AAA - M. Pignataro, N. Rizzi and A. Luongo Auth. Stability, Bifurcation and Postcritical Behaviour of Elastic Structures 1991 PDF
AAA - M. Pignataro, N. Rizzi and A. Luongo Auth. Stability, Bifurcation and Postcritical Behaviour of Elastic Structures 1991 PDF
AAA - M. Pignataro, N. Rizzi and A. Luongo Auth. Stability, Bifurcation and Postcritical Behaviour of Elastic Structures 1991 PDF
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
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
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.
PREFACE
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
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
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
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)
x{ = xt+i (t = 1, 2 , . . . , n - 1)
(1.2.4)
xn = j [Xj 5 t)
x + k (x2 - 1) x + x = 0 (1.2.5)
ii = x2
(1.2.6)
x2 = - k [x\ - 1) x2 - Xi
x = X ( x ; t) (1.2.7)
x = X(x) (1.2.8)
x2
*1
X\ — a n ^1 + #12 %2
(1.3.1)
x<i — a>i\ X\ + a22 ^2
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
A2 - ( a n + a 22 ) A + ( a n a 22 - a i 2 a 2 i ) = 0 (1.3.3)
(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.
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.
(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.
a) b)
Fig. 1.9 - Degenerate stable node.
a) b)
(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
Xi = dij Xj (1.3.9)
Xi = X2
(1.3.10)
i2 = — a2 Xi — 2 b x2
(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.
x = X(x) (1.4.1)
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)
&i = x2
(1.4.3)
X2 = -f{xi)
dx\ x<i
(1.4.4)
dx2 f{xl)
\A +I f{xi)dx1 = H (1.4.5)
l
-x\ + P{x1) = H (1.4.7)
12
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.
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).
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.
(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
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
V=k
Fig. 1.17 - Plane of phases.
x — — y — x3
(1.6.5)
y = x - y*
At a point which is arbitrarily near the origin we have V < 0, and so the
origin is asymptotically stable.
19
i = V3 + x5
y= x +y
The function V(x,y) = x 4 — y4 satisfies the conditions of Chetayev's the-
orem
(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
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.
ii = a{j Xj + Ri ( x i , . . . , x n ) (1-7.1)
say, ||R||/||x|| tends to zero with ||x||. This fact can be expressed by
||R(x)||=o(||x||) (1.7.2)
Xi = a,ij Xj (1.7.3)
y = Px = PAP_1y + PR (1.7.4)
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.
(a) The Xh roots are all negative. The following function of Liapunov is
assumed
V = yl + y\ + • • • + y2n (1.7.7)
21
from which
(b) Some of the Xh roots, for example Ai, A 2 , . . . , Ap (p < n) are positive
and the rest negative. This time we take
from which
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
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
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
x = 2x + 8y
(1.7.14)
y = - x- 3y
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
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
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
REFERENCES
[1] H. Poincare: "Sur I'equilibre d'une masse fiuide animee d'un mouvement
de rotation", Acta Math., 7, 1885, 259.
[6] H.H.E. Leipholz: Stability theory, Academic Press, New York, 1970.
[8] C.L. Dym: Stability theory and its applications to structural mechanics,
Noordhoff, Leyden, 1974.
27
Chapter 2
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
(2 2 n)
JtM-^ = Qi^q) --
29
q = f(q,q) (2.2.12)
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)
. _ 3H
dpi
(2.2.18)
Pi = - -^ + Qi ( p , q)
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:
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)
• 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.
