary condi-


tions

UA = 0 , vA = 0 , vB = 0 (5.3.2)

with uA,vA,uB,vB node displacements. From the stationarity of (5.3.1) with


respect to w we have

n'<5w = [ [EI(p'6(p'-(cos <p-l)6R + {Rsin<p-S cos<p)6<p


Jo
- sin <p 6S] dx+(R + \)6uB + uB6R + uB6\ = 0 (5.3.3)

from which, by equating the coefficients of 6uB to zero, we have t h a t R = — A


corresponding to the equilibrium. Equation (5.3.3), in consequence, simplifies
to

II'(5w = / [El(p18(pf + (cos <p - 1)<5A - (A sin <p + S cos <p)6(p


Jo
-sm<p6S}dx = 0 (5.3.4)

Note t h a t (5.3.4) could have been derived by starting from (5.3.1) and
imposing a priori that the equilibrium of node B be satisfied. This would
lead to R = — A and therefore to the elimination of the last two terms on the
right of (5.3.1).
In the study of beams subjected to various types of constraint, we shall
always suppose that the equilibrium of node B in the direction of the axis x
has been satisfied a priori and so shall make use of the functional M T P E in
the simplified form

I I ( w ; A) = / [ - EI<p'2 + A(cos (p - 1) - 5 sin <P dx (5.3.5)


io L2
163

except to modify the term in 5 if VB ^ 0. Comparing (5.3.5) with the


functional T P E of classical theory

(w;A) = |Q [ i £ V 2 + A ( c o s v ? - l ) dx (5.3.6)

we find t h a t , whilst the first of the boundary conditions (5.3.2), UA = 0, has


been used in writing the potential energy of the loads, the equation

v{t) = [ sin<pdx = 0 (5.3.7)


Jo
which comes from the second and third relations of (5.3.2) and the second
of (5.2.22), has been ignored. It is therefore evident t h a t if in the study of
the stationarity of the functional (5.3.6) we take into account the auxiliary
condition (5.3.7) by means of Lagrange multipliers, we obtain the functional
(5.3.5). It should be noted here t h a t (5.3.6) coincides with (4.4.23) if v(x) is
used as the only displacement function for the description of the kinematics
of the internally constrained beam. The reader may now ask himself why, in
Section 4.3, mention of the presence of auxiliary conditions to be accounted for
in the study of the stationarity of the functional (4.3.23) was not made. The
reason, as will be seen in the following, is t h a t in the case under examination
5 = 0 and therefore (5.3.5) and (5.3.6) coincide. This is not true in general
(see Section 5.5), and therefore we shall always analyse the beams starting
from the modified functional (5.3.5).
Along the fundamental path, it is simple to verify t h a t

<p(\) = 0 , 5(A) = 0 (5.3.8)

For evaluation of the critical load and buckling mode we use (4.5.13), where
the T P E functional is replaced by the modified functional M T P E . We have

II"u6u = / \EI(p'6<p' — A cos <p<p6(p — 5 cos (p6(p

+ 5 sin <p(p6 <p — cos <p <p 6S] dx (5.3.9)

from which

Jl'^uc6u=f (Elip'Jip' -\c(pc6<p-Sc6<p-<pc6S)dx = 0 (5.3.10)

having used (5.3.8). The variational equation (5.3.10) leads to the differential
equation

EItptte + \e<pe + Se = 0 (5.3.11)


164

and to the boundary conditions

JW e (0)=0, EI<p'e{l)=0 , f <pcdx = 0 (5.3.12)


Jo

The solution gives

* . = " ^ ) ( «

X
<pc(x) — cos 7r — , Sc — 0 (5.3.14)

having normalised <£>c(x) in such a way that <pc{0) = 1. Note that Ac coincides
with the value furnished by (4.3.33). For the calculation of Ac, we use (4.5.19).
We have

IT'"iiu<5u = / fA sin <p<p<pS<p + S sin <p<p6<p + S sin tptp6<p

-\- S cos <p<p<p S <p + sin (p<p<p6 S) dx (5.3.15)

from which, by first identifying ii and 6u with li and then li with v , u and
6u with ii, we have respectively

n ' " ii 3 = I (A sin <p <p3 + 3 S sin <p <p2 + S cos <p <p3) dx (5.3.16)

II"' v ii 2 = / (A sin (p <p (p2 + 2 S sin <p<p<p + S cos (p<p<p2

+ sin <p S (p2 ) dx (5.3.17)

In correspondence with the bifurcation point, (5.3.16) and (5.3.17) both


vanish by virtue of (5.3.8). From (5.3.9) we have besides

n"ii£u = - / cos <p<p6(pdx (5.3.18)


Jo

which leads to

fl':ul = -Jot<pldx = -^ (5.3.19)

having used the first of equations (5.3.8) and (5.3.14). By substituting in


(4.5.19) the results obtained, we have Ac = 0. For the calculation of u c we
use (4.5.14), written in the simplified form

n c ' ii c 6 u - - n'c" iic2 6 u (5.3.20)


165

The left side coincides with (5.3.10) if quantities with one dot there are
replaced by quantities with double dots. The right side, instead, is obtained
from (5.3.15), identifying u with u and calculating the functional at the bi-
furcation point c. We have

/ (EI<P'C6<P'-\c<pc6<p-Sc6<p-<pc6S)dx = 0 (5.3.21)

as H"' ill 6u = 0, which is simple to verify. The variational equation (5.3.21)


leads to a differential equation identical to (5.3.11) with the same boundary
conditions. Therefore we have the solution

<pc = 0 , Sc = 0 (5.3.22)

having normalised in such a way that <pc is orthogonal to <pc. In order to


evaluate Ac we observe that from (5.3.21), by identifying 6u with u , we im-
mediately have IT" ii* = 0. Furthermore

K" *c = K T & dx = I Xc t (5.3.23)


./O o

From (4.5.21) we therefore have

Ac - \ Xc (5.3.24)
4

We observe that from

<p{x9t) = £c(*,0) 6 + i v5 c (x,0) 62 (5.3.25)

by virtue of the normalisation for <pc and <pc we have

6 = p(0,e) = 0 (5.3.26)

for which (4.5.8) yields

A = ! + I e2 (5.3.27)
Ac 8

which represents the projection of the bifurcated equilibrium p a t h on the


plane A,0. Finally, from (5.3.14) and (5.3.22), we have 5 = 0 up to the
second order.
166

5.3.2 The cantilever


Let us consider the cantilever in Fig. 5.12.

2HA

x=0
s s
VA
Is
R R
l «vfr I u(x
^ R R
t
A

Fig. 5.12 - Cantilever.

The geometrical boundary conditions are

uA = 0 , vA = 0 , tpA = 0 (5.3.28)

As VB 7^ 0, the potential energy (5.3.5) is modified as

n ( w ; A) = / I - EI<p'2 + A(cos <p - 1) - S sin <p\ dx + SvB (5.3.29)


Jo L2 J
from which we immediately get S = 0 by imposing stationarity with respect
to Vs- Equation (5.3.29) therefore reduces to the form

fri
n ( w ; A) = / \- EI(p'2 - A(cos <p - 1)1 dx (5.3.30)
./o L2 J
which coincides with (5.3.6). By following the procedure applied in the previ-
ous Section, we come to the differential equation for calculation of the critical
load and of the buckling mode

EI(p"e + \c(pe =0 (5.3.31)


with the boundary conditions

<pc(o) = o , Ei<p'c{e)=o (5.3.32)

The solution is
n2EI
Ac = (5.3.33)
4£ 2
■K X
<Pe{x) = sin
si (5.3.34)
~2l
having normalised <pc{x) in such a way that <pc{£) = 1- As in the previous case
IIJ? ii^ = 0 , n'c" v c ii 2 = 0, as is easily verified, and furthermore f["c ii 2 = -1/2.
Therefore from (4.5.19) we have Ac = 0. The differential equation for the
calculation of u c is identical to (5.3.31) with the same boundary conditions.
167

We arrive at the solution

<pe{x) = 0 (5.3.35)

having normalised in such a way that <pe is orthogonal to <pc. Finally, we can
immediately verify t h a t IT" u* = 0 and II"" ii* = 3A c £/8 from which we obtain

Xc = - \e (5.3.36)

and

^ = l + -02 (5.3.37)
Ac 8
w i t h 0 = <p(l).

5.3.3 Approximate methods of solution: the Ritz


method
The critical load can be obtained in a very simple way, and with sufficient
accuracy, by using approximate methods of solution such as the Ritz method.
This consists of the representation of the buckling mode by means of a linear
combination with arbitrary constants a = {ai, a 2 , . . . , a n } of functions each of
which satisfies the geometrical boundary conditions, and of determining the
values of the constants which make the quadratic functional II" a 2 positive
semidefinite, t h a t is, which satisfy the variational problem

II" a 6 a = 0 (5.3.38)

Equation (5.3.38) represents an algebraic eigenvalue problem similar to


that for discrete systems, the solution to which gives Ac and the constants
a 1? a 2 , . . . , a n , defined to within one of them.
The expression II" u 2 for problems governed by the functional (5.3.5) under
the hypotheses (5.3.8) and for t;(0) = v(l) = 0 is

- I I " u 2 = f V - EI <p'2 - - A (p2 - S <p) dx (5.3.39)


2 Jo \ 2 2 /
having used the series expansion up to the second order, cos(p = 1 — <£>2/2. If
the function <p(x), as pointed out above, is chosen in such a way as to satisfy
the geometrical boundary conditions

v(l) = [ ip[x) dx = 0 (5.3.40)


168

deduced from (5.3.7) by putting sin<£> = <p, equation (5.3.39) reduces to

- II" u 2 = - f (EI <p'2 - A <p2) dx (5.3.41)

Equation (5.3.41) coincides with the expression of II" u 2 deduced from


the energy functional (5.3.30), which governs the class of problems of the
cantilever type.
In problems regarding beams, it is often more simple to give an approxi-
mate expression of the buckling mode in terms of v(x) instead of <p(x). It is
then necessary to substitute for (3.5.41)

- II" u 2 = - f(EI v"2 - A v'2) dx (5.3.42)

obtained by replacing tp with (p = arcsin(t/) = t/. This comes from the second
of relations (5.2.22) by means of a series expansion up to the first order.
Let us now solve the problem of the cantilever assuming
v{x) = ax2 (5.3.43)
where a is an arbitrary constant. Function (5.3.43) satisfies the geometrical
boundary conditions v(0) = i/(0) = 0. From the substitution in (5.3.42) we
obtain

-n"a2 = 2 [ EIa2dx-2X f a2 x2 dx
2 Jo Jo
e?
= 2EIa2 t-2\a2 — (5.3.44)
from which, by applying (5.3.38), we have

\al I El A] 6a = 0 (5.3.45)

which gives
3EI
(5.3.46)
(2

Comparing (5.3.46) with the exact solution (5.3.33), we find an error of about
20%. We repeat the analysis assuming
v(x) = ax x2 + a2 x3 (5.3.47)
With the substitution in (5.3.42) and by calculating the integrals, we obtain

i n " a2 - (2EI^^ex)a2+UEIt2--t4\\a1a2

+ UEIlz - — t \ \ a\ (5.3.48)
169

from which, by applying (5.3.38)

UEIL-- £3\\ai + UEII2- - e*\\a2 = 0 (5.3.49)

UEU2-- e4x)ai + (nEie3-- e5x\a2 = o (5.3.50)


By equating the determinant of the coefficients to zero we get

Ac = 2.49 ^ / (5.3.51)

with an error of about 1% relative to the exact solution. By substituting


(5.3.51) in (5.3.49) or in (5.3.50), we obtain

a2 = - 0 . 2 9 6 °j (5.3.52)

from which

v(x) = ax x2 (l - 0.296 - \ (5.3.53)

The constant a\ can be eliminated by choosing a convenient normalisation.


Note t h a t the values of Ac furnished by (5.3.46) and (5.3.51) are higher than
the exact value. This is a general feature of the method, and arises because
the introduction of a buckling mode which is different from the exact one is
equivalent, from a mechanical point of view, to introducing constraints in the
b e a m with consequent increase in the critical load.

5.3.4 Beams under various edge conditions


We shall consider a series of classical problems for which we indicate the
expression of total potential energy, the geometrical boundary conditions and
the value of the critical load. The critical mode is qualitatively shown by a
dashed line in the figures. The interested reader may carry out calculations
for verification [3,4,6,8,9,12,13,16,22].
Let us consider the beam with clamped edges shown in Fig. 5.13. The

*V//A

x=0 x=(

Fig. 5.13 - Beam with clamped edges.


170

geometrical boundary conditions are

uA = 0 , vA = 0 , <pA = 0 , vB = 0 , *?B = 0 (5.3.54)

and the total potential energy of the beam is given by (5.3.5). For the critical
load we have
7T2£J
Ae = (5.3.55)
(o.5 iy
Let us now examine a beam with a clamped edge A and simply supported
at B (Fig. 5.14).

B
S3
x=0

Fig. 5.14 - Beam with a clamped edge and a simply supported edge.

We have the geometrical boundary conditions

uA = 0 , vA = 0 , <pA = 0 , vB = 0 (5.3.56)

The total potential energy is again given by (5.3.5). The critical load is

c *'EI (5.3.57)
(0.6991) 2

Let us next consider a beam with a fixed edge A and a vertically sliding
edge B (Fig. 5.15). The geometrical boundary conditions are

UA = 0 , vA = 0 , (pA = 0 , (pB = 0 (5.3.58)

x= 0 1
■x=I
Fig. 5.15 - Beam with clamped and sliding edges.
171

The total potential energy is equal to that of the cantilever and is given by
(5.3.30). For the critical load we obtain

n2EI
Xc = —p- (5.3.59)

The critical load values for the different types of beam examined so far
can be calculated from the formula

n2EI
A, = - ^ (5.3.60)

where at represents the distance between two adjacent nodal points of the
buckling mode. This permits us to calculate in a simple way the critical load
of a continuous beam with n equal spans of length £ (Fig. 5.16).
If we assume, as is easy to demonstrate, t h a t the buckling mode is given by
<p(x) = a cos7rx/£, then the critical load is given by (5.3.60) with a — 1.
We now wish to study the problem of the Euler beam, elastically con-
strained at A and B, as shown in Fig. 5.17. Depending on the values of
the constants of the springs kA and /;#, we obtain some of the cases exam-
ined previously. For example, if kA — kB = 0 we have the Euler beam, if
kA = kB = oo we have the beam with clamped edges, while if kA — oo and
kB = 0 we obtain the beam with a clamped edge and a simply supported
edge. The geometrical boundary conditions are

uA = 0 , vA = 0 , vB = 0 (5.3.61)

Fig. 5.16 - Continuous beam with n equal spans.

?2
x=0

Fig. 5.17 - Beam with elastic supports.


172

The total potential energy is obtained from (5.3.5), with the addition of
the elastic energy of the springs. We have

TT(w ; A) = - I \E I <p'2 + A(cos <p - 1) - S sin <p] dx + - kA <p\


2 io L J
2
+ -kB<p2B (5.3.62)

from which, by means of (4.5.13), we get the differential equation with the
boundary conditions for calculation of the critical load and the buckling mode.
By introducing the solution into the boundary conditions we obtain a homo-
geneous system of linear algebraic equations from which, by imposing that
the determinant of the coefficients vanish, we get the characteristic equation

[l - on - a2 - ttift2(H)2] kls'm kt

+ [2 + ax{kl)2 + a2{kl)2] cos kl - 2 = 0 (5.3.63)

where we have put


El EI ,« Ac . x
a = h= 5 3 64
"' = £ « • ' ^l- El <'- >
For « ! = a2 = a (5.3.63) reduces to

kl
tan — = -akl (5.3.65)

If the beam is simply supported at the ends without elastic constraints,


then kA — kB — 0 and a —► 00. From (5.3.65) we get kl = 7r, to which
(from the third of equations (5.3.64)) Ac = TT2EI/12 corresponds. If the beam
is clamped at the ends a —* 0 and (5.3.65) gives the non-trivial solution
kl = 27r, to which Ac = 4n2EI/l2, already given by (5.3.55), corresponds. If
the beam is clamped at x = 0 and simply supported at x = £, then we have
a i = 0 and a2 —> 00. Dividing (5.3.63) by a2 and letting a2 —► 00, we have

k I ( - sin k I + k I cos k I) = 0 (5.3.66)

from which we obtain kl = 0 or tanfcl = kl. Discarding the trivial solu-


tion, the second equation gives kl = 1.43, to which the previously found
Ac = 2.04 El/I2 corresponds. Solutions to equations (5.3.63) and (5.3.65)
cannot be given in closed form; an adequate approximate solution to (5.3.63)
based on numerical calculations is provided [2] by

( f c 0 W (« 1 + 0.4)(a,+0.4)
v ; v
( a ! + 0.2)(a 2 + 0.2) '
173

x=0
m^/////////. x=/
k
f

Fig. 5.18 - Beam on elastic foundation.

to which there corresponds


_(a1 + 0.4)(a 2 + 0.4) ^EI
K
~ (a1 + 0.2)(a2 + 0.2) ~1T (5-3-68)
Let us conclude this Section by analysing a beam on an elastic foundation.
We consider a simply supported beam resting on an elastic foundation of the
Winkler type (Fig. 5.18). The reaction provided by the foundation is equal to
kj v(x), where kj is the foundation rigidity which we assume to be constant.
A Winkler elastic foundation model can therefore be thought of as an
infinite series of springs without shear coupling. In the most general case the
Winkler model is inadequate and the foundation must be treated as an elastic
continuum. The geometrical boundary conditions are

uA = 0 , vA = 0 , vB = 0 (5.3.69)
For calculation of the critical load, we can start directly from the expres-
sion of the total potential energy (5.3.39) written up to the second-order
terms, modifying it with the addition of the term relative to the energy of the
reactions of the foundation. Since the reactions are expressed as a function of
v(x), it is convenient to express also the other terms of (5.3.39) as functions
of v(x), we therefore have

- n " u 2 = / Y - EI v"2 + - kf v2 - - A t/ 2 - S v') dx (5.3.70)


2 Jo \ 2 2 2 /
to which corresponds the Eulerian equation of the problem
EI v"" + A v" + kf v = 0 (5.3.71)
with boundary conditions
v(0)=0, v(£)=0, Elv"(0)=0, Elv"{l)=0 (5.3.72)
It is interesting to note that the solution to problems (5.3.71) and (5.3.72)
is independent of the parameter S which is determined by the boundary
conditions. This is obvious if we realise that the field variable v(x) in the
place of <p(x) satisfies the auxiliary condition (5.3.40) a priori, and therefore
the last term of the energy functional (5.3.70) does not appear.
174

We can verify that a solution of the type


TITV x
v(x) = a sin (n = 1 , 2 , 3 , . . . ) (5.3.73)
I
satisfies boundary conditions (5.3.72). Furthermore, by substituting (5.3.73)
in (5.3.71), we can see that the differential equation is satisfied if A is given
by

If kf — 0, the value of n which minimizes Ac is unity, and therefore equation


(5.3.13) is obtained. If kj ^ 0, the value of n which satisfies this condition
depends on £, El, and kf, and in general is different from unity. For a given
beam, an increase of kj causes a decrease in the wavelength of the buckling
mode. Equation (5.3.74) can be put into the form

2M P (5.3.75)
XE

with XE = TT2EI j!} and /? = i4kf/7r4EI represented in Fig. 5.19 for a number
of values of n [17]. For 0 < /? < 4 the critical load belongs to the straight
line A c l /A£, for 4 < /3 < 36 to the straight line A c2 /A£, etc. The value of j3
corresponding to which the buckling mode passes from n t o n + 1 half waves,
is given by
2
/? == [n(n + l)] (5.3.76)

¥\
AE Ac 2 _4 + P
4
AE
18 -

16
14 --
/
12 -
i --"""" ^ —
10 -
A£_2v/g"
8 - t ^ ^ ^ AE

6 -

4 "'/ /
2 -

4 8 12 16 20 24 28 32 36 40 44 48 52

Fig. 5.19 - Critical load for beam on elastic foundation.


175

If we consider n as a continuous variable, then the minimum of Ac is ob-


tained for n 2 = y/fi and (5.3.75) gives

^ = 2 yffi (5.3.77)

The expression (5.3.77) is represented in Fig. 5.19. We can see t h a t the


error made by using the approximate relation (5.3.77) decreases with increase
of /?. Such a procedure is widely used in analysing the stability of plates and
shells. Another type of procedure, finally, is to minimize Ac by a trial and
error method.
We shall not enter into further investigation of the vast class of problems,
as the aim of this book is essentially limited to indications on the methodology
of analysis.
The specialist in a hurry interested in the solution to particular technical
problems can refer to numerous classical texts [3,4,6,8,9,22] or to manuals [12],
if his interest is limited to knowledge of the critical load. In the following
Section we analyse, by means of the Ritz method, some problems of great
practical interest, for which the exact solutions are extremely complex.

5.4 PARTICULAR PROBLEMS

5.4.1 B e a m subject to distributed load


Let us consider the cantilever subjected to its own weight A, which is constant
if the cantilever is of constant cross-section and made of one material only
(Fig. 5.20) [4]. The total potential energy II"u 2 , which we shall use to calcu-
late the critical load and buckling mode, is obtained from (5.3.42), modifying
the term relating to the potential energy of the loads. We have

- II" u 2 = - jl[EI v"2 - \{t - x) v'2] dx (5.4.1)

x=/

A=const.

x=0

Fig. 5.20 - Cantilever subjected to its own weight.


176

From the condition of stationarity (4.5.5) we obtain

II" u 6 u - f€ [EI v" 6 v" - \{i -x)v'6 v'] dx = 0 (5.4.2)

from which, integrating by parts, we have the differential equation

EI v"" + A(£ - x) v" - A v' = 0 (5.4.3)

with boundary conditions

v (0) = 0 , v'{0) = 0 , EIv"{l) = 0 , EIv'"{t) = 0 (5.4.4)

Here (5.4.3) is a differential equation with variable coefficients, which can


be integrated giving

EIvm + \(l-x)v' = 0 (5.4.5)

This can be put in the form


2
dluu 1 du
du ( 1 \ ,

where

dv 2 / A ,„ .3
u = Tz, * = - y — «-x)* (5-4.7)
Expression (5.4.6) is a Bessel equation, which can be solved by using Bessel
functions. The arbitrary constants are then determined from boundary con-
ditions (5.4.4) expressed in terms of the new variable u [l],
The problem was studied for the first time by Euler and solved about a
hundred years later by Greenhill in 1881. The solution to the problem [4]
gives

P£)c = 7 - 8 3 7 i r (5-4-8)
u(z) = a z-'l* ( l - j | z 2 + ^ z* + ■ ■ •) (5.4.9)

We now wish to give an approximate solution to the problem, using the


Ritz method. Let us choose (5.3.34) as an approximate solution, that is, the
buckling mode of the cantilever subjected to end load. If we use the total
potential energy in the form (5.4.1), we must express (5.3.34) in terms of
v(x). As <p = (p = t/, we have to the first order

t,(x) = a ( l - cos | | ) (5.4.10)


177

By substituting in (5.4.1) we obtain

- n"a2 = - / El — -4 cos 2 — dx
2 2 Jo 16£ 21

/ XU-x) —— sin 2 — dx
2 Jo v ;
U2 21
2
- — -£■/—---ATT + - A (5.4.11
v ;
8 \8 P 4 2 /
from which, using the variational equation (5.3.38), we get

( A£ ) C = 7 89
- 7T (5-4-12)
which exceeds the exact solution (5.4.8) by about 0.7%. With equal simplicity
we can solve the problem of the cantilever subjected to its own weight and to
a concentrated load A applied to the end [4].
The total potential energy is then

- II" u 2 = - fl\EIL V"2 - \(t - x) v'2 - A vl2J} dx (5.4.13)


2 2 Jo
As an approximate solution we again use (5.4.10). By substituting in (5.4.13)
we obtain

- II" a 2 = — [- El —-- \— - - A TT2 + - A K(5.4.14)


2 8 \8 £3 2 £ 4 2 ) )

Each combination of values of A and A which makes the right side of (5.4.14)
vanish, thus making II"a2 positive semidefinite, leads to a critical condition
of equilibrium. If A = 0 then Ac = n2 EI/4£2, as previously obtained. The
distributed load A obviously reduces the critical load Ac = TT2EI/412. By
putting

K = m ^ (5.4.15)

the factor m equals 7r2/4 for A = 0 and diminishes with increase of A. If we


introduce the relation

(5 4 16)
"-idlWe --
the values of the coefficient m of equation (5.4.15) corresponding to different
values of n are those collected in the following table [4].
178

TABLE I - Values of m in equation (5.4.15)

n 0 0.25 0.50 0.75 1.0 2.0 3.00 3.18 4.0 5.0 10.0

m 7T2/4 2.28 2.08 1.91 1.72 0.96 0.15 0 -0.69 -1.56 -6.95

In calculating the effect of the uniform load A on the value of Ac, we obtain
a sufficiently accurate approximation by assuming t h a t the effect of A£ is
equivalent to a load 0.3 XI applied at the end of the cantilever. The critical
load is therefore

AC = ^ | ^ - 0 . 3 A £ (5.4.17)

If the distributed load exceeds the value given by equation (5.4.8), then Ac
becomes negative. This implies that the critical condition of equilibrium
occurs in correspondence with traction loads applied at the ends.
The problem of the Euler beam can be treated in the same way, if the
distributed load A is added to the concentrated load A (Fig. 5.21) [4]. Let us
introduce as an approximate solution the buckling mode (5.3.14) of the rod
subjected only to concentrated load. In terms of v(x) we have

v(x) = a sm —— (5.4.18)

By substituting in the functional of the total potential energy given by (5.4.14),


we obtain
V f f J _ XI _ l £ 2 >
n " a 2 = a2 (5.4.19)
~2i 2 4

c^/

t A=const.

x=0

Fig. 5.21 - Simply supported beam with concentrated load at the end and
distributed axial load.
179

from which we can see t h a t , as in the case of the cantilever, the distributed
load A reduces the critical load Ac = n2 EI/l,2. By taking (5.4.15) and intro-
ducing the relationship

\l
n = (5.4.20)
n EI/l2
2

we can calculate the values of the coefficient m corresponding to different


values of n. The results are collected in the following table [4].

TABLE II - Values of m in equation (5.4.15)

n 0 0.25 0.50 0.75 1.0 2.0 3.0

m 7T2 8.63 7.36 6.08 4.77 -.657 -4.94

From (5.4.19) we find in particular t h a t for A = 0

n2EI
(A£) c = 2 (5.4.21)
I2

This implies t h a t in evaluating the effect of the uniform load A on the value
of Ac within the framework of the present approximate solution, we can sub-
stitute for the distributed load A£ a concentrated load equal to X£/2 applied
at the end of the beam. Therefore instead of making use of the table, we can
use the approximate formula

n2EI XI
A. = (5.4.22)
I2 2
If XI is equal to the value given by (5.4.21), then Ac = 0 and for higher
values Ac < 0. Therefore, the critical condition of equilibrium is obtained in
correspondence with traction loads applied at the end of the beam.

5.4.2 C a n t i l e v e r w i t h variable c r o s s - s e c t i o n
With the aim of decreasing the weight of members under compression, rods
with variable cross-sections are often used. The differential equations for the
determination of the critical load were derived by Euler for beams of various
forms such as the pyramid frustum and cone frustum. In much more recent
times the problem of the rod in which the moment of inertia of the section
varies according to a power of the distance along the axis of the beam has been
analysed. Let us consider the cantilever in Fig. 5.22 in which the moment of
180

I1
N\\\\M
irf-n
IUUJ
x=0

a)

Fig. 5.22 - Cantilevers with variable cross-sections.

inertia varies according to the law [4]

7 7
- '(1-rh)' (5.4.23)

with Ii the moment of inertia at x = 0. For n = 1 we obtain the case of


a column whose cross-section is rectangular with t constant (Fig. 5.22(a)).
For n — 2, (5.4.23) represents with sufficiently good approximation the case
of a cantilever formed by four angles connected by braces. In this case the
area of the cross-section remains constant whilst the moment of inertia is
approximately proportional to the square of the distance of the centroid of
a single angle from the axes of symmetry of the section (Fig. 5.22(b)). For
n = 4, we obtain the cone frustum or pyramid frustum. Using (5.4.23), we
write the expression of the total potential energy (5.3.42)

(5.4.24)
2 2 Jo [ \ l + dj J
from which the Eulerian of the problem is

[£/'(I-^)"»"]"+A""=° (5.4.25)

with boundary conditions

v(0)=0, t/(0)=0, v"{e)=0,


(5.4.26)
0
{["■('-^M'-Mr
181

Equation (5.4.25) can be solved by means of the Bessel function for any
value of n. For n = 2 it gives [4] the exact value of the critical load in the
form

A, =
mEh (5.4.27)
e2
The factor m depends on the ratio djl and its values are given in the table
[4]. Note that for hjh —► 1 with 72 the moment of inertia at x = t, m tends
tO 7T 2 /4.

TABLE III - Values for the factor m in equation (5.4.27) for n = 2

h/h 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

m 0.250 1.350 1.593 1.763 1.904 2.023 2.128 2.223 2.311 2.392 * 2 / 4

Let us now look for an approximate solution to the problem by choosing


(5.4.10) as the solution. Substituting in (5.4.24) and assuming n = 2, we have

-n"a 2 = Eh
2 4£ 2 2+£UT<* J U JTV

\t
(5.4.28)
l + d \4 W
By choosing h/h = 0-5 from (5.4.23) we get d = 2A I. By substituting in
(5.4.28) and requiring the coefficient of a 2 to vanish we obtain

A, = 2.064
Eh (5.4.29)

which exceeds the exact value by 2%.

5.5 SIMPLY SUPPORTED BEAM AXIALLY


LOADED AT MID-SPAN
In this Section we make a complete analysis of the simply supported beam
axially loaded at mid-span (Fig. 5.23), evaluating the critical load and the
buckling mode and studying the post-critical behaviour. We shall show that,
contrary to what happens for the Euler beam and the cantilever examined in
Section 5.3, an analysis performed by means of the classical potential energy
leads to wrong results. This makes it necessary to use the modified potential
energy (5.2.41), which leads to the correct solution.
182

1 x
' * -• * 2 1
/ /
— 1 —
*

Fig. 5.23 - Simply supported beam axially loaded at mid-span.

Let

^(w;A) = fl lEItp'ldx^ I" \EI<p'l<Lx2


JO l JO Z

+ A / (cos <p2 — 1) dx2 (5.5.1)


Jo
be the total potential energy of the beam, obtained from (5.2.24) expressing
the displacement uB of the point at which the force is applied in terms of <p2.
The quantities with index 1 and 2 in (5.5.1) refer respectively to the interval
0 < xi < t, OT 0 < x2 < £. The flexural rigidity El is assumed constant in the
beam. We want to analyse the problem of the critical load and post-critical
behaviour, using the energy functional (5.5.1). From the variational problem
(4.5.13) we obtain
_ n2EI
(5.5.2)
7T X
(p2c - cos
where use has been made of <£>i(A) = 0, <p2(\) = 0 and the normalisation
(p2c(0) = 1. The shape of the buckling mode is shown in Fig. 5.24 and is
evidently incorrect, as the boundary condition at x\ — 0 is violated. The
post-critical equilibrium curve is
_ TT2EI

where 0 — <p2(0).
H') (5.5.3)

mm
Fig. 5.24 - Buckling mode which does not respect the boundary conditions.
183

M I ^ Vl(Xl) t v„
uB u 2 (x 2 ) x2
—•- I *■

^2(x2)
v
z (x 2 ) I y2

Fig. 5.25 - Components of the displacement.

Let us now solve the problem by making use of the modified functional
(5.2.41). After dividing the beam as indicated in Fig. 5.25, we write the
expression of the total potential energy for the problem under examination
[26]

II(w ; A) = / \ - EI<p'l - Ri(cos <pi - 1) - Si sin <pi dx\


Jo L 2 J

-f / - EI(p'2 — i2 2 (cos <p2 — 1) — ^ sin <p2\ dx2


Jo L 2 J
+ i2i (ujg - uA) -{- S1vB - R2 uB - S2vB - X uB (5.5.4)

having used the boundary conditions

VA = 0 , uc = 0 , vc = 0 (5.5.5)
From the variation with respect to UA , uB and t># we get R\ = 0, R2 = —A,
Sx = S2 = S and therefore the functional (5.5.4) is modified in the form

IT(w;A) = / - EI<p'*dx1+ [ l-EI<p'2dx2


Jo 2 Jo 2

+ A / (cos <p2 — 1) dx 2

— S I / sin <pi dxi + / sin <p2 dx2 I (5.5.6)

The remaining geometrical boundary conditions, other t h a n (5.5.5), which


must be satisfied are

U l (£) = - u 2 (£) , Vl{i) = - v2(i) , Vl{l) = <p2{l) (5.5.7)

T h e fundamental path is given by

V?i(A)=0, *>2(A)=0, 5(A) = 0 (5.5.8)


184

whilst the differential equations and boundary conditions which govern the
eigenvalue problem (4.5.13) are

EI<p"u + Se = 0
(5.5.9)
EI<p"ie + \e<p2e + Se = 0

EI<p'lc{0) = 0 , EI<p'2c{0) = 0 , tple{t) = tpu(t)


(5.5.10)
EI<p\e{t) + EI<p'2c{£) =0, f <p lcddx
<Pu Xl+x + f / <p2c dx2 = 0
/o
Jo Jo
Jo

We obtain the solution


/ x s i n kl /1,«>9 1 ,-> „? \
<Pu{x) = T-» r-7^ ~ k2x2-2-- k2t2)
^lcK J kl + smkl \2 6 /
, x k£ cos k x + sink £
*■<*> = M + rinM '"•"'
• sinki
Sc = kt + sinkt
with the characteristic equation

k212
klcotkl= 3 (5.5.12)

having written k2 = \C/EI and used the normalisation <£>2c(0) = 1. The


lowest root of the characteristic equation gives
VT
Ac = 4.6665 — (5.5.13)

As n'c"ii^ - 0 , n'c"vciic2 = 0 , fi;'iic2 + 0 , Ac = 0 is obtained from (4.5.19).


Furthermore, as n"'u 2 <$u = 0 , the variational equation (4.5.14) in ii c is iden-
tical to (4.5.13) and therefore the relevant differential equations in u c and
boundary conditions for this problem are obtained from (5.5.9) and (5.5.10),
substituting for the quantities with one dot, quantities with two dots. If we
choose u c orthogonal to ii c , we have

Vu = 0 , <p2e = 0 , Sc = 0 (5.5.14)

By using (4.5.21), which in this specific case is reduced to Ac =


2
- I I ™ u * / 3 n' c 'ii c , we obtain

Ac = 0.12982 Ac (5.5.15)
185

from which

F1 T
A = 4.6665 — ( l + 0.06491 0 2 ) (5.5.16)

with 0 = <P2{0). Contrary to what happens for the Euler beam, it is obvious
t h a t the results obtained here by making use of the potential energy in the
classical form of (5.5.1), or modified form (5.5.6) in which we take into account
the non-linear equation of constraint, are completely different in t h a t t h e
Lagrange multiplier S is different from zero.