2.5 T H E O R E M S OF L I A P U N O V A N D CHETAYEV
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
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
q+ c q3 = 0 (2.6.3)
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)
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
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
dm = 0 (2.7.10)
Ui = Ui(s) (2.7.12)
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
U! = l , ui = xii = 0 (2.7.19)
ut = l , ut: = 0 (t = 2 , 3 , . . . , n) (2.7.20)
P = dni (2.7.21)
n
P = rfim + 6 dinfi<+ 3 52da(ui)2 (2.7.22)
i=2
u{ = - ^ (i = 2 , 3 , . . . , n ) (2.7.24)
da
(no s u m with respect to i), from which
37
and therefore conclude that the sufficient condition for the total potential
energy to be positive definite is that
din = 0 (2.7.27)
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)
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)
u 2 = - \ u\ (2.7.33)
38
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
:>J I*
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
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
6 = 0 VAT (2.8.7)
d2P 1
— = k-NL+-NL02 (2.8.9)
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
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
m £ 2 0 - f {k-mgt)0 = 0 (2.8.16)
43
A = ± \mgl
<2-8-i8>
—k ,
v-b-
and so the general solution is
P(0) = ±(k-mge)e2 + ^ O i
(2.8.20)
Q
6{t) = 0Q s i n u t + — cos CJ t (2.8.21)
(jj
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.
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
stable eq.
unstable eq.
k
^ T
N
stable eq.
unstable eq.
Nc=k/
/ *
-1
■jr/2 TC/2
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)
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
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)
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)
f(a)
deduce
+ Nl(l-cos0Q) (2.9.18)
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.
kl
have
dN 9_o
= kt\-6 + (2.9.21)
dO 02
from which
el'3 (2.9.22)
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
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
and as
a = (2.10.2)
sin<p
and
a— 1+ | + 2 sin 0 (2.10.3)
V tan<p \^tan<£>
we have
I*
Let us assume for simplicity that <p — n/4 and substitute in (2.10.4) the
series expansion around 0 = 0. We then have
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
d2
P 1 , ,2 2 - 3 0 02
Nt\l (2.10.9)
M* ~ 4 2
stable eq.
- unstable eq-
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
and
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.
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/
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.
K }
(2.11.2)
2 Vcos0 o cos0/
ki2 / 1 1 \2
P{0) = + 2A(tan0-tan0o) (2.11.4)
V cos 0O cos0/
A = sin0 f —J (2.11.6)
\ COS 0 O COS0/
-jr/2
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
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
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\
stable eq.
unstable eq.
F!
Ac c
B
/ \
V* J
°\ / °\
•- ± \ A
D
~K
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
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
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.
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
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.
6 (c - 1) - (c + 3) 6\ + (3 c + 1) 0\ = 0 (2.12.9)
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
%
\ y
\ /
1 \
\ stable eq.
C) unstable eq.
/F E
/ \
\
/ \
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
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)
kx (1 - 4 p + 3c) > 0
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
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
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* / \'
'*,
I'P
v-\\3\ n
^i i3; ^ *
\\ % /
\\ perfect stable
\\ / //
\ perfect unstable
\) imperfect stable
/ /\ imperfect unstable
/ /
//
r*2
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.
[4] H.H.E. Leipholz: Stability theory, Academic Press, New York, 1970.
Chapter 3
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
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.
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
too = tt (0)
(3.2.3)
A0 = A(0)
Pii(qj(t);\(t)) = 0 Vt (3.2.4)
and in particular
L0v0 = 0 (3.2.9)
p(Lo) = n (3.2.10)
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
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.
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)
L0v0 = b0 (3.2.14)
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)
x T L0 v0 = x T b 0 = 0 Vx (3.2.17)
81
%i = a, xa Via (t — 1 , . . . , n) (3.2.18)
with aa = 1.
Similarly, the oo 2 solutions to system (3.2.9) are
\Potijk [Oi Cj ck v2p0 + (at- Cj dk + at- cjt dy) v^o v^o + «t- dy djk t)20J
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
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)
Potij QJO + Po,i\ Ao + -Po.iy* Qjo Qko + 2 Po,ij\ fyo Ao + Po,ixx A0 = 0 (3.3.2)
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
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
t = /.-(*-*o)+f(A-A0) (3.3.6)
1 = *.•*(*)+**(<) (3-3-8)
0 = liSi{t)+li{t) (3.3.9)
0 = lili{t)+li{t) (3.3.10)
0 = /,• h0 + / Xo (3.3.12)
0 = /,- L + / *o (3.3.13)
t = A - A0 (3.3.14)
1 = A0 , 0 = X0 , 0 = X0 (3.3.15)
86
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
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
2
^W = */o* + ^ o ' + o(*2) (3.3.19)
Vo = bi 9.o + b A0 (3.3.23)
ho = bi ho + bi0 (3.3.24)
z <lio{bjqjo + b\0)
9,"o = /(bjqjo + bXo)2 (3.3.26)
i Oj qjo + o A0 .
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
13 \P0tijk a>i qfo Qko + Po,ij\ «• Ufo A0 + lj0 ^o]) + Pojxx a. ^>2) A0
(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;
With respect to the previous case in which p(L 0 ) = n, we now find ourselves
in the presence of the following substantial differences:
(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.
Qi = 9i{t)
(3.4.1)
A - g(t)
L(t) v = 0 Vt (3.4.2)
<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
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
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)
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)
for any £(e), that is, for any i(e) , i(e) , . . . and this obviously implies
Ptij Uj = 0 (3.4.24)
fi
hi i + hit iy *k + 2 Ptij u, i = 0 (3.4.25)
P,i^l,t) = -^{P{qi{m{t),u,);X{t))}
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}\
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
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)
PcijUjc = 0 (3.4.33)
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
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
We now denote by
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.
te = - - PcMkUicUicUkc ^ (3 4 4 2 )
2
Pctijk Uic Ujc ()kc + Pctij\ Uic Ujc Xc
+ Pc,ijX Uic Ujc A c ) j JZ (PCyijk Uic Ujc he + Pc.ijX Uic Ujc K) (3.4.43)
become
ft = U (A) (3.4.44)
w
= ( "' ) ( 3 - 4 - 46 )
Qu = fi (Ac) (3.4.48)
Ui = 0
(3.4.49)
A = A(e)
Ui = Ui (e)
(3.4.50)
A = A(e)
Finally, the eigenvalue problem for the search for the bifurcation points is
rewritten as
Pty(/,(A);A)uy = 0 (3.4.54)
tr. — A-
^C A* — 1
(3.4.55)
te = \c xc = o
We obtain
Pe,ij uje + PCiijk ujc iikc + 2 Pe,ijk 9ke uje \c + 2 Pe>ij u]e Xc = 0 (3.4.57)
Xc = - 1 Pc^cUicUkc ( 34 5 Q )
+ 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)
i i ,
1
I
I
+
# 0 +a
0) (A
<W oo O
-o
1 o
-J
L C\J
^ -M i
O Q.
+u
5£
t/>.
r
o '
'
+a
o
Lsin (# G )
Lsin@0+|8)
h ■ H
In Fig. 3.7 the symbols adopted are indicated. It is easy to see that
m = sa (3.5.3)
F = Jfci A + k2 A 2 (3.5.4)
The equations of equilibrium (3.1.2), in the case under examination, are writ-
ten as
a = t = gx(t)
P t = a
= *M (3.5.7)
_ 23 2t _
A
" Tsm(0o + t)-9(t)
103
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
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
which is called the characteristic equation of the problem. This equation can
be simply factorised in the form
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
(3.5.17) yields
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
This implies
1 + cos 2 (0O + tc) - A* cos {0O + te) = l - cos 2 ($0 + tc) (3.5.22)
105
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)
ule = 1 (3.5.26)
uu = 0 (3.5.27)
5lc= 0 (3.5.28)
uu = 1
(3.5.29)
u 2c = -1
P^px = 0 (3.5.35)
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
Pc.ijk w,c Ujc 9kc = Pc,aaa u2u occ + Pc>aap (u2lc (5c + 2 uu u2c &c)
= -3.8719 5 (3.5.49)
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
ab{e) = t{e)+u1{e)
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
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
Fig. 3.8 - Projection of the linearised bifurcated path on the plane a, /?.