5.6 STABILITY OF BEAMS


Discussion of the stability of the beam subjected t o a load A applied a t the
end, on the basis of the functional M T P E (5.3.5) or T P E (5.3.6), seems diffi-
cult. However, if use is made of a beam model derived from a 3-dimensional
continuum axially deformable and shear undeformable in which the kinematic
relations are provided by (7.3.27), we arrive at the following expression of en-
ergy

P(W) =
I liSn^ + E^fjdV
dx (5.6.1)

after using the relation Sn — —A/A, with A the beam cross-section. In


correspondence with the critical load, the second variation of energy

- P " u 2 = - jl (-Au/3 +EAu'2 + EIw"2)dx (5.6.2)

is positive semidefinite and therefore we must have u(x) = 0. It follows t h a t


at bifurcation

Pc(w) = i fl EAw'dx (5.6.3)


o JO

The absence of third-order terms in (5.6.3) implies t h a t the post-critical


behaviour is symmetrical, since from (4.5.19) Ac = 0. Furthermore, as
(5.6.3) is positive definite, we deduce t h a t the equilibrium corresponding t o
the bifurcation is stable and the concavity of the bifurcated p a t h is upwards.
186

5.7 SYMMETRIC, SIMPLY SUPPORTED, TWO


BAR FRAME

Let us consider the system in Fig. 5.26 and let I and El be the length and
flexural rigidity, respectively, of each rod.

Fig. 5.26 - Symmetric two bar frame.

Figure 5.27 shows in detail the local and global systems of coordinates
adopted [24].

Fig. 5.27 - Global and local systems of coordinates.


187

The M T P E of the system on the basis of (5.2.41) is

l
n ( w ; A) = \ \ Eltp'l dxx + \ I' EI<p'l dx2
L Jo 2 Jo

+ Ri\u cos 0 + v sin 0—1 (cos <pi — 1) dx\


[ Jo

+ S\ — u sin 0 + v cos 0 — sin (p\ dx\


| Jo

+ #2 u cos(7r — 0) + v sin(7r — 0) — I (cos <p2 — 1) dx


Jo

+ S2 — u sin(7r — 0) + v cos(7r — 0) — sin <p2 dx


Jo

+ XFv (5.7.1)

where we have used explicitly the constraint conditions uA — vA = uc = Vc —


0. We can immediately verify that the fundamental p a t h takes the form

Rx = R2 = - \ F/2 sin (5.7.2)

all the other quantities vanishing.


The differential and algebraic equations which determine the buckling
mode are obtained through (5.2.42):

2 sin 0

2 sin 0

uc cos 0 + vc sin 0 = 0
(5.7.3)
uc cos(7r — 0) — vc sin(7r — 0) — 0

— uc sin 0 + vc cos 0 = / <pic dx


./o

— ti c sin(7r — 0) + t)c cos(7r — 0) = / <pdx


2c

With these equations are associated the boundary conditions obtained from
188

(5.2.47)

Ru cos 0 6u + Ric sin 0 6v — Su sin 0 Su + Su cos 0 Sv

+ EI<p'lc{?) 6 <pi{e) - EI<p\e{0) 6<P!{0) + R2c COS(TT - 0) Su

+ R2c sin(7r — 0) S v — S2c sm(ir — 0) Su + S2c COS(TT — 0) Sv

+ EI<p'2e(e)6<p2(l) - £ J # e ( 0 ) «p,(0) = 0 (5.7.4)

and, from the explicit boundary condition £>i(^) = V?2(^)>

<Pu{l) = <P*M) (5.7.5)

Taking this into account, equation (5.7.4) gives the following boundary
conditions

(RU - R2c) cos 0 - (Su + S2c) sin 0 = 0

(RU + R2c) sin 0 + (Sle - S2c) cos 0 = 0


(5.7.6)

EI<p'u(0)=EI<p'2e(0)=0

By putting A:2 = Ac F/2 El sin 0 and solving the system of differential and
algebraic equations (5.7.3) with boundary conditions (5.7.5) and (5.7.6), we
obtain the characteristic equation

{k I - tan k t) A; £ sin 0 = 0 (5.7.7)

whose lowest positive root is

kl = n (5.7.8)

The critical load

XCF = 2TT2 El sin 0/t? (5.7.9)

and the asymmetrical buckling mode

£>lc(z) = (P2c{x) = COS7Tx/£

Rlc = 5 l c = i2 2c = 5 2 c = 0 (5.7.10)
u c = vc = 0

correspond to this root. The buckling mode (5.7.10) has been obtained by
using the normalisation <Pic{fy = — 1. It is easy to see t h a t the expressions
(5.7.10) make n'c"u* vanish, and thus Ac = 0 follows from (4.5.19).
189

The second-order perturbation equations obtained from (5.2.44) are

2 sin 0

EI<pl + -^—<p2c + S2c = 0


2 sin 0

r 7T X
uc cos 0 + vc sin 0 = — / cos 2 —— dx
Jo t
(5.7.11)
Uc COs(7T — 0) + Vc sin(7T — 0) — — \ COS2 — — dx

— uc sin 0 + vc cos 0 = <Pu dx


Jo

— uc sin(7r — 0) + ijc cos(7r — 0) = / <p2<

and the boundary conditions obtained from (5.2.45) read

Ru cos 0 6u + .Rlc sin 0 6v — Su sin 5 6u + 5i c cos 0 6v

+ EIip'lc(l) 6 (pxil) - EI(p'le{0) 6ipxiti) + R2c COS(TT - 0) 6u

+ R2c sin(7r — 0) 6 v — S2c sin(7r — 0) 8 u + S2c cos(7r — 0) 6v

+ EI<p'2c(t)6<p2(t) -EI<p'2c(0)6<p2(0) = 0 (5.7.12)

To these is added the explicit boundary condition

Vu{t) = $M (5-7.13)

From (5.7.12) we get

[ku - R2c) cos 6 - (Sle + S2c) sin 0 = 0

(Ru + R2c) sin 0 + [Slc - S2c) cos 0 = 0


(5.7.14)
EI<p'u(l) + EI<p'u(t)=0

EI <p'u{0) = EI <p'2c{0) = 0
190

The solution is symmetrical and is written as

1 / 7T X \
<Plc{x) = ~ <P2c{x) = ~ ~ (COS — + 1J COt 0

Ru = R2C = -TT2EI cot 2 0/2 i2

Slc = -S2C = TT2EI cot 0/212 (5.7.15)

I
vc = - 2 sin 0
uc = 0

The secondary mode has been normalised by putting <Pu(t) = 0. The


calculation of Ac is left as an exercise for the reader. The result is

(5 7 i6)
~K=\K G ~ cot2e ) --
Bearing in mind the adopted normalisations <Pic{£) = — 1 and <f>ic{£) = 0,
from the asymptotic expression of <pi(t)

<p1(t)=<Pu(t)e + ±<ple(l)e2 (5.7.17)

we get

<p1(l)=-e (5.7.18)

The dependence of the load on the parameter, corrected up to the second


order, is

A F = ( A C + A C 6 + U C £ 2
) F (5.7.19)

that is, by virtue of (5.7.16) and (5.7.18),

2n2EI
[l + JQ-cot 2 0)^ 2 ]
^ .
x
AF = sin (5.7.20)
having put (p = <Pi(£) = <P2^)-
Note that for 0 = 0O = 26.565° we have Ac = 0, and therefore the analysis
of stability should be carried out by examining higher order terms. The value
0o = 26.565° represents a turning point, since for 0 > 0Q we have Ac > 0 and
the behaviour of the frame is stable, whilst for 0 < 0O we have Ac < 0 and the
behaviour of the frame is unstable.
191

5.8 HINGED SYMMETRICAL PORTAL FRAME


Let us consider the structural system in Fig. 5.28 and let Eli and EI2 be
the flexural rigidities of the two columns and the beam respectively; h is the
height of the columns and I the length of the beam [24].

AF AF

© ©
D

Fig. 5.28 - Hinged symmetrical portal frame.

Figure 5.29 shows in detail the beam elements and nodes which make up
the system, also illustrating the local systems of coordinates for the parame-
ters of each element and the global system of coordinates to which the nodal
displacements are referred.

uA
R,
1 v 2 (x 2 ) S2

R2 9z(*2)
-^
Si 1 2 (x R2 S3
"IU^ U
3 ( X3)
VIIX^^H

V 3 ( X 3) \

X1 I Sa
:
' — ►

Ri 1 ? R3

vc = 0 V n- 0

Ur = 0 uD-0
0 X

Fig. 5.29 - Global and local coordinate systems.


192

The M T P E of the system is obtained from (5.2.41), noting t h a t in this


case 0i = 0 3 = 7r/2, 02 = 0, and putting uc = Vc = V>D = vp = 0.
We get

n(w;A) - ±J (Eh<p'* +EI1<p'*)dx1 + - f Eh<p'l dx2

VA — / (cos (pi — 1) dxi


0

— VA— sin <pi dxi


./o

^B — UA~ / ( c o s £>2 ~~ 1) ^ x 2
0

+ S2 VB — VA — / sin tp2 dx2


Jo

+ i23 VB - / (cos <p3 - 1) dxi


Jo

— uB — sin <£>3 dxx


./0
+ AF(t;A+t;B) (5.8.1)
Also in this case the fundamental path is very simple, since the only pa-
rameters different from zero are
R1 = R3 = -\F (5.8.2)
The equations of the buckling mode are immediately obtained from (5.2.42)
EI1<p'le + \eF<ple + Sle = 0

EI2<p'le + S2e = 0

EI1<p"3e + \eFip3c + S3e = 0

VAc = 0

UBC -UAC = 0 (5.8.3)


vBc =0
rh
-UA = / (Pledx
Jo

VBc -VAC= / <£>2c dx


Jo

-U>Bc= / <P3c dx
Jo
193

With these are associated the boundary conditions obtained from (5.2.47)

Ru * vA - Slc 6 uA + E h <p'u{h) 6 <p1{h) - E h <p'le{0) 6 <Pi{0)


+ R*c S{uB - uA) + 5 2 c 6{vB ~vA) + E h <p'2e{l) S <p2{l)
-EI2 <p'2e(0) 6<p2{0) + R3c6vB-S3c6uB + Eh <p'Sc{h) 6<p3{h)
-EI1<p'ze{0)6<p3{0)=0 (5.8.4)

and the boundary conditions on <p(x) which immediately derive from the
geometry of the problem

<Pu{h) = <P2c{0)
(5.8.5)
£>3c(/0 =<P2c{l)
By requiring 6(pi(h) = S<p2{0) and S(ps(h) = 6ip2(l), equation (5.8.4) provides

-EJ1<p'le{0)=0
-EI1<p'Zc{0)=0

EI1<p'le{h)-EI2<p'ie{0) = 0
EI1<p'ae{h) + EIi<p'ie(l)=0
(5.8.6)
- Sic — R-2c — 0
- S3c + R2c = 0
- S2c + Rlc = 0
S2c + Rzc = 0
Solving system (5.8.3) with the boundary conditions (5.8.5) and (5.8.6),
we obtain the buckling mode

<Pu{x) = <P3c(z) = - . , , coskx


sin kh
, x k2 h2 ( x2 x 1

Ru = ~Rsc = S2c = 2 Ac F h/l (5.8.7)

Sic — Ssc — R2c — 0


uAc = UBC = h
VAc =VBC = 0
194

where the normalisation UAC = h is chosen and we have put k2 = Xc F/EIi,


a = Eli h/EIi I. The critical load corresponding to the buckling mode (5.8.7)
is obtained from the transcendental equation

kh t a n B = 6 a (5.8.8)

Figure 5.30 shows the dependence on a of the two non-dimensional pa-


rameters

Ai1) = k2 h2 (5.8.9)

kh
Ai2> = 6 J^LjtU^lAW (5.8.10)
tan kh h E h a
which represent the critical value of the load parameter A when we write
respectively
Eh
FW = (5.8.11)
~~h?~

F = F& = EI2 (5.8.12)


hi

By writing out the expression of n"'u^ we find that it vanishes when eval-
uated in correspondence with the buckling mode (5.8.7), and so we conclude
that Ac = 0. The equations which govern the secondary mode are obtained

A c =6
6.0

5.0 \- \

4.0

3.0 k
^£1r /4
2

2.0

r 9*
1.0

10-3 10 10 10*

Fig. 5.30 - Dependence on a of the load parameter Ac


195

from (5.2.44) as

EIl<p"u + \eF<pu + Sie = -4\eF - - — - r cos kx


I sin kh
EI2<p2'c + S2c = 0

EI1<p'ic + XcF<pZc + S3c = 4XcF - — — cos kx


I sin kh
h k2h2 ( sm2kh
VAc 2 1+
2 sin )fc/ii 2kh
2 2
£ k h

h k2 h2 ( sin 2
"Be = - - . 22 , , 1 +
2 sin A: /i V 2 A:h )

./o

VBc -VAC= / <p2c dx


Jo
rh
-UBc
= / ^Scdx
./o

with which are associated the natural boundary conditions derived from
(5.2.45)

-EI1tp'le{0)=0
-EI1tp'9e{0)=0

EIlip'le{h)-EI2<p'2e{0)=0
ElMh)+EI2tp'2e{l)=0
(5.8.14)
- 5 l c — R2c = 0

- SZc + R2c = 0

- S2c + Rlc = 0

S2c + R3c = 0

Finally, the explicit boundary conditions on <p are


<Pu{h) = ^2c(0)
(5.8.15)
<Pzc{h) =<p2c{£)
196

The solution to the system corresponds to a symmetrical deformation of


the frame and is
1 — / COS k X
fiuix) — — &3c{x) — - 0 tan k h 2 cot k h H—;——
3 v sin k h
h kh
+ 2 I tanA:/i 1 + k h(cot k h + tan k h)
cos kx ( kh \ _ kx sin kx
cos kh \ tan kh) cos kh

<p2c(x) = 0 (l - 2 j}

R\c — R$c — Sic — o (5.8.16)

S\c — — S:3c — Roc — —2\CF

+ - — [1+k h(cot k h + tan k h)}


I tan k h
k2hH
u>Ac = - uBc =10 tan 2 k h

h k 2h 2 sin 2 kh\
vAc = VBC = - 2- s'm'kh
1 + 2k hT)

where we have put

h k2h2 cot kh
0 = <p2c{O) = -3 1 +
I 1 + tan ifc h/2 k h kh

+ [
lo^-l]cot2kh (5.8.17)

The calculation of Ac does not present any difficulties, but it is not possible
to express the result in a concise form.
To identify the parameter e, the following observations are made. The
buckling mode is the first-order approximation of the bifurcated p a t h and
therefore the equations of kinematic compatibility which must be satisfied
are (5.2.25), from the first of which we get ii(x) = constant. In particular,
therefore, u 2 c (x) = u2c{£/2) — *i2c(0) = h. Furthermore, as the secondary
mode is symmetrical, it will be U2c(^/2) = 0, so by writing the asymptotic
expression of the displacement 6 of the mid-span point of the beam exact up
to the second order we have

6 = u2{l/2) = ti 2c (£/2) 6 -f - ue{l/2) e2 = he (5.8.18)


Li
197

from which £ = 6/h. The load-displacement relationship which defines the


bifurcated p a t h will therefore be

A F = Ac + K
( \ *') F = K F I1 + A2(<5//l)2 l (5 8 19)
'-
having put

** = \ye (5-8-2°)
In all of the examples examined so far, and for the sake of simplifying
the analysis, use has been made of a beam model with internal constraints
of axial and shear undeformability. Apart from cases in which the axial
and shear rigidity are important factors, the simple analysis of this model,
attributed to a drastic reduction of the unknown displacement components,
reaches a limit in the occurrence of non-linear boundary conditions. This is
what happens, for example, if we want to adopt a method of discretisation
such as that of finite elements, for complex cases in which it is not possible
to arrive at a closed form solution to the equations of the problem. For these
reasons, therefore, it is worth examining the problem of the portal frame
by using the unconstrained beam model. Referring to Figs. 5.28 and 5.29,
we shall adopt the same systems of coordinates and the same names for the
variables which appear in the problem. It is only necessary here to define the
axial and shear rigidity for the three rods, EA\y EA2, EA3, GA\, GA2y GA3
respectively, which now have finite values (with EA\ = EA3, GA\ = GA3).
The T P E of the system is written

2
P(w;A) - i | {^^[(l + u'Jcos^i+^sin^i-l]

+ G Ai [v[ cos <pi — (1 + u[) sin tpi] + E Ix <p[ \ dxx

+ - / l\EA2 [(1 + u2) cos <p2 + v'2 sin ip2 - l ] 2


2 Jo

+ G A2 [v'2 cos (p2 - (1 + u'2) sin <p2] + EI2 (p'2 | dx2

+ - / {EA 3 [(1 + u'3) cos <ps + V3 sin <ps - l ] 2


2 Jo l

+ G A3 [t>3 cos (ps - (1 + u'3) sin <p3}2 + E J 3 (p% | dx3

+ AF(u! + u3) (5.8.21)


198

It is easy to see that along the fundamental path the only quantities dif-
ferent from zero are

ui(x) = u3(x) = - —— x
EAi (5.8.22)
Ni[x) = N3{x) = -XF
Once the fundamental path has been identified, it is possible to write
the perturbation equations of the first order for the case under examination.
They are obtained from (5.2.9) and (5.2.10) and give the system of differential
equations
EAiu'l^O
Nu<p'lc + GA1{v,u-<plc)' =0
(Nlc - G A,) (v[c - <plc) -Eh <p'[c = 0

EA2uuie = 0
GA2{v2c-<p2e)' =0 (5.8.23)
GA2{v'2e-tp2e)-EI2tp"2e =Q

EA3u'i =Q
N3e<p'3c + GA3{v3c-<p3c)' =0
(N3c - G A3) (v'u - <p3e) -Eh <p"3c = 0

and the relation

E A, u'lc(£) 6 Ul(t) + Nu <plc{t) 6 Vl{e) -GAi \tpu{t) - v\M)\ ?> »i(<)


+ Eh V>'le{l) 6<Pi{l) -EAi ti'lc(0) 6Ul{0) - Nlc<pu{0)6Vl{0)
+ GAX \<plc{0) - «ie(0)] * t>!(0) - Eh<p'u{0) SipriO)
+ EA2u'2c(l)Su2{t) -GA2 \<p2c{l) - v'2c{l)]6v2{t)
+ EI2<P'2M S<p2{t) -EA2 u'2c{0) 6u2(0)
+ GA2 \<p2c{0) - v'2c{0)}6v2{0) -EI2<p2c{0)6<p2{0)
+ EA3 u'3c(i) 6 u3{£) + N3c <p3c(C) 6 v3{t)
-GA3 \<p3c(e) - v'3c(t)} 6 v3(i) +EI3 <p'3c(L) 6 <p3(t)
-EA3 u'3c{0) 6 u3(0) - N3c <p3c{0) 6 v3{0)
+ GA3 \<p3c{0) - v'3c{0)} 6 v3(0) -EI3 <p'3M 6 <p3{0) = 0 (5.8.24)
199

We must also consider the geometric boundary conditions, which now in-
volve the three functions u,v, <p

«ic(0) = 0 u3c(0) = 0
»i«(0) = 0 t>3c(0) = 0

uu{f) = v2e{0) u3e{l) = v2c{l) (5.8.25)


vie(l) = - ti2c(0) i>3c{t) = - u2e{£)
<Plc{£) = <f2c{0) <P3c(t) = <p2e(l)

From (5.8.24), bearing in mind (5.8.25), we obtain the natural boundary


conditions

EI1<p'lc{0)=0
EI3<p'3c{0)=0

EAi u'u{l) -GA2 \<p2c(0) - v'2c(0)} = 0


Nlc tpu{l) -GA1 \<pu(L) - v'u{l)\ -EA2 u'le(0) = 0
(5.8.26)
EIl<p'u(l) + EIi<p'iM = 0
E A3 u'3c{l) + GA2 \<p2c{l) ~ v'M] = 0
NSc <pSc{t) -GA3 [^(l) - v'3c(t)} + EA2 u'2c{l) = 0

Eh<p'ae(t)-EI2<p'2e{l) = 0

By solving the system of differential equations (5.8.23) with boundary


conditions (5.8.25) and (5.8.26), we obtain the characteristic equation

4/i 2 I2 \ k
KEAxt
- 2 + 6EI2J \EIl
+ GA -t sin k h = cos A; h (5.8.27)

having put

By adopting the normalisation

<Pu{h) = <PTC{0) = 1 (5.8.29)


200

we obtain the buckling mode

ti l c (x) = ti 3c (x) = - 2 — — - ta,n kh (I- x)x


£j A\ t
sin A;x
vu(x) = v*(x) = i (l - ^ - )
cos A: h
cos A;x
<Plc(X) = ^ScW =
cos A:/i
Ne
«2e(a:) = - £ ( l - ^ - tan kh) (5.8.30)

1 ( I3 ix2 x3
«2e(a;) = - 2 +
Eh '2 V 24 4 6
A: £
EL 1 — tan A: /i hx
£ 2

<p2c{x) = 1 - y y T tan fc /i (£ - x) x

Note t h a t at the limit for EA\, EA2, GA\^ GA2 - ^ o o w e get


2
k = XeF/EIl (5.8.31)
and the characteristic equation becomes
FT h
k h tarn kh = 6 —y- - (5.8.32)
E11 t
We note that (5.8.32) coincides with (5.8.8), which is the characteristic
equation obtained by using the internally constrained beam model. We also
see t h a t , in the characteristic equation (5.8.27), the extensional rigidity EA2
of the beam does not appear. This depends on the fact t h a t the buckling mode
is asymmetrical and therefore the horizontal displacement component of the
points of the beam is constant, as shown by the fourth of equations (5.8.30).
As the kinematics of the buckling mode is described by the linearisation of
(5.2.4) it is e2c = u'2c = 0, and therefore the extensional rigidity of the beam
has no influence on either the critical load or the buckling mode. By inserting
into the characteristic equation (5.8.27) the values of EA\, GA\^ EI\, GA2y
EI2y h, I which characterise each problem, we can numerically determine the
lowest value of k which satisfies (5.8.27). From this the buckling mode is
completely identified by (5.8.30), whilst from (5.8.28) we get the value of the
critical load

= -Nc = -
1 -
hik'^i
2
(5.8.33)

GAi
201

To provide an idea of the critical and post-critical behaviour of the frame


analysed, we give in the following some numerical results (Fig. 5.31) obtained
in parametric form for a frame in which

EAt = EA2 = EA3 = EA (5.8.34)

GAx = GA2 = GA3 = GA = EA/2 (5.8.35)

Instead of considering the critical load, it is more effective to refer to the two
non-dimensional quantities

\W = -Nch2/EI1 (5.8.36)

\W = -Nch2/EI2 (5.8.37)

Among all the results which it is possible to obtain, we believe it interesting


to give diagrams which show the dependence of Xy' and A^2' on the ratio
EI2/EIi excluding, however, the two cases Eli — 0, EI2 = 0 and Eli = oo,
EI2 = oo for which EI2/EI\ is undetermined. To examine all the possible
situations, we consider separately the cases,

(a) Eli — constant, EI2 — [0, oo[

(b) EI2 = constant, Eh = [0, oo[

r
Ac =2Ar-
24.
l/h=.25
©
A=-N c h 2 /EI,
©
16. — - A^-Nch2/EI2

°)
4
1—>io
I— 1 ° 32
10
2
w \\103 EAh 2 /EI 2 =10
EAh 2
El2 "
io\ 10 \Vi 04 AV=247

Fig. 5.31 -i _ _3

10" 10" 10 102 EI,/EI2


202

10" 10_1 1 10 102 EI,/EI 2

l/h=5. >104
Ac= 2.47 r10 3
A®-Nch2/EI,

10~2 10-' 1 10 102 EI,/EI 2

Fig. 5.31 - Dependence of the critical load on the stiffness ratio EIi/EI2.
203

In case (a) only the values of A ^ will be shown in diagram form, and in
case (b) only the values of X^\ corresponding to a number of values of the
non-dimensional ratio EAh2/EI2. The analysis is repeated for a few values
of the aspect ratio l/h (Fig. 5.31) [18].
The expressions for A^1) and A^2) are obtained from (5.8.33), (5.8.36),
(5.8.37). They are

1 -Jl + 2k2h2 Eh 2
EAh
(5.8.38)
Eh
EAh2
Eh
A< 2) = (5.8.39)
EI2

We can see that at the limit for ► oo, we have


Eh
AW = k2 h2 (5.8.40)

Ai8) = * * * ' § £ (5-8-41)


Note t h a t whilst (5.8.40) coincides with (5.8.9), (5.8.41) differs from (5.8.10)
by the factor l/h. Therefore, the curves in Fig. 5.30 coincide with those ob-
tained here for l/h — 1 in the limit case EAh2/EIi —» oo.
From the diagrams we can see how the axial rigidity EA plays a secondary
role in the interval of values which is most commonly met in technical ap-
plications. This becomes even more important with increase of the flexural
rigidity of the beam. In analysing the expressions of the buckling mode rela-
tive to the case under examination and those obtained from the undeformable
model, we must remember that here it is normalised according to <Pu{h) — 1,
whilst in the previous case 6 — h was used. The steps of the procedure,
and the results, for determination of the secondary buckling mode and the
parameters A c , Ac are omitted.

5.9 A P P L I C A T I O N OF T H E M E T H O D OF FINITE
ELEMENTS TO P R O B L E M S OF B I F U R C A -
TION IN P L A N E F R A M E S
From the previous examples we have seen how the use of the internally un-
constrained beam model, even though increasing the number of differential
equations to be solved in comparison with the constrained model, makes the
204

V, *
"1

9\ (F 4^2
^ 2
V = */le

V=0 1=1

Fig. 5.32 - Beam element and displacement components of nodes.

treatment of boundary conditions more simple. This results in a number of


computational simplifications when a numerical approach to buckling prob-
lems is adopted.
We illustrate here an unconstrained compatible finite element beam model
and discuss its use for non-linear problems of buckling for plane frames [19].
The following interpolation functions are assumed for the displacement
fields (Fig. 5.32)

u(rj) = (1 -rj)ui +r)u2

v{rj) = (1 - 3 T)2 + 2 rj3)Vl + tt[r) - 2 rj2 + T J V I

+ {3rJ2-2r13)v2 + le{-ri2 + rJz)<p2


(5 9 l}
+ 4 ( ^ - 3 ^ + 2^)70 - -

<p{v) = - J {V - V2)vi + (1 - 4 V + 3 r72)<pi + - {rj - r)2)v2

+ ( - 2rj + 3 T 7 2 ) ^ 2 - 6(r/ - r)2)l0

From the second and third of equations (5.9.1) we have v'(rj) — <p(rj) =
7o, for which 70 represents the linear part of the shear deformation of the
element, assumed constant. The boundary conditions are expressed simply by
identifying the nodal parameters Ui, vi, <pi and u 2 , v2? £>2 of each element with
the corresponding points of the adjacent elements, after making an obvious
reference frame change. By using (5.9.1), (5.2.7) can be rewritten in the form

A (w(q e ) 5 A) = We (w(q e )) - Ve ( w ( q . ) ; A) (5.9.2)

where

qf = { u i , vi, <pi, u2, v2, <p2, 70} (5.9.3)


205

The potential energy function (5.9.2) depends on a discrete number of


parameters. This therefore describes a discrete mechanical system which
approximates the continuous system described by (5.2.7). The problem for-
mulated in this way allows us to pass from analysis of the bifurcation of a
continuous system to that of a discrete system, to which all the considera-
tions made in Chapter 3 are applied. It is interesting to note at this point
how, working with successive approximations, the problem of the solution of
a system of non-linear differential equations deriving from the condition of
equilibrium (4.5.2) of a continuous system is transformed by the perturbation
process into that of a succession of systems of linear differential equations,
and hence, because of the discretisation, into that of the solution to a succes-
sion of linear algebraic systems. We leave all mathematical considerations on
the consistency of the procedure shown to specialised texts. Here, we wish
only to illustrate briefly an application to the case of plane frames.
Using (5.9.2) the TPE of a system composed of any number of elements
can be written in the form

P(w(qc);A)-^Pc(w(qe);A) (5.9.4)
e=l

Equation (5.9.4) is an approximated form of (5.2.21). Using (5.9.2), the


perturbation problem of the first order (4.5.13) for a single element is written
as

P e "iMu= \U gFe(\)6Zedxe (5.9.5)


Jo
where

( «e,x(»?) ^

ve,x(»?) - <Pt{vi)
Uv) =

(5.9.6)
( EAe ■ -Te{\)
GAe JV.(A)
*.(A) =
-Te{\) JV.(A) Nt{\)

V EL • J
By virtue of (5.9.1) we can express the vector £,.(»?) in terms of the nodal
parameters q e in the form

&(»?) = Se(»j) q e (5.9.7)


206

where the matrix S e (jj), which is generally known as the matrix of interpola-
tion of the element, assumes in our case the form

(27,-1)

2
3T7 2 - 4 77 + 1 (3r/-2)

Slfo) = (5.9.8)

6^
(217-1)

Zn2 -2n

6(n*-V) 7" ( 2 i ? - 1 )
V J
Denoting by q the vector containing all the nodal parameters of the system
expressed in one convenient reference frame we can, for each element, define
a matrix T„ such that

Te q (5.9.9)

Note t h a t the vector q contains the information concerning the mutual


constraints existing between all elements of the system (internal constraints)
and between the system and the body of reference (external constraints). By
substituting (5.9.9) in (5.9.7) and then (5.9.5) we obtain

lU gFe(\)6Zedxe = rqrTfSf(7?)Fe(A)Se(f?)T^qdxe
Jo Jo

= qTTjKe{X)Te6q (5.9.10)

where

K.(A) = / ^ Sffa) F.(A) S.(IJ) dx. (5.9.11)


Jo
Now substituting (5.9.10) in (5.9.4), we have

P"u6u = £qTTfK,(A)Te<5q

= qTK(A)£q = £qTK(A)q (5.9.12)


207

where

K(A) = f : T r K < ( A ) T e (5.9.13)


e=l

Note that in (5.9.12) we have

q T K(A)<5q = £ q T K ( A ) q (5.9.14)

by virtue of the symmetry of the matrix K(A).


Up to this point we have omitted any reference to the type of fundamental
path along which we are looking for the bifurcation. However, since for most
common plane systems of beams this can be considered linear, it is useful
from now on to make this assumption, given that it allows the introduction
of great simplifications. In this case, the internal forces of a beam along the
fundamental path are

AT (A) = \EAu'
T{\) = \GA{v' -tp) (5.9.15)
M(\) = \EI<p'
By making the dependence on A explicit, the matrix F e (A) can be written as

F , ( A ) = F 0 e + AF G e (5.9.16)

where

/ EAa ■
GA,
F«.= (5.9.17)

El

GA{v' -<p)
EAu'
FGe = (5.9.18)
GA{v'-<p) EAu' EAu'

Consequently, (5.9.11) can be written

K.(A) = /'*Sf(i7)Fo.S.(i ? )<te. + A [teSl{ti)FGtSt{r,)dzt


Jo Jo
= K 0 c + AK Ge (5.9.19)
208

and (5.9.13) becomes

K(A) = f^TfK0eT4 + A f^T[KGeTe


e=l e=l

= KQ + A KG (5.9.20)

The approximation, in discrete form, of the first-order perturbation equa-


tion (4.5.13) is therefore written as

q T K(A)<5q = 0 V6q (5.9.21)

from which, by virtue of (5.9.20), of (5.9.14) and of the arbitrary nature of


6 q, we have

(K 0 + A K c ) q = 0 (5.9.22)

The essential ingredients of (5.9.22) are the element matrices Koc and Kc c ,
which are shown in the following

K 11 Kg'
K 0e — (5.9.23)
Tf21 ■R-22

K
Ge K£
KGC = (5.9.24)
K& KE
where
/ EA
I
K
El El
o! = ■
12 (5.9.25)
T
El El
e?
/ EA
209

/ EA
I
EI
• -6
XT21 IF (5.9.27)
a E I EI
*~F 2 £
EI EI
12
V T 6 £2 )

/ EA
~T~
EI £7 EI
• 12 -nr 6 12
■K-22 £3 £3 IF (5.9.28)
£7 £7 a& E I
—6 — 4 £ 1F
£3
#7 EI EI
12 -=- 6 GAt+12
V £3
£3 £2 )

and

-GA
£2
6
K& = -EA^ ± E A \ (5.9.29)
£2 £2 10
1 r, . *
— EA - — EAe
10 £ 15

12 Tf \
GA
£ -°Ai
GA 6 „ . e 1 7
l
12 1 ^ , e
K Gt (5.9.30)
~IEAP 10 £ 5 £

EAe £Ae
10 £ 30 10

G 4
- i
GA
£2 5 £2 10 £
K 21
Ge
_
- (5.9.31)
£j4e
10 £ 30
7 £Ae
-GA
5 £ 10
210

(
-°Al • OAl

~GAl
1 „ , e
■K-22 _
5 £2 10 £

-±EA± — £Ae
15
10 £
G A
1 ^ , e — £Ae - - £Ae
K J 10 5
(5.9.32)

where we have put

e = u'
(5.9.33)
^ = v' — <p

and the writing of the index e for all the quantities in the matrices has been
omitted.
We can now determine the lowest value of A satisfying (5.9.22), which is
taken as the critical load of the system. The buckling mode of the system is
the name given to the associated eigenvector.
To study the post-critical behaviour it is necessary to rewrite in discrete
terms, using (5.9.1), all the terms which appear in (4.5.14) and (4.5.15). Then,
by means of (4.5.19), Ac is calculated, whilst by using (4.5.14) and (4.5.20) we
can determine the secondary critical mode and the coefficient Ac. We omit the
writing of the explicit expressions of these terms because, from a conceptual
point of view, there is nothing to add to what was necessary to construct
the expression P"u6u. We wish only to remind the reader (see Chapter 3)
that for discrete systems the algebraic system for the determination of the
secondary buckling mode takes the form

(K 0 + Ac KG) q c = / ( q e ; Ac) (5.9.34)

where / ( q c ; A c ) is a known term which depends on the buckling mode. The


solution q c , Ac depends on an arbitrary parameter which is eliminated by using
a condition of normalisation. From the point of view of automatic analysis,
this can be obtained simply by introducing a fictitious constraint.

In the following we report some numerical results obtained by using the


finite element model illustrated. The examples are chosen from those already
solved in closed form, to allow a direct comparison of results [19].
211

Euler b e a m

No. elements Ac Ac/A. Ac/2Ac


1 12.00 0- 0.10000
2 9.9438 0.- 0.12085
4 9.8747 0.- 0.12488
8 9.8699 0- 0.12499
exact solution 9.8696 0.- 0.12500

Portal frame with rigid beam

No. elements Ac A c /A c Ac/2Ac


1+1+1 2.4860 0.- 0.24396
2+1+2 2.4687 0.- 0.18205
4+1+4 2.4675 0.- 0.17828
8+1+8 2.4674 0- 0.17804
exact solution 2.4674 0- 0.17803

Portal frame with rigid columns

No. elements Ac A c /A c \c/2\e


1+1+1 6.0000 0.- 0.52524
exact solution 6.0000 0.- 0.52524

These results have been obtained by putting El = constant, EA —► oo,


GA —► oo, P = El ft7, and adopting the normalisations illustrated in Fig. 5.33 .

1 / h
//
I d
XZZZZZZZZZZZZL ^ 3
11
'/ I I
//
Vti Vitil /

V,
Fig. 5.33 - Sample structures.
212

Finally, it is interesting to pass to the analysis of bifurcation for some


plane frames which can be representative of structural types currently used
in construction. We present the results obtained for four frames without
braces, two symmetrical and two non-symmetrical [21]. Figures 5.34 and
5.35 show the geometry of the structures and the distribution of the flexural
rigidity of the elements as suggested by practical cases of steel structures.

30 El 10EI
►-

El
- 3m
>-

-
2 El

4 El
>-
8 El *
j3m
< 1 0 m
, t6mf>6m»

Fig. 5.34 - Symmetrical multi-storey frames.

15EI 15EI
3m
El

2 El

4 El

8 El
3m
■ 10m 6m 12, 4m
-*—f-

Fig. 5.35 - Non-symmetrical multi-storey frames.

The external loads have been divided into dead loads g and living loads
p equal to 1500 kg m" 1 and 1000 kg m _ 1 respectively. The multiplier A is
obviously of interest for living loads only. The values of the moments of
inertia / for the four cases examined have been chosen in such a way that by
213

subjecting each frame to the action of horizontal forces applied to each storey
of a size equal to (g + p)/80, we obtain a horizontal displacement of the top
storey (calculated using linearised theory) equal to If/500, where H is the
total height of the frame. The values of the moment of inertia which allow
us to obtain the required results, for the four cases, are

h = 1000 cm 4 , I2 = 600 cm 4 , J3 = 1500 cm 4 , J4 = 1700 cm 4

Normalising in such a way that e — 5, where by 6 we indicate the horizontal


displacement of the top storey of the frame, the following values for A c , A c , Ac
are obtained

Frame type K K Ac
1 9.32 0.- 0.175
2 10.48 0.- 0.157
3 15.86 0.019 0.248
4 19.48 0.053 0.517

For frame type 4 we also give the graph of the bifurcated path. Despite
the fact that this is the case with the largest value of Ac, the tangent to the
bifurcated path is practically horizontal.

X
200

18.0

160

14.0

12.0

100

80 4
6.0

40 4
2.0 J

—l—
0.5 1.0 1.5 2.0
*W
Fig. 5.36 - Equilibrium curve for a multi-storey frame.

In conclusion we can state that, however non-symmetrical plane frames


commonly used in construction may be, the post-critical path may well be
represented by a horizontal straight line.
214

REFERENCES
E. Janke, F . Emde: Tables of functions, 4th ed., Dover, New York, 1945.

N.M. Newmark: "A simple approximate formula for effective end-fixity


of columns", J. Aeronaut. Sci., 16 (1949), 116.

F. Bleich: Buckling strength of metal structures, McGraw-Hill, New York,


1952.

S.P. Timoshenko, J.M. Gere: Theory of elastic stability, 2nd ed.,


McGraw-Hill, New York, 1961.

M.R. Home, W. Merchant: The stability of frames, Pergamon Press,


Oxford, 1965.

A.S. Volmir: Stability of deformable systems, NASA AD 628 508, 1965.

J. Roorda: "Stability of structures with small imperfections", J. Eng.


Mech. Div., Proc. Am. Soc. Civ. Eng., 91, No. E M I (1965), 87.

V. Franciosi: Strength of materials (in Italian), Liguori, Napoli, 1967.

M. Como: Theory of elastic stability (in Italian), Liguori, Napoli, 1967.

W.T. Koiter: "Post-buckling analysis of a simple two-bar frame", in


Bertram Broberg et al. (editors): Recent progress in applied mechanics,
Almqvist & Wiksell, Stockholm, 1967, p. 337.

J. Roorda, A.H. Chilver: "Frame buckling: an illustration of the pertur-


bation technique", Int. J. Non-Linear Mech., 5 (1970), 235.

Handbook of structural stability, Column Research Committee of Japan,


Corona Publishing Company, Tokyo, 1971.

M. Como: "Stability of structures" (in Italian), in R. Baldacci, G. Cera-


dini, E. Giangreco (editors): Dynamics and stability (in Italian), Italsider
technical-scientific series for the design of steel structures, Vol. l i b , 1971.

S.J. Britvec: The stability of elastic systems, Pergamon Press, New York,
1973.

E. Pozzo: "A post-buckling behaviour investigation on an imperfect sim-


ple portal frame", Report No. 11, Istituto di Scienza delle Costruzioni
of the University of Cagliari, Cagliari (1974).

D.O. Brush, B.O. Almroth: Buckling of bars, plates and shells, McGraw-
Hill, New York, 1975.
215

[17] J.G. Simitses: An introduction to the elastic stability of structures,


Prentice-Hall, New Jersey, 1976.

[18] M. Alessi, P. D'Asdia, M. Pignataro: "Post-buckling analysis of some


sample frames", Report 11-189, Istituto di Scienza delle Costruzioni of
the University of Rome "La Sapienza", Rome (1976).

[19] R. Casciaro, A. Di Carlo, M. Pignataro: "A finite element technique for


bifurcation analysis", Report 11-192, Istituto di Scienza delle Costruzioni
of the University of Rome "La Sapienza", Rome (1976).

[20] S.S. Antman: "Bifurcation problems for non-linearly elastic structures",


in P.H. Rabinowitz (editor): Application of bifurcation theory, Academic
Press, New York, 1977.

[21] F.M. Mazzolani, A. Di Carlo, M. Pignataro: "Post-buckling behaviour of


multistorey steel frames", Proc. International colloquium on stability of
structures under static and dynamic loads, Structural stability research
council, Washington D.C. (1977).

[22] L. Corradi: Structural instability (in Italian), CLUP, Milano, 1978.

[23] P.A. Kirby, D.A. Nethercot: Design for structural stability, Granada
Publishing, London, 1979.

[24] M. Pignataro, N. Rizzi, A. Di Carlo: "Symmetric bifurcation of plane


frames through a modified potential energy approach", J. Struct. Mech.,
8 (1980), 237.

[25] N. Rizzi, A. Di Carlo, M. Pignataro: "A parametric post-buckling anal-


ysis of an asymmetric two-bar frame", J. Struct. Mech., 8 (1980), 435.

[26] A. Di Carlo, M. Pignataro, N. Rizzi: "On the proper treatment of ax-


ial and shear undeformability constraints in post-buckling analysis of
beams", Int. J. Non-Linear Mech., 16 (1981), 221.

[27] M. Pignataro, A. Di Carlo, R. Casciaro: "On non-linear beam models


from the point of view of computational post-buckling analysis", Int. J.
Solids Struct., 18 (1982), 327.
217

Chapter 6

THIN-WALLED BEAMS WITH OPEN


CROSS-SECTION

6.1 INTRODUCTION

In the previous Section we analysed the mechanical behaviour of plane sys-


tems of beams and determined the conditions under which the critical state
of equilibrium manifests itself. Each beam element was bent in one of the
principal planes of inertia of the cross-section and subjected to normal and
shear forces and to a bending moment.
The results of the analysis are applicable to systems of beams with con-
straint conditions which allow deflection in only one of the two principal
planes of inertia, as well as to beams with compact cross-section for which
the deflection corresponding to the critical load takes place in the plane with
the lowest rigidity. However, when the torsional rigidity of the beam is very
small, as happens in thin-walled open cross-sections, buckling can take place
with a rotation of the cross-sections around a longitudinal axis, and may be
accompanied by deflection in one or both of the principal planes of inertia.
The first case is known as torsional instability and the second as flexural-
torsional instability, and in general they appear for load values which are
much below the lower of the two Eulerian critical loads. In such a case, the
omission of the torsional effect from design considerations leads to a danger-
ous over-estimation of the load-bearing capacity of the structure.
The figures which follow, taken from [15], show the results of some exper-
imental tests. Figure 6.1 shows the deformation under load of a symmetric
angle cross-section. The two flanges bend according to a sinusoidal law, with
maximum deflection at the free edge. If a horizontal section is made at mid-
height, we can see that the change of shape of the cross-section is insignificant
in relation to the rigid rotation. The mode is therefore essentially torsional.
Figure 6.2 shows a thin-walled square cross-section with a longitudinal
cut along two adjacent plates. The critical load this time is manifested by
a corrugation of the two flanges adjacent to the free edge, so t h a t the cross-
section loses its original shape. This phenomenon is substantially different
from the previous one, and is characterised by a local buckling of the beam.
In this case the load under which local buckling of the flanges occurs is lower
than that which causes overall deflection of the beam. If the ratio between the
218

gjj^g&*»»"~

Fig. 6.1 - Torsional buckling of a thin-walled beam.

p§&


l\i
UJU

Qp-

Fig. 6.2 - Local buckling of a thin-walled beam.


219

thickness and the width of the flanges increases, the overall behaviour prevails.
The phenomena of local buckling can be avoided in practice by reducing the
extreme slendemess of the flanges which make up the thin-walled beam. This
can be achieved either by increasing the thickness of the plates, or by using
transverse stiffeners which decrease the distance between two adjacent nodal
lines of the deflected shape of the flanges.
Figure 6.3 shows another type of buckling known as lateral buckling. This
figure shows an I-section fixed at one end and loaded by a weight applied
to the centroid of the free end. By increasing the load the cantilever first
deflects in the vertical plane and then, for a certain value of the load, it
suffers a displacement in the horizontal plane and rotation around its axis.
On the pulley fixed to the end section it is possible to read the torsional angle
of rotation of the beam. Figure 6.4 shows the same phenomenon for a beam
constrained at both ends.
The examples illustrated show clearly how, to study the phenomenon,
it is necessary to abandon the plane beam model and formulate a three-
dimensional model which allows us to analyse the effects of interaction be-
tween bending and torsion. The problem is much more complex than t h a t

Fig. 6.3 - Lateral buckling of a Fig. 6.4 - Lateral buckling of a b e a m


cantilever. constrained at both ends.
220

discussed so far and is today still a subject for study. We therefore limit our-
selves here principally to analysis of the linearised problem of stability, t h a t
is, to the evaluation of the critical loads and buckling modes. An outline will
then be given of the post-critical behaviour of thin-walled beams.
The work of Vlasov [2] is followed in the formulation of a mathematical
beam model, and then the particular problems are solved by means of the
general theory illustrated in Chapter 4.

6.2 H Y P O T H E S E S ON THE M E C H A N I C A L BE-


HAVIOUR OF THIN-WALLED B E A M S
The theory of thin-walled open cross-section beams is founded on two basic
kinematical hypotheses:

(a) the beam cross-sections are not deformable in their own planes;
(b) shear deformation on the middle surface is negligible.

The analysis is conducted under the hypothesis of small displacements (lin-


earised kinematics).
As a consequence of the previous hypotheses, the cross-sections move in
their own planes as rigid bodies and they warp, leaving the straight angles
on the middle surface between the generatrices and the cross-section middle
line unaltered.
The first hypothesis allows us to describe the motion in the plane, accord-
ing to the laws of the kinematics of rigid bodies, and the second to express
the motion outside the plane as a function of t h a t in the plane, to within a
rigid longitudinal displacement. The generic configuration of the system is
therefore identified by the position of the points of a longitudinal axis as well
as by the rotation of the cross-sections around the axis [1-4,6,7,9,12,13].
It is well known that if, in a beam subjected to torsion, the twisting moment
is variable along the beam axis or the warping is impeded in one or more cross-
sections, self-equilibrated normal stresses which are variable along the axis
arise on the cross-section. Tangential stresses are associated with these, in
addition to those deduced from classical theory. As a result the beam exhibits
a torsional rigidity which is greater than t h a t deduced from the theory of
De Saint Venant, with smaller torsional deformations, accompanied however
by longitudinal deformations which are not negligible. On the other hand,
longitudinal loads or couples may cause torsional deformations associated
with self-equilibrated states of tangential stresses.
If these effects are secondary for beams with compact cross-sections, in
t h a t they are rapidly damped with distance, they cannot be neglected in thin-
walled open cross-sections. As a consequence, the principle of De Saint Venant
221

Fig. 6.5 - Thin-walled beam subjected to eccentric axial load.

is not valid for thin-walled beams. In fact, two equivalent systems of longi-
tudinal forces applied to the same cross-section of a beam do not produce
the same state of stress and deformation in the solid; in other words, a self-
equilibrated system of forces induces a state of stress which is different from
zero. The example in Fig. 6.5, taken from Vlasov [2], serves to make these
circumstances clear. The longitudinal force P is applied at the free end of
a cantilever of I cross-section (Fig. 6.5(a)). This can be thought of as the
sum of the four systems of forces shown in figures (b)-(e). As a result of the
deformation, the forces produce a uniform stretching (b) and two bendings
in the principal planes of inertia (c,d) in accordance with classical theory.
The fourth system (e) is equivalent to zero and, according to the principle
of De Saint Venant, should not induce deformation except in the immediate
vicinity of the cross-section on which it is applied. In the case under exam-
ination, on the other hand, the forces cause bending of opposite sign in the
two flanges, with consequent warping and rotation of the cross-sections and
222

not negligible deformations, which decay very slowly along the beam axis.
The example shows clearly how such types of solids require the development
of a theory which is independent of the classical theory of beams. On the
other hand, what is presented here is limited by the kinematic hypotheses
mentioned. A more correct formulation of the problem would require the
study of the beam as a cylindrical shell, thus abandoning the hypothesis of
undeformability of the cross-section. Nevertheless, the results reached are
validated by experience, and are without doubt applicable to technical use.

6.3 KINEMATICS OF THE BEAM


Let us consider a thin-walled open cross-section member of constant thickness.
Let S be the middle surface of the beam and P the generic point belonging
to it. We take an orthogonal Cartesian system of coordinates as shown in
Fig. 6.6, with x and y the principal axes of inertia of the cross-section and
z the centroid axis, and denote by C the curve of intersection of S with the
plane z — constant. By choosing an arbitrary origin M on C, it is possible
to introduce a curvilinear abscissa s which identifies the position of the point
P. Let up,vp,wp be the displacement components of P in the Cartesian
coordinate system and £p,r)p,wp the components in the intrinsic coordinate
system of the curve (Fig. 6.7). We study first of all the motion in the plane.

Fig. 6.6 - Curvilinear coordinates Fig. 6.7 - Intrinsic system of coor-


and displacement com- dinates.
ponents.
223

Under the hypothesis of undeformability of the cross-section, the kinematics


of rigid bodies gives

up{z,s) = u{z) - 0{z) {y - yc)


(6.3.1)
vp{z, s) = v(z) + 6(z) (x - xc)

in which u(z) and v(z) are the displacement components of an arbitrary


point C of coordinates x c ,y c and 0{z) is the rotation of the section around
C, positive if the rotation is anti-clockwise; x = x(s) and y = y(s) are
the coordinates of the point P (Fig. 6.7). Projecting (6.3.1) in the intrinsic
coordinate system we get

£p(z, s) = u(z) cos a(s) + v(z) sin a(s) + 0(z) r(s)


(6.3.2)
rjp{z, s) = — u{z) sin a(s) + v{z) cos a(s) — 6(z) t[s)

where r(s) and t(s) are the coordinates of C in the local system of coordinates
and a(s) is the angle which the tangent forms with the axis x.
Let us now pass to an examination of the motion outside the plane. Under
the hypothesis of no shear deformation on the middle surface, we get

7..M = ^ ^ + ^ ^ = 0 (6.3.3)

from which it is possible to obtain wp(z,s) by integration. Making use of the


first of equations (6.3.2) and indicating differentiation with respect to z by a
dash, we have
rP

JM
rP
= f (z) - / [u'(z) cos a(s) + v'(z) sin a(s) + #{z) r(s)] ds
J hi
(6.3.4)
in which f (z) is an arbitrary function of z. From Fig. 6.8 we get

cos ads — dx
sin ads = dy (6.3.5)
r ds = du

doj being twice the shaded area in the diagram, which is called the sectortal
area.
Using (6.3.5), (6.3.4) yields

wp (z, s) = $(z) - v!{z) x{s) - v'{z) y{s) - 0'{z) u{s) (6.3.6)


224

Fig. 6.8 - Element of sectorial area.

in which f = f + u'xm + v'ym is again an arbitrary function of z and

w (s) = f r (a) ds (6.3.7)


JM

We observe that the sign of u given by (6.3.7) is positive if the sectorial


area is spanned by the radius vector CP in an anti-clockwise direction.
Relations (6.3.1) and (6.3.6) describe the displacement field of the points
belonging to the middle surface of the beam in terms of the four independent
functions u(z),v(z), 0{z), $(z). If the beam is identified by its middle surface,
t h a t is, if the variations of the functions in the thickness are neglected, the
relations describe the whole displacement field. We shall return later to the
consequences of this last hypothesis.
Let us now examine the axial displacement wp{z,s), which is the sum of
four terms. The first represents a longitudinal rigid motion of the generic
cross-section, variable along the axis, and is associated with an extensional
deformation of the beam. In the following we assume f (z) = 0. It is possible to
show t h a t such a hypothesis does not influence the results which are obtained
here. The second and third terms represent rigid rotations of the cross-
section around the principal axes of inertia, and are associated with flexural
deformations under the hypothesis of shear undeformability. All three terms
can be deduced from classical theory. The fourth term characterises the
behaviour of the solid under examination, representing a distortion of the
section outside the plane. The warping which results is proportional to the
torsional curvature 0'(z) and varies on the section according to the value of
225

Fig. 6.9 - Sectorial area.

the sectorial area, given by the function u(s). Figure 6.9 shows a qualitative
plot of the function CJ(S) for a generic cross-section.
The extensional deformation ezz(z,s) = w'p(z,s) corresponds to the dis-
placement (6.3.6) and is given by

ezz (z,s) = - u"{z) x{s) - v"{z) y{s) - 0"{z) u{s) (6.3.8)

which shows the dependence of the deformation on the curvatures u"(z) and
v"(z), as well ELS on the variation of the torsional curvature 0"(z). Equation
(6.3.8) includes, as a particular case, the De Saint Venant pure torsion, in
which we have 0'(z) = constant. The extensional deformation ezz(z,s) rep-
resents the only non-vanishing component of the tensor of the infinitesimal
deformation. This, at first sight, is in contrast with the De Saint Venant
theory of torsion, from which we get a linear distribution of shear deforma-
tion along the thickness, with a zero value corresponding to the middle line.
The problem arises from the fact that neglecting the displacement variations
along the thickness is equivalent to identifying the beam by its middle surface,
which, by hypothesis, is free of shear deformation. Consequently we have a
mechanical model for which the De Saint Venant problem of torsion cannot
be formulated. On the other hand, there are cases in which such an effect
has a determining role in analysis of the behaviour of the beam, especially
for those sections in which u (s) = 0. Therefore, in conformity with classical
theory, this is taken into account in an approximate way, adding the torsional
deformation energy of De Saint Venant to the energy of deformation of the
Vlasov model. A more rigorous theory without this formal inconsistency re-
quires more sophisticated assumptions on the displacement iu(x,y), defined
over the domain of the whole cross-section. If on the one hand this procedure
appears more correct, on the other it loses the simplicity of formulation of
226

the rule of the sectorial area, without perhaps adding greatly to the accuracy
of the results.
We now pass to a closer examination of the choice of the point C to be
assumed as a pole of the sectorial area. By expressing the distance r(s) in
terms of the coordinates of P and C, we have (Fig. 6.7)

du = rds — — [xc — x) sin ads + (yc — y) cos ads

= (x - xc) dy - (y - yc) dx (6.3.9)

where we have used (6.3.5). Denoting by uc — uc(s) and ud = cJd{s) two


functions referred to two points C and D (Fig. 6.10) we have, by applying
(6.3.9),

duc = (x - xc) dy - (y - yc) dx


(6.3.10)
dud = (x - xd) dy- (y - yd) dx

from which

d(uc - ud) = (yc - yd) dx - (xc - xd) dy (6.3.11)

and therefore by integration, as o;c(0) = wd{0) = 0 , we have

uc = Vd + {yc - Vd)(x - xm) - {xc - xd)(y - ym ) (6.3.12)

in which x m and ym are the coordinates of the origin M of the sectorial area.
Equations (6.3.12) show the variation of the sectorial area u(s) with variation
of the pole.

ik
V
^•D
Q^d/2/
. -^ C
M
S >v /
^©c/2

P
\ X

Fig. 6.10 - Sectorial area^ referred to two poles.


227

For reasons which will later become clear, it is convenient to determine


the point C for which the two conditions

f ucxdA = 0 , f ucydA = 0 (6.3.13)


JA JA

are satisfied, A being the area of the cross-section. By substituting (6.3.12)


in (6.3.13) and carrying out the integration, we have

z c = x,iH-— udydA, yc=yd~Y ^d^dA (6.3.14)


1X JA ly JA

in which Ix and Iy are the moments of inertia with respect to the principal
axes. Equations (6.3.14) provide the coordinates of C if we know the function
Wd{s)> calculated with respect to an auxiliary pole. In applications, it can be
convenient to assume D is coincident with the origin O. Equations (6.3.14)
then give

xe = — u>0ydA , ye= - — u0xdA (6.3.15)


lx JA ly JA

Point C is known as the principal pole. It is simple to verify t h a t if the


cross-section has an axis of symmetry, then the pole lies on the axis.
Up to now an arbitrary point has been chosen for the origin M of the
curvilinear abscissa. In what follows, it is convenient to make reference to an
abscissa for which

uc dA = 0 (6.3.16)
/ .A
From Fig. 6.11 we obtain

CJ C (S 2 ,S) = w c (si,s) - ue{sus2) (6.3.17)

Supposing t h a t CL>C(S2,S) satisfies (6.3.16), we have by using (6.3.17)

ue:{sus2) = — / ue(sl9s)dA (6.3.18)

Equation (6.3.18), when the function C J C ( 5 1 5 5 ) with origin N is known, allows


us to determine the abscissa s of the unknown origin M. Function C J C ( S 2 , 5 )
is therefore

w c (s2 ? s) =UJC(SUS) - — 0Jc(sus)dA (6.3.19)


A JA
from which we observe t h a t this is obtained from the function u; c (si,6), with
arbitrary origin, by subtracting its average value. The point M is called the
null point of the sectorial area. As is simple to verify, if the cross-section has
an axis of symmetry, all the points of the cross-section which belong to the
228

(Dc(S„S2)/2

c(»2.S)/2

Fig. 6.11 - Sectorial areas referred to two origins.

axis are null points; of these, that nearest to the principal pole is called the
principal null point.
The principal pole and the null point of the sectorial area are points
whose positions depend exclusively on the geometrical properties of the cross-
section, such as the centroid and the principal directions of inertia, and as-
sume a relevant role in the theory of torsion as the centroid and the principal
axes in the theory of bending. In fact, the point C coincides with the twisting
centre and so is the centre of rotation of the cross-section in the torsional
deformation. The point M is the origin of the function u(s) which describes
the warping. To verify this, a beam subjected only to torsion is considered.
By referring the function u(s) to the unknown centre of rotation C, we have
from (6.3.6)

wp(z, s) = f (z) - 6'{z) tuc(s) (6.3.20)

in which u;c(s) is measured from an arbitrary origin M and f (z) — constant.


To the displacement wp(z,s) there corresponds a longitudinal deformation
and a normal stress, respectively

ezz(z,s) = -6"(z)uc(s) (6.3.21)

ozz{z,s) = -E6"{z)ue{s) (6.3.22)

where E is the elastic modulus and the Poisson effect is neglected. As, by
hypothesis, the axial force and bending moments are zero, we must have

/ ozz dA = 0 , / azz xdA = 0 , / ozz y dA = 0 (6.3.23)


JA JA JA
229

from which, for (6.3.22),

JucdA = 0, I ucxdA = 0 , fucydA = 0 (6.3.24)

Equations (6.3.24) allow us to determine the centre of rotation C and the


unknown origin M. As these coincide with (6.3.16) and (6.3.13), our state-
ment is proved. In the following, the sectorial area function which satisfies
(6.3.24) is indicated by the symbol u(s).
Finally, it is possible to verify that the twisting centre coincides with the
shear centre. The two systems of forces indicated in Fig. 6.12, applied to the
same beam, are considered for this purpose. The force T is applied to the
shear centre and by definition does not cause rotation of the cross-sections;
therefore, the work done by the couple M is zero. According to the theorem
of Betti the work done by the force T, because of the rotation caused by the
twisting moment M , is also zero. From this we deduce t h a t the axis of the
shear centres CC has no displacement due to the torsional deformation of the
beam, and thus it is also the axis of rotation.

Fig. 6.12 - Shear centre and twisting centre.

6.4 F O R M U L A T I O N OF T H E LINEARISED P R O B -
LEM OF STABILITY
It is known from the general theory presented in Chapter 4 t h a t the critical
condition of equilibrium arises from the solution to the variational problem
(4.5.5), which expresses the stationarity of the second variation of the total
230

potential energy. The eigenvalue problem which appears is generally non-


linear, depending on the type of the fundamental equilibrium path. It has
been seen, however, that if the path is trivial, as happens for inextensible
beams subjected to axial load, then the eigenvalue problem is linear.
In the following we shall look for the critical condition of equilibrium of
beams which, in the fundamental configuration, are subjected to compres-
sion, bending or both types of stress, and are therefore characterised by a
non-trivial fundamental path. With the aim of linearising the problem, we
introduce the hypothesis that the pre-critical deformations are negligible,
therefore identifying the critical configuration with t h a t which is unloaded.
Writing the second-order terms of the series expansion of the total potential
energy PQ'U2/2 beginning from the configuration u = 0, we can then deter-
mine the eigenvector and eigenvalue which make the functional stationary.
The second variation of the total potential energy results, according to
(4.3.8), from the sum of three terms

^Po'u2 = \ I Eijhk e,-y ehk dV + \ f S{j uhyi uhJ dV - \ Ttf u 2

= W2 + W;-V2 (6.4.1)

The first term W2 is the elastic potential energy of linear theory, where Eijhk
are the components of the elastic tensor and ety the components of the tensor
of infinitesimal deformation; the second term W2* is the work of the stresses
Sij acting in the fundamental state due to the second-order components of
the additional deformation (4.3.1), and is usually called the geometric term,
as it does not depend on the elastic properties of the system; the third term
V2 is the second-order part of the work of the external loads.
In (6.4.1) the displacements u, are measured from the fundamental config-
uration. In the problem under examination, we have u = {u(z), v(z),0(z)}.
Due to the kinematic relations (6.3.1) and (6.3.6), it is possible to express
(6.4.1) in terms of the components of the vector u . By then imposing sta-
tionarity with respect to the three independent functions, we arrive at the
eigenvalue problem (4.5.5).
We can therefore begin to calculate the first term of (6.4.1). From (6.3.8),
by omitting the independent variables, we have

W2 = - J Ee222dV = - j E{-u"x-v"y-0"uf dV

= - [dzE [ {u"2x2 + v"*y2 + 0"2u2+2u"v"xy

+ 2u"0"LJx + 2v"6"uy)dA

= - J (EIy u"2 + EIX v"2 +ET 0"2) dz (6.4.2)


231

being

J xydA = 0, J ojxdA = 0 , fuydA = 0 (6.4.3)

since (x, y) is a principal coordinate system and OJ is referred to the twisting


centre. In (6.4.2) I is the length of the beam, A the area of the cross-section,
Ix and Iy the moments of inertia with respect to the central axes and

T= f u2dA (6.4.4)
JA

is the warping constant. It must be remembered t h a t in (6.4.4) the function


UJ = u(s) has its origin at the null point of the sectorial area; if u is known
instead with reference to an arbitrary point, we subtract from this its average
value according to (6.3.19).
The table which follows shows, for several types of commonly used cross-
sections, the position of the twisting centre and the warping constant V [3].
For cross-sections consisting of narrow rectangles, the medians of which all
converge on the same point such as in T, L,V and cross-shaped sections,
we have r = 0. The twisting centre is, in fact, the point of intersection of
the medians, for which the function u is identically zero. It is the same for
the narrow rectangular cross-sections. This is obviously the consequence of
having neglected the variations of wp along the thickness. A more rigorous
analysis supplies non-zero values of T which are, however, very small. For the
narrow rectangular cross-section, for example, we have T = 6 3 /i 3 /144 where
6 and h are the sides of the cross-section. In accordance with what was said
in Section 6.3, the De Saint Venant energy of torsional deformation must be
added to the energy H>2- This leads to

W2 = - f (EIy u"2 + EIX v"2 + E r 6"2 + G J 0'2) dz (6.4.5)

GJ being the torsional rigidity furnished by classical theory. For cross-


sections consisting of narrow rectangles of thickness hi and length 6,-, and
for sections of slowly variable thickness, we can assume, respectively,

J = lj2bih*, J = lJch3(s)ds (6.4.6)


where the sum is extended to all the rectangles and the integral to the middle
line of the cross-section.
Let us now consider the second term of (6.4.1), W2*- This depends on the
state of stress acting on the beam in the fundamental configuration, and has
therefore to be specified for each single problem under examination.
We consider in the following a state of stress 5t;- generated by axial and
shear forces and bending moments in the absence of torsion. From the hy-
pothesis, it follows t h a t plane cross-sections of the beam before deformation
232

Geometrical Properties of Thin- Walled Open Cross-Sections

b
,- r.
2btf + fctj
/= if tf=tw=t

li.
tw
cr
3
h
r 24~
J = i - ( 2 6 + h)
1=—
bi e= if tf=tw=t
r—1 1 -
tw
(t>i+b2)t3f + ht3w t3
h /= J= — (bx + b2 + h)
|<D 3 o
II ;
bT .J r 12 6? + 6 |
2
36 ty

-v
e= if tf = tw=t
66 fy + /if,,,
N- 362
26t 3 + fcti g —
© IT *-i -r
/= 66 + /i
o h
3

-f, ty6 3 /l 2 36ty + 2htw


tu»L^
w
. _. j
T
1-
r
12 66ty + fctw *6 3 /i 2 36 + 2/i
p -
12 66 + h

2btf + /if ^
t-.
b
r. /= if
♦"1 3
tf-tw-t

J=j(2b + h)
t cj
h
r = [2tf(b2 + bh+ h2) +
Y
b3h2 tb3h2 6 + 2/i
Vb~ + 3twbh]
r -
T
12(26 + h) 2 12 26 + h

sina — acosa
e = 2a : if 2a = 7T
a — sin a cosa

^^t c
J=
2aat3
3
e = —
4a
7T
mt3
/ - - ■ - —
3
5 5 3
2ta
r 3
r = 2 f3—
a , TT

4
12

TT '
v

e
ur*^ -' r 6(sina - a cosa
x
L
a 3 _
a — sina COSa '-}
233

remain plane also after a deformation which takes the solid from the natural
state to the fundamental state, and therefore the stresses 5ty can be deter-
mined through the theory of De Saint Venant. T h e torsional behaviour and
consequent warping of the cross-sections of the beam are taken into account
only in passing from the fundamental configuration to the adjacent generic
configuration. Obviously, the hypothesis of the absence of torsional deforma-
tion in t h e fundamental state is satisfied only if the external loads are suitably
applied. It is, first of all, necessary that the lateral loads pass through the
twisting centre of the generic cross-section, but such conditions are not suf-
ficient. We have in fact seen, in Section 6.2, how torsion can develop also
in the presence of only longitudinal forces and bending couples. Mention
was also made of the fact that equivalent systems of longitudinal forces do
not produce the same effects. Since in the following we shall consider the
case of axial forces and bending couples applied to the ends of the beam, it
is understood t h a t they must result from a distribution of forces such t h a t
no torsion is generated. This can, for example, be realised by distributing
the longitudinal forces uniformly on the cross-section and requiring t h a t the
planes of the couples pass through the twisting centre. If these conditions
do not occur, a more rigorous analysis requires the determination of a com-
plementary state of normal and tangential stresses, which are added to that
furnished by classical theory.
Bearing in mind the previous hypotheses, we can pass to the calculation of
the geometrical term. For simplicity, we determine separately the contribu-
tions to W2* arising from the normal and tangential stresses. Let us consider
first the normal stresses 5 3 3 = ozz. Figure 6.13 shows the sign conventions
assumed for the axial force N and for the moments Mx and My. T h e force N
is constant along the axis, but the moments are in general variable because
of the effect of the lateral loads. The state of stress at the abscissa z is

<? N Mx My

having taken the tensile stresses as positive. T h e work W£a of the normal
stresses S33 is expressed as

*>*I = \ I S33 « + < + < ) dV

* \ j v SM{[U' - 0'(y - yc)}2 + [v' + 0'(x - xc)}2} dV

=
\ hi Szz { ^ + "' + 6'2 I(X ~ ^ +{V~ y ' )2 l
+ 2 e'{v'x - u'y) + 2 0'{u'yc - v'xc)} dA dz (6.4.8)
234

Fig. 6.13 - Positive direction of the generalised stresses.

where we have neglected the term £331^ with respect to Ew'p present in W2,
as £33 <C E. By substituting for 5 3 3 its expression (6.4.7) and carrying out
the integrations on the area of the cross-section, we have

N
ic2a =
y - f (u'2 + v'2 +r2J'2 +2u'0'yc -2v'9'xc) dz

+ f Mx (Cx f 4- u 6') dz + j My (Cy 6'2 - v' 0') dz (6.4.9)

in which r\ = Ic/A, with Ic the polar moment of inertia with respect to the
twisting centre, and Cx and Cy are two coefficients defined as

C, = Vc-^r fy(x2 + y2)dA


A ix JA
(6.4.10)
Cy = xc- —— / x(x2 + y2) dA
Z lv JA

If the axis x is an axis of symmetry then Cx = 0; analogously, if y is an axis


of symmetry, Cy — 0.
235

Fig. 6.14 - Positive direction of stresses and generalised stresses.

We now examine the contribution W2*r OI* ^ n e tangential stresses S13 = TXZ
and 523 = ryz. With reference to the sign conventions indicated in Fig. 6.14,
it is

L TZX dA = Tx=M'v,
L ryz dA = Ty = M'x (6.4.11)

We therefore have

Wr = J [Sl3 {UP>1 Up>3 + Vpti VPf3 + WPfl Wpt3)

+ $23 {Up,2 UPt3 + VPf2 Vp,3 + Wp%2 Wpyz)\ dV

^ j y {TZX 0 W + ff[x - xc)\ - rzy 0 [u1 - 0'(y - yc)}} dV

= fdz00' f [rzxx + Tzyy)dA

+ f M'x0{u' + 0' yc) dz- f M'y 0{v' - 0' xc) dz (6.4.12)

where uPti = dup/dx — 0 and vPf2 = dvp/dy = 0 from (6.3.1) and we have
used (6.4.11). The terms Sis wP,i wP)3 and 523 wPi2 wP$ have been neglected
with respect to the elastic term (6.4.5).
Let us now consider the first of the three terms in W2T. We have the
identity

rzxx + Tzyy = - — [(x 2 + y 2 ) TZX] + - — [(x 2 + y 2 ) rzy]


2 dx 2 dy
1 2
(x + yy 2 ) ( ^ + ^ (6.4.13)
2K ' \ dx dy

The integral over the cross-section of the sum of the first two terms on the
right side vanishes, because the cross-section contour is free of forces. In fact,
236

from the divergence theorem, we have

L{ik[v+**-] +&*+*'-]}"
= j [(** + V2) {nx rzx + ny rzy)] ds = 0 (6.4.14)

where nx and ny are the direction cosines of the outer normal.


Taking into account (6.4.14) and the indefinite equation of equilibrium in
the direction z
drzx drzv dozz . x
+ (6 16)
aT ~H + "af = ° "
we have

J {rzx x + Tzyy)<LA = - J (x2 + y2) -—^ dA

= M'x{Cx-yc) + M'y{Cv-xc) (6.4.16)

having used (6.4.7) and (6.4.10). By substituting t h e integral (6.4.16) in


(6.4.12) we obtain

W2; = j [cx M'x 09' + Cy Ml 09' + M'x 0 u' - M'y0 v') dz (6.4.17)

which expresses the second-order part of the work of the tangential stresses.
Adding up the two contributions (6.4.9) and (6.4.17) we have

= - f ft{u'2+v'2+r2c0'2+2u'0'yc-2v,0'xc) dz

+ / [CX{MX 0)'0' + Cy{My 0)' 0' + u'{Mx 0)' - v'{My 0)'} dz


(6.4.18)

Once W2 and W2* have been determined it remains to express the third term
V2 of (6.4.1), which is the second-order term of the Taylor series expansion of
the work of the external loads. The loads acting in the axial direction make
a work which is linear in wp and therefore, from (6.3.6), linear in u, v, and
0, and so do not appear in V2- If the lateral loads are applied t o the axis of
the twisting centres we have V2 = 0? m t h a t the work is still a linear function
of the displacements u and v of the axis. If, on the other hand, the loads are
237

applied at a point which is not the twisting centre, the work is a non-linear
function of the rotation of the cross-section 0(z).
Let us suppose that the load q(z), of components qx(z) and qy(z), is applied
to a line which is parallel to the beam axis and fixed with the beam, and let
Q = {xqiVq) be the intersection of the line with the plane of the cross-section
identified by the abscissa z. Let us assume t h a t the load direction passes
through the twisting centre C (Fig. 6.15).
Let
Ox — »Eq A>c
(6.4.19)
K = Vq- Vc

be the components of the distance vector CQ on the axes. From the ef-
fect of the rotation of the cross-section around C, the point Q undergoes a
displacement of components

\iq = p [cos(a + 0) - cos a] = - by 0 bx 02 + o(02)


Li
(6.4.20)
2 2
vq = />[sin(a + 0) - s i n a] =bx0- - bv0 + o{0 )

where p and a are shown in Fig. 6.15 and having expressed the circular
functions sin(a + 0), cos(a + 0) as series expansions of 0 with initial point
0 = 0. The second-order part of the work of the loads is therefore

V 2 = - " j(qxbx+qyby)02 dz (6.4.21)

Fig. 6.15 - Thin-walled beam subjected to lateral pressure.