= 0.74894 6 (3.5.55)
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
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)
(3.5.71)
= -1.04566 s (3.5.77)
= 3.54777 s (3.5.79)
= 15.6786 5 (3.5.84)
3 *c -Pc>«jib fitc fijc 9kc — 3 ^Pc,aaa file file &c + -PC,aa^ file file A:
= -7.6725 5 (3.5.85)
113
= -7.6725 5 (3.5.86)
&
+ 3 Pc,a00a *>U u\c c + 3 Pc,a000 UU \i\c (3c
= 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\
6 t\ Xc Pc>ijkX Uic Ujc Qkc = 6 t\ Xc [Pc,aaaX u\c 6cc + PC,aa0X u\c (5c
= 4.6954 5 (3.5.90)
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
XI = - 1.5819 (3.5.93)
ic = - 1.4403 (3.5.94)
a"e = te + uu = -1.4403
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
V(e)=y(e)-K (3.5.97)
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
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
PfeiA) = Po,AA + - P o , , ; ^ g i + P o , , A g . A + - P 0 , A A A 2
+ 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
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
-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
te = 0.1763
A* = 2.0308 (3.6.12)
UU = U2c
tc = 0.16128
K = 1.8575 (3.6.13)
uu = -<l"ic
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°
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.
(4.2.1)
where
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
Making use of the inequality (4.2.3), the following two basic theorems have
been proved [6]
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
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
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
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^ u 3 = i j
l
y Eijhk ei5 ulth u,,4 dV - i X « 3 (4-3.9)
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
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]
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
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
Pc"u2=° (4.3.19)
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
PW-jf 1 „, »"*
+ x(Vi 1)
dx (4.3.23)
130
P'cv =0 (4.3.24)
P"'vs = 0 (4.3.26)
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
P?u* = 0 (4.4.2)
Pc"X>0 (4.4.3)
u = aiii + w (4.4.4)
P c "u 2 = 0 (4.4.6)
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
24 c lie + - ue
+
= - e2 P" ii 2 + - e3 P"' ii 3 + - e3 P" u u + — e4 P'"' ii 4
2 6 2 24
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
8 4 4 4
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)
u* = a i i + ii p (4.4.18)
P c "[ii;-ii;]<5u = 0 (4.4.19)
from which
Pc'"ii3 = 0 (4.4.25)
T"u2 = e2 (4.4.27)
T"ii 2 = 1 (4.4.29)
T"iiu = 0 (4.4.30)
a = -T"iiup (4.4.32)
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)
„,„ . , 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)
= 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.
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
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
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)
5?{P'(v(c)+u(6);A(€))*u} = 0 Ve (4.5.9)
- 3 P'" [Ac VC ii c + Ac v c ii c + A* v c ii c + ii c u c ] 6 u
-3\2cP'luc6u-3\cP'c'uc6u (4.5.15)
vc - Acvc (4.5.16)
vc = A c v c + Ac2^c (4.5.17)
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
*N A2<0
a) b) c)
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)
consider the term P"(v(A); A)ii* and make the series expansion in terms of A
starting from A = Ac:
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.
R.T. Shield, A.E. Green: "On certain methods in the stability theory of
continuous systems", Arch. Ration. Mech. Anal., 12 (1963), 354.
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
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
P«0 F«0
vt&rn w^r
i *
r&gfim ^rtygjpm
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)
+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)
+ Xc (u'c<pc + u'c<pc)}'
{[-Nc(<p2c+2\c<pc<pc)
N+dN
M+dM
-M' + Ni-Te-T = 0
From these the first- and second-order incremental equations are obtained
f — N cos <p + 2 N <p sin <p + N <p sin <p + N <p2 cos <p
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
(-AT c + T c ^ c ) ' = 0
(-fc-Nc<pc)' =0 (5-2.18)
-M'c + Ncic-Tcec-fc = 0
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
v* = sin (p (5.2.22)
X = <P*
M = EIX (5.2.23)
A(w;A) = I f''EI<p'2dx
L JO
Obviously, this fact does not appear in linear analysis because then (5.2.22)
becomes
= 0
v1 — <p (5.2.25)
X = <P
v(x)
<Kx).
R x=Q
^ u(xj x=L
A
ne(w;A) = i llEI<p'2dx
I JO
-EI& + Rc<pc-Sc = 0
uc{t)-uc{0) =0 (5.2.28)
Jo
with the boundary conditions
and the secondary critical mode is described by the system obtained from
(4.5.14)
[i -ft (5.2.30)
uc(l) - uc(0) = - <pcdx-2\c (pc<pcdx
Jo Jo
The explicit expressions of the terms which appear in the relations which
give Ac and Ac are
v(x e ) / j
u(x e )
-R
<p. "e
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),
1 /■«.€
nc(wc;A) = - f EIt<p'l dxe
L JO
therefore be
-EIe<pl + Rec<pec-Sec = 0
(5.2.44)
fie
(ujc - uic) cos 0e + (vjc - vic) sin 0t = - / <pec dxe
Jo
. fie
- 2 Ac / <ptc <pec dxe
Jo
(5.2.40), its variations can be obtained by first finding the variations of the
single terms and then adding them together. In particular
+ 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
-R.
B X
x=0
Si v(x) st
R
A T <K
A Kx) * U(x) B
UA = 0 , vA = 0 , vB = 0 (5.3.2)
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
(w;A) = |Q [ i £ V 2 + A ( c o s v ? - l ) dx (5.3.6)
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
from which
having used (5.3.8). The variational equation (5.3.10) leads to the differential
equation
* . = " ^ ) ( «
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
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)
which leads to
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)
<pc = 0 , Sc = 0 (5.3.22)
Ac - \ Xc (5.3.24)
4
6 = p(0,e) = 0 (5.3.26)
A = ! + I e2 (5.3.27)
Ac 8
2HA
x=0
s s
VA
Is
R R
l «vfr I u(x
^ R R
t
A
uA = 0 , vA = 0 , tpA = 0 (5.3.28)
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
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
<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).
II" a 6 a = 0 (5.3.38)
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
Ac = 2.49 ^ / (5.3.51)
a2 = - 0 . 2 9 6 °j (5.3.52)
from which
*V//A
x=0 x=(
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.
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
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)
?2
x=0
The total potential energy is obtained from (5.3.5), with the addition of
the elastic energy of the springs. We have
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
kl
tan — = -akl (5.3.65)
( 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
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
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
^ = 2 yffi (5.3.77)
x=/
A=const.
x=0
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)
- 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
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)
(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
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)
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
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)
7 7
- '(1-rh)' (5.4.23)
(5.4.24)
2 2 Jo [ \ l + dj J
from which the Eulerian of the problem is
[£/'(I-^)"»"]"+A""=° (5.4.25)
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.
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
-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)
1 x
' * -• * 2 1
/ /
— 1 —
*
Let
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
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]
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
+ A / (cos <p2 — 1) dx 2
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
k212
klcotkl= 3 (5.5.12)
Vu = 0 , <p2e = 0 , Sc = 0 (5.5.14)
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.
P(W) =
I liSn^ + E^fjdV
dx (5.6.1)
Let us consider the system in Fig. 5.26 and let I and El be the length and
flexural rigidity, respectively, of each rod.
Figure 5.27 shows in detail the local and global systems of coordinates
adopted [24].