238

It is worth observing that, occasionally, M2 = 0 is written in the literature


and the term relative to the loads is determined as the work of the stresses
S^ in the fundamental state W2* [7,9]. The procedure involves evaluation
of the stresses cxx and ayy and therefore calculation of the corresponding
deformations of the second order. The result arrived at is the same.
Once the three terms W2, W2* a n d ^2 have been expressed as a function of
the unknown vector u = {u(z),v(z),0(z)}, it is possible to give an explicit
form to the functional (6.4.1). By substituting (6.4.5), (6.4.18) and (6.4.21)
we have

ip<X = i I (EIxv"2 +EIyu"2 + ET0"2 + GJ0'2)dz

-— [ (u'2 + v'2 + r\ 0'2 + 2u'0'yc -2v'0'xc) dz

+j \CX{MX ey & + cv(My ey & + U'(MX ey - V'{MV ey\ dz

+ - f d2 {qx bx + qy by) dz (6.4.22)

By requiring its first variation to vanish we obtain

P^uSu = [ {EIxv"6v"+ EIvu"6u" + ET6"60"

+ 6u'[-N{u' + yce') + {Mxey}

+ 6 v' [- N{v' - xj') - (My 0)'}

+ 66,[GJ0' - N(r\ ff + ycu'- xe v') + CX{MX 0)'

+ CX MX 6' + Cy(My 0)' + Cy My ff + u' M, ~ V' My\

+ 60 [Cx0'M'x + Cy0'M'y + M'xu' - M'yv'

+ {Qx bx + qy by) 0] } dz = 0 (6.4.23)


Integrating by parts and imposing the condition t h a t (6.4.23) is satisfied for
every kinematically admissible 6u, we obtain the three equations
E Iy u"" + N{u" + yc 0") - (Mx 0)" = 0

E Ix v"" + N(v" - xc 0") + [My 0)" = 0

E T 0"" + N{r2c 0" + yc u" - xc v") - CX(MX 0)" - Cx Mx 0" (6.4.24)


+ Cy [My 0)" -CyMy0"-GJ O" ~ MX U" + My V"
+ [qx K + qyby)0 =O
239

and the twelve boundary conditions in z = 0 , I

EIyu"6u' = 0, EIxv"6v' = 0, ETO"60' =O

[E Iy u'" + N(u' + yc 6') -{Mx0)'}6u =O

[E Ix v'" + N(v' - xc 6') + {My 6)'} 6v = 0 (6.4.25)

[E T 0'" + N{r2c ff + ycu'- xc v1) - CX{MX 0)> - Cx Mx ff


- Cv{My6)'-CyMy6' -GJ6' -Mxu' + Myv'}6$ = 0

The differential system (6.4.24), (6.4.25) constitutes an eigenvalue problem


which allows calculation of the critical load and corresponding buckling mode.
Some applications will be presented in the following.

6.5 THIN-WALLED BEAMS SUBJECTED TO UNI-


FORM COMPRESSION
Let us consider a thin-walled beam loaded by an axial compressive force acting
on the centroid, constant along the axis of the beam. We wish to determine
the value of the load under which we reach the critical equilibrium condition.
From the general equations (6.4.24) and (6.4.25), taking Mx = Mv — qx —
qy = 0, we get

EIyu"" + Nu" + Nyc6" = 0


E Ix v"" + N v" - N xc 6" = 0 (6.5.1)
ET0"" - {G J - N r2c) 6" + N ycu" - N xcv" =0

with the boundary conditions in z — 0 , I

EIyu"6u' = 0 , EIxv"6v' = 0 , ET0"66' = 0


{EIyu'" + N{u' + yc0')}6u =O
(6.5.2)
[EIxvm + N{v' -xc0')}6v =O
[ET ff" -GJff + N(r] 0' + ycu' - xc v')} 60 = 0

Equations (J6.5.1) and (6.5.2) are immediately solved in the particularly


simple case in which the end section rotations around the axis z are prevented
while rotations around the axes x and y as well as warping are not constrained.
The geometrical boundary conditions are in this case

u = v = 0= 0 in z = 0,1 (6.5.3)
240

from which follow, through (6.5.2), the natural conditions

E Iv u" = 0 , E Ix v" = 0 , E T 0" = 0 in z = 0, t (6.5.4)

Equations (6.5.1) and boundary conditions (6.5.4) are susceptible to me-


chanical interpretation; the first two equations express conditions of equilib-
rium in the x and y directions in a configuration which is very close to the
fundamental configuration, the third equation is a condition of equilibrium
for rotation around the twisting centre. Consequently, the natural boundary
conditions are intended as mechanical boundary conditions which, in cor-
respondence with the constraints, impose the vanishing of the two bending
moments and a third quantity, the bi-moment, associated with the warping
deformations.
The three functions
7T Z 7T Z 7T2
u = Au sin — , v = Av sin 0 = Aa sin (6.5.5)

satisfy the boundary conditions (6.5.3) and (6.5.4). By substituting (6.5.5)


in (6.5.1), all the derivatives being of even order, it is possible to factorise the
term sin nz/l. By imposing the condition that the system is satisfied for any
value of z we obtain

( r T \

** -ef-N)A»-Ny<Ae = ° (6.5.6)
(TT2 * ^ - N ^ A V + NxcAe =0
-NycAu + NxcAv+[7r2^- + GJ-NrfjAe =0

which is a homogeneous system of algebraic equations in the unknown quan-


tities Au, Au, Af. Let us write

EIX
Nx = n*
£2
N
1 v
= 7 T 2

£2
^ Ns = Pf+Oj) (6.5.7)
where Nx and Ny are Eulerian critical loads corresponding to lateral deflection
around the axes x and y and N$ is a purely torsional critical load, as will be
shown later. With the relations (6.5.7), system (6.5.6) is rewritten as

>
f Ny-N
Nx-N
■ -Nyc
Nxc
( Au
) ( 0°] (6.5.8)
Av =
Nxc A
V N
~ Vc rl(Nt - N) j { * J ^ J
241

f(N)

— * , 1 1 ^.

Nc\ NX Ny N^ N

Fig. 6.16 - Representation of the characteristic equation.

In order that the system has a non-trivial solution, it must be

2 2
{Ny - N){NX - N){Ne - N) - N2 V-\ [Nx - N) - N2 ^ {Ny - N) = 0
r r
c c
(6.5.9)

which is obtained by requiring the determinant of the coefficients to vanish.


The cubic (6.5.9) is the characteristic equation of the eigenvalue problem:
the smallest root TV which satisfies it is the critical load. This equation is
of the type f(N) — 0. Let us plot the function / , supposing for example
t h a t Nx < Ny < Ne (Fig. 6.16). For N = 0 we have / > 0; for N = Nx,
/ < 0. Since f{N) is a continuous function, it vanishes for a value of Nc <
min{7Vx, iVy, A^}, that is, for a value of the load which is smaller than the
smallest of the three critical loads (6.5.7). It is possible to verify by means of
an analogous procedure that such a property is general, and does not depend
on the values of Nx, Ny, N$. It can therefore be concluded t h a t the flexural-
torsional coupling leads to a reduction of the critical load if compared with
the Eulerian critical load, with a consequent dangerous overestimate of safety
if the analysis if conducted with reference to the plane beam model.
Once the first eigenvalue Nc is determined from (6.5.9), equations (6.5.8)
furnish the corresponding eigenvector a = {Au, Av, A$} from which we can
calculate, by means of (6.5.5), the buckling mode u = {u, v, 0}. In general,
all three components of u are different from zero, which shows how the beam
deflects on a plane which does not coincide with a principal plane of inertia,
and rotates around the axis of the twisting centres.

Let us now examine two particular cases which frequently occur in appli-
cations.
242

(a) Cross-section with two axes of symmetry


The twisting centre coincides with the centroid and is therefore xc = yc = 0.
The characteristic equation (6.5.9) in this case has the roots

Ni = Nv, N2 = Nx , N3 = N$ (6.5.10)

and the critical load corresponds to the smallest of the three critical loads
(6.5.7). The system of equations (6.5.8), the matrix of coefficients of which
is diagonal, furnishes the three normalised eigenvectors

ai = { 1 , 0 , 0 } , a2 = { 0 , 1 , 0 } , a3 = { 0 , 0 , 1 } (6.5.11)

corresponding to the three eigenvalues. The problem is completely uncoupled,


for the thin-walled beam buckles by deflecting in a principal plane of inertia
if N is equal to one of the two Eulerian critical loads, or by rotating around
the centreline for a load equal to the torsional critical load N$.

(b) Cross-sections with only one axis of symmetry


This is the more commonly used case of thin-walled beams. If, for example,
y is the axis of symmetry, then xc — 0 and the characteristic equation (6.5.9)
becomes
,,2
{Nx - N) {Nv - N){Ne - N) - N 2 Vc
= 0 (6.5.12)

This has the solutions


1
Nx {Ny + N6) T ^{Ny + Ne)2-4KNyN$
(6.5.13)
N2 = Nx
in which

K = 1 (6.5.14)

is the flexural-torsional coupling factor. The smaller of the two roots (6.5.13),
the first being assumed with a negative sign, is the critical load Nc. If Nc =
7V2 = 7VZ, then the second equation of (6.5.8) is identically satisfied. To
satisfy the other two, it must necessarily be that a 2 = { 0 , 1 , 0 } , that is, the
mode consists of a deflection in the plane of symmetry (y,z). If Nc = JVl5 we
get Av = 0 from the second equation (6.5.8) and Au/A$ = Niyc/(Ny — N\)
from the first. The eigenvector is then a x = {Niyc/(Ny — Ni),0,1}, having
normalised according to A$ = 1, and the buckling mode consists of a deflection
in the plane (x,z) and of a rotation.
The critical load iV\ is always smaller than both Ny and N$ and the reduc-
tion is t h a t much greater because the coupling factor K is smaller. Figure
243

k/
NV
/A.O N,= N Y

1.6

ST V-™

Fig. 6.17 - Critical load of thin-walled beams with an axis of symmetry.

6.17 [3] shows the dependence of Ni/Ny on Ne/Ny for different values of l/K.
For K — 1, t h a t is for a cross-section which is doubly symmetrical, we have
Ni = min{iV y , Ne}, and we return to the case already discussed. For de-
creasing K and constant values of Ny and Ne, Ni diminishes, that is, the
critical load is greatly affected by interaction. If however N$ and Ny differ
greatly, t h a t is, if we have Ne <C Ny or No ^> Nyj then N\ is about equal
to the smaller of the two, whatever the value of K\ in this case there is no
interaction between bending and torsion.
Let us consider, for example, a thin-walled beam consisting of a cylindrical
tube cut along a generatrix (Fig. 6.18). The table in Section 6.4 shows the
warping constant V and the position of the twisting centre for an arc of a circle
with a semi-opening angle a. For a = 7r we obtain T = (27r 2 /3 — 4)7riZ5/i,
xc — 0, yc = — 2R. The torsional moment of inertia is calculated by using
(6.4.6), and we obtain J = 2TTRII3/3. The radius of gyration relative to the
twisting centre is given by r\ — R2 + (2R)2 — 5i? 2 . The coupling factor, in
accordance with (6.5.14), is K = 1/5.
Knowing the geometrical properties, it is possible to calculate from (6.5.7)
the three critical loads NT Nv,Ne

hRz
AT, Nv = Ttz E
£2
(6.5.15)
R3 1 t2h2
Ne = v'Eh —
iG--«) + 2
15(1+1/) 7T R4 \

where v is the Poisson ratio. For h/R = 0.1, t/R — 100 (slenderness of
the beam) and v — 0.3 we have N$/Ny = 1.035, t h a t is, the three critical
loads are about equal. Note that, due to the particular form of the cross-
244

C(0,-2R)

Fig. 6.18 - Open circular cross-section.

section, the contribution to the critical load N$ from the warping rigidity ET
is considerable. In the numerical example in question this represents about
50% of the resulting value.
Once the critical loads and the coupling factor are known, the first of
relations (6.5.13) yields Ni/Ny = 0.537. The bending-torsion interaction
therefore leads to a drastic reduction in the load with respect to the Euler
critical load, equal to about 50%.
Let us now examine the buckling mode. The eigenvector which corresponds
to the eigenvalue JVX is

Nl/Ny
ai = yc , 0 , l \ = { - 2.320 R ,0,1} (6.5.16)
- Ni/Ny

from which we can see that the buckling mode consists of a deflection in the
plane (x, z) and a rotation around the twisting centre. All the points belong-
ing to the axis of symmetry move in the x direction and the displacements
result from the first of equations (6.3.1) and (6.5.5)

u p - [Au-Ae{y-yc)}s'm ^ = - [2.32 i? + (y + 2R)] sin ^ (6.5.17)

Imposing up = 0 we find y = — 4.32 R, which identifies the position of the


axis of the twisting centres.
245

6.6 THIN-WALLED BEAMS SUBJECTED TO EC-


CENTRIC AXIAL LOADS
In the previous Section we examined the case of a thin-walled beam subjected
to axial load applied at the centroid. We now wish to generalise the problem
by considering the case of eccentric forces both of compression and of tension.
The general equations (6.4.24), (6.4.25) are specialised in this case by setting
Mx — Nev, My = Nex, qx = qv = 0, where ex and e„ are the coordinates
of the point of application of the load. We obtain the following differential
equations
E Iy u"" + Nu" + N{yc - e„) 0" = 0
E Ix v"" + N v" - N(xc - ex) 0" = 0
(6.6.1)
E T 0"" - (G J + 2 N Cxev + 2 N Cy ex - N r\) 6"
+N{yc - ev) u" - N{xc - ex) v" = 0
with the boundary conditions
E Iy u" 6 u' = 0 , EIxv"6v' =0 , E T 0" 6 0' = 0
[E Iy u'" + Nu' + N{yc - ey) 0'}6u = O
\E Ix v'" + N v' - N{xc - ex) 0'} 6 v = 0 (6.6.2)
[E T 0'" -GJ0' + N{r2c 0' - 2 CX ey 0' - 2 Cv ex ff)
+N{yc - ey) u' - N{xc - ex) v')66 = 0
to be imposed at the two ends z = 0 and z = I.
Equations (6.6.1) and (6.6.2) are easily solved for the boundary conditions
previously discussed. From the geometrical boundary conditions (6.5.3) the
mechanical boundary conditions (6.5.4) again follow. By putting the solution
in the form (6.5.5) and following the same procedure as before, we arrive at
the following algebraic system
/ N„-N - N(yc - e v )
( Au \ ( 0 \
Nx-N N(xc - ex)
Av 0
r*e(N, -N) +
- N{yc - e„) N(xc - ex)
2 N{CX ty + Cy ex)
(6.6.3)

in which Nx, Ny, Ng are given by (6.5.7). By requiring the determinant of


the matrix of coefficients to vanish, we obtain the characteristic equation
(Nv - N){NX - N) [r2c(Ne -N) + 2N(CX ey + Cy exj\

- N2 [Nx - N){yc -ty)2- N2 [Nv - N){xc - ex)2 = 0 (6.6.4)


246

With the same reasoning as that made previously, we can verify t h a t


Nc <mm{Nx,NViN$}.
Let us now examine some particularly noteworthy cases.

(a) Axial load applied on the twisting centre


In this case we have ex = x c , ey = yc and all the terms of the matrix of
system (6.6.3) off the principal diagonal vanish. The equations are uncoupled
and furnish the three eigenvalues

».=*.. * = * . , * . - , , _ , 4 * , ^ <•*•»>
to which the three eigenvectors (6.5.11) correspond. The buckling mode con-
sists of either a simple pure bending or a pure rotation around the axis of the
twisting centres. The case treated generalises that of a compressed centrally
loaded thin-walled beam of double axis of symmetry, in which uncoupling
of the modes occurs when the point of application of the load, the twisting
centre and the centroid coincide. Finally, we note that whilst the eigenvalues
N\ and N2 are positive, N$ can be either positive or negative. This demon-
strates that buckling may occur by pure torsion even if N is a force of traction
(tension).

(b) Cross-sections with an axis of symmetry and external load applied on the
axis
Let this axis, for example, be y. We therefore have xc = Cy = 0 for
the symmetry and ex — 0 for the hypothesis on external load position. The
characteristic equation is simplified to

(Nx - N) {(Ny - N) [r2c(Nd - N) + 2NCxey] - N2 (yc - ey)2} = 0


(6.6.6)

from which we have Nc = Nx or f(N) — 0, where f(N) is the term of (6.6.6)


in braces. Correspondingly, the buckling mode consists of a deflection in the
plane of symmetry or of a deflection accompanied by rotation.

(c) Cross-sections with two axes of symmetry and external load applied on
one of the axes

In this case xc — yc = Cx — Cy — 0 for the double symmetry; also ex = 0


if the external load is applied to a point of the axis y. The characteristic
equation is

{Nx - N) [r2c {Nv - N) {Ne - N) - N2 ej] = 0 (6.6.7)


247

and the critical load either coincides with Nx or is a root of the quadratic
equation

^ 2 f1 - ^ ) - N
(Nv + N*) +NyNe=0 (6.6.8)

In the first case we have pure bending, in the second torsion and bending as
in the case of a single axis of symmetry. If also ev — 0, t h a t is, the load is
applied to the centroid, then we have a case which has already been treated.
A simple examination of (6.6.8) shows the possibility of negative N roots.
In fact, the coefficient of the term linear in N is negative, the known term is
positive, and the coefficient of the quadratic term in TV can be either positive
or negative, according to whether the eccentricity ey is smaller or larger than
the radius of gyration r c . From the Descartes rule of signs we deduce that if
the eccentricity is small then the roots of (6.6.8) are both positive, and the
critical conditions can be reached only if the external load is compressive;
for large values of eccentricity, instead, one root is negative and the other
positive, and critical equilibrium occurs as much for a force of traction as for
compression. We remark that, under the action of a force of traction, we can
reach the critical equilibrium only if the eccentricity of the force is sufficiently
large. If, in fact, the point of application is inside the core of the section, we
have tensile stress only and the fundamental configuration of equilibrium is
stable. If, instead, the point is outside the core, then we have compression
over part of the section, which is greater the larger the eccentricity. We can
therefore understand that there exists a minimum eccentricity below which
the critical condition of equilibrium does not occur whatever the intensity of
traction TV, and above which, instead, it is satisfied by an ever-decreasing load.
Such a minimum eccentricity describes, in the plane of the section, a closed
curve which contains a stable zone. If the load is applied inside the region
of stability, critical equilibrium can occur only for forces of compression; if
the load is applied outside, we may have critical equilibrium also for forces of
traction.
The stable region can be determined by imposing the condition that, when
the force of traction is acting on its boundary, the critical load becomes in-
finite. With reference, then, to the general case (6.6.4), dividing the charac-
teristic equation by TV3, we obtain

(£-)(f-)KfH-< c —']
- ( ^ - i) (»• - ' . ) ' - ( f - 0 (*. - '■)' = o C-6-9)
and passing to the limit for N —> oo we have
- (r\ - 2 Cx ev - 2 Cy ex) + {yc - evf + {xc - exf = 0 (6.6.10)
248

By writing (6.6.10) in the form

e\ + e\ - 2 [ex(xc - Cv) + ey(yc - Cx)\ = r2 (6.6.11)

in which r\ — r\ — {x\ + y2c) = (Ix + Iv)/A is the polar radius of gyration


with respect to the centroid of the cross-section, we can see t h a t the region
of stability is a circle of radius p given by

P = sjrl + (xc - C v )» + (yc - Cxy (6.6.12)

and centre Q, of coordinates

x q = xc - C y , yq = yc- Cx (6.6.13)
As an example, we again consider the cross-section shown in Fig. 6.18.
From the definitions (6.4.10), as x2 -\- y2 — R2 = constant, we have Cx —
yc = —2J?, Cy = xc = 0. Equations (6.6.13) give xq = yq = 0, t h a t is the
centre Q of the region of stability coincides with the centroid of the section;
equation (6.6.12) instead gives p — r 0 = R, which shows that the region of
stability coincides with the cross-section itself. For forces of traction acting
outside the cross-section, the equilibrium of the fundamental configuration
can become critical. For example, if the force is applied at the twisting centre,
remembering that r\ — 5i2 2 , we have from the third of relations (6.6.5)

iV3 = - H N9 (6.6.14)

If the force is of compression, then the critical load is given by (6.5.15), if the
force is of traction, the critical load is furnished by (6.6.14).

6.7 THIN-WALLED BEAMS UNDER P U R E BEND-


ING
Let us consider a thin-walled beam loaded at its ends by a bending moment
of components Mx and My on the principal axes of inertia. The differential
equations of critical equilibrium and the relative boundary conditions can be
deduced from (6.4.24), (6.4.25) by putting Mx = constant, Mv — constant,
N = qx — qy = 0. Here, however, we limit ourselves to considering the par-
ticular case of boundary conditions expressed by (6.5.4) which have already
been treated in Sections 6.5 and 6.6. The characteristic equation can then be
obtained directly from that relating to the problem of eccentric load by let-
ting the axial force TV tend to zero and the eccentricities ex and ey to infinity,
and then requiring that

lim Nex = My , lim TV ey = Mx (6.7.1)


N-+0 AT—0
ex—♦oo ey—'00
249

By imposing the limit conditions (6.7.1) on system (6.6.3) we obtain


Mx \ I \ ( 0 \
Nx -M„ Av 0 (6.7.2)
\ MZ -My rlNB + 2CxMx + 2CyMy J A6
to which the characteristic equation

(2 Cx Mx + 2 Cy My + r\ Nt) Nx Ny - Nx M\ NyM2y=0 (6.7.3)


corresponds.
Equation (6.7.3) describes an ellipse on the plane Mx,My, with axes par-
allel to the straight lines Mx — 0, My — 0. The region bounded by the ellipse
represents a domain of stability. Points which lie inside this area correspond
to a stable fundamental configuration, points on the boundary are represen-
tative of the critical state and points outside represent states of generalised
stresses for which the fundamental configuration is unstable. Knowing the
law of variation of Mx and My, the point of intersection of the corresponding
curve with the ellipse of stability furnishes the critical condition of equilib-
rium. If Mx and My increase proportionally with a parameter A starting from
zero, then the critical load can be determined by putting Mx = A Mx and
My = A My, where Mx and My are the coordinates of a point on a straight
line through the origin. Substituting in (6.7.3) and solving the quadratic
equation in A, the critical load multiplier A,, of the bending moments Mx and
My is obtained as the smallest root of the equation.
The case of simple bending is obtained by writing separately My = 0 and
Mx = 0 in (6.7.2) and (6.7.3). We thus obtain four critical values

Mxc = CXNV± y/{Cx Nyy + r\ Nv Ne


(6.7.4)
Myc = CVNX± yJ{Cv Nx)* + r\ Nx Ne
for positive and negative bending moments. The corresponding buckling
modes consist of a torsion associated with bending in which the beam axis
deflects in the direction of the moment vector. If, in particular, the cross-
section is symmetrical with respect to the x or y axes, then Cx — 0 and
Cy — 0, respectively, and the positive and negative critical moments coincide.
Equations (6.7.4) yield, using (6.5.7),

7T
Mx = ± EIy[-ET + GJ
M (6.7.5)
■K
MLye = ± EIX ET + GJ
vc
~n P
250

If the warping rigidity ET vanishes, or is negligible, then equations (6.7.5)


are reduced to
Mxc = ±j yjEIv G~7 , Myc= ±jy/EIxGJ (6.7.6)

Equations (6.7.6) provide the Prandtl critical loads.


We finally observe that, depending on the coefficients Cx and Cy, positive
and negative critical moments can be very different. If we again consider the
beam in Fig. 6.18 we have, from the first of equations (6.7.4),

Mxc = Nu Cm±slCl + tj± = NVR 2 ± W 4 + 5


(6.7.7)
^ J

as Cx = yc = — 2R and r\ — 5R2. For Ne/Ny ~ 1, as in the numerical example


in Section 6.5, we have
Mxc~(-2±3)NyR (6.7.8)
which shows that there is a 1:5 ratio between the critical moment which
compresses the fibres near to the longitudinal cut, and the opposite critical
moment.

6.8 FLEXURAL-TORSIONAL STABILITY OF


BEAMS SUBJECTED TO LATERAL LOADS
Study of the stability of thin-walled beams subjected to simple bending has
shown how the critical value of the bending moment depends on the torsional
properties of the cross-section, as well as on the moment of inertia with re-
spect t o the axis which lies in the plane of the couple. If the cross-section
has the shape of a thin rectangle or an I section with narrow flanges, t h e
two principal moments of inertia are very different from each other and the
torsional rigidity is very small. If, therefore, the beam is bent in the plane of
maximum rigidity, then the fundamental configuration can become unstable
for very small values of the moments and the beam may find new positions
of equilibrium by rotating and bending outside the plane of the couples. An
analogous behaviour is observed if the beam is subjected t o a lateral load
acting in the direction of maximum rigidity of the beam. The phenomenon is
very dangerous, and requires adequate devices such as braces and stiffeners to
reduce the spans of lateral deflection and to oppose the rotation of the beam
sections. Here we limit ourselves to formulation of the general equations of
the problem, to solving some simple cases and furnishing a limited number
of technically interesting results, leaving the numerous and more particular
cases to specialised texts [1-4,6,7,9,13].
Let us consider a beam which is subjected to lateral loads and bent, in the
fundamental configuration, in the principal plane of inertia (t/,2). Let qv(z)
be the load applied to an axis in the plane of bending at a distance by from
251

ku)

bv
Y

O} -
Fig. 6.19 - Thin-walled beam subjected to lateral loads.

the twisting centre (Fig. 6.19). T h e differential system which furnishes the
critical equilibrium condition is deduced from the general equations (6.4.24)
and (6.4.25), applied to the specific problem under examination. By writing
N — My — qx = 0 we obtain a differential equation in v[z) with boundary
conditions in z — 0,£
EIxv"" = 0 (6.8.1)

E Ix v" 6 v' = 0 , EIX v'" 6 v = 0 (6.8.2)

and a system of two differential equations in u(z) and 0(z)


Elyu"" - (MZ0)" = O
(6.8.3)
ET6"" -GJ0" - CX{MX 0)" - CxMx 0" - Mx u" + qyby0 0
with the boundary conditions
EIyu"6u' =0 , ET0"60' =0
[EIy u'" - [Mx 0)'} 6u = 0 (6.8.4)
[E T 0'" -GJ0'-Cx(Mx0)'- CX Mx 0' - Mx u'} 6 0 = 0
to be imposed at the ends z = 0, z = I.
Note t h a t the differential system in the three unknowns u ( z ) , v(z), 0{z) is
uncoupled since the equation in v(z) and the relative boundary conditions are
independent of the other equations. T h e problem in (6.8.1), (6.8.2) allows as
the only solution the trivial one v = 0. T h e significant variables are therefore
only the u(z) and 0(z) which can be deduced from the differential system
(6.8.3), (6.8.4). It is appropriate to note that, unlike the cases treated so far,
the system in general is one with variable coefficients, as the moment Mx is
252

variable with z. This implies that, even in very simple cases, the search for a
solution can present great difficulties. If we have qy = 0, Mx = constant we
return to the case of pure bending already treated in the previous Section.
We wish finally to observe that in solving a problem or in applying the
results furnished by handbooks, it is always necessary to consider correctly
the position of the point of application of the lateral load; indeed, the value
of the critical load varies with this position.
Figure 6.20 shows the cross-section of a beam on which a load qv(z) acts,
applied at three different points: (a) at the twisting centre, (b) at the intrados
and (c) at the extrados of the cross-section. If we interpret the equations
(6.8.3), (6.8.4) as conditions of equilibrium in a configuration adjacent to the
fundamental configuration, we can understand how in case (a) the load qy
does not contribute to the equilibrium to rotation around the twisting centre
C, whilst the effect is stabilising in case (b) and destabilising in case (c).
Under the same conditions, the critical load in case (c) is the smallest of the
three and becomes smaller as by increases.
As an application of equations (6.8.3), (6.8.4) let us consider the cantilever
in Fig. 6.21, loaded by a force F applied at the twisting centre of the end
section. If the section is symmetrical with respect to the axis x and the
warping rigidity is zero or negligible (narrow rectangular section, cruciform
section), then we have Cx = 0, T = 0. The differential equations (6.8.3),
bearing in mind that qy = 0, are simplified to
Elyu""- {Mx6)" = 0
(6.8.5)

bY-0 by<0

a) c)

Fig. 6.20 - Lateral load applied at the twisting centre, at the intrados and
extrados of a thin-walled cross-section.
253

Fig. 6.21 - Cantilever subjected to bending and shear.

and the boundary conditions (6.8.4) for z = 0,1 become

E Iy u" 6 u' = 0

[EIyUm-{Mt0)'}6u =O (6.8.6)

[GJ0' + Mxu'}66 = 0

being

Mx = -F{i-z) (6.8.7)

Note that, in the case in which the force F is applied at a point different
from the twisting centre, the third of relations (6.8.6), written for z = I, is
modified by the addition of a corrective term. This term is easily obtained
by adding the contribution of the second-order part of the work of the force
F to the functional (6.4.21).
Returning to the problem under examination and imposing the geometrical
boundary conditions

u(0) = 0 , u'(0) = 0 , 0(0) = 0 (6.8.8)

we have from (6.8.6) the natural boundary conditions

EIyu"{i) =0
EIvum{l)-\Mx(l)6{l))' = 0 (6.8.9)

GJ0'{L) = 0

We now integrate the first of the two equations (6.8.5) twice. Integrating
the first time, we have

Ei„um-{Mxey = cl (6.8.10)
254

where cx is an arbitrary constant. By writing (6.8.10) for z — t and taking


into account the second of conditions (6.8.9), we get C\ = 0. Integrating again
we have

EIvu" -Mx0 = c2 (6.8.11)

where c 2 is another arbitrary constant. By evaluating (6.8.11) at z = £, from


the first of (6.8.9) and from (6.8.7), c2 = 0 follows. Specifying Mx by means
of (6.8.7), we get from (6.8.11)

u"= -JT{1-Z)e (6.8.12)

By substituting this expression in the second of equations (6.8.5) we obtain

0 ,+ (6 8 13)
' GTEL{i^z)2e = O
- '
which is a differential equation of the second order in the only variable 0(z).
By solving (6.8.13) with the boundary conditions furnished by the third of
equations (6.8.8) and (6.8.9), we determine the eigenvalue F and the eigen-
function 0. Finally, by integrating (6.8.12) twice with the first two boundary
conditions (6.8.8), we can also determine u(z)y thus completing the solution
to the problem.
To solve (6.8.13) it is convenient to introduce the change of variable

< = ' - ! (6.8.14)

anLd rewrite the equation in the form

£ + ^-o (6.8.15)

in which

A - FH* (6.8.16)
' GJ Ely
The third of relations (6.8.8) and (6.8.9) become in the new variable £

~d6
0(1) = 0 , - 0 (6.8.17)
d£J £ = 0
We note t h a t (6.8.15) is a Bessel equation which can be easily solved by
expressing the unknown function 0 as a series of powers of the independent
variable £
CO

*(0 = £ A . £ B (6-8.18)
n=0
255

where An are unknown coefficients. By substituting in the differential equa-


tion, we have
oo

J2 [n(n - 1) An C~2 + XAn £"+2] = 0 (6.8.19)

Equating to zero the coefficients of the generic power k of £, we obtain

[k + 1) {k + 2) Ak+2 + A Ak-2 = 0 (6.8.20)


from which we have the relationship between the coefficients

^ = ~ (k + l)(k + 2) A
" (6-8-21)
From (6.8.18) and (6.8.21) it is possible to obtain two independent 0(f) solu-
tions. The general solution is the sum of the two

■ A ( C ^ _ £5 , _ ^ ^ _ £9 * ^_ A 13
1
V* 4-5 ^ 4-58-9^ 4 - 5 8 - 9 12 • 13 ^
(6.8.22)

where AQ and A\ are two arbitrary constants. By imposing the boundary


conditions (6.8.17), we obtain from [dO/d£\z=o = 0

Ax = 0 (6.8.23)
and from 0(1) — 0 we have

+ 7 — r ^ - ^ - ^ - 7 ^ r ^ TV - T7 + ' • • = 0 (6.8.24)
3-4 3-47-8 3 - 4 7 - 8 11-12
where (6.8.24) is the characteristic equation. By truncating the series at
the second term, we obtain A = 12 as a first approximation of the lowest
root; truncating at the third term we find A ~ 17.42, and so on. The series
converges to the exact value Ac = 16.104. The critical value of F is therefore,
from (6.8.16),

JGJ EIy
v
Fc = 4.013 -* - (6.8.25)
l
l
The eigenfunction 0(f) is obtained from (6.8.22) for A± = 0 and A = Ac; by
integration we then determine u ( f ) .
In the problem treated it is supposed that, in addition to the symmetry
of the cross-section, the warping rigidity is negligible. This hypothesis has
led us to a lower order system of differential equations and, consequently, to
a more simple solution. The case f / 0 for symmetrical cross-sections has
256

been solved by Timoshenko [3]. The critical load is again put in the form

SJGJ EIy
Fc=n (6.8.26)
I?
of the type (6.8.25). In this, however, 7X is a function of the ratio i.2GJ/ET
which decreases monotonically with increase of the ratio, until it reaches the
value previously calculated in (6.8.25). The following table shows some values
of 7i [3].

TABLE I - Values of 7X in the formula (6.8.26)

p2 GJ 0.1 1 2 3 4 6 8
L
ET

1\ 44.3 15.7 12.2 10.7 9.76 8.69 8.03

p2 GJ 10 12 14 16 24 32 40
C
ET

11 7.58 7.20 6.96 6.73 6.19 5.87 5.64

Finally, we report some results for problems which are frequently met in
applications [3],[7].

(a) Simply supported beam of I section loaded at mid-span (Fig. 6.22).

The critical value of the load is given by

yjGJEIy
Fe=l2 (6.8.27)
I2

111 1 111 111TTT1

WMk

Fig. 6.22 - Simply supported beam Fig. 6.23 - Simply supported beam
subjected to concentrated subjected to distributed
load. lateral load.
257

The diagram in Fig. 6.24 [7] shows the behaviour of the coefficient 72 as a
function of the non-dimensional ratio t2GJ/£T, for three different situations:
the load is applied at the twisting centre (a), at the intrados (b), or at the
extrados (c) of the cross-section. Furthermore, the curve (a) is valid for sec-
tions different from I sections, provided that they have an axis of symmetry.
An application presented in the following Section will serve to clear up the
question.

(b) Simply supported beam with I section subjected to uniform lateral load.
(Fig. 6.23)

We have

y/GJ EIy
M ) c = 73 (6.8.28)

The diagram in Fig. 6.25 shows the behaviour of 73 for different positions of
the load. As in the previous case, the curve (a) is also valid for cross-sections
with an axis of symmetry.

ii ii
/fej
FT
III IIn
II \
V u \

1
100 IN 100 \
1 ^
V
U \ \
\ K \ y
lr \\ \ \ \ b]

10
1 N
\\ } \a)
lb)
l
inl r« dos
10
\ >
\
\1
^
N\ nt ra dc»s

LC cente ^
\ 1
I K N
S.C er tc r 1 ex tra dc S
i
1
e xtracios
s l vs
s

0.1 1—^ 0.1 —^


0 10 20 30 40 50 60 70 y2 0 10 20 30 40 50 60 70 /3

Fig. 6.24 - Representation of 72 as a Fig. 6.25 - Representation of 73 as a


function of i2 GJ/ET. function of £2 GJ/ET.
258

(c) Beams with different boundary conditions, symmetrical cross-section and


no warping rigidity.
Figure 6.26 illustrates several cases: case 1 is the P r a n d t l problem already
discussed; case 2 refers to a beam with constraints at the ends which prevent
rotation and bending outside the plane; case 3 to a cantilever with uniform
lateral load applied at the axis of the centroids; cases 4 and 5 are deduced
from the previous (a) and (b) for r = 0; the last case 6, like case 2, again
refers to a beam with flexural and torsional clamps.

Vrr
D Mc =j [/Elv GJ

«n 23V
f /ET~GJ

12,85 ,
3)
r n T r m r r n q (q')c = "15—
I2 ]/^7GJ

16,94

¥
4) |/El y GJ
/2

28,3 , /_, ..
5) -jr T r-*-» T *-#—*~r T *-«r
_L J_

F=3
6) Fc = 26,6
^ /ii-ej
^

Fig. 6.26 - Critical load for different types of beam.

6.9 M E T H O D S OF D I S C R E T I S A T I O N
For all the problems discussed in this Chapter we have arrived at exact solu-
tions, in closed form, to the systems of differential equations which embody
the critical conditions of equilibrium. This has been possible because of the
simplicity of the cases treated, characterised by very particular conditions of
constraint, relating to single bay beams with regular load distribution. In
applications, however, it is common to meet problems for which we either do
not know the exact solution or which discourage, because of their complexity,
an exact analysis. It is then necessary to use approximation methods which
reduce the difficulty of solving the problem. Of these, we mention the methods
259

of discretisation which transform the original differential problem into a more


simple algebraic problem, reducing the infinite degrees of freedom of a system
to a finite number. Here we take up again two methods, the Ritz method
and the finite elements method already illustrated in previous chapters, to
calculate the critical load of open thin-walled beams.

(a) The Ritz method.


The discretisation of the system is obtained by restricting the class of func-
tions which describes all the kinematically admissible configurations into a
sub-class, in which the unknown functions are expressed as linear combina-
tions with unknown coefficients of known functions. Let us put
u(z)=<f>T(z)a, v(z) = T
X (z)b, 0(z) = tl>T(z)c (6.9.1)
z z
where the vectors <j){z), x[ )-> V>( ) each describe n functions defined over
the whole domain occupied by the system such as to satisfy the geometrical
boundary conditions, and the vectors a, b , c contain the unknown coefficients.
Equations (6.9.1) describe a 3n-dimensional space contained within the space
of admissible functions of dimension oo 3 . By substituting (6.9.1) in the total
potential energy functional, we obtain a function of the 3n parameters a, b ,
c. By equating to zero the first variation of the second-order terms we arrive
at an algebraic eigenvalue problem, whose solution provides the critical load
and the buckling mode. The eigenvalue and eigenvector found in this way
make te functional P Q ' U 2 / 2 positive semidefinite in the sub-space spanned by
(6.9.1), but not necessarily in the space of all admissible functions. Therefore
if the exact buckling mode can be expressed in the form (6.9.1) the solution
we reach is exact; otherwise, the critical load determined is an upper bound
of the correct value.
In the following, we again refer to the functional (6.4.22) to illustrate the
application of the method; more complex cases, such as a beam elastically
constrained at points or on part of the domain (Winckler type foundation)
can easily be treated by using the potential energy of the constraints.
By substituting (6.9.1) in (6.4.22) we have

ip<X = ^(£/2bTx,/X/'Tb + ^/yar0'>''Ta + ^ r c ^ , > ' , T


c

+ G J cV V' T c) d z - - [ ( a r <jJ / a + b T X' x'Tb

+ r\ cT V' V>'Tc + 2 yc a T <f>' 0'''c - 2 xc b T x ' V ^ c ) dz

+ j ^ \cT 4, \cx (Mx V> T )' + Cx (My V)'] c

+ a V (M x ^ r ) c - b V (M v i& T )'c} ^

+ - I {Qx K + qy by) cT V> 4>T c dz (6.9.2)


z J1
260

in which the integrals are extended to the whole beam. We can now define
the following square matrices of order n

K°u = J(EIv<t>"<t>"Tdz, KZ = ftEIxX"x"Tdz


(6.9.3)
T T
K°e = f (ETxl>"il>" + G J 4>'xl>' ) dz

and the matrices of the same order

KG _
- N jt<t>'<t>'T dz, KGV0 = -N j i X ' x ' T dz

K% = - N f V' ^ dz + 2 f V [Cx (M, fi>T)' + C„ (MV V>T)'l <*z

+ / [qx bx + qv by) ip xpT dz (6.9.4)

KS, = - N jyc <j>' i>'T dz + I 0 (Mx tl>T)'dz

KGe = - N j t xc X' V-'T dz - j t x' (My if)' dz

With the positions (6.9.3), (6.9.4) the potential energy can be written in a
matrix form

2
^ ' u = iqr(K 0 + AKG)q (6.9.5)

where A is the multiplier of the loads. The matrices or order 3n which appear
in (6.9.5)

( K° O O \
K„ = O K?„ o (6.9.6)
V o O K
J
( K?.. O K* A
K, O K K (6.9.7)
\ Ku# ~Kv6 ~Kee J

are called elastic and geometrical matrices, respectively, by analogy with the
corresponding energy terms; the vector of order 3n

T T T T
q=(a b c ) (6.9.8)

is the vector of Lagrange coordinates.


261

The critical condition of equilibrium is determined by imposing t h a t the


first variation vanish for each kinematically admissible displacement
P^uSu = 6qT ( K 0 + A K G ) q = 0 V <5q (6.9.9)
whence
(Ko + AKG)q = 0 (6.9.10)
Here (6.9.10) is a homogeneous system of algebraic equations; the smallest
root of the characteristic equation
d e t ( K 0 + AKG) = 0 (6.9.11)
is the critical multiplier of the loads Ac and the corresponding eigenvector,
given by (6.9.10), is the buckling mode.
As an application of the method illustrated we consider the system in
Fig. 6.27, already seen in Section 6.8, which consists of a simply supported
beam prevented from torsional rotation at the ends, subjected to a load q
uniformly distributed, applied along a line at distance b from the twisting
centre. The section of the beam is assumed symmetrical with respect to the
axis x, and therefore Cx — yc — 0. Since this is a problem of lateral stability
v(z) = 0 identically, as seen in Section 6.8, and it is therefore necessary to
approximate only the functions u(z) and 0(z). The corresponding algebraic
system is obtained from (6.9.10) by cancelling the n rows and n columns
relating to the b coordinates.
In the problem under examination we assume
3

G)'
u{z) = z a = 4>{z) a
(6.9.12)
0(z) =
(§r + c = ip{z) c

f f Y T T f T T T T T T T l

wz%.

Fig. 6.27 - Simply supported beam subjected to uniform load.


262

that is, the functions u(z) and 6{z) are approximated by means of a single
shape function where, in particular, (f)(z) = ip(z) is chosen. The problem
is reduced from oo 2 degrees of freedom to only two. Note that the shape
functions satisfy not only the geometrical boundary conditions (6.5.3) but
also the mechanical boundary conditions (6.5.4). The choice is particularly
appropriate, but not always easy to realise. In the most common cases, in
fact, even to satisfy only the geometrical boundary conditions can bring great
difficulties. From the positions (6.9.3), taking into account (6.9.12), we have
24 EL
K = f EIy<j)"2dZ =
(6.9.13)
2 2 24 ET 17 GJ
K< = / ( £ T ^ " + G J tl>' ) dz =

For evaluation of the geometrical terms it is necessary to determine the state


of stress in the fundamental configuration. We have
i2
M,{z) = q - My(z)=0 (6.9.14)

and from (6.9.4)

(6.9.15)

Kc =
2
- / ' a bib dz = qbl
H
Jo 630 H
having taken into account that qy = —q, according to the conventions in
Fig. 6.15. The eigenvalue problem (6.9.10) is obtained by substituting the
matrices (6.9.13), (6.9.15) in matrices K 0 and K G (6.9.6). We have after
summation of the matrices
/ 24 EL 11 „ \
/ 0 \
5 lz 210 *
(6.9.16)
11 24£T 17 GJ 31
qi
V 210 7 T5 ~~F
l z + bql
3535 I~T~ 630 V J
\oJ
By requiring the determinant of the coefficients to vanish and solving with
respect to the eigenvalue qi, we determine the critical load as

{qt)e = -43.04—n

ElyGJ ET \
1852.5(^*6)' 849.7 + 8397.2
+ PGJJ
(6.9.17)
263

By writing (q£)c in the form (6.8.28) we have

i EL b2
,3 = - 4 3 . 0 4 , / ^ -

EL b2 ET .
1852 5 849 7 + 8397 2 (6 9 18)
- ~GJI^ - - ¥GJ ' -
which shows the dependence of 73 on the two non-dimensional parameters
EIvb2/GJl2 and EY/t2GJ.
Let us consider some particular cases. If the load is applied at the twisting
centre, then 6 = 0, from which 73 = ±73 with

7£ = ^ 8 4 9 . 7 + 8397.2 -^j (6.9.19)

For T = 0 the approximate solution gives 73 = 29.15, whilst the exact solution
is 73 = 28.3, an error of 3% excess. For T ^ 0 the values given by (6.9.19)
have to be compared with those deduced from curve (a) of Fig. 6.25, valid for
any cross-section with an axis of symmetry. For example, for t2GJ/ET — 10,
we get 73 = 41.10 which is in very good agreement with the exact solution.
Note finally that the positive or negative sign of 73 depends on the fact that
the load is applied to the axis of symmetry, and therefore the positive and
negative critical values coincide.
We now examine the influence of the point of application of the load by
assuming b ^ 0 in (6.9.18). If the cross-section is rectangular and b is the
order of magnitude of the height of the cross-section, then EIyb2 /GJ£? <C 1
since Iy/J = 1/4. Therefore, by neglecting the first term in the second root
of (6.9.18) with respect to the other two, and bearing in mind t h a t r = 0, we
rewrite equation (6.9.18) as

73 = 73° ± 1 - 0.738 J - - (6.9.20)

If q > 0, that is, if the load is directed downwards in Fig. 6.27, then the
positive sign is taken in (6.9.20). If the point of application is above the
twisting centre (6 > 0), then 73 < 73 and the critical load decreases; if it is
below (6 < 0), it increases. The opposite occurs if the load is negative.
To evaluate the importance of this effect, we consider a narrow rectangular
cross-section in steel with E/G — 2(1 +v) — 2.6. For b/l = 1/20, for example,
(6.9.20) gives 73 = 73 ( ± 1 — 0.059), with a variation in the critical load of about
6% with respect to the case of load applied to the centroid.
A particularly noteworthy case is that of the I section with equal flanges
and with the load applied either to the extrados or the intrados. Since, in
264

this case, we have Iy ^> J , the approximation made for the rectangular cross-
section is no longer valid and it is necessary to make reference to (6.9.18).
Note that for the I section we have the identity Iy = 4T/h2 (see table in
Section 6.4), where h is the height of the cross-section. As b — ± / i / 2 , the two
non-dimensional parameters which appear in (6.9.18) are equal and therefore
we have

l3 = - 43.04 y - ^ j ± y 849.7 + 10249.7 - ^ (6.9.21)

This explains why the curves (b) and (c) in Fig. 6.25, like those in Fig. 6.24,
are valid only for I sections as, in this case, 73 is a function of the non-
dimensional parameter ET/E2GJ only. As an example we again consider the
case (?GJIET — 10; equation (6.9.21) gives the two values 73 = 29.69 and
73 = —56.91, very similar to those in Fig. 6.25. The differences between the
values of 73 and the value 73 = 41.10 are 27.8% and 38.5% respectively.

(b) Finite elements method.


This is basically analogous to the Ritz method, with which it has in com-
mon the description of the variables through a linear combination with un-
known coefficients of known functions. However, unlike the Ritz method, in-
stead of the shape functions (interpolation functions) being defined over the
whole domain of definition of the solution, they are defined independently
in sub-domains which are known as finite elements. The set of interpolation
functions provide a piecewise approximation of the variables which allows us
to overcome the difficulties of satisfying the geometrical boundary conditions.
This difficulty is one of the major limits to application of the Ritz method.
Let us consider the generic element c e' of a system of open thin-walled
beams (Fig. 6.28) [8]. Within the element the unknown coefficients of the
displacement functions u(z), v(z), 0(z) are expressed in terms of the values
which the variables and their first derivatives assume at the end sections
(element nodes). Let c/5t and t/>t (i = 1,2) be the rotations around x and y
of the nodes, indicated in the figure in the positive direction, related to the
displacements by the relations

<t>i = ~v'i9 ^ = - u\ , (* = 1, 2) (6.9.22)

where v\ and u\ are the derivatives of v(z) and u(z) calculated at z = 0 and
z — lt. By putting

Xi = -0[ , (* = l , 2) (6.9.23)

where 6\ — dO/dz in z = 0,£ e , the unknown functions can be represented in


265

C, u, %

Fig. 6.28 - Finite beam element.

the form
u{z) = h{z)u1 + f2{z)u2 + f3{z)4>1 + fi{z)fa = fTh)

v{z) = f1{z)v1 + f2{z)v2 + f3{z)(f>l + fi{z)<l>2 = fT (J J (6.9.24)

= f T
o(z) = f1(z)el + fi{z)02 + Mz)Xi + f4(z)x2 (x)
The fi(z) (i = 1 , . . . , 4 ) are interpolation functions. They describe the dis-
placement field inside the element for a generalised unit displacement of the
node to which they refer, all other nodal displacements being equal to zero.
As interpolation functions we can, for example, choose the exact solutions to
the purely flexural problem. In this case we have
,2 / , X 3

AW = i--y«y
z\>
2
AW = *(f) - ( f (6.9.25)
AW = - ( f -

AM = " 4 ( f
266

It must be noted that with the choice (6.9.25), the expression of the angle
of rotation 0{z), unlike the displacements u{z) and v(z), does not satisfy the
corresponding homogeneous equation of equilibrium
ET0" GJ0" = O (6.9.26)

In other words, the description of the displacement components u(z) and v(z)
is more accurate than that of 0(z). A better approximation could be made
by assuming, for the rotation 0(z), the exact solution to the homogeneous
problem (6.9.26). This, however, would complicate the expressions (6.9.25).
Equations (6.9.24) are formally analogous to (6.9.1) used in the Ritz meth-
od. Therefore, by following the same procedure, already illustrated, we arrive
at the definition of two matrices for the element: the elastic and the geomet-
rical local stiffness matrices. They are of order twelve and are associated with
the vector of local displacements
q e = ( u i , u2, tpi, V>2, vi , v2, <t>i, 4>2, 0 i , 02, X i , X2)T (6.9.27)

By assembling the two matrices separately, taking into account the connec-
tions between the elements and the orientation in a global system of coordi-
nates (see Section 5.9), we again arrive at the eigenvalue problem (6.9.10).
The global matrices K o and K ^ of the system, which in the case of the Ritz
method are generally full, are now banded, thanks to the local approximation

torsional instability -•£


# = 0 # = 0

Mc Mc

lateral
- -A--
instability
^^ vmt. /
#=0 #=0
v =0 v =0

SI
v =0
;i if
v'=0
#=0 #'=0

1 2 3 4 5 6
Number of elements

Fig. 6.29 - Error convergence curves for different cases.


267

of the functions of interpolation, and are therefore more suitable for solving
the numerical problem.
We do not give here the rather complex explicit expression of the matrices
of rigidity, which can be found in [8]. We report only the convergence curves
relating to some test problems, for which the exact solutions are known, in
order to show the accuracy of the results obtainable. Figure 6.29 shows, for
the three examples indicated, the error in the evaluation of the critical load as
a function of the number of discrete elements into which the b e a m is divided.
Note t h a t in all the cases, only two elements are required to obtain an error
lower than 1%.
Notice too that with a single element, the largest error occurs for the
purely torsional buckling mode, in accordance with the observation made on
the interpolation functions. Figure 6.30, relating to a simply supported beam
with different load conditions, shows an analogous behaviour of the error. In
conclusion, we note how all the curves converge monotonically to the exact
solution with an increase in the number of elements and how the error is
always positive (in excess). This property is assured by the fact t h a t the
finite element described is compatible and complete.

% Error
i l°
36' "IHIUtt
I-TU0
3 5 j•
25' - #=o v
v= 0
24
Uniform load applied to the extrados
23 -
(l s e c t i o n )
22 * U n i f o r m load applied to the intrados
0
(l section)
7
Uniform load applied to the shear center
6 - (rectangular section)

5 - Concentrated load applied to the extrados


(l section)
4 k
Concentrated load applied to the intrados
3 h (l section)
Concentrated load applied to the shear center
2 k (rectangular section)

1 \-
-#-^ 6---,..
Number of elements

Fig. 6.30 - Error convergence curves for simply supported beam.


268

6.10 OUTLINE OF THE POST-CRITICAL BE-


HAVIOUR OF THIN-WALLED B E A M S
The post-critical behaviour of open thin-walled members can be analysed by
adapting the general theory of continuous systems presented in Chapter 4, as
was done in Chapter 5 for systems of plane beams. With respect to the latter
treatment, however, a complication arises which is related to the formulation
of a correct mathematical model which accurately describes the mechanical
phenomenon. The model studied so far allows us to write, though incom-
pletely, the total potential energy up to the fourth order in the displacements
u(z), v[z), 0{z), if we assume a quadratic measure of deformation such as
the tensor (4.3.1). This, however, is not adequate for a non-linear analysis of
the type previously carried out. The results at which we arrive, in fact, are
greatly influenced by the assumption of undeformability of the cross-section,
by the law of warping, and by the linearisation of the kinematics. A more rig-
orous analysis would require a reformulation of the model, which is outside
the limits of the present treatment. In the following we present a treat-
ment taken substantially from [14], in which we use the results from linear
analysis to write the third-order terms of the total potential energy which,
although incomplete, is sufficient for our present purposes. The analysis leads
to the determination of the tangent to the equilibrium path at the bifurcation
point, allowing us to distinguish whether or not the post-critical behaviour
is symmetrical. However, in the affirmative case, it does not give an answer
regarding the stability of the bifurcated path (see Fig. 4.4).
We again consider the field of displacements (6.3.1) and (6.3.6), rewritten
here excluding the inessential function f (z)
up{z,s) = u(z) - 6(z) (y - yc)

vp(z, s) = v(z) + 6{z) (x - xc) (6.10.1)

wp(z,s) — — u'[z) x — v'(z) y — 0'(z) u


The components of the deformation tensor (4.3.1) are all homogeneous quad-
ratic in the displacements u(z), v(z), 0(z) with the exception of 733, which
also has a linear part e 33 . It has already been observed, in fact, that the linear
components of the deformation 7^ all vanish, with the exception of e 33 . From
this it follows that, as the elastic potential energy is quadratic in 7ty, there is
a third-order term which comes from 7I3. From (4.3.1) we have

733 = ™p+2 {UP+VP+WP) ~ -u"x-v"y-0"u

+ \ { K - <t{v ~ Vc)}2 + W + 0'(x - xc)}2} (6.10.2)


269

having neglected w'p with respect to wp. By writing the total potential energy
in the form
2
P(u) = i ^ ' u + ip»'u3 (6.10.3)

we get

^ PS' u 3 - I E 6 33 M33 rfF (6.10.4)

in which /x33 is the second-order part of 733. By making use of (6.10.2) and
carrying out integrations, we have
,2 2
i P»'U3 = - l - E L Xcu r
[e"0 dz--* * EICX fu"0'2 dz
6 "° " 2 h 2 Ji

- EIcy fv"6'2 dz + EIx [u'v"6' dz


2
EL y Jv'u"6'dz (6.10.5)

in which we have taken into account that x and y are principal axes of inertia
and t h a t the function CJ is referred to the twisting centre; also, we have the
following

lex = I [{x - xcf + (y - y c ) 2 ] x cL4

7cy = JA[{x-xc)2 + {y-yc)2]ydA (6.10.6)

Jew = y [{x - xc)2 + [y - y c ) 2 ] CJ dA

Equations (6.4.22) and (6.10.5) substituted in (6.10.3) provide an explicit


expression in terms of u of the total potential energy up to the third order.
In the critical condition of equilibrium we have P"VL\ — 0, where ii c is
the buckling mode. The necessary condition for the stability of the critical
configuration is that P"'u^ = 0. If P^'ii 3 . ^ 0 then the configuration is
unstable and the tangent to the equilibrium path at the bifurcation point is
given by (4.5.19)
1 P ' " i ic 3
Ac = ^— - (6.10.7)

being v = 0, since the fundamental path is trivial. Therefore, by substitut-


ing in the functional (6.10.5) the buckling mode ii c for the problem under
examination, already calculated in previous Sections for some simple cases,
we obtain information on the post-critical behaviour of the system.
As an example, we now examine the case of a thin-walled beam subjected
to simple compression with boundary conditions expressed by (6.5.3), (6.5.4).
270

The buckling mode, from (6.5.5), is

/ u \ ( Au\
7T Z
u„ Av sin (6.10.8)

\'J \ A
e )

in which Au, Av and Ae are solutions to the homogeneous system (6.5.8). By


substituting (6.10.8) in (6.10.5), noting that all the integrals are of the type

2 7T Z . TV Z 2 7T
cos —— sin —— dz — — (6.10.9)
/: 1 1 3 P

we have

pill ,-,3 7T2£


u; \lCUJ AQ + [lcx Au + lcy Av) AQ
3 I2

+ 2{Iy-Is)AuAvA$\ (6.10.10)

The functional is therefore reduced to a homogeneous polynomial of the third


degree in the parameters of the buckling mode A u , Av, A$, which are appro-
priately normalised.
In the discussion in Section 6.5 it was implicit that the three roots of the
characteristic equation are distinct. It can happen, however, that for par-
ticular values of the geometrical properties of the cross-section, two or all
three roots coincide, or assume very close values. If, for example, we have
Ix — 7 y , from (6.5.7) Nx — Ny results, and two roots of the characteristic
equation for a doubly symmetrical thin-walled section coincide. As a conse-
quence, the buckling modes which correspond to the coincident roots occur
simultaneously. In fact, if an eigenvalue has a multiplicity n, all the solutions
are obtained as a linear combination with arbitrary coefficients of n linearly
independent eigenvectors. We therefore have

Uc = Ci Uc (i = 1 , 2, . . . , n) (6.10.11)

where n, in the problem under examination, can at the most be equal to 3.


The phenomenon of coupling of modes is called interaction of simultaneous
buckling modes and will not be treated in detail here. In this regard we
only point out that the interaction between the modes in general produces a
remarkable reduction in the load-carrying capacity of the imperfect structure,
as observed in the model of two degrees of freedom studied in Section 2.12.
We now proceed to an examination of the two cases discussed in Section
6.5.
271

a) Cross-sections with two axes of symmetry


For the double symmetry we have

-*cx -*cy 'cw U (6.10.12)

ind (6.10.10) becomes

2 TT2E
pi" tf = ; ^ 2 (/, - ix) AU A„ A, (6.10.13)
3 £
The three critical loads are N\ — Ny, N2 = Nx, N3 = Ng and are defined by
(6.5.7). The associated eigenvectors are

I Au\ / 0 \ / 0 \
uei = 0 , uC2 = \ Av , iiC3 = 0 (6.10.14)
V 0 ) { 0 ) \At j
If the three eigenvalues coincide, any linear combination of the type

( Au \ ( 0 \ ( 0 \
Uc = Ci 0 + c2 Av + c3 0 (6.10.15)
V 0 ; V 0 ; \A6 )
is a solution to the system (6.5.8). Equation (6.10.15) is equivalent to

( l\ ( Q\ ( 0 \
ur 0 + Av 1 + A$ 0 (6.10.16)
v1 y
If only two eigenvalues coincide, then the associated eigenvectors can be de-
termined by (6.10.11) for n = 2.
Three sub-cases can be distinguished.
(1) The three eigenvalues are distinct. The buckling modes are given by
(6.10.14), that is, only one of the three amplitudes is non-zero. Equation
(6.10.13) therefore gives P'" uzc = 0 for each mode and the post-critical
behaviour of the beam is symmetrical.
(2) Two eigenvalues coincide and are distinct from the third. For example,
let Nx = Ne y^ Ny. The buckling modes associated with Nx — N$ are
obtained by linearly combining iiC2 and ii C3 ; the buckling mode associated
with Ny is ii C l . In all cases, we again have P"' u^ = 0.
(3) The three eigenvalues coincide. The eigensolutions are described by
(6.10.16) and Au, Av, Ad are all different from zero. However, as
AL 7Vy, we also have Ix Iy and (6.10.13) again yields P™ u3c = 0.
272

In all cases the post-critical behaviour of the thin-walled beam is symmetrical.


To decide upon the stability of the critical configuration, an analysis up to
fourth-order terms is necessary.

(b) Sections with one axis of symmetry


Suppose for example that y is the axis of symmetry. We have

*cx =
lew — V (6.10.17)

and (6.10.10) is simplified to

7T2E
pin
u„ [ley Av A) + 2{IV - Ix) Au Av Ae] (6.10.18)
31 2
The critical loads are provided by (6.5.13) and the buckling modes are

(° Avi
A
u„
/ A " m2 A
0
( AuxU3 \
0 (6.10.19)
A
; A
V <>2 J V e, J
the first corresponding to Nx, the other two to a combination of Ny and iV^
according to the first of relations (6.5.13).
Three sub-cases are distinguished.
(1) The three eigenvalues are distinct. By separately substituting the eigen-
vectors (6.10.19) in (6.10.18) we always have P™ usc = 0.
(2) The two eigenvalues associated with iiC2 and iiC3 coincide and are different
from the third. As Av = 0 or Ae — 0, we again have P"1 li^ = 0.
(3) The eigenvalues associated with ii Cl and iiC2 (or iiC3) coincide. In this
case, the corresponding buckling modes are given by

/ oA ( AU2 \
uc C\ + c2 0 (6.10.20)
0 J V A02 J
V
that is, the three amplitudes are different from zero. In this case we have
P"' u^ y£ 0. An analogous result is obtained if the three roots coincide.

In conclusion, the post-critical behaviour of the thin-walled beam is sym-


metrical if there is no interaction between the purely flexural mode and the
flexural-torsional mode. In this case, the bifurcated curve has a tangent which
is different from zero at the bifurcation point and the critical equilibrium is
unstable.
273

REFERENCES
F . Bleich: Buckling strength of metal structures, McGraw-Hill, New York,
1952.

V.Z. Vlasov: Thin-walled elastic beams, 2nd ed., The Israel program for
scientific translations, Jerusalem, 1961.

S.P. Timoshenko, J.M. Gere: Theory of elastic stability, 2nd ed.,


McGraw-Hill, New York, 1961.

A.S. Volmir: Stability of deformable systems, NASA AD 628 508, 1965.

A.H. Chilver (editor): Thin-walled structures, Chatto and Windus, Lon-


don, 1967.

V. Franciosi: Strength of materials (in Italian), Liguori, Napoli, 1967.

M. Como: Theory of elastic stability (in Italian), Liguori, Napoli, 1967.

R.S. Barsoum, R.H. Gallagher: "Finite element analysis of torsional and


torsional-flexural stability problems", Int. J. Numer. Methods Eng., 2
(1970), 335.

M. Como: "Stability of structures" (in Italian), in R. Baldacci, G. Cera-


dini, E. Giangreco (editors): Dynamics and stability (in Italian), Italsider
technical-scientific series for the design of steel structures, Vol. l i b , 1971.

Handbook of structural stability, Column Research Committee of J a p a n ,


Corona Publishing Company, Tokyo, 1971.

D.O. Brush, B.O. Almroth: Buckling of bars, plates and shells, McGraw-
Hill, New York, 1975.

L. Corradi: Structural instability (in Italian), CLUP, Milano, 1978.

W.T. Koiter: Theory of elastic stability, lecture notes, University of


Technology, Delft, academic year 1978-79.

A. Grimaldi, M. Pignataro: "Post-buckling behaviour of thin-walled


open cross-section compression members", J. Struct. Mech., 7(2) (1979),
143.

P.A. Kirby, D.A. Nethercot: Design for structural stability, Granada


Publishing, London, 1979.

J. Rhodes, A.C. Walker (editors): Thin-walled structures, Granada Pub-


lishing, London,1980.
275

Chapter 7

ANALYSIS OF PLATES A N D SHELLS

7.1 INTRODUCTION
By a shell we mean a solid spanned by a segment whose middle point belongs
to a regular surface and which moves so as to remain orthogonal to the surface.
The length h of the segment is defined as the thickness of the shell. If,
during motion, the length of the segment remains constant, we have a shell
of constant thickness] in the opposite case the shell is said to be of variable
thickness. The surface to which the middle point of the segment belongs is
called the middle surface. In the theory of shells it is usually assumed that
the thickness is small relative to a dimension typical of the middle surface,
for example, a radius of curvature. In this case we speak of thin shells. If
the middle surface belongs to a plane, we have a plate. The mathematical
model of shells will be presented in a very concise form; the interested reader
is referred to classical texts [4,9,11,14,16,19,21,24]. We introduce the subject
with some elementary notions of surfaces and shells.

7.2 SOME B A S I C RESULTS OF T H E THEORY OF


SURFACES A N D SHELLS
Let x, (z = 1,2,3) be the Cartesian coordinates of a point in space. We
define as a surface the set of points whose coordinates are regular functions
of two independent parameters; these functions are the parametric equations
of the surface

Xi = fi ( a x , a 2 ) (7.2.1)

By means of (7.2.1) each point of the surface is uniquely determined by two


numbers a\ and a 2 and we can therefore call these quantities coordinates
on the surface. By keeping a2 constant and making a x vary, we obtain a
family of curves which we call coordinate lines o^. Analogously, we define the
coordinate lines a2. Each point of the surface can therefore be considered as
the intersection of two coordinate lines ai and a 2 - There will obviously exist
some choice of coordinates for which the formulae characterising the surface
take the simplest form. This happens when the coordinate lines are the lines
of principal curvature of the surface, characterised by their being mutually
276

orthogonal at every point. In the following we shall adopt this intrinsic system
of coordinates.
By writing (7.2.1) in vector form

r = r(a1,a2) (7.2.2)

we indicate by dr the increase of the vector position corresponding to an


increase of the a coordinates. We have
3r <9r ,
dr = - — dax + —— da2 (7.2.3)
OOLi UCLi

from which

ds2 = Hi da\ + Hi d a\ (7.2.4)

where ds is the modulus of dr, having made use of the condition of orthog-
onality of the vectors d r / d o ^ and d r / d a 2 > tangents to the coordinate lines,
and having put

2 / dx\ \ I dx2 \ I dx$

If in particular only one of the coordinates varies, then we have

ds\ = H\ d ot\
(7.2.7)
ds2 = H2 d (X2

The quantities Hi and H2 are defined as Lame parameters. They are given by
the ratio of the increase of the arc of the coordinate lines to the corresponding
increases in the curvilinear coordinate.
In thin shells, starting from the two-dimensional system of curvilinear
coordinates, we can construct a three-dimensional system, taking the normal
to the surface at a generic point P (Fig. 7.1). The position of a point M ,
belonging to the normal, will be defined by its distance z from P and by the
coordinates c*i and a2 of P . The system of coordinates cq , a2 , OLZ — Z forms
an orthogonal system which we call the shell coordinates. Each point in the
space which is occupied by a shell is an intersection of the three coordinate
lines a i , a 2 , a 3 where a j and a2 are outside the middle surface.
To determine the Lame parameters of this new system of coordinates, we
refer to their definitions. Indicating by Rx the radius of curvature, we have
277

Fig. 7.1 - Intrinsic system of coordinates for a shell.

from Fig. 7.2


ds[z) _ Rj + z
(7.2.8)
ds\ Ri
from which

dsP=ds1(l+-j^ (7.2.9)

and by making use of the first of equations (7.2.7)

ds[z) = H, ( l + ^ ) da, (7.2.10)

Equation (7.2.10) can be written in the form

ds[z) = H[z) don (7.2.11)

Fig. 7.2 - Shell segment.


278

having put

H[')=H1{l + -^-) (7.2.12)

Analogously,

(7.2.13)

Finally, we have

H3 = 1 (7.2.14)

because along the axis z, the length of the arc of the coordinate line coincides
with the coordinate itself.
Let us now look for the specific Lame parameters for shells of revolution.
For the surfaces of revolution the principal curvature lines are the meridians
and the parallels, that is, the intersection lines of the surface with the planes
containing the axis of revolution and with the planes perpendicular to this
axis [14]. Consequently, the angle 0 between the normal to the surface and the
axis of revolution and the angle <p which determines the position of a point
on the corresponding parallel, can be assumed as curvilinear coordinates of
the surface (Fig. 7.3). Let us put a x = 0, a 2 = <P and indicate by R\ the
radius of curvature of the meridian, and by R^ the second radius of curvature
equal to the segment of the normal to the surface between the surface itself
and the axis of revolution. The elements of the meridian and parallel arc are,

Fig. 7.3 - Revolution surface.


279

respectively,

dsx = RxdO
(7.2.15)
ds2 — R2 sin 6 d(p

from which

H x = Rx , H2 = R2 sin 0 (7.2.16)

Furthermore we have

H[* = * ( l + £)
V RlJ
(7.2.17)
J^z) = fl2 sin 6
('♦*)
For a sphere, (7.2.15)-(7.2.17) remain unaltered except t h a t we put R\ =
R2 — R.
In a cylinder, the principal curvature lines are the directrices and the
generatrices. Consequently, we can assume as curvilinear coordinates of the
surface the distance x of a generic point from a directrix of reference, and the
lengths of the arc in a circumferential direction measured from a generatrix
of reference. By writing a.\ — x, a2 — 6, we immediately have from ds\ = dx
and ds2 = ds

H! = 1 , H2 = 1 (7.2.18)

Also, as R\ = oo and R2 = i2, with R the radius of the cylinder, we have

H[z) = 1 , H[z) = 1 + - | (7.2.19)

For the plate we have finally

Hx = 1 , H2 = l (7.2.20)

H[z) = 1 , H{2z) = 1 (7.2.21)

as can easily be verified.

7.3 KINEMATICS OF SHELLS


Let P be a point on the middle surface and s the displacement vector of P ,
which is a function of a x and a2. Let k x and k 2 be the unit vectors tangent
280

to the coordinate curves a\ and a2 in P and k n the unit vector orthogonal


to k i and k 2 defined by

k„ = k x x k 2 (7.3.1)

where the symbol x indicates the vector product. Let us denote by uyv,w
respectively the componentes of s in the frame ( k i , k 2 , k n ) . Let us now take
the normal to the middle surface in P and consider a point Pi on it, at a
distance z from P . The positive direction of z is assumed to coincide with
that of k n . Let us suppose that during motion the segment P P i , orthogonal
to the undeformed middle surface, remains straight, undeformed and orthog-
onal to the deformed middle surface [4]. On the basis of this hypothesis the
displacement s ^ of point Pi is a function of a i , a 2 and z and its components
uW, wW, wW can be expressed as a function of u,v,ti;.
To this end let us write the vector equation (Fig. 7.4)

z k n + sM = * k ; + s (7.3.2)

where kj^ represents the unit vector orthogonal to the deformed middle sur-
face. From (7.3.2) we obtain

gW = s + z (K - k„) (7.3.3)

where it can be seen [14] that

k ; = k n + ^ i k i + £>2k2 (7.3.4)

Fig. 7.4 - Kinematics of a surface element.