l
n ( w ; A) = \ \ Eltp'l dxx + \ I' EI<p'l dx2
L Jo 2 Jo
+ XFv (5.7.1)
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
With these equations are associated the boundary conditions obtained from
188
(5.2.47)
Taking this into account, equation (5.7.4) gives the following boundary
conditions
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
kl = n (5.7.8)
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
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
Vu{t) = $M (5-7.13)
EI <p'u{0) = EI <p'2c{0) = 0
190
1 / 7T X \
<Plc{x) = ~ <P2c{x) = ~ ~ (COS — + 1J COt 0
I
vc = - 2 sin 0
uc = 0
(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)
we get
<p1(l)=-e (5.7.18)
A F = ( A C + A C 6 + U C £ 2
) F (5.7.19)
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
AF AF
© ©
D
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
^B — UA~ / ( c o s £>2 ~~ 1) ^ x 2
0
EI2<p'le + S2e = 0
VAc = 0
-U>Bc= / <P3c dx
Jo
193
With these are associated the boundary conditions obtained from (5.2.47)
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
kh t a n B = 6 a (5.8.8)
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?~
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*
from (5.2.44) as
h k2 h2 ( sin 2
"Be = - - . 22 , , 1 +
2 sin A: /i V 2 A:h )
./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
<p2c(x) = 0 (l - 2 j}
h k 2h 2 sin 2 kh\
vAc = VBC = - 2- s'm'kh
1 + 2k hT)
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
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]
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
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
EI1<p'lc{0)=0
EI3<p'3c{0)=0
Eh<p'ae(t)-EI2<p'2e{l) = 0
4/i 2 I2 \ k
KEAxt
- 2 + 6EI2J \EIl
+ GA -t sin k h = cos A; h (5.8.27)
having put
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
= -Nc = -
1 -
hik'^i
2
(5.8.33)
GAi
201
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)
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
l/h=5. >104
Ac= 2.47 r10 3
A®-Nch2/EI,
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
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
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
where
P(w(qc);A)-^Pc(w(qe);A) (5.9.4)
e=l
( «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
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)
lU gFe(\)6Zedxe = rqrTfSf(7?)Fe(A)Se(f?)T^qdxe
Jo Jo
= qTTjKe{X)Te6q (5.9.10)
where
P"u6u = £qTTfK,(A)Te<5q
where
q T K(A)<5q = £ q T K ( A ) q (5.9.14)
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'
= KQ + A KG (5.9.20)
(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)
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
Euler b e a m
1 / h
//
I d
XZZZZZZZZZZZZL ^ 3
11
'/ I I
//
Vti Vitil /
V,
Fig. 5.33 - Sample structures.
212
30 El 10EI
►-
El
- 3m
>-
-
2 El
4 El
>-
8 El *
j3m
< 1 0 m
, t6mf>6m»
15EI 15EI
3m
El
2 El
4 El
8 El
3m
■ 10m 6m 12, 4m
-*—f-
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
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.
REFERENCES
E. Janke, F . Emde: Tables of functions, 4th ed., Dover, New York, 1945.
S.J. Britvec: The stability of elastic systems, Pergamon Press, New York,
1973.
D.O. Brush, B.O. Almroth: Buckling of bars, plates and shells, McGraw-
Hill, New York, 1975.
215
[23] P.A. Kirby, D.A. Nethercot: Design for structural stability, Granada
Publishing, London, 1979.
Chapter 6
6.1 INTRODUCTION
gjj^g&*»»"~
p§&
"¥
l\i
UJU
Qp-
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
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.
(a) the beam cross-sections are not deformable in their own planes;
(b) shear deformation on the middle surface is negligible.
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.
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)
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
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
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)
from which
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
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
uc dA = 0 (6.3.16)
/ .A
From Fig. 6.11 we obtain
(Dc(S„S2)/2
c(»2.S)/2
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)
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
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
= 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
+ 2u"0"LJx + 2v"6"uy)dA
being
T= f u2dA (6.4.4)
JA
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
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
=
\ 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
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
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
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
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
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
L{ik[v+**-] +&*+*'-]}"
= j [(** + V2) {nx rzx + ny rzy)] ds = 0 (6.4.14)
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
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
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
u = v = 0= 0 in z = 0,1 (6.5.3)
240
( r T \
** -ef-N)A»-Ny<Ae = ° (6.5.6)
(TT2 * ^ - N ^ A V + NxcAe =0
-NycAu + NxcAv+[7r2^- + GJ-NrfjAe =0
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
2 2
{Ny - N){NX - N){Ne - N) - N2 V-\ [Nx - N) - N2 ^ {Ny - N) = 0
r r
c c
(6.5.9)
Let us now examine two particular cases which frequently occur in appli-
cations.
242
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)
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-™
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$/Ny = 1.035, t h a t is, the three critical
loads are about equal. Note that, due to the particular form of the cross-
244
C(0,-2R)
section, the contribution to the critical load N$ 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)
».=*.. * = * . , * . - , , _ , 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
(c) Cross-sections with two axes of symmetry and external load applied on
one of the axes
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
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).