281

having written

1 dw u

(7 3 5)
i a - -
1 OW V

H2 da2 R2

From (7.3.4) we can write


S W = s + z [lfl k x + <p2 k 2 ) (7.3.6)

or in scalar form

u(z) = u + z <pi
vW = v + z<p2 (7.3.7)

Equation (7.3.6) or (7.3.7) represents the required relations which express s ^


as a function of s. From (7.3.4) we can see t h a t <pi and <p2 are the components
of k^ on k i and k 2 .
The next step is to determine the components of the deformation tensor
(4.3.1) in the form

W-tf + lW+WW+uti) (7-3-8)


where v^ = (uitj - u t J ) / 2 . S ince for the stability of thin shells ety <C u\y <C 1,
as is observed in experiments, it is possible to simplify (7.3.8) into

-$ = if + \ «i? 41 (7-3.9)
A justification for the validity of the approximations (7.3.9) is furnished by
Koiter [16]. A further simplification follows from the observation t h a t the
rotation around k n (rotation in the tangent plane) is much smaller t h a n the
rotation around k x and k 2 (rotation outside the tangent plane), t h a t is,

UJU <C W i s , CJ23 (7.3.10)

Thus in the second member of (7.3.9) we shall neglect, with respect to the
other terms, all the u^i ^hj products in which one of the u factors is uii2.
Using the displacement components (7.3.7) in the expressions for e,-y and uty
in general curvilinear coordinates which are not given here, and remembering
282

that in shells i / 3 = 1, we reach the following expressions of the deformations


[14] to within higher order terms

/^*> en + z
lxi - w
622+z x2
e
22 ~
( - S) ei2 "
en + z Xl2 +
e
12 -
- U ft) 2 . (7.3.11)
e
33 = 0

e
13 = 0

—- 0

and the expressions of the rotations

M = A - A + *(Xi-5f,)+*(^-^)
u 12
1 dw v
CJ..W _ ^23 (7.3.12)
J?2

23 H2 da2
W _ 1 dw u
u;13 — Hi octi Ri

In (7.3.11) and (7.3.12) we have put

du dHx w
en
Hx dax
+' Hx H2 da2 V +
~R[
1 dv 1 d#2 w (7.3.13)
^22
1 u H
# 2 ##2 #1 #2 d#l #2
2e 1 2 ^ 2 j 9 _ (jv_\ Hi _d_ /_u_\
+
ffi a«i \ # J #2 3«2 v f r j

1 du;w ajix
Xl = 77" 13 w
a;.23
fTi 3 a i HXH2 da2
dJ$ 1 dH2 , ,W
X2
H2 da2 + Hi H2 W
dax 1S
(7.3.14)
1 W
du;23 i aw13w
2X12 = -
Hi dax H2 da2
dHi 3#2
a45))
w

Hi H2 V d a 2 a; 13
<9ai
283

1 dv 1 dHt
Pi = — u
Hi dcti Hi Hi da2 (7.3.15)
1 du 1 dH2
H2 da2 Hi H2 da.i

_ = _ j _ du£_ i_ dih w
Xl
" ~ * Oai ~ Hi H2da2^ {? g ^
_ = J _ du4£ 1 dff2 (,)
X2
H2 da2 Hi H2 dai "2Z

We can see by comparing the second and third of equations (7.3.12) with
(7.3.5) that we have <pi = o;13 and <p2 = — u23. The expressions

en 1 1
Xi "
Ri
(7.3.17)
^22 i 1
X2 -
i?2 i?2 ~R2
are called variationas of curvature, whilst

is defined as the variation of the curvature of torsion. In many problems


it is possible to identify the variation of curvature with the curvature itself,
without introducing appreciable errors in the solution. In this case the first
three of relations (7.3.11) reduce to

cvii ^ e n + z Xi

e$ ^ e22 + zX2 (7-3.19)

2e$ ^ 2e 1 2 + 2 z X i 2
This type of simplification is possible in shallow shells or in shells in which
the displacement components are given by rapidly varying functions along the
curvilinear coordinates. Experience shows that in these cases the terms in u
and v can be neglected with respect to the gradient of w in the expressions
for rotations, and therefore we can write
( Z) _ _ 1 dw
^23 - ^23 - T7- ^ —
H
> ?"' (7-3.20)
1 dw
W _ _
284

from which simplified expressions of (7.3.14) follow

1 d 2w 1 dHi dw 1 dHi dw_


Xl +
~ ~ Hj !h% Hj d^~ ~d^ ~ Hx Hi d^~ ~d^2

1 d 2w 1 dH2 dw_ 1 dff 2 dw


X2 + _ 2
~ ~]^6>o! tffao7ao^ #! #2 "^7 a^ ( •• )
1 d2w 1 affi dt^ 1 dH2 dw
Xu + 2 +
~ ~ HiH2 dax da2 H H2 ~d^ d ^ 7 HiH\ ~d^ ~d^~2

The kinematic relations (7.3.13), (7.3.20) and (7.3.21), together with (7.3.19),
lead to the so-called Donnell-Mushtari-Vlasov equations of shallow shells
[9,11,
14,15,29].
In the case of plates, by putting «i = x, and a 2 = y and taking (7.2.20)
into account, the linear kinematic relations are reduced to

^23 = ^23 = W,y


(7.3.22)
w
13 = wis = w, y

Xl = - W,zx

X2 = -™,vv (7.3.23)

Xl2 = -w,xy

en = «,i

t22 = w
* (7.3.24)

ei2 = j (".«' + V
.»)

e
n = u, x - ziu^x

e
22 = v
,w - * w,vv (7.3.25)

2 eiv = «,„ + K , - 2 Z wtXy


285

By using (7.3.9), (7.3.10) and (7.3.19) we can finally write the measures of
deformation for shells

(Z) 1 o

7n = en + z x i + - wj s

M , >* 2

722 = ^22 + 2X2 + - W23

1
M { 2 , 2 \
2
(7.3.26)

7l2} = e
12 + * Xl2 + ~ Wis ^23

7# = 0

1® = 0
Analogously, from (7.3.9), (7.3.10) and using (7.3.22) and (7.3.25), we have
for plates
M . 1 2
7ll = Utz ~ * ™,xx + ~ W%x
M , 1 2
722 = V.V-* ™,VV + o ^ f y

(7.3.27)

7i2 = 2 (U'v +
^ ~~ z W xy
' +
2 W,x W,v

1® = 0

7# = 0

7.4 ELASTIC S T R A I N E N E R G Y
In deriving the expression for the elastic strain energy we make the hypothesis
that the stress <% normal to the middle surface is zero. This situation is
satisfied to a sufficient approximation if we allow t h a t h/R <C 1, where R
is the smallest of the principal radii of curvature of the undeformed middle
286

surface. From the generalised Hooke's law

E
+ 7
°u ~ l + v{ l n
l-2*/ )

7
°22 - 1 + uV22 +
1-2!/ )
u
J°) - JL_ (J*) + 4*)\
°3Z ~ l +v V 3 i +
l - 2 v 1
) (7.4.1)

=
<?23 ^ G 723

an( =
with 7W = 7ii +722 +733 i ^ E/2(l + i/) we get, under the assumption
aW=0

which, substituted in the first two of equations (7.4.1) leads respectively to

(7.4.3)

The coefficients i? and i/ are the Young's modulus and the Poisson ratio,
respectively. Using the strain energy in the form

{l\f) = \ / /_*" ag 4:1 Ry ^ d^ ds* dz («. )9 = 1.2) (7.4.4)

(to within the terms S^UMU^ /2) and introducing equations (7.4.3) and the
fourth equation of (7.4.1), we arrive at

W = \LC[T^W^^W
+ 4G"$* Hi H2 dSl ds2 dz (7.4.5)

from which, resorting to the first, second and fourth of (7.3.26) and integrating
287

over the thickness, we obtain

w =
Y Is ( 7 " + 722 + 2 vlu 722 + 2^ " ^ 7 ' 2 ) ^ ^2 d51 d52
+ -y / 5 [ x i + X2 + 2 i / X i X 2 + 2 ( l - i / ) x i 2 ] ^ i ^ 2 ^ 1 d 5 2 (7.4.6)

In (7.4.6) we have made use of


1
2
e w
7n = n + 2 i3

C
722 = 22 + - ^23 (7-4-7)

1
712 = ^12 + - CJi3 CJ23

and of De = Eh/(1 - v1), Df = Eh3/12(l - z/ 2 ), De being the extensional


stiffness and Df the bending stiffness.

7.5 STABILITY OF PLATES


Let us consider a plate of constant thickness /i, whose middle surface lies in
an (x, y) plane. The state of stress, acting in the fundamental configuration,
is described by the three stress components 5 n , S12, 522- In the generic
adjacent configuration the total potential energy, omitting the term V(u;A)
because it is bilinear in u and A and therefore does not contribute to successive
variations, is

p(u) = f / J ^ + I ^ ' + ^ + I.!) 1


+ 2 i/(u,,+ -«;?,) («,» + £ «£)
1 V
~ 1 \2
dxdy
H — (u,„ + «,, + w,z w,v)

+
^2~ Is I""'11 + ^ + 2U W,xx W,vv + 2
^ ~ ^ "^"l dXdV
+ - f [Su w2x + 2 512 wiX wiV + S22 w\] dx dy (7.5.1)

Equation (7.5.1) is obtained by adding to the expression of elastic energy


(7.4.6) the contribution given by the stresses in the fundamental state 5,y,
and then using the equations (7.4.7), the kinematic relations (7.3.22)-(7.3.24)
288

and equations (7.3.20). Collecting terms of the same order in displacements


u,i> and w, (7.5.1) can be written

P(u) = \ P'J u2 + I P1:1 U 3 + 1 Pc"" u< (7.5.2)

where

^ p; u * = Y fs I"2* + w* + 2"u,x t,,v + X~T~(u,y + u , l ) 1 d xdy


+ - y / s [u/fss + ti;2,,,, + 2i/ t« i i s w,m + 2 ( 1 - 1 / ) «>2X|/] dx dy

+ - f [snw*x + 2SuwiXwiV + S22wl]dxdy (7.5.3)

e = -f f [u,, (w2x + 1/ u;f„) + vtV (w2y + t/ «;2X)] dx dy (7.5.4)


6

^ P c "»u 4 = ^ j s (w'x + t, 2 v ) 2 d x d y (7.5.5)

The second-order term (7.5.3) consists of the sum of three contributions: the
first represents the work done in a purely extensional motion and therefore
depends only on the in-plane displacements u and t;; the second is associated
with a purely flexural deformation and depends only on the lateral displace-
ment w; while the third expresses the work done by the stresses present in
the fundamental state and also depends on displacements w only. The last
circumstance is a consequence of the simplifications made in the study of
the kinematics of the system described in Section 7.3. The result is t h a t the
terms in u and v are uncoupled from the terms in w and are independent of
the state of stress 5 t ; . As we also have — 1 < i/ < 0.5, the extensional term
represents a quadratic functional which is positive definite. On the other
hand, in the critical condition of equilibrium, we must have P c "u 2 /2 = 0,
and therefore it must necessarily be that u = v = 0. The buckling mode
consists then only of the lateral displacements w, and the deformation of
the plate is purely flexural. Corresponding to the critical state we have also
P c '"u 3 /6 = 0, P""u4/24 > 0, as the in-plane displacements are identically
zero and the functional (7.5.5) is positive definite. We can conclude that the
critical configuration of equilibrium of the plate is stable [31].
Let us now pass to a determination of the equilibrium equations, using the
functional (7.5.3). From an analytical point of view, this is accomplished by
requiring the vanishing of the variations of the functional with respect to the
three functions u, v, w. This leads to three differential equations with relevant
289

boundary conditions. Considering the particular structure of the functional,


however, the differential equations which contain the unknown functions u
and v are uncoupled from the third and are homogeneous with homogeneous
boundary conditions, as is easy to verify. Therefore, the only possible solution
is the trivial solution u = v = 0. Thus the problem consists of the solution of
the only equation in w which we shall now derive in detail. Taking the first
variation of the functional (7.5.3) with respect to w and equating it to zero,
we have

Pl'uSu = Df / {{W}XX+ VWtyy)6WtXX + {Wtyy +VWtXX)6Wtyy


Js

+ 2(1 - v) wtXy 6 wtXy] dx dy


+ h [{SnwiX +Si2wiV) 6wiX
Js
+ {S22 wtV + S12 wiX) 6 wiV] dxdy = 0 (7.5.6)

As already done for one-dimensional problems, it is necessary to integrate


the various terms by parts. From the Green theorem we have the integration
formulae

<f>tpxdxdy — — \ T\) (j>>x dx dy + <f> </> ip dy


(7.5.7)
<f)tpydxdy = — / x/j <f>tV dx dy — <f> <j) V> dx

where the curve C is the boundary of the domain 5 .


By limiting ourselves to consider only rectangular plates with x, y axes as
shown in Fig. 7.5 and with the axis z oriented upwards, we have on applying

Fig. 7.5 - Rectangular plate.


290

(7.5.7)

P" u 6 u = Df / [wtXXXX + 2 u;jXxyy + w tyyyy ] Swdxdy


Js

- h i [(Sn u;|X)iX + (5x2 ti;|X)iy + (5 1 2 wtV)tX + (S 2 2 w,v),v] d x d y


Js
fb
W ( ix ~h ^ ^,yy) ^ ^ , z | x = 0 ^ ^

£ ) dx
+ /y IKvv + ^™xz)£Wy|y=0
a _.
/
| [ ^ ( S l 2 tW|X + 5 2 2 Wty)-Df {Wtyyy + (2 ~ ^W^y)} Sw\VyZQ d.X
rb
+ / | [ ^ ( 5 n u ; | X + Si2u; iy )--jD / (u; iXXS + (2-i/)u; i y y x )]«u;|* = ody
Jo
\x=a
+ ||2Z?,(l-»/)u;lSy*ii;|p|!t=o = 0 (7.5.8)

In order that equation (7.5.8) is valid for any kinematically admissible 8w,
the single integrands must vanish separately. In particular, we get

W xzzz + 2 W xxyy
xzutv +" W
™,yyyy = ^jy- [ ( S l l M ; , « ) i « + ( 5 12W | «) iy

+ ( ^ i 2 ^ y ) , x + (5 22 ti;, y ) >v ] (7.5.9)

which simplifies to

W,ZZZZ + 2 W > x x y y + W.yyyy = — [S\\ W|XX + 2 5 i 2 U7|Xy + 5 2 2 tl7ftfy) (7.5.10)

by virtue of the equations of equilibrium in the fundamental configuration


'S'll.z + ^12,y — 0
(7.5.11)
S\2tX + 5 2 2 | y — 0

By writing for convenience

Nx = - h S n , Ny = -hS22, Nxy = -hSl2 (7.5.12)

with Nx and Ny compressive forces and 7Vxy the in-plane shear force per unit
length of the fundamental configuration, we can rewrite (7.5.10) in the form

w
1
W,xxxx + 2 W,zxyy + ,yyyy = ~ "fT" {Nx W,xx + 2 Nxy WyXy + Ny W)J/y)

(7.5.13)
291

The boundary conditions of equation (7.5.13) are obtained by equating to


zero the other terms of (7.5.8). With the positions (7.5.12) we have

(w,zz + vwtVy) 6 wtX = 0 in x = 0 , x = a

{Wm + V W}XX) 6 Wty = 0 Ul t/ = 0 , t/ = 6

WtZXX + ( 2 - I/) W>yyX + — (JVX WfX + NXy W j) 6w = 0

in x = 0 , x = a (7.5.14)

1
w
,yyy + ( 2 ~ u) w^xV + 77- ( N i y u;ia. + Ny W)V) <5t/; = 0

in y = 0 , y = b

2Df{l- v) wtXy6w = 0 in (0,0), (a,0), (a, 6), (0,6)

The eight conditions (7.5.14), to be imposed along the edges of the plate,
and the last of the (7.5.14) conditions, to be imposed at the corners, complete
the differential problem. The solution allows us to determine the buckling
modes and the critical loads.
Note t h a t the differential equation (7.5.13) can be interpreted as the con-
dition of equilibrium, written in a configuration adjacent to the fundamental
configuration. The natural boundary conditions obtained from the variational
procedure are therefore interpreted as mechanical boundary conditions. In
particular, in the case of simply supported edges, the first two equations
(7.5.14) provide the conditions for the vanishing of the bending moment and,
in the case of a free edge, the vanishing of the shear is obtained from the
third and fourth equation. From the last equation, in the case of free corners,
comes the Kirchhoff condition which prescribes the vanishing of the vertical
reactions.

7.6 CRITICAL LOAD OF RECTANGULAR PLATES


We present here some applications concerning the determination of the critical
load and buckling mode of rectangular plates, by means of the integration of
equation (7.5.13). In all cases, the stresses Nx, Ny, Nxy in the fundamental
state are assumed constant inside the plate, so t h a t the differential equation
has constant coefficients [7,8,10,12,15,17,18,22,23,25,26,28-31].

(a) Simply supported plate uniformly compressed in one direction


With reference to Fig. 7.6, let us consider a simply supported rectangu-
lar plate compressed by the stress Nx, which is assumed positive, uniformly
distributed along the edges x — 0 and x — a.
292

v4

Nx N>

Fig. 7.6 - Uniformly compressed simply supported plate.

The differential equation (7.5.13) is in this case


Nr.
w. + 2 to. + w. w, (7.6.1)
D,
By imposing the geometrical boundary conditions

w(x,0) = w(x,b) = w(0,y) = w(a,y) = 0 (7.6.2)

on the four edges, equations (7.5.14) furnish the natural boundary conditions

wtXX + v wfVy = 0 in x = 0 , x = a
(7.6.3)
Wtyy + V W)XX = 0 in y = 0 , y = b
so bearing in mind that wtVV = 0 along the edges x = 0, a and wjXX = 0 along
the edges y = 0,6, we have

w. = 0 in x = 0 , x = a
(7.6.4)
w
,vv — ° in y = 0 , y = b
The conditions at the corners are identically satisfied by virtue of (7.6.2). Note
that both in the differential equation (7.6.1) and the boundary conditions
(7.6.2) and (7.6.4), only derivatives of even order appear. The solution can
therefore be put in the form
A . mn x . nn y
w(x,y) Ann sin sin —-— (7.6.5)

in which m and n are any two positive integers. Equation (7.6.5) satisfies all
of the boundary conditions. By replacing w in the differential equation we
obtain the condition

r/ra7r\2 /ri7r\2| Nx /m7r\2


+
[(-) (T)J -5,{—) M--° (7.6.6)
293

Equation (7.6.6) has the solution Amn = 0 and a non-trivial solution on


condition t h a t
I '
+
\TO7r/ {—) (T) (7.6.7)

Equation (7.6.7) gives the eigenvalue Nx as a function of the two parame-


ters TO and n. To determine the critical value of iVz, it is therefore necessary
to minimise (7.6.7) with respect to the two variables. Note t h a t Nx increases
monotonically with n and therefore, for any a and 6, the minimum is obtained
for n = l. Thus we have

+
(—) u)
/rmv\2 /7r\2
(7.6.8)
\TO7r/

or, equivalently

(7.6.9)

where

jf=( m *+ijy (7.6.10)


\ a mo)
The non-dimensional coefficient i f is a function of the ratio a/b and the
parameter TO. For a given a/b it is necessary to determine the value of TO
which makes K minimum. The reader will note the analogy between this
problem and t h a t of the beam on an elastic foundation treated in Section 5.3.
The factor K can be plotted as a function of the ratio /? = a/b for different
values of TO (Fig. 7.7); for a given abscissa, the lowest curve gives the value
sought. If, for example, /? = 1, then the lowest value of K is given by the
curve TO = 1; if /? = 2, by the curve m — 2. The curve K = K(/3), of practical
relevance, is therefore that which is shown in the diagram by the solid line. In
correspondence with integer values of /? the curve assumes the value K — 4,
and stays increasingly close to this value with increase of /?. For sufficiently
long plates it can therefore be assumed t h a t K = 4 for every /?.
Note t h a t by varying /3 the shape of the buckling mode changes, depending
on the parameter TO. SO, whilst the plate always deflects with a half-wave in
the direction orthogonal to the load (n — l ) , the number of half-waves in the
direction of the load is a function of (3. In particular, if /? is an integer, then
TO = /?, and the plate deflects by forming square regions of side 6, the critical
load being equal to that of the square plate (see Fig. 7.7). It is worth noting
the analogy with the problem of the continuous beam with equal spans in
Fig. 5.16. If (3 is not an integer, then K > 4 and the plate deflects by
forming regions which are as near as possible to the square form (Fig. 7.8).
294

4
1 \flT 2 V6~ 3 Vl2~ V20" 5
/»="

Fig. 7.7 - The coefficient if as a function of (3 and m.

For /? = yjm(m + 1) there are two buckling modes corresponding to the same
load value, and therefore Nx is a multiple eigenvalue.

(b) Simply supported plate uniformly compressed in two directions


We consider the plate in Fig. 7.9 uniformly compressed in two directions
by positive normal forces Nx and Ny. The differential equation (7.5.13) is

w,xxzz + 2 w)Xxyy + wtyyyy = - — (Nx w)XX + Ny wtVV) (7.6.11)


v
f
with the boundary conditions (7.6.2) and (7.6.4).
As in the previous case, the solution is of the type (7.6.5). By replacing w
in the differential equation and requiring that this is satisfied for Amn =fi 0,
we have

Ar /m7T\2 Ar /tt7T\2 _ \fm7T\2 /M7T\


N +N =D + (7.6.12)
'{—) »{T) '\{—) (T)

j8=1 f m = l f k=4 |3 = 3,m=3,k = 4 |3=1.5 f m=2,k=4.34

Fig. 7.8 - Buckling modes of uniformly compressed simply supported plates.


295

I I I I 1 I I 1 i 1 |Ny
Nx N,

111 I f 1 1 I I 1 HNy

Fig. 7.9 - Simply supported plate uniformly compressed in two directions.

In (7.6.12) the two load parameters Nx and Nv appear. If the parameters


are not independent, it is convenient to introduce their ratio

(7.6.13)

and solve, for example, (7.6.12) with respect to Nx. We obtain

7T2Df
NX = K (7.6.14)
62

where

1 2
m +n
V (7.6.15)
K=
m
j] +Pn

The non-dimensional factor K depends on the aspect ratio (3 — a/b, on


the load ratio p and on the parameters m and n. To determine the critical
value of the load, given (3 and p, it is necessary to minimise K with respect
to m and n. For example, for a square plate and for various p, proceeding by
trial and error we find the values of ra, n and K reported in the table below
[28].
296

TABLE I -- Values of K as a function of p for the square plate

P m n K

1 1 1 2
0 1 1 4
-1 2 1 8.33

For p = 0 we again have the case treated previously; for p = 1, t h a t is,


for a uniform compression equal in the two directions, we have a reduction
of 50% in the critical load without change in the buckling mode; for p = — 1,
t h a t is, for a compression associated with a traction of equal magnitude, there
is an increase of the critical stress above 100%, accompanied by a change of
the corresponding buckling mode. The presence of traction therefore has a
stabilising effect, as suggested by mechanical intuition.
The procedure followed is onerous if calculation of the critical load for
various values of p is required, or if the two parameters Nx and Ny vary
independently. In such cases it is convenient to plot the stability domain,
which is the region in the plane Nx,Ny bounded by the family of straight
lines (7.6.12) (for ra,n fixed). The points inside the domain represent load
conditions for which the configuration of the plate is stable. Figure 7.10 shows
the domain relative to the square plate. On the coordinate axes the ratios

NX/N,

Ny/Nyc

Fig. 7.10 - Stability domain of a simply supported square plate compressed


in two directions.
297

Nx/Nxc and Nv/Nyc are shown, where Nxc and Nyc are the critical stresses
corresponding to uniform compression in one direction [10,28].

(c) Plate uniformly compressed in one direction and subjected to various


boundary conditions
In the two previous cases the boundary conditions of simple support on
the four edges of the plate suggested the solution (7.6.5). This expression
is evidently no longer suitable for solving problems with different boundary
conditions. However, if the plate is simply supported along the loaded edges
it is still possible to find a simple solution to the problem. Let Nx be the
applied load and let the plate be simply supported on the edges x = 0 and
x = a. The function

m 7T X
w = f(y) sin (7.6.16)

with m any positive integer satisfies the boundary conditions on the simply
supported edges. By substituting (7.6.16) in the equation (7.6.1) and re-
quiring t h a t this is satisfied for any x, we arrive at the ordinary differential
equation with constant coefficients in the unknown function f(y)

+
'""-KT)> [(7)'-f;(^)]/ = o (7.6.17)

where a dash indicates differentiation with respect to y. Equation (7.6.17) is


easy to solve for arbitrary boundary conditions on the edges y = 0 and y = b.
The associated characteristic equation

m 7rV TTITT Nx /m7r\2


A4-2 A2 + (7.6.18)
a J a ~D~f \~a

is biquadratic in A and has the four roots

,1/2
mn fmn Nx
Ai 2,3,4 (7.6.19)
~a~ \~a~ V £V
Note that whatever the constraint conditions on the edges y — 0 and y = 6,
we have

Nx /mn
(7.6.20)
ITf ~ \~a~
the equality sign applying only in the case of free edges. Two of the four
298

roots (7.6.19) are therefore, in general, complex conjugate. Writing


I 1/2
/mn\2 77Z7T N*_
(7.6.21)
V a J a
1/2
Nx
Or)' +
mn (7.6.22)
Q =
D

we have the four roots Ai = p, A2 = —p, A3 = iqy A4 = -iq. The general


solution to (7.6.17) can therefore be put in the form

f(y) = C\ sinh p y + C 2 cosh py + C 3 s i n g y + C 4 c o s ^ y (7.6.23)

By imposing the boundary conditions at y = 0 and y — b we arrive at a


homogeneous system of algebraic equations in the unknown constants C l 9 C 2 ,
C3, C4. The vanishing of the determinant of the matrix of coefficients leads
to a transcendental equation in p and </, from which the critical eigenvalue
Nx can be calculated. The solution to the system gives the corresponding
eigenvector of coefficients C t [i = 1 , . . . ,4) with which / ( y ) is identified and
therefore, from (7.6.16), the buckling mode. As an example we consider the
plate in Fig. 7.11, clamped on both edges y — 0 and y = b. The boundary
conditions on these edges are geometrical only

u;(x,0) = wtV(x,0) = w(x,b) = wiV(x,b) = 0 (7.6.24)

and therefore from (7.6.16)

/(0) = / ' ( 0 ) = / ( 6 ) = / ' ( 6 ) = 0 (7.6.25)

By imposing the two conditions on the edge y = 0, from (7.6.23) we get

C 4 — — C2 C3 = Cx (7.6.26)

N
Nx M^^^^ *

^mmmmmwmmmr
Fig. 7.11 - Clamped simply supported plate under uniform compression.
299

and the function / ( y ) is rewritten as

P
f{y) = Ci ( s i n h p y s i n g y ) + C2 (coshpy - c o s g y ) (7.6.27)

By imposing the two conditions on the edge y = 6 we obtain the system

I sinhpfc sing6 coshpb — cosqb / f \


q I / 1 \
(7.6.28)
cosh p b — cos q b s i n h p t H — sing 6 \C2 j 0
V P
By requiring that the determinant of the coefficients vanish we obtain the
following transcendental equation

2 ( 1 — cosh p b cos q b) sinhp6 sing6 (7.6.29)

where, in virtue of (7.6.21), (7.6.22) it is


1/2
mnb mnb Nxbi
pb — (7.6.30)
+ a \ D

1/2
mnb mnb NTb*
qb = (7.6.31)
+ a \ D

-1/2
q p 17 a \2 Nxb2 (7.6.32)
- - - = -2
[Vm7r6/ Dj

From equation (7.6.29), with (7.6.30)-(7.6.32), we can determine the non-


dimensional factor Nxb2/Df as a function of the plate aspect ratio a/b and
of the parameter m. The critical value is determined, for a given a/6, by
minimising the solution with respect to m.
The calculation is of course carried out numerically, and can be difficult;
however, there are graphs and tables which provide a great number of cases
already solved [8,23]. Figure 7.12 [28] shows some results for problems which
are frequently met. Expressing the critical load in the form

7T2Df
NX = K (7.6.33)
b2
the diagram gives the factor K as a function of /? = a/6. The curves drawn
with a solid line relate to plates simply supported on the loaded sides x = 0
and x — a, and different boundary conditions on the other edges. The dashed
curves are for plates clamped on the loaded edges x — 0 and x = a, and
300

16
'/////////////, V///////////A

A B J
14
V////////////A
'//////////////,
c D
L

»-f
Fig. 7.12 - Coefficient i f as a function of /? for different boundary
conditions.

the same boundary conditions as in the previous case on the other edges.
From the diagram we can see, under the same boundary conditions on the
longitudinal edges, that the critical load for plates clamped on the edges
x — constant is greater than that of the simply supported plate, as was to be
expected. The difference between the two values tends, however, to zero for
sufficiently large /?. This is because the longer the plate, the less the effect of
the constraints on the short edges. For /3 = 5 the two values of the critical
load, to a good approximation, can be considered coincident with the critical
load of the infinite plate.

(d) Plate subjected to shear


Let us consider a plate loaded by tangential forces uniformly distributed
along the edges (Fig. 7.13). The equation (7.5.13) takes the form
Nx
^,xxxx i & W,xxyy i W\yyyy — ~ ^ j ^ " w,xy (7.6.34)

Although it has constant coefficients, equation (7.6.34) is difficult to solve;


the unknown function, in fact, appears in the equation with derivatives of
301

Vf

Nxy

Nxyj

H
Fig. 7.13 - Plate subjected to shear.

even and odd order with respect to each variable, so t h a t solutions of the
sine and cosine type cannot be used. The only exact solutions available are
due to Southwell and Skan [l]. They are for an infinitely long plate, and are
obtained by a procedure analogous to t h a t shown in item (c).
If, for example, b is the finite dimension of the plate, since the solution has
to be periodic in x it will be of the type

w = f(y) eiXxx/b (7.6.35)

where A is an unknown parameter related to the longitudinal wave length of


the buckling mode. By introducing (7.6.35) into (7.6.34) we get the ordinary
differential equation in the unknown function f(y)

Ajr
r ATT
r + 2 ^ , ^ ] f +
T / =o (7.6.36)

where a dash indicates differentiation with respect to y. Equation (7.6.36)


has constant coefficients, with an integral of the type
b
f(y) — Jn*vl (7.6.37)

with fi real. By substituting (7.6.37) into (7.6.36) we obtain an algebraic


equation of the fourth degree in /x

b2K xy
/x4 + 2 A2 /x2 - 2 \ n + A4 = 0 (7.6.38)
7V2Df

with four roots /z l5 /X2? M3> A*4 which depend on A and on the non-dimensional
factor Nxyb2/Df. The solution to the differential equation (7.6.36) is then

f[y) = d ei*ilTy/b + C 2 e t>2 * y/6 + C 3 e t > 3 T y / 6 + C4 e t > 4 T y / 6 (7.6.39)


302

By imposing the boundary conditions on the edges y = 0 and y = b and


requiring the determinant of the matrix of coefficients to vanish, we get the
eigenvalue Nxy as a function of the parameter A. By minimising the solution
with respect to A, we can determine the critical load.
Putting the eigenvalue in the form

7T2Df
NXV = K b2 (7.6.40)

we find K — 5.35 for a simply supported plate, and K = 8.98 for a clamped
plate.
Knowing the exact solution, we shall now show how use of the Ritz method
can provide results to a good approximation with a modest computing effort.
Let us consider, for example, the infinitely long plate in Fig. 7.14, simply
supported on the y = constant edges. We assume that the nodal lines of
the buckling mode are straight and do not intersect the long edges at right
angles; using a solution of the type [31]
. n . y
(7.6.41)
w — A sin — [x + m y) sin TT -
the nodal lines are then represented by the family of parallel straight lines
given by

x + my — nt (n = 0 , l , 2 . . . ) (7.6.42)

where t is the unknown longitudinal wave lenght. The second-order term of


the total potential energy is, from (7.5.3) and (7.5.12),

+ 2 V W xx W vv + 2(1
\ P" "2 = Ss{^2 [W-xx + ^ ' ' " ") ""'I
- Nzy wtX wiV\dxdy= I I —}- [(wtXX + wiVy)2

+ 2(1 -v){w\v
^ , i i ^,yy)j ^xy ^,z ^,y j dx dy (7.6.43)

N XY

Fig. 7.14 - Infinite simply supported plate subjected to shear with straight
nodal lines.
303

For a polygonal plate simply supported along the edges we have

/ (™,2xy - w.sz wtVV) dxdy = 0 (7.6.44)

as can be demonstrated by applying the divergence theorem. Equation (7.6.43)


therefore simplifies to

i Pc" u 2 = j s [^ (wtXX + wtVy)2 - Nxy wtX wA dxdy (7.6.45)

Substituting the assumed solution (7.6.41) and carrying out integrations over
the length t equal to a half-period, we have

Ip,u2 Df ^A2
=
l + (l+ra2) + 4ra2-
2 c 8 ft3

Nxyn2A2^ (7.6.46)

In the critical state we must have P"u2/2 = 0, from which

7T2Df 1 1
N* + 6 m - f — 4- — ( 1 + m. 22 A)2 2
2 (7.6.47)
6 2
2 m 6 m m £

To obtain the smallest eigenvalue it is necessary to minimise with respect to


[b/t)2 and to m. By minimising with respect to (b/t)2 we obtain {b/l)2 =
1/(1 + ra2) from which

1\ 7T2Df
Nxu = 2[2m + (7.6.48)
mj b2

Minimising again with respect to ra, we find ra = \ / 2 / 2 and therefore

7T2Df
Nxy = \\[2 (7.6.49)
b2
By comparison with (7.6.40), we find K = 5.66 which differs by less t h a n 6%
from the exact solution.
Note t h a t the function (7.6.41) satisfies the geometrical boundary condi-
tions, as required by the Ritz method, but not the mechanical conditions
w
,vv = 0 along the simply supported edges. This is due to the fact t h a t the
nodal lines do not intersect the long edges at right angles and they therefore
provide a certain restraint to rotation. The assumption of curved nodal lines
such as those shown in Fig. 7.15 permits us to obtain even better results. The
chosen function is now
n n
y (7.6.50)
w = A sin — [x + <p[y)\ sin ——
304

Nxv
tP\j;
t "~

Fig. 7.15 - Infinite simply supported plate subjected to shear with curved
nodal lines.

The nodal lines are represented by the family of curves

x + <p(y) =n£ [n = 0 , 1, 2 . . . ) (7.6.51)

and the function <p must be such that d(p/dy = 0 on the edges of the plate.
We can assume
d(p . ny ,
-j- = - m s i n -p (7.6.52)
ay b
and determine (p by integration
mb t ny \
<p= (cos -y - 1 j (7.6.53)

Proceeding as in the previous case we find K = 5.41, with an error of only


i%.
The cases examined so far relate to an infinitely long plate and can be used
in applications for sufficiently large values of the ratio a/b. For low values of
the ratio a/b it is not possible, in general, to obtain closed form solutions and
it is therefore necessary to use approximate solutions.
Figure 7.16 [7] shows the factor K appearing in (7.6.40) as a function of the
ratio 1//3 — b/a for a simply supported plate (a) and for a clamped plate (b).
The solid curves represent the solutions obtained by the authors indicated in
the figure; the curves shown by dashes are parabolae which approximate the
"exact" solution and are given by

K = 5.34 + ± (7.6.54)

for the supported plate and

K = 8.98 + -j- (7.6.55)

for the clamped plate. For (3 —> oo the expressions (7.6.54) and (7.6.55) tend
to the exact solutions for the infinitely long plate.
305

100
I
u
934
Seydel-^
90
Nxy
r n
80 I:* h f—
i
|Nx,
r
//

-^-SteirvNeff
70

6.0
^ -Parabol,a 534*^2

Z£^'
^o P
534
<
*S5iJthwel-Skan —►
5.0
0.2 0.4 0.6 0.8 1.0 0.2 0.4 Q6 08 1.0

1/0= b/& Vjj=b/a

Fig. 7.16 - Coefficient i f a s a function of /? for (a) simply supported plate


and (b) clamped plate.