7T
Mx = ± EIy[-ET + GJ
M (6.7.5)
■K
MLye = ± EIX ET + GJ
vc
~n P
250
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)
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
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
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
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
£ + ^-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
^ = ~ (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)
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].
p2 GJ 0.1 1 2 3 4 6 8
L
ET
p2 GJ 10 12 14 16 24 32 40
C
ET
Finally, we report some results for problems which are frequently met in
applications [3],[7].
yjGJEIy
Fe=l2 (6.8.27)
I2
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
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
^
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
+ a V (M x ^ r ) c - b V (M v i& T )'c} ^
in which the integrals are extended to the whole beam. We can now define
the following square matrices of order n
KG _
- N jt<t>'<t>'T dz, KGV0 = -N j i X ' x ' T 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)
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%.
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 =
(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
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
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
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
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.
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)
C, u, %
the form
u{z) = h{z)u1 + f2{z)u2 + f3{z)4>1 + fi{z)fa = fTh)
= 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
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
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)
1 \-
-#-^ 6---,..
Number of elements
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
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
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
/ u \ ( Au\
7T Z
u„ Av sin (6.10.8)
\'J \ A
e )
2 7T Z . TV Z 2 7T
cos —— sin —— dz — — (6.10.9)
/: 1 1 3 P
we have
+ 2{Iy-Is)AuAvA$\ (6.10.10)
Uc = Ci Uc (i = 1 , 2, . . . , n) (6.10.11)
2 TT2E
pi" tf = ; ^ 2 (/, - ix) AU A„ A, (6.10.13)
3 £
The three critical loads are N\ — Ny, N2 = Nx, N3 = Ng and are defined by
(6.5.7). The associated eigenvectors are
I Au\ / 0 \ / 0 \
uei = 0 , uC2 = \ Av , iiC3 = 0 (6.10.14)
V 0 ) { 0 ) \At j
If the three eigenvalues coincide, any linear combination of the type
( Au \ ( 0 \ ( 0 \
Uc = Ci 0 + c2 Av + c3 0 (6.10.15)
V 0 ; V 0 ; \A6 )
is a solution to the system (6.5.8). Equation (6.10.15) is equivalent to
( l\ ( Q\ ( 0 \
ur 0 + Av 1 + A$ 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
*cx =
lew — V (6.10.17)
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.
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.
D.O. Brush, B.O. Almroth: Buckling of bars, plates and shells, McGraw-
Hill, New York, 1975.
Chapter 7
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.
Xi = fi ( a x , a 2 ) (7.2.1)
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)
from which
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
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
dsP=ds1(l+-j^ (7.2.9)
having put
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,
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)
Hx = 1 , H2 = l (7.2.20)
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)
k ; = k n + ^ i k i + £>2k2 (7.3.4)
having written
1 dw u
(7 3 5)
i a - -
1 OW V
H2 da2 R2
or in scalar form
u(z) = u + z <pi
vW = v + z<p2 (7.3.7)
-$ = 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,
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
/^*> 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
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
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
cvii ^ e n + z Xi
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
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
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)
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
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
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
(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
(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
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)
C
722 = 22 + - ^23 (7-4-7)
1
712 = ^12 + - CJi3 CJ23
+
^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)
where
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
(7.5.7)
£ ) 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
which simplifies to
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
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
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.
v4
Nx N>
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
+
(—) u)
/rmv\2 /7r\2
(7.6.8)
\TO7r/
or, equivalently
(7.6.9)
where
4
1 \flT 2 V6~ 3 Vl2~ V20" 5
/»="
For /? = yjm(m + 1) there are two buckling modes corresponding to the same
load value, and therefore Nx is a multiple eigenvalue.
I I I I 1 I I 1 i 1 |Ny
Nx N,
111 I f 1 1 I I 1 HNy
(7.6.13)
7T2Df
NX = K (7.6.14)
62
where
1 2
m +n
V (7.6.15)
K=
m
j] +Pn
P m n K
1 1 1 2
0 1 1 4
-1 2 1 8.33
NX/N,
Ny/Nyc
Nx/Nxc and Nv/Nyc are shown, where Nxc and Nyc are the critical stresses
corresponding to uniform compression in one direction [10,28].