7.7 CRITICAL LOAD OF S T I F F E N E D PLATES


The critical load of a plate, whatever the conditions of constraint or load,
is proportional to the flexural rigidity Df = Eh3/12(1 - is2), as can be de-
duced from equation (7.5.13). It can therefore be increased by increasing the
thickness, which means an increase in weight and in the cost of the struc-
tural element. A more advantageous technical choice is t h a t of keeping the
thickness small and reinforcing the plate with stiffeners [7,10,28].
With the aim of understanding the mechanical behaviour of the stiffened
plates, we consider the simple example illustrated in Fig. 7.17. The plate

Fig. 7.17 - Stiffened plate.


306

is supported along the edges, compressed in the longitudinal direction x and


stiffened by a longitudinal beam situated at mid-span.
We consider first the limit case in which the flexural and torsional rigidi-
ties of the beam are negligible. The system is then reduced to t h a t already
studied, consisting of an unstiffened plate: the critical load is inversely pro-
portional to 6 2 , according to a factor K dependent on the ratio a/6, and the
buckling mode has only one half-wave in the direction y.
Let us now consider a second limit case in which the flexural rigidity of
the beam is very large whilst the torsional rigidity remains negligible. Under
these conditions the beam behaves as a simple support; the buckling mode
has two half-waves in direction y and the critical load is inversely proportional
to (6/2) 2 , according to a factor which depends on the ratio 2a/6. The critical
load of the system is therefore about four times larger than the value found
in the preceding case.
Apart from the two limit cases just considered, it is of interest to investigate
the behaviour of the stiffened plates for finite values of the inertia of the
stiffeners.
The formulation of the problem requires writing the total potential energy
of the system obtained by adding the energy of the beams to t h a t of the
plate. By imposing the compatibility conditions on displacements and rota-
tions along the junction lines between each beam and the plate, and requiring
the vanishing of the first variation of the second-order terms of the energy, we
get the system of differential equations and the related boundary conditions
which govern the problem.
If, for simplicity, we suppose t h a t the cross-sections of the beams are sym-
metric and that the centroid lies on the middle plane of the plate, then we
have u = v = 0 and we arrive again at a differential equation in the sole un-
known w. Even for simple problems, however, integration of the differential
equation is very hard, so it is preferable to use approximate methods such as
the Ritz method.
As an example, suppose we want to find an approximate solution to the
problem in Fig. 7.17. The procedure may be extended to the case of more
stiffeners, which is treated in [9].
Let us assume as a solution the function
rrtTT X / . 7Tt/ . 2 7T J/ >
w = sin (AX sin ^y + M sin -y-\ (7.7.1)

which satisfies all boundary conditions. With the situation (7.7.1) the number
of degrees of freedom of the system is reduced to two, represented by the
unknowns A\ and A2; m is left as a parameter. The total potential energy of
the plate, up to second-order terms and taking account of (7.6.44), is

\P>:Pu2 = [f [5t {Wxx + Wyvf - !± „£ ] dxdy (7.7.2)


307

Similarly, the total potential energy of the beam, neglecting the torsional
term, is

i *■•-£[¥<-«■ =»/» 2
( \*
6/2 dx (7.7.3)

where El is the flexural rigidity and F the axial force acting in the funda-
mental configuration. From (7.7.1), by differentiating, we have

(w(xx + w,w)2 = sin 2


mn x
h (-^-+^jsm -b
fm27r2 4TT2\ . 2 7r 1
+
2 m 2 7T2 2 m TTX / . Try . 2?rt/\2
w ^ — —-— cos Ai sin ——h A2 sin —:— 1 I7-7*4]
a V 6 by
4 ^4
ra 7r mix x
(w.x*)l= 6/2 sm

ra2 7r2 2 mux


K*)y=6/2 = —-— cos

Bearing in mind t h a t

fa fb . 2 Pn x . qny . r7ry c ab
— sm —-— sin ——- ax ay = o^ — (7.7.5)
/ / sin b b 4
Jo Jo a
where ^ is the Kronecker symbol, the total potential energy (7.7.2) of the
plate is given by

1 ^n o ab ^ m 7T TV m 2 7T 2 4 7T 2 V
- p" u2 = — Df A\ + ^r\ +Ai
2 cp 8 f
ad b2

ab , T m 27rJl
— N. (A{ + At) (7.7.6)
a*

and t h a t of the beam from (7.7.3) by

\n*
1 an^A,(n^_EI_F,
4 a2 {""-') (7.7.7)
308

The total potential energy of the system is the sum of the two contributions
(7.7.6), (7.7.7).

m 2 7r 2
^ >e
2
2
= ^
M 8
~
**»'I £ + £)'-« x
2

a m 2 7r2 / „ m 2 7r2 \ 1
+
ab 4r. /m* 4\2 Ar m 2 7r 2
(7.7.8)

The energy expression (7.7.8) is a diagonal quadratic form. In order to be


positive semidefinite, the coefficients of A\ or A\ must vanish. By equating
the coefficient of A\ to zero, we obtain

*(»>&)-¥$+$ {$+>¥{")***
By requiring the coefficient of A\ to vanish we have

7r2 Df (mb 2a
Nx = 4 — 2^ — (7.7.10)
6 2a mb

Introducing the non-dimensional quantities

El A_
/? = 7 = b~D~, 6 = (7.7.11)
bNT, ~bh
where A is the area of the cross-section of the beam, equations (7.7.9) and
(7.7.10) may be written as

7T2Df 1 m 0_ m\
N™ = + 27 (7.7.12)
62
1 + 26 J m

7T2Df m 2/3
JVJ2) = 4 (7.7.13)
b2 \2~J3 +
~m~

where we have marked the two different solutions with a superscript. Equa-
tions (7.7.12) and (7.7.13) furnish the two eigenvalues iVj1) and N^ as func-
tions of the aspect ratio /?, of the non-dimensional rigidities 7 and 6, and of
the parameter m. By first minimising with respect to m and then choosing
the smaller of the two solutions, we obtain the critical load.
309

The buckling modes corresponding to the two eigenvalues are, respectively


mnx . ny . . .
w\ — A\ sin sin —— (7.7.14)
a b
mnx . 2 ny ,nnir\
w2 = A2 sin sin —— (7.7.15)
a b
t h a t is, the buckling mode has one half-wave or two half-waves in direction y.
This follows from the approximation (7.7.1). By considering a larger number
of terms in the series expansion, we determine buckling modes which are a
linear combination of sinusoids with an odd number of half-waves, and modes
consisting of a single sinusoid with an even number of half-waves. In the
present case, if the deflection is w\, the beam bends according to a sinusoid
and the corresponding critical load depends on 7 and 6; if the deflection is
u>2, the b e a m remains straight and the critical load depends only on /? and
coincides with t h a t of the plate of width 6/2.
Let us now examine the behaviour of the system as the rigidity of the beam
varies. We fix, for example, £, and increase 7 from very small initial values.
At first we get N^ < N^ and the buckling mode consists of a half-wave
in direction y. With increase in 7 the critical value of the load increases,
in t h a t the rigidity of the system increases. When JVW = N?\ the two
buckling modes (7.7.14), (7.7.15) coexist. A further increase in rigidity leads
to NJ?) < NJ^\ and the deflection is represented by (7.7.15) only. The beam
remains straight whilst the plate bends, forming two half-waves in direction
y. N]f) then represents an upper bound for the critical load whose value, in
agreement with (7.7.13), depends on the parameter /?.
It seems interesting at this point to determine the values of 7 and 6 which
give N^ = N^. From this we obtain the maximum value of the critical
load with a minimum use of material. In reality, such considerations arise
from study of the critical state of equilibrium and ignore the mechanical
behaviour of the system in the post-critical range. In all cases studied so far,
it has been verified that the occurrence of simultaneous buckling modes under
the same critical stress causes instability in the post-buckling range even if
the single buckling modes are associated with stable post-critical behaviour.
This has been observed in particular in a study of the model of two degrees
of freedom presented in Section 2.12. Therefore, those values of 7 and 6 to
which N^ = NJ?) corresponds, and which apparently optimise the structure,
are to be avoided.

7.8 I N F I N I T E CYLINDER S U B J E C T E D TO LAT-


ERAL P R E S S U R E
Consider a cylinder of circular cross-section and infinite length subjected to
a constant lateral pressure p always orthogonal to the surface of the cylinder
310

tn p=const

Fig. 7.18 - Infinite cylinder subjected to hydrostatic pressure.

(hydrostatic pressure) (Fig. 7.18), and let us evaluate the critical load and
buckling mode. To do this, we write the total potential energy up to second-
order terms. Note that as the problem is independent of x, then we have from
(7.3.13), (7.3.14)
w (7.8.1)
en = 0 , e22 = v. + e12 = 0
R '

Xi=0, X2 = - W,38 + Xl2 = 0 (7.8.2)


R '
having used (7.2.18) and the relations Ri = oo, R2 — R and having put a\ =
x, a2 = s. It is assumed, as is simple to verify, t h a t u(x,s) = 0. The term for
the state of stress in the fundamental state St-yCdj£'wj£'/2 reduces to S^c^s /2>
as all the components of the tensor St-y = 0 (i, j ^ 2). Equilibrium equations
immediately give S22 — —pR/h and also, from the second of relations (7.3.12),

U>23 w.» - — (7.8.3)

By using (7.4.6) and the results above we can write, up to second-order terms,
pR
-W"uc 2
2
=
~2~ ) ds +
T n?> + -)ds

EL f ds (7.8.4)
+
2 Jo
The potential energy of the external loads which, in the case of lateral pres-
sure, contains second-order terms, is now to be evaluated. It can be shown
that the hydrostatic pressure p is a conservative load and t h a t the potential
energy is equal to the product of p with the difference AV between the vol-
ume of the cylinder in the adjacent and fundamental configurations. Such a
change of volume per unit length is [28]
r2*R
AV = / w+ + + W V,i vw, ds (7.8.5)
./o ~R
311

From this we obtain the second-order term of the potential energy of the
pressure p
1
n,* u 22 = - -P /f2irR
- y"
w v
—- + — 4- w v s - v w A\ Jds (7.8.6)
2 c 2Jo \R R '* '7
If we put
,. A 2ns 2ns
v ; — — cos
vis) w(s) = A sin (11 = 1,2,...) (7.8.7)
2 R
we arrive at
1
^ P cc" u 2 = nA2
2 R3 V 2/ 2R K
' 2 V 2/

-f( l "-i)] P-8-8)


The value of p which makes (7.8.8) positive semidefinite is

Pc = 2n 2n (7 8 9)
^ { ~l) --
and reaches a minimum in correspondence with n = 1. Note t h a t in this case
(7.8.7) describes an inextensible mode. We have
3Df
Pc = (7.8.10)
i23
The case n = 1 corresponds to ovalisation of the cylinder (Fig. 7.19). With
such a type of deformation, the energy term due to the hydrostatic pressure
plays an essential role due to the effect of decrease of the cross-sectional area
in passing from the circular shape to the inextensible oval shape [2]. If this
term is ignored we obtain
4Df
(7.8.11)

Fig. 7.19 - Ovalisation of the section.


312

with an error of 33%. Note that with increase of n this error tends to dis-
appear as AV approaches zero, and therefore the contribution to the energy
of the external pressure becomes negligible relative to the contribution of the
stresses 5 tJ -. Such contributions are given in explicit form by the fourth and
third terms on the right of equation (7.8.8) respectively, and their ratio tends
to zero if n —> oo. Therefore, in all problems in which n is a large number,
we can always ignore the term expressing the potential energy of the external
pressure, without introducing an appreciable error in the results.

7.9 CYLINDER OF FINITE LENGTH SUBJECTED


TO LATERAL PRESSURE
The problem of the determination of the critical load of a cylinder of finite
length subjected to lateral pressure is more complex than t h a t examined in the
previous Section, in that the variable x is present in the problem. However, a
great simplification can be made if we accept, as is evident from experimental
results, t h a t the phenomenon of instability is manifested with a large number
of corrugation waves. This fact allows us to use the kinematic relations of
shallow shells and also to neglect the potential energy of the external pressure
relative to the elastic energy associated with the state of pre-stresses 5 t y.
By using (7.2.18) and remembering that Rx — oo, R2 = i2, we obtain from
(7.3.13), the first of (7.3.20) and (7.3.21)
w
e n = utX , e 22 = vtS + — , 2 el2 = ut8 + vtX (7.9.1)

X l = ~ ™,xx , X2 = - W,33 , Xl2 = - W,Z8 (7.9.2)

u4? - wtS (7.9.3)

having put a\ = x, a2 — s. Let us assume that the cylinder is constrained


at its ends by thin plates which prevent radial and tangential displacements,
but permit the rotation wtX and the axial displacement u. This type of con-
straint, which we shall call simple support, implies t h a t in the fundamental
configuration we have a membrane state of stress in the whole cylinder, with
the exception of small regions around the edges where the cylinder is bent, as
can be seen from the deformation indicated in Fig. 7.20(a). For simplicity of
analysis, we shall ignore the presence of these regions (Fig. 7.20(b)) and take
the only non-zero pre-stress to be constant for the whole cylinder and given
by

(7.9.4)
313

w
p = const J

o) b)

Fig. 7.20 - Finite cylinder subjected to hydrostatic pressure.

The error introduced by this simplification, which leads to differential equa-


tions with constant (instead of variable) coefficients, is discussed in [28].
By using (7.4.6) and (7.9.1)-(7.9.4), we write the total potential energy up
to second-order terms as

",,2 _ pR
P'u j w2sdS +\nejs \u2x + v2s + 2i/utXvt8
~2
1-v 1-v w
+ < +
V
,x + (1 - " K . V | X + ^ J

+ 2 ^ v>3 + 2 v ^ tt|X] dS + -Dfjs [{wiXX + wt88)2

- 2(1 - is){wiXX wi3S - w2xs)] dS (7.9.5)

For determination of the critical load and buckling mode we solve the varia-
tional problem in the form (4.5.5), which from now on we shall prefer to the
form (4.5.13) as the absence of the dot on the variables simplifies the writing.
We have the equilibrium equations

1-v 1+ v
-De[u V
XX I 7T ^ , S3 i ~ U.X8 I
R "-) = °
- De lvt33 + — — viXX + ——- utX3 + — wt3J = 0 (7.9.6)

A De fw \
DfV4w + — ( ^ + V.* + I/M
.«) +pR*>%89 = 0

where V 4 = d4/dx4 + 2d2/dx2ds2 + d4/ds4, with the boundary conditions


314

v{0) = 0 , v{i) = 0 , w{0) = 0 , w{t) = 0

D, (u, +- ,+|i)=° (7-9J)

to be satisfied in x — 0,£. T h e third of equations (7.9.7) represents t h e


vanishing of the bending moments at the ends of the cylinder and is analogous
to t h a t of plates, whilst the second of (7.9.7) represents t h e vanishing of the
stress Nx = oxxh generated by t h e deformation. Equations (7.9.6) are t h e
Donnell-Mushtari-Vlasov (DMV) equations of the shallow cylinder [29]. For
R —► oo these are reduced to t h e equation of t h e plate. Such equations
are coupled, though Donnell has demonstrated t h a t they can be partially
uncoupled and rewritten in the form [12]

V 4 u - — ( - v wtXXX + wiX33) = 0
K

V 4 v + i [(2 + v) w>XX3 + wtSS3] = 0 (7.9.8)

o Eh A
Df
~R2" ™,*zxz + p i 2 V V a 5 = 0

where V 8 = (V 4 ) 2 . Let us write the solution to t h e third of equations (7.9.8)


in the form
mnx ns
w(x,s) = wmn sin — — sin — (7.9.9)
with wmn as arbitrary coefficients. The solution satisfies the third, fourth and
last of the boundary conditions (7.9.7) and is also periodic in s. T h e first two
of equations (7.9.8) also require t h a t
mnx ns
u{x,s) = umn cos —-— sin —
l K
(7.9.10
, x mn x ns
v(x,s) = vmn sin — — cos —

be associated with (7.9.9). T h e coefficients umn and vmn can be expressed


in terms of u; mn by means of linear relationships, by substituting (7.9.9) and
(7.9.10) in the first two of (7.9.8). It is then easy to check that (7.9.9) and
(7.9.10) satisfy t h e first two and t h e fifth of the boundary conditions (7.9.7).
By substituting (7.9.9) in the third of equations (7.9.8), we obtain the critical
value of the lateral pressure
(m 2 + /? 2 ) 2 12 Z2
(7
*■=—w - —r^^ 1 +
-"I)
315

having p u t

Rl2 (7.9.12)
n2D,

nl
P = VR Rh
VI (7.9.13)

The parameter Z is known as the Batdorf parameter. It is obvious that for


m = 1 equation (7.9.11) reaches its minimum, and we then have

_ (1 + /? 2 ) 2 , 12 Z2
Pc= 55 + 7T 4 /? 2 (l+/? 2 ) 2 (7.9.14)

It follows t h a t a circular cylinder subjected to external pressure always buckles


with only one half-wave in the axial direction. By minimising pc with respect
to j3 we obtain pc as a function of the only parameter Z. T h e results of this
analysis are shown in Fig. 7.21 [6]. For relatively long cylinders (Z > 103)
the analytical procedure of minimisation is not very accurate, as n is small.
This means t h a t /? cannot be treated as a continuous variable and the DMV

1000

t-fip*
P.

100

k
10

I 1 1 1 1 1 1 1 1 1 1 1 Mill 1 1 11 1 1 1 1 i | 1 1 l i 11

10 100 1000 10 000


z _/2 [z-v^
Rh

Fig. 7.21 - Critical load for cylinders subjected to lateral pressure.


316

equations valid for shallow shells are no longer acceptable. In particular,


taking into account (7.9.13), we have in this case

Pc-P2 = ^ CM*)
and assuming n = 2, as happens with infinite cylinders, we obtain

4£ 2
(7.9.16)
y
' 7T2 R2

From (7.9.16) we deduce that, by virtue of (7.9.12),

P. = ^ (TA17)

which, when compared with (7.8.10), shows an error of 33%. For short cylin-
ders, by assuming Z ~ 0, we obtain

(1 + /? 2 ) 2 [
Pc - -^± (7-9.18)

from which, by minimising with respect to /?, we obtain pc — 4 for (3 = 1.


From the first of (7.9.13) we get n — 7ri2/£, which indicates that in short
cylinders the circumferential waves have a half-wavelength equal to the height
of the cylinder. This is analogous to the case of a long plate subjected to
compression along the short side.

7.10 CYLINDERS S U B J E C T E D TO AXIAL P R E S -


SURE
The analysis of cylinders subjected to axial pressure proceeds in a way parallel
to t h a t in the previous Section. Confident of the experimental observations,
we continue to use the kinematic relations for shallow shells. By supposing
t h a t the cylinder is constrained by two thin plates at its ends, as previously,
the state of stress in the fundamental configuration will not be purely of mem-
brane type in that close to the supports, the cylinder deforms as indicated in
Fig. 7.22(a), giving rise to bending stresses. As before, we assume a mem-
brane state of stress in the whole cylinder with a deformation characterised
by uniform dilation (Fig. 7.22(b)). This greatly simplifies the analysis for the
determination of the critical load in t h a t the differential equations of DMV
which govern the problem have constant, rather than variable, coefficients.
The error in the results that this simplification entails is discussed in [28].
By neglecting the effect of dilation in the fundamental configuration, we
have Sn — —p, with p the external axial pressure, whilst all the other stresses
317

p=const.
f

w
b)

Fig. 7.22 - Cylinder subjected to axial pressure.

vanish. The energy due to the state of pre-stress is therefore, from the second
of relations (7.3.20),

U'«
Pk
u[f f 2 ,c
'13 dV 2 Js ,x (7.10.1)

whilst the elastic strain energy is the same as in the previous case. The
equilibrium equations of the problem are therefore

+ V V
- De \uiXX H — u}SS + - W +
Y~ - RW
^ / l-i/ l + i/ 1 \
- De \vtSS H — v>xx + — — utXS + — wi8\ = 0 (7.10.2)

De fw \
Df V 4 w + —- I — + vtS + vutXJ - ph wtXX = 0

with boundary conditions furnished by (7.9.7). The first two of equations


(7.10.2) are respectively identical to the first two of (7.9.6), whilst the third
differs from its correspondent in (7.9.6) in the last term. With a procedure
analogous to t h a t of the previous Section, we can transform the first two of
(7.10.2) into the corresponding equations of system (7.9.8) whilst the third
becomes

r. ^ 8 Eh
7 w xxxx -\-phV4wtXX = 0 (7.10.3)
~R2~ >

Let us put the solution to this equation in the form


mn x ns
W[X,S) = Wmn Sin — Sill — (7.10.4)

with wmn as arbitrary coefficients. Equation (7.10.4) satisfies the third, fourth
318

and last of the boundary conditions (7.9.7) and is also periodic in s. By


substituting (7.10.4) in (7.10.3) we have

(m2+/?2)2 12 Z2m2
1
Pc = ~ r- - + 4/ 2 , /ms 7.10.5
m2 7T4(ra2 + /? 2 ) 2 v
'
having written
hi2
(7.10.6)
n2Df
and with /3 and Z given by (7.9.13). To each value of m and n there cor-
responds a different eigenvalue pc. For cylinders of average length we can
obtain, to a sufficient approximation, the smallest value of pc by minimising
(7.10.5) with respect to (m 2 + /? 2 ) 2 /m 2 . We have

4N/3 0 ^ - ^ ( 7 1 0 ? )
v = V1 — vL
Pc
7T2 Rh
that is
E h
(7.10.8)
>/3(l-«/J) ^

corresponding to
I2
(7.10.9)
' Rh
where we have used the expression for Z in (7.9.13). For v — 0.3, (7.10.8)
gives
Eh
pc = 0.605 — - (7.10.10)
XL

If we take the square root of the first and second members of (7.10.9), we
obtain
m2{nR/t)2 + n2
yj2cR/h (7.10.11)
m(n R/l)

with c = \ / 3 ( l — J / 2 ) , having used the first of equations (7.9.13). Equation


(7.10.11) is that of a circle (Fig. 7.23). This shows t h a t a cylinder subjected
to axial pressure, in correspondence with the critical load, manifests a great
number of buckling modes simultaneously [31]. This fact leads to catas-
trophic consequences in the behaviour of the cylinder in t h a t the equilibrium,
corresponding to bifurcation, is strongly unstable and the bifurcated curve is
asymmetrical and unstable [5]. The reader will remember t h a t in the anal-
ysis of the model of two degrees of freedom presented in Section 2.12, the
319

mnR/i

Vic^/h

Fig. 7.23 - Diagram of function (7.10.11).

1000
/

[Mc /2h Mc

100

I
10

y^ i i i i 11 ii i i i i 11 I .I i i i i i 1 11 1 1 ■ i ■r1 l

10 100 1 000 10 000

Rh

Fig. 7.24 - Critical value of the pressure for a cylinder subjected


to axial load.
320

occurrence of two simultaneous buckling modes led to unstable post-critical


behaviour even though each of the two modes is stable. In cylinders subjected
to axial pressure, the presence of a large number of simultaneous modes mag-
nifies the phenomenon, giving rise to a bifurcated equilibrium curve with a
strongly inclined tangent at bifurcation. It follows t h a t for very small initial
imperfections, we have a collapse load p* which varies between 25% of pc for
industrially produced cylinders and 80%-90% of pc for cylinders produced in
specialised laboratories.
As m and n are positive numbers, it is impossible to satisfy (7.10.11) for
Z < 2.85. Thus it is necessary to use (7.10.5) and solve it by trial and error.
The calculations are simplified by use of the diagram in Fig. 7.24 [6]. For
Z > 2.85 the values of pc furnished by the diagram and by (7.10.7) coincide.
When the radius of the cylinder tends to infinity and Z tends to zero, we
have from the diagram pc = 1, that is, from (7.10.6)

Pch = —^ (7.10.12)

If the cylinder is very long, the buckling mode is that of a beam. The Donnell
formulation leads in this case to a wrong result, in that the critical load is
that of the Euler beam where I = nhR3.

7.11 CYLINDERS S U B J E C T E D TO T H E A C T I O N
OF C O M B I N E D LOADINGS
In this Section we shall examine the problem of determining the critical load
for a cylinder of finite length subjected to the action of lateral pressure pi
and axial pressure pa. The treatment will necessarily be concise, given the
similarity of this case to the problems treated in the two previous Sections,
and therefore only the essential points of the analysis and the results will
be discussed. With reference to the theory of shallow shells, the first two of
equations (7.9.8) remain unaltered, whilst in place of the third we have

F1 h
D y 8 w y 4 w
f + -HT w."" + (PtR w
,°° +P«h ,xx) = 0 (7.11.1)

Equations (7.9.7) remain as the boundary conditions. Putting Sn/S22 — «,


we have

= a (7 1L2)
T T -
Pi h
having used (7.9.4) and Sn = — pa. By virtue of (7.11.2), equation (7.11.1)
321

can be rewritten as

D/Vsw + — w>xxxx + RptV4 (wt8S + awtXX) = 0 (7.11.3)

Assuming a solution of the type (7.10.4) which, as seen previously, satisfies


the appropriate boundary conditions, we arrive at
2 12 Z2
K + /?2\2
) (7.11.4)
am2 +/?2 +
"(*♦£)■ ( a m 2
+ /? ) 2

where pc is given by (7.9.12). For a = 0, (7.11.4) leads back to (7.9.11).


Fixing the value of a, we obtain the value of pc corresponding to a pair of
values m and n. The minimum eigenvalue is obtained by trial and error. A
particularly interesting case is represented by a = 1/2. From (7.11.2) we have
Q = 7vR2pi, where Q — 2nRhpa is the total axial load. This case corresponds
to the cylinder subjected to pressure pt acting uniformly on the lateral surface
and at the ends. The results given by (7.11.4) for a = 1/2 are shown in Fig.
7.25 [6] as a function of the Batdorf parameter Z. Also represented in the
diagram, for comparison, are the results for lateral pressure alone and for
axial pressure alone.

1000

\ /2R
100

[ /
y \ a x ial
pr essure

/lateral
pressure.
10
[ /

I un form pressi ire


_i i i i i 1111 i i i t 1111 i i i i i 111 i i i 11111

10 100 1000 10000


Z=il(1-
Rh

Fig. 7.25 - Comparison of critical loads due to uniform lateral


and axial pressures.
322

7.12 COLLAPSE OF CYLINDRICAL SHELLS

The problem of the post-critical behaviour of cylinders will not be treated


in this volume. We shall limit ourselves to describing the phenomenon and
showing some experimental results.
Cylinders subjected to the load conditions examined in the previous Sec-
tions, and those subjected to torsion, all have unstable asymmetrical post-
critical behaviour of the type illustrated in Fig. 2.14 for a model of one degree
of freedom. The slope of the tangent to the bifurcated curve at the bifurca-
tion point depends essentially on the type of load, as well as on the boundary
conditions and on geometric properties. Under axial load the worst situation
arises, due mainly to the fact that in correspondence with the critical load,
different buckling modes occur simultaneously. There is a very high sensitiv-
ity to initial imperfections, with a drastic reduction in the collapse load with
respect to the critical load. In laboratory tests in which the collapse load is
measured, a large dispersion of results is obtained due to the random distri-
bution of initial imperfections. The results are also well below the valu,e of the
critical load, for the above reason. Figures 7.26-7.29 [8] show the theoretical
curves for the critical load of a cylinder subjected to various load conditions as
a function of Z , together with the experimental results for the collapse load.
It is obvious that the largest discrepancy between the two types of results is
for the cylinder subjected to axial pressure. The results of the diagram in Fig.
7.26 refer to a cylinder with built-in ends, those of Fig. 7.27 to a cylinder of
infinite length, and the other diagrams to a simply supported cylinder. The
value of fc on the ordinates in Fig. 7.27 is defined by rc — (f.2h/7T2Df)rc, Z
is provided by the second of relations (7.9.13) and the pc quantities of Figs.
7.26, 7.28 and 7.29 have been defined in the respective Sections. Note that
the diagrams are on logarithmic scales and therefore mitigate the differences
between the theoretical and experimental results.
In designing cylinders, we take the theoretical critical load as a basis and
multiply it by a reduction factor evaluated empirically for each type of load.
The value of this factor depends on the average difference between the theo-
retical and experimental results and on the range of the dispersion of exper-
imental results.
Let us determine, for example, the factor of reduction for a cylinder sub-
jected to axial compression with Z > 2.85. For such a value of the Batdorf
parameter, we have from (7.10.10) pc = 0.605 Eh/R, whilst the experimental
results in Fig. 7.26 indicate that it is necessary to use a coefficient much lower
than 0.605. For this purpose we make use of the diagram in Fig. 7.30 [28],
where we have the experimental results of the collapse load as a function of
R/h together with an interpolation curve recommended to designers. For
R/h — 500 we read on the ordinates the factor 0.24 which, multiplied by
Fig. 7.26 - Comparison between theoretical and experimental results for a
cylinder subjected to axial pressure.

1000

100

Fig. 7.27 - Comparison between theoretical and experimental results for a


cylinder subjected to torsion.
1000

100

Fig. 7.28 - Comparison between theoretical and experimental results for a


cylinder subjected to lateral pressure.

1000

100

Fig. 7.29 - Comparison between theoretical and experimental results for a


cylinder subjected to hydrostatic pressure.
325

1.1

1.0

Pcs = exp. critical pressure


net Pct = theor. critical pressure
0.8J = 0.605 Eh/R

0.6

0.4

0.2

500 1000 1500 2000 2500 3000 3500


R/h

Fig. 7.30 - Distribution of experimental results for a cylinder subjected


to axial pressure.

0.605, yields pc — 0.15Eh/R, a reduction of 75%. We shall not go any further


into the study of cylinders, on which thousands of works have been written.
The classical results are presented here for some of the more common types of
load conditions. Numerous other results for the determination of the critical
load can be found in [8] and [23].
326

7.13 S P H E R E S U B J E C T E D TO HYDROSTATIC
PRESSURE
The analysis presented for the determination of the critical load for a sphere
subjected to hydrostatic pressure is based on the hypothesis, supported by
experimental results, that ELS a result of the buckling mode characterised by
short half-waves, the sphere is transformed into a facetted surface with plane
faces so small that they can be confused with the curved surface t h a t they
subtend. This allows us to write the kinematic relations in the form
w W
en = utX - R ' ^22 = V%y - 2 e 1 2 = utV + vtX (7.13.1)
R '

Xi = - w f S X2 = ~ W i X12 = - w xy (7.13.2)

w
M W ^23 = W
A (7.13.3)
13 = ,x ,

Equations (7.13.l)-(7.13.3) are analogous to (7.9.l)-(7.9.3) derived for the


cylinder, except for the addition of the term w/R to the expression for e n
and the introduction of the rotation u[z3'. The coordinate system adopted has
the axis z orientated inward (Fig. 7.31). This explains the presence of the
minus sign in the first and second of relations (7.13.1) in place of the plus
sign which appears in the second of (7.9.1). The stress in the fundamental
configuration is

1 R
Sn — 522 — P (7.13.4)
2 h
where p is the uniform hydrostatic pressure. As a result of the hypotheses
on the buckling mode, the second-order part of the energy of the external
pressure is negligible in comparison with the elastic energy associated with

Fig. 7.31 - Area element of a sphere.


327

the state of pre-stress (7.13.4). By using (7.4.6) and adding the contribution
of (7.13.4), we can write the total potential energy up to second-order terms

2 c
2 /.[(•--s)'+("-s)"
+ 2u (u, s - ^ ) (i>,„ - | ) + 2(1 - i/)(tt,v + v,x)2 dxdy

+ - y / s [«£, + wfw + 2i/ w>xs to,,,, + 2(1 - v) w ' J dx dy

- - / a h (w2x + w2y) dx dy (7.13.5)

From (7.13.5) come the differential equations of the problem


W
T^ \ 1—1/ 1 + 1/ , \ X]
De \uiXX + —j- utVy + - y ~ vtXy - (1 + v) -^- I = 0
_ [ 114-1/
+ 1/ 1 — 1/
l-i/ . , \ wwtV y]
Df I ~Y~
—z— uu™TAI+H~Y~:—V'xx
v.M+ +V'vvv.™ + (7.13.6)
~ -' (1 +^ v\~£~—±r I = 0

DfV*w +^ [ 2 (1 + 1/) ^ - (1 - i/)(u , + v,) + <jh(wtXX + wtVy) = 0

As a solution we assume
u x cos cos
{ iV) — A ^^ 0V
v(x,y) = J5 sin ax sin /?t/ (7.13.7)
iy(x,y) = C sin a x cos /?y

with
mn nn
a = , /? = (7.13.8)
R ' ^ 72
By substituting in the first two of equations (7.13.6) we obtain

a 2 + ± (1 - „)/?2 A + l [l + v)aPB + {l + v) § C-0


2 XL
(7.13.9)
2 2
{l + v)aPA + i(l-z/)a + /? ]j3 + (l + l / ) | c = 0

from which

R a 2 + P2 ' (!+")£i? a2+/?2


(7.13.10)
328

From the third of (7.13.6) we have, on the other hand,

E Eh2
oc =
i2 2 (a 2 + /?2) + 12(1 - ^ 2 ) ("2+/H (7.13.11)

having used (7.13.10). By minimising (7.13.11) with respect to (a2 + P2) we


arrive at
Eh
Or = (7.13.12)
~cR
with

c = ^ 3 ( 1 - i/») (7.13.13)

from which
_ 2Eh2
Pc (7.13.14)
~~c~W
having made use of (7.13.4). In order for the material to remain in the elastic
range we must, from (7.13.12), have R/h of the order 1000 for E — 2.1 X
10 6 kg/cm 2 and v — 0.3. The associated external pressure, from (7.13.14), is
pc ~ 2.6 kg/cm 2 . The value of (a 2 + ft2) to which the minimum eigenvalue
(7.13.12) corresponds is given by

2 , 2 2c R
(7.13.15)
** ~h
where we have used (7.13.8).
Equation (7.13.15) represents a circle with centre at the origin and radius
equal to y/2c/Rh (Fig. 7.32) and therefore, like the situation of cylinders
subjected to axial pressure, a large number of buckling modes occurs simul-
taneously in correspondence with the critical load [31]. In his analysis of the
post-critical behaviour of a sphere subjected to hydrostatic pressure, based

V2cRAr2h

Fig. 7.32 - Diagram of function (7.13.15).


329

a
R
b

u
i a
H-

I I I I I I I I I I

Fig. 7.33 - Plate on elastic foundation subjected to uniform pressure.

on more rigorous kinematic relations than (7.13.1)-(7.13.3), Koiter [20] shows


that the equilibrium at bifurcation is unstable and t h a t the post-critical be-
haviour is asymmetrical unstable.
Reissner has shown t h a t a simply supported plate on an elastic foundation
of the Winkler type, subjected to uniform pressure a along the four edges
(Fig. 7.33), behaves like a sphere subjected to hydrostatic pressure. In fact,
denoting by kf the foundation stiffness per unit surface, we have the total
potential energy

2 P" U' =
~2 L ^xx + W™ + 2 U W'xx W'vv + 2(1 " ") W « l dx dy

+ \ l8[-« + »\) + \kwl dxdy (7.13.16)

having omitted the terms in u and v, because in the plate u(x,y) = v(x,y) —
0. By assuming w(x,y) = A sin ax sin f3y with a = mn/a and /? = mn/b (a is
the half-wavelength in the direction x and b is the width of the plate) and by
substituting in (7.13.16), we find that the value of a which makes the energy
functional semidefinite is
Eh2
(7.13.17)

which is completely analogous to (7.13.11).

7.14 P O S T - C R I T I C A L B E H A V I O U R OF PLATES
A N D SHELLS
We shall not give details of the analysis of the post-critical behaviour of plates
and shells in this Section, as it is a complicated subject and therefore outside
330

the aims of this book, which is intended only as an introduction to the study
of the stability of elastic structures. For this reason we limit ourselves to
providing only some theoretical results, with the support of experimental
results where possible.
In Section 7.5 we presented a qualitative analysis in which it was concluded
t h a t the equilibrium of the plate, in correspondence with the critical load,
is stable. This implies that a perturbation analysis leads to a bifurcated
path of the type presented in Fig. 4.4(b). This behaviour is qualitatively
analogous to that of beams (see Section 5.6), but differs quantitatively in
that the curvature of the bifurcated path of plates, in correspondence with
the bifurcation, is much larger than that of beams. As a consequence, using
a current terminology, the post-critical behaviour of plates is more stable
than t h a t of beams. The mechanical explanation of the phenomenon lies in
the presence of a membrane state of stress which increases the load-carrying
capacity of the plates beyond bifurcation.
No conclusions of a general character are possible for shells, due to the
large variety in their geometry and in the loads applied to the surfaces. We
therefore limit ourselves to considering two classical problems: the cylinder
subjected to axial pressure, and the sphere subjected to hydrostatic pressure,
for which there exist both the analytical solutions and experimental results.
It has been seen in Sections 7.10 and 7.13 that, corresponding to the critical
load, both structures present a large number of buckling modes to which a
certain number of bifurcated paths correspond. We shall not discuss this, as
in the present book we have excluded the analysis of simultaneous buckling
modes. It is demonstrated in [20] that the structure has a strongly asymmetric
bifurcated path with a large value for the coefficient Ac, and therefore the
behaviour of the imperfect structure for a given imperfection is characterised
by a value of the collapse load A*/Ac <C 1.
A different type of representation of the equilibrium curves of the cylin-
der which are met in practice is provided by Figs. 7.34 and 7.35, where the
axial load N is reported on the ordinates and the shortening A£ or the spe-
cific shortening A£/£ on the abscissa. The first figure refers to an isotropic
cylinder, the second to a cylinder with longitudinal stiffeners on the outside
surface. The solid line curves indicate in both the figures the theoretical
behaviour of the cylinder, as results from perturbation analysis. The experi-
mental results have been indicated by dashes in the first figure and by circles
in the second [27].
In Figs. 7.36-7.40 we show photographs of specimens subjected to different
load conditions [27].
331

25 7000-
N

20 6000- — theor. results


p-°-o exp. results
5000-
15
—theor results/ 4000-

L
--exp.n 3SUltS/
10 3000-

2000-

V\
1000- j

0
0.005 0.010 0.015 0020
0.2 0.4 0.6 0.8 1.0
Ai/l

Fig. 7.34 - Equilibrium curves for Fig. 7.35 - Equilibrium curves for a
an isotropic cylinder cylinder with longitudi-
subjected to axial load. nal stiffeners on the out-
side surface, subjected
to axial load.

Fig. 7.36 - Cylinder subjected to Fig. 7.37 - Cylinder subjected to


lateral pressure. torsion.
332

Fig. 7.38 - Cylinder subjected to axial load with:


(a) external stifFeners, (b) internal stiffeners.

Fig. 7.39 - Isotropic cylinder sub- Fig. 7.40 - Stiffened cylinder sub-
jected to axial load. jected to axial load.
333

Figure 7.41 shows a photograph of a cylinder stiffened with rings subjected


to axial load [32]. Figures 7.42 [32] and 7.43 [33] show two silos of cylindrical
form damaged by vertical forces of friction on the inside wall, due to the
settling of grain.

Fig. 7.41 - Ring-stiffened cylinder subjected to axial load.

Fig. 7.42 - Buckling of a silo.


334

Fig. 7.43 - Collapse of a silo.

Figures 7.44 and 7.45 refer to a sphere subjected to a hydrostatic pressure p.


In the first figure we have in solid lines the theoretical equilibrium curve with
p/pc on the ordinate and AV/AV C on the abscissa; p represents the current
pressure and pc the critical pressure, whilst AV — V — V0 and AVC = VC-V0
with V the current volume, Vc the volume under critical pressure and V0
the volume of the unloaded sphere. The curve with circles refers to the
experimental results. In Fig. 7.45 we have a photograph of the sphere on
which the test was carried out [27].

0.4 0.8 1.2 1.6 2.0


ZIV/JV C

Fig. 7.44 - Equilibrium curve for a sphere subjected to hydrostatic pressure.


335

Fig. 7.45 - Sphere subjected to hydrostatic pressure.


336

REFERENCES
R.V. Southwell, S.W. Skan: "On the stability under shearing forces of a
flat elastic strip", Proc. R. Soc. London, Ser. A, 105 (1924), 582.

G. Krall: Stability of elastic equilibrium (in Italian), Zanichelli, Bologna,


1939.

T. von Karman, H.S. Tsien: "The buckling of thin cylindrical shells


under axial compression", J. Aeronaut. Sci., 8 (1941), 303.

A.E.H. Love: A treatise on the mathematical theory of elasticity, Dover


Publications, New York, 1944.

W.T. Koiter: On the stability of elastic equilibrium (in Dutch), Disserta-


tion, Delft, 1945; English translation published as NASA T T F-10, 833,
1967 and AFFDL Report T R 70-25, 1970.

S.B. Batdorf: "A simplified method of elastic stability analysis for thin
cylindrical shells", NACA Report 874 (1947).

F. Bleich: Buckling strength of metal structures, McGraw-Hill, New York,


1952.

G. Gerard, H. Becker: Handbook of structural stability, Part III, Buckling


of curved plates and shells, NACA T N 3 7 8 3 , 1957.

S.P. Timoshenko, S. Woinowsky-Krieger: Theory of plates and shells,


McGraw-Hill, New York, 1959.

S.P. Timoshenko, J.M. Gere: Theory of elastic stability, 2nd ed.,


McGraw-Hill, New York, 1961.

W. Fliigge: Stresses in shells, Springer-Verlag, Berlin, 1962.

G. Gerard: Introduction to structural stability theory, McGraw-Hill, New


York, 1962.

W.T. Koiter: "Stability of equilibrium of continuous bodies", Techn.


Report No. 79, Brown University (1962).

V.V. Novozhilov: Thin shell theory, Noordhoff, Groningen, 1964.

A.S. Volmir: Stability of deformable systems, NASA AD 628 508, 1965.

W.T. Koiter: "On the non-linear theory of thin elastic shells", Part III,
Proc. K. Ned. Akad. Wet., Ser. B, 69 (1965).
337

[17] V. Franciosi: Strength of materials (in Italian), Liguori, Napoli, 1967.

[18] M. Como: Theory of elastic stability (in Italian), Liguori, Napoli, 1967.

[19] A.E. Green, W. Zerna: Theoretical elasticity, Clarendon Press, Oxford,


1968.

[20] W.T. Koiter: "The non-linear buckling problem of a complete spherical


shell under uniform external pressure", Proc. K. Ned, Akad. Wet. Ser.
B, 72 (1969), 40.

[21] W.T. Koiter: "On the foundation of the linear theory of thin elastic
shells", Parts I and II, Proc. K. Ned. Akad. Wet. Ser. B, 73 (1970).

[22] M. Como: "Stability of structures" (in Italian), in R. Baldacci, G. Cera-


dini, E. Giangreco (editors): Dynamics and stability (in Italian), Italsider
technical-scientific series for the design of steel structures, Vol. l i b , 1971.

[23] Handbook of structural stability, Column Research Committee of J a p a n ,


Corona Publishing Company, Tokyo, 1971.

[24] P.M. Naghdi: "The theory of shells and plates", in Handbook der Physik,
vol. V I a / 2 , Springer-Verlag, Berlin, 1972.

[25] C.L. Dym, L.H. Shames: Solid mechanics: a variational approach,


McGraw-Hill Kogakusha Ltd., Tokyo, 1973.

[26] C.L. Dym: Stability theory and its applications to structural mechanics,
Noordhoff, Leyden, 1974.

[27] M. Esslinger, B. Geier: Post-buckling behaviour of structures, Springer-


Verlag, Vienna, 1975.

[28] D.O. Brush, B.O. Almroth: Buckling of bars, plates and shells, McGraw-
Hill, New York, 1975.

[29] L.H. Donnell: Beams, plates and shells, McGraw-Hill, New York, 1976.

[30] L. Corradi: Structural instability (in Italian), CLUP, Milano, 1978.

[31] W.T. Koiter: Theory of elastic stability, lecture notes, University of


Technology, Delft, academic year 1978-79.

[32] J.G.M. Wood: "Thin-walled silo structures, failures, testing and design",
in J. Rhodes and A.C. Walker (editors): Thin-walled structures, Granada
Publishing, London, 1980, p. 339.
338

[33] M.M. Ghobrial, G. Abdel-Sayed: "Stability of orthotropic cantilever


cylindrical shells", in J. Rhodes and A.C. Walker (editors): Thin-walled
structures, Granada Publishing, London, 1980, p . 351.
339

Appendix

THE CALCULUS OF VARIATIONS

A.l INTRODUCTION
A brief outline of the calculus of variations is given in the following pages to
acquaint the reader with concepts and operations frequently used in previous
chapters.

A.2 STATIONARY VALUES OF D E F I N I T E INTE-


GRALS. B A S I C P R O C E D U R E S OF T H E CAL-
CULUS OF VARIATIONS
The analytical problems of mechanics often require the determination of the
stationary value of a functional. The branch of mathematics which deals with
this problem is known as the calculus of variations.
A functional is a law which establishes a correspondence between functions
and real numbers. A definite integral therefore comes under this definition.
Let
rb
1=1 F{u,u' Jx)dx (A.2.1)
Ja

where F represents a function whose arguments are another function u(x)


and its first derivative u'(x) defined over the interval [a, 6], and x itself. From
(A.2.1) we can see t h a t J is a function which associates a real number to every
u(x) which is continuous and has a continuous first derivative.
To introduce the concept of the stationarity of a functional, we shall use
that of a real-valued function of real variables which belongs to the class of
functionals.
We say that a function f(x) is stationary at x0 when

m -■
that is

/(X + £) /(Xo)
lim ° " = 0 (A.2.3)
€—0 f
340

Let us now consider the function

u(x) = u(x) + e<p(x) (A.2.4)

which is also continuous and has a continuous first derivative if such is <p(x),
with 6 a real number.
The expression

(A.2.5)

is defined as the derivative of the functional I.


The relation (A.2.5) obviously depends on the function u(x). If there is a
function u 0 (x) for which the derivative of / vanishes V <p(x), we say that I is
stationary in u 0 (x), that is, u0(x) makes I stationary.
It is useful at this point to look again at the concepts shown previously in
the traditional form of the calculus of variations.
Let us suppose that we know the function u = u0(x) which makes the
integral (A.2.1) stationary. To prove that we have a stationary value, we must
evaluate this integral in correspondence with the slightly modified function
(A.2.4) and show that the variation of the integral vanishes.
Assuming that 6 is a very small quantity which at the limit approaches
zero, we have the possibility of modifying the function Uo{x) by an arbitrarily
small quantity.
Let us now compare the values of the modified function u(x) with the values
of the original function u0(x) in a point x making the difference between u(x)

u(x)

a x x+dx b x

Fig. A.l - Representation of a generic function u(x) and of the varied


function u(x).
341

and Uo(x). This difference is known as the variation of the function UQ(X)
and is indicated by

6 u — u(x) — u0(x) = 6 <p(x) (A.2.6)

The variation of a function is characterised by two basic properties. It repre-


sents an infinitesimal change in the assigned function u0{x) since the para-
meter e tends to zero, and it is a change which can be chosen in an arbitrary
way.
Note the basic difference between 6u and du. Both the quantities represent
infinitesimal changes, but whilst du refers to the difference of values of the
assigned function Uo(x) corresponding to an infinitesimal change dx of the
independent variable, 6u represents an infinitesimal change of u 0 (x) in each
point x to which a new function u0 + 6u is associated.
If the two limiting values u0(a) and u0(b) of the function u0{x) are prescri-
bed they cannot be varied. This means that we must have
6u(a) = 0
(A.2.7)
6u(b) = 0

In this case we talk about variations between definite limits.

A.3 COMMUTATIVE PROPERTIES OF T H E VA-


RIATIONAL P R O C E S S
The variation of the function u(x) is defined through a completely new func-
tion e<p(x), of which we can calculate the derivative. On the other hand, we
can calculate the derivatives of the functions u(x) and u(x) and evaluate the
difference. In the first case we have the derivative of the variation

—- 6 u = — [u(x) — u(x)} — —- e (p(x) = e<p'(x) (A.3.1)


dx dx dx
and in the second we have the variation of the derivative

6 4~ u{x) = u'(x) - u'(x) = {u' + e <p') -u' = e <p'{x) (A.3.2)


dx
From comparison of (A.3.1) and (A.3.2) we obtain
d d x
— 6u = 6 — u (A.3.3)
dx dx
which shows that the derivative of the variation is equal to the variation of
the derivative.
Let us now suppose t h a t we wish to calculate the variation of a definite in-
tegral. This operation is carried out by subtracting from the definite integral,
342

calculated in accordance with the modified function / , the definite integral


associated with the original function, that is

6 f f{x)dx = f J{x)dx- f f{x)dx


Ja Ja Ja

= f \f{x)- f{x)]dx = f 6f{x)dx (A.3.4)

6 f f{x)dx = f 6f{x)dx (A.3.5)


Ja Ja

Equation (A.3.5) shows that the variation of a definite integral is equal to the
definite integral of the variation.

A.4 STATIONARY VALUE OF A D E F I N I T E IN-


TEGRAL FROM THE CALCULUS OF VARIA-
TIONS
By making use of the calculus of variations, we determine here the stationary
value of the definite integral (A.2.1). To solve the problem, we analyse the
variation of the integral caused by the variation of function u — u(x). Let
us start by writing the variation of the integrand F(u,u' ,x) associated with
the variation of u, remembering that F is an assigned function of u,u',x and
t h a t this functional dependence is not altered by the variational process. We
have

A F (u , u', x) = F (u + e<p , u' + e<p', x) — F (u , u', x)

= <5F + 0(e 2 ) (A.4.1)

with

cmr I ^ l dF Fd
i\ /*
6F{u,u',x) = e [ _du< p + _du'
<p'j (A.4.2)

where the name variation or first variation is given to the first-order terms
of the increment of F. The variation of integral (A.2.1) is therefore equal to

rb' rb rb I dF dF \
61 = 6 Fdx= / 6Fdx = e \ — {p + —ip'\ dx (A.4.3)
Ja Ja Ja \OU OU1 J
We can see t h a t the expression (A.4.3) can be written, by virtue of (A.2.5),
as 61 = el'(u)<p and therefore, for the definition of stationarity, we must have
61 = 0.
343

To obtain the rate of variation of the integral we divide by e

(A.4.4)
6 Ja \du du' )

Using integration by parts we can write

dF
dx dx (A.4.5)
Ja du1 ^ Ja dx \ du1

for which (A.4.4) is reduced to

e
dF
lb

+
a du
A. ^L
dx du1
<pdx (A.4.6)

Note t h a t because of the integration by parts (p'(x) has disappeared from


(A.4.4), reducing this expression to (A.4.6) where only <p(x) appears.
If we carry out the variation between definite limits, the function <p(x)
vanishes at the two end points a and b and (A.4.6) is simplified to

I _ fb (dF d dF\ (A.4.7)


—— I <p dx
€ Ja \\dudu du11J
dx du
For the sake of brevity, let us introduce the notation
„, , dF d dF
Elx) = (A.4.8)
du dx du1
and write the condition for the stationary value of J in the form
rb
j E(x) (p(x) dx = 0 (A.4.9)
Ja

It is not difficult to see that the integral can vanish for any arbitrary
function <p{x) only if E(x) vanishes everywhere between a and 6. In fact, let
us choose a function <p(x) which vanishes everywhere, except in an arbitrarily
small interval around the point x — f. Within this interval E(x) is practically
constant and we can write
8
J- E{Z)l ri+p p(x)dx (A.4.10)
e

The error made tends to zero when p approaches zero. Since the integral on
the right side does not need to vanish, the condition 6I/e = 0 requires that
E(£) — 0. Since the point x = f can be chosen arbitrarily in the interval
a < x < 6, we obtain for the entire interval the differential equation
5F
(A.4.11)
du dx du1
344

The equation (A.4.11) is known as the Euler-Lagrange equation and represents


a necessary and sufficient condition for the vanishing of 61.
The results so far obtained can easily be extended to the case in which the
function F which appears in (A.2.1) is a function of n functions and of their
first derivatives, that is, F = F{ux,... ^un^u\^... ,u^,x).
The Euler-Lagrange equations which permit us to determine the unknown
n functions U i ( x ) , . . . , un(x) associated with a stationary value of I are in this
case

^ ! ^ - | ^ = 0 , (.- = l , . . . , n ) (A.4.12)
dx du\ oui

Given the definite integral

J = f F (u , u ' , u" , x) dx (A.4.13)


the reader can prove without difficulty that the Euler-Lagrange equation as-
sociated with the stationary value of / is
dF d_d]^ d2 dF _
1 (A.4.14)
du dx du dx2 dun
to which must be added the terms relating to the boundary conditions
lb lb
dF dF
= 0 (A.4.15)
~du~'
V + du" <P ~dx du" ^
if we do not carry out the variation between definite limits.
Finally, given the definite integral
rb rd
1= / ^(u,u|Z,u>y,t;,t;fX,t;>y,u;,u;|Z,wiy,ti;>xz,u;>yy,u;>xy,x,y)dxdj/
Ja Jc
(A.4.16)
with ux — du/dx, etc., where F depends on the three functions u(x,y),
v ( x , y ) , w(x,2/), on their derivatives as indicated in (A.4.16) and on x, y, the
Euler-Lagrange equations associated with the stationary value of / are
dF d dF d dF
= 0
du dx du)X dy dutV
d dF d_ dF_
= 0
dv dx dviX dy dvtV
(A.4.17)
d dF d dF dF
dw dx dw) dy dwy + dx2 dw

d2 dF dF
= 0
+ dxdy dw)Xy + dy2 dwtVy
We leave it as an exercise for the interested reader to derive the terms relating
to the boundary conditions.
345

A.5 VARIATIONS OF AN INTEGRAL WITH AUX-


ILIARY HOLONOMOUS CONDITIONS
Let us return to the problem in the previous Section which consists of the
search for the stationary value of the definite integral
rb
1= F{ui,u[,x)dx , (i = 1 , 2 , . . . , n) (A.5.1)
Ja

and assume t h a t the functions u t (x) are not independent but restricted by a
certain number of auxiliary holonomous conditions expressed in the form

fj(uijx)=0 , (i = 1 , . . . , n ; j = 1, . . . , m) (A.5.2)

with m < n. It is sometimes possible to eliminate m of the u t in terms of


the remaining functions and thus reduce the problem to n — m independent
variables. The calculus of variations shown in the previous Section then
permits us to deduce the Euler-Lagrange equations. The inconvenience of this
procedure lies in the fact that the elimination of the variables is, in general,
an impossible task. The method of Lagrange undetermined multipliers offers
a suitable and elegant solution to the problem.
By varying (A.5.2) we obtain from (A.4.2)

6fj = ^ 6 u i = 0 (A.5.3)
OU{

where a repeated index denotes summation.


In applying the method of Lagrange multipliers, we multiply each of the
equations (A.5.3) by an undetermined function A t (x). The problem is there-
fore presented in the following form: instead of equating to zero the given
variation of the integral / , we modify it in the following way

61* = 6 f Fdx+ f Xj6fjdx = 0 (A.5.4)


Ja Ja

This is obviously possible, since we have added zero to the variation of J. Let
us now carry out the integration by parts of the first term and reduce it to
the form

6 f Fdx= [ Ei{x)6Uidx (A.5.5)


Ja Ja

where for brevity we write

„. , dF
dF d dF ,. r „.
dui dx du
346

From (A.5.4), collecting the terms which have the same £u,, we get

/•» (dF d dF , dfA r , ,


Ja \OUi (IX OU\ OUi)

where i = 1 , . . . , n ; j = 1 , . . . , m. By making use of equations (A.5.3) we


can now eliminate m of the <f>ut variations, for example the last m's. We can
arrive at the same result by choosing Ay in such a way that
dF d dF dfj ,. , tk

The remaining 8u{ variations are independent and can be chosen arbitrarily.
This leads to the conclusion that (A.5.7) is satisfied only if the coefficients
of these variations vanish. In the final result we can therefore see that the
coefficients of each <5ut vanish, independently of whether the u t are dependent
or independent variables. The resulting equations are
dF d dF dfj ,. . N
— —+Ay - ^ = 0 , t = 1 , . . . , n ; j = 1, . . . , m A.5.9
OUi dx ou\ OUi

which, when linked with (A.5.2), form a system of n + m equations in the


n + m unknowns u\,..., u n , A i , . . . , A m .
The procedure so far is equivalent to the following variational problem.
Instead of considering the stationarity of the definite integral

1= / F dx (A.5.10)
J a
with auxiliary conditions (A.5.2), we consider the stationarity of the modified
integral

I* = [ F* dx (A.5.11)
J a
where

F* = F + X1f1 + --- + Xmfm (A.5.12)

in which the auxiliary conditions are imbedded. By considering ut- and Ay as


free functions, we obtain from (A.5.11) and (A.5.12)

SI* = f [SF + XjSfi + fjSX^dx


Ja
£ +A 6 i+/
=r[( ' 'S) " '"' dx = 0 V Sui, SXj (A.5.13)

having made use of (A.5.6). From (A.5.13) we get the n equations (A.5.9)
and the m equations (A.5.2).
347

A.6 E X T R E M U M VALUES OF A D E F I N I T E IN-


TEGRAL. S U C C E S S I V E VARIATIONS OF A
FUNCTIONAL.
The Euler-Lagrange equations furnished by the variational procedure solve
the problem of the search for a stationary value of an integral. In accor-
dance with the solution, the rate of variation of the functional vanishes for
every possible variation of the independent variables. This is however a nec-
essary but not a sufficient condition for the functional to have a minimum or
maximum.
In many mechanical problems it is sufficient to know only the stationary
value of a functional but in others, such as in the analysis of stability, the
search for the extremum values is of fundamental importance.
To solve the problem, we again consider the function F which appears in
(A.2.1) and rewrite (A.4.1), this time considering also the second-order terms
in 6

A F ( u , u', x) = F (u + e <p , u' + e <p', x) — F (u , u', x)


'dF dF ,\ 1 2 /^F 2 d2F ,
du du' / 2 \ du2 du du'

+du'S^I+-
2 (A-6-1)
Equation (A.6.1) can be rewritten as

AF = F'6u+- F"\6u}2 + --- (A.6.2)

having introduced

F'6u = 6F = ^ Su+^-t { 6u' (A.6.3)


du du

and recalled t h a t 6u = e(p.


Equation (A.6.4) is known as the second variation of the function F. The
variations of higher order can be defined in a similar way. Correspondingly,
the increment in the integral (A.2.1) is written

AI = I'6u + - 7"[<5u]2 + . . . (A-6-5)


348

where

rb ( dF dF \
I'Su = 6I= [ — 6u + — 1 6u')dx (A.6.6)
Ja \ OU OU J

and

(A.6.7)

Let u 0 (x) be the function which satisfies the Euler-Lagrange equation


(A.4.11). The first variation <$/, obtained by calculating (A.6.6) for u{x) =
u0(x), is equal to zero for every Su which satisfies the boundary conditions.
The increment A J , in the neighbourhood of u0(x), is therefore governed by
the sign of the secon variation; if I" [t>u] / 2 > 0 V Su ^ 0, Jo is a minimum;
if IQ [SU] / 2 < 0 V <5u 7^ 0, Jo is a maximum; if I" [6 u] = 0 for a particular
6u, then higher order terms must be analysed.
Generalising, in the case of a functional of n variables,
rb
I = F(ui,u'i,x)dx (i = 1, 2 , . . . , n) (A.6.8)
Ja

and the increment in the integral is written as

1
AI = I'6u + - I"[6uY + .>- (A.6.9)

with

rb f dF dF \
r6u = 6I= [— 6ui + — Su'Adx (A.6.10)
Ja \OUi 0U\ J

, ( +i ( ,
\'"^' = fA{i&t
'u, dUj « '>' i£k « ''<-
cfUi cfu',

d2F
+ (A 6 n)
*wis<Su'ir --
The condition 61 = 0 gives the Euler-Lagrange equations (A.4.12) from which
we can determine the solution u 0 ( x ) . If I" [Su] / 2 is definite in sign, that is,
if it assumes only positive or only negative values for all admissible Su, then
I0 is an extremum; if / " [<5u]2/2 is indefinite, then I0 is a stationary value of
/ but not an extremum.
349

A.7 STATIONARITY OF A FUNCTION OF n


VARIABLES
The problems discussed in the previous Sections are simplified when we are
looking for the stationarity of a function of n variables

F = F{Xi) , (i = l , . . . , n ) (A.7.1)

By defining the first variation of F in the form


dF
6F=—- Sxi (A.7.2)
OX{

we say, by analogy to (A.4.3), that the function F is stationary at a point P


if

6F=*^- 6xi = 0 (A.7.3)


OXi

Given that £x t are arbitrary, it follows from (A.7.3) that


a necessary and sufficient condition for a function F of n variables to have a
stationary value at a point P is that the n partial derivatives of F with respect
to all the n variables vanish at the point P, that is

dF
^ = 0 , (i = l , . . . , n ) (A.7.4)

If the variables of the function F(xi) of which we wish to find the stationarity
are not free but subject to m constraint equations

fj{xi) = 0 , (t = 1, . . . , n ; j = 1 , . . . , m) (A.7.5)

with m < n, then it is possible to eliminate from (A.7.5) m of the x t , for ex-
ample, the last ones. Expressing these variables as functions of the remaining
n — m and replacing them in (A.7.1), we get a function of n — m free variables.
For the stationarity, we therefore again write (A.7.4) where i varies from 1
to 7i — m. However, it is possible to proceed quite differently by using the
method of Lagrange multipliers. By writing the variations of (A.7.5)

6fj = ^-6xi = Q (A.7.6)


OXi

the condition of stationarity (A.7.3) is still satisfied if we add the m equations


(A.7.6) multiplied by m undetermined Ay factors, which are functions of the
variables x t
350

By using (A.7.6) we can now eliminate m of the <5xt-, for example the last
ones. We can arrive at the same result by choosing the Xj in such a way that
we have

1
— + Xj■ - = 0 , (t = n - m + 1, . . . , n; j = 1, . . . , m) (A.7.8)
axt uX{

The remaining <5xt variations are independent and this, taking (A.7.8) into
account, leads to the conclusion that equations (A.7.7) are satisfied if

— + Ai-^- = 0 , (* = l , . . . , » ; y = l , . . . , m ) (A.7.9)
OX{ OXi

From (A.7.9) it appears that the coefficients of each (5xt- vanish as if all
the x t variables were free variables. Equations (A.7.9), together with (A.7.5),
comprise a system of n-\-m equations in n + m unknowns x t , Xj, and therefore
the problem of constrained stationarity is solved.
The procedure used so far can be seen in a more general way. By introdu-
cing the function

F* = F + Xjfj (A.7.10)

we can calculate the stationarity of F* with respect to x t and A; considering


all of them as free variables. We have

6F* = 6F + \i6fi + fi6Xi=(^i+Xi^6xi + fj6Xi =0

V 6xi,6Xj (A.7.11)

from which come the n equations (A.7.9) and the m equations (A.7.5).
351

REFERENCES
[l] R. Courant, D. Hilbert: Methods of mathematical physics, Interscience,
New York, 1953.

[2] C. Lanczos: The variational principles of mechanics, 4th ed., University


of Toronto Press, Toronto, 1970.
353

Name Index

Augusti, G., 64, 70 Ritz, W., 167, 175, 176, 259, 264,
Batdorf, S. B., 315, 321, 322 266, 302, 303, 306
Bessel, F . W., 176, 181, 254 Sewell, M. J., 83, 123
Brantman, R., 101, 108 Skan, S. W., 301
Chetayev, N. G., 17, 18, 19, 20, 21, Southwell, R. V., 301
31, 33 Sylvester, J. J., 16, 67
Clausius, R., 124 Thompson, J. M. T., 33, 123
Dirichlet, P. G. L., 30, 31, 32, 33, Timoshenko, S.P., 256
41, 125 Van der Pol, B., 2, 3
Donnell, L. H., 284, 314, 320 Vlasov, V. Z., 220, 221, 225, 284,
Duhem, P., 124 314
Einstein, A., 9 Winkler, E., 173, 329
Euler, L., 134, 135, 141, 162, 171,
176, 178,179, 181,185,211,
244, 320, 344, 345, 347, 348
Fourier, J. B. J., 124
Frechet, M., 126
Fredholm, I., 132, 139
Greenhill, A. G., 176
Hamilton, W. R., 27, 29
Hunt, G. W., 33
Hurwitz, A., 23, 24
Kerr, A. D., 101, 108
Kirchhoff, G., 291
Koiter, W. T., 123, 125, 128, 281,
329
Lagrange, J. L., 27, 28, 29, 30, 31,
32, 33, 4 1 , 57, 6 1 , 62, 125,
145,155,156,163,185,260,
344, 345, 347, 348, 349
Lame, G., 276, 278
Liapunov, A. M., 1, 14, 15, 16, 17,
18, 19, 20, 21, 22, 30, 31,
33, 4 1 , 43, 125
Malkin, I. G., 20
Mushtari, K. M., 284, 314
Newton, I., 1
Poincare, H., 1
P r a n d t l , L., 250, 258
355

Index

adjoint homogeneous problem, 132 conservative forces, 28, 33, 123, 310
asymptotic method, 136 conservative systems, 11, 31, 62, 75
asymptotic stability, 16, 17, 18, 62 critical case of equilibrium, 33
asymptotically stable, 4, 6, 9, 14, critical load, 40, 129
19, 20, 21, 22, 23 for traction, 247
Augusti model, 64, 70 P r a n d t l , 250
autonomous system, 3, 10, 14, 19 torsional, 240, 242
critical mode, 129
beam
secondary, 110
cantilever, 166-167, 168, 171,
critical point, 3, 59, 95
219, 221, 252, 258
cylinder
subjected to its own weight,
Batdorf parameter, 315
175-179
of finite length, long, 315
with variable cross-section,
of finite length, short, 316
179-181
of infinite length, 309
clamped, 169-170
ovalisation, 311
clamped-simply supported, 170
post-critical behaviour, 322-325
clamped-sliding, 170-171
under axial pressure, 316-320
Euler, 161-165
under combined loads, 320-321
elastically supported, 171-173
under hydrostatic pressure,
loaded at mid-span, 181-185
309-312
on elastic foundation, 173-175
under hydrostatic pressure po-
stability, 185
tential energy, 310
beam model
internally constrained, 154-161 discrete systems, 27
internally unconstrained, 148- n degrees of freedom, 27
154 one degree of freedom, 9
bi-moment, 240 two degrees of freedom, 37, 101
bifurcation point, 40, 4 1 , 135, 136
elastic foundation, 173, 259, 293
secondary, 65
energy
buckling mode (see critical mode)
geometrical, 230, 233
calculus of variations, 339-350 kinetic, 27
centre, 6 potential, 27
characteristic equation, 4 total, 12
collapse load (see limit load) equilibrium
condition of normalisation, 84, 85, asymptotically stable, 14
97, 134, 139, 210 neutral, 64
condition of orthogonality, 81, 87, stable, 14
132, 139 unstable, 14
356

Euler-Lagrange equation, 344 tangent stiffness, 33

finite element method node


for beams, 203-213 stable, 4
for thin-walled beams, 264-267 stable degenerate, 6
interpolation functions, 264, 265 unstable, 4
interpolation matrix, 206 unstable degenerate, 8
focus stable, 6 non-autonomous system, 3
unstable, 6 non-conservative forces, 28, 32
free oscillations, 9, 63
damped, 10 path
function bifurcated, 40, 41, 47
first variation, 342, 349 extrapolated, 118
second variation, 347 fundamental, 40, 47
functional fundamental, approximation, 117
derivative, 340 natural equilibrium, 45, 50
stationary value, 339 non-linear fundamental, 57, 101
unnatural equilibrium, 45, 46,
Hamilton equations of motion, 29 50
Hamilton variables, 29 perturbation analysis (see pertur-
bation method)
initial imperfections, 44, 49 perturbation equations
sensitivity, 51
first-order, 150, 159, 198, 208
instability degree, 34
second order, 150, 159, 189
integral curve, 3
perturbation method, 79, 83, 136,
intermediate integrals, 11, 63
161, 330
Lagrange equations of motion, 28 change of parameter, 114
Lagrangian coordinates, 57 plane frame
Lagrangian multipliers, 345 hinged symmetrical, 191-203
Lame parameters, 276 multi-storey, 144, 212
Liapunov direct (second) braced, 144
method, 1 non-symmetrical, 212
Liapunov first method, 1 symmetrical, 212
Liapunov function, 16 unbraced, 144
limit load, 51, 70 two-bar, 186-190
limit point, 51, 59, 79, 136 plate
compressed in one direction,
matrix 291-294, 297-300
elastic, 260 compressed in two directions,
elastic local, 266 294-297
geometrical, 260 infinitely long, 304
geometrical local, 266 Kirchhoff condition, 291
of inertia, 28 post-critical behaviour, 329-335
357

stability, 288 of the equilibrium (point), 10,


stiffened, 305 14
under shear, 300-305 qualitative analysis, 1, 185, 288
post-critical behaviour region, 247
asymmetrical, 54 stationary value of a definite inte-
stable symmetrical, 42 gral, 342
unstable symmetrical, 48 stationary value of a function of n
variables, 349
Ritz method Sylvester criterion, 67
for beams, 167-169
for plates, 302 theorem of Hurwitz, 23-24
for thin walled beams, 259-264 theorem of Lagrange-Dirichlet, 30
shape function, 262 theorem on instability, Chetayev, 18,
31
saddle point, 5 theorem on instability, Koiter, 125
scleronomous system, 28 theorem on stability, Koiter, 125
sectorial area, 223 theorems of Liapunov
null point, 227 on asymptotic stability, 16
pole, 226 on instability, 17, 31
principal null point, 228 on stability, 16
principal pole, 227 thin-walled beams
shell flexural-torsional coupling fac-
bending stiffness, 287 tor, 242
extensional stiffness, 287 flexural-torsional instability, 217
kinematics, 279-285 geometrical properties, 232
post-critical behaviour, 329-335 kinematical hypotheses, 220
shallow, 283 lateral buckling, 219
simultaneous buckling modes, 70, lateral load, point of applica-
270, 309 tion, 252
simultaneous buckling modes, great local buckling, 217
number, 318, 320, 328 post-critical behaviour, 268-272
snap-through, 61 torsional instability, 217
space of the phases, 3, 10, 12, 63 under eccentric axial load, 2 4 5 -
sphere 248
post-critical behaviour, 329-335 under lateral loads, 250-258
under hydrostatic pressure, 326- under pure bending, 248-250
329 under uniform compression,
stability 239-244
coefficients, 34 trajectory, 14
domain, 296 stable, 1
linear approximation, 20 unstable, 1
of motion, 1 twisting centre, 228
358

twisting centre and shear centre co-


incidence, 229

variation of a definite integral, 341


variation of the derivative, 341
variations between definite limits,
341

warping constant, 31

You might also like