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)
,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
C 4 — — C2 C3 = Cx (7.6.26)
N
Nx M^^^^ *
^mmmmmwmmmr
Fig. 7.11 - Clamped simply supported plate under uniform compression.
299
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)
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
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.
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
Ajr
r ATT
r + 2 ^ , ^ ] f +
T / =o (7.6.36)
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
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)
+ 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
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)
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 £
1\ 7T2Df
Nxu = 2[2m + (7.6.48)
mj b2
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.
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)
K = 5.34 + ± (7.6.54)
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
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
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
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*
\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)
*(»>&)-¥$+$ {$+>¥{")***
By requiring the coefficient of A\ to vanish we have
7r2 Df (mb 2a
Nx = 4 — 2^ — (7.7.10)
6 2a mb
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
tn p=const
(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 '
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/
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)
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.
(7.9.4)
313
w
p = const J
o) b)
",,2 _ pR
P'u j w2sdS +\nejs \u2x + v2s + 2i/utXvt8
~2
1-v 1-v w
+ < +
V
,x + (1 - " K . V | X + ^ J
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
V 4 u - — ( - v wtXXX + wiX33) = 0
K
o Eh A
Df
~R2" ™,*zxz + p i 2 V V a 5 = 0
having p u t
Rl2 (7.9.12)
n2D,
nl
P = VR Rh
VI (7.9.13)
_ (1 + /? 2 ) 2 , 12 Z2
Pc= 55 + 7T 4 /? 2 (l+/? 2 ) 2 (7.9.14)
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
Pc-P2 = ^ CM*)
and assuming n = 2, as happens with infinite cylinders, we obtain
4£ 2
(7.9.16)
y
' 7T2 R2
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)
p=const.
f
w
b)
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
r. ^ 8 Eh
7 w xxxx -\-phV4wtXX = 0 (7.10.3)
~R2~ >
with wmn as arbitrary coefficients. Equation (7.10.4) satisfies the third, fourth
318
(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)
mnR/i
Vic^/h
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
Rh
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)
= 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
1000
\ /2R
100
[ /
y \ a x ial
pr essure
/lateral
pressure.
10
[ /
1000
100
100
1000
100
1.1
1.0
0.6
0.4
0.2
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 ,
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
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
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
from which
E Eh2
oc =
i2 2 (a 2 + /?2) + 12(1 - ^ 2 ) ("2+/H (7.13.11)
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
a
R
b
u
i a
H-
I I I I I I I I I I
2 P" U' =
~2 L ^xx + W™ + 2 U W'xx W'vv + 2(1 " ") W « l dx dy
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)
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
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. 7.39 - Isotropic cylinder sub- Fig. 7.40 - Stiffened cylinder sub-
jected to axial load. jected to axial load.
333
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.
S.B. Batdorf: "A simplified method of elastic stability analysis for thin
cylindrical shells", NACA Report 874 (1947).
W.T. Koiter: "On the non-linear theory of thin elastic shells", Part III,
Proc. K. Ned. Akad. Wet., Ser. B, 69 (1965).
337
[18] M. Como: Theory of elastic stability (in Italian), Liguori, Napoli, 1967.
[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).
[24] P.M. Naghdi: "The theory of shells and plates", in Handbook der Physik,
vol. V I a / 2 , Springer-Verlag, Berlin, 1972.
[26] C.L. Dym: Stability theory and its applications to structural mechanics,
Noordhoff, Leyden, 1974.
[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.
[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
Appendix
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.
m -■
that is
/(X + £) /(Xo)
lim ° " = 0 (A.2.3)
€—0 f
340
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)
u(x)
a x x+dx b x
and Uo(x). This difference is known as the variation of the function UQ(X)
and is indicated by
Equation (A.3.5) shows that the variation of a definite integral is equal to the
definite integral of the variation.
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
(A.4.4)
6 Ja \du du' )
dF
dx dx (A.4.5)
Ja du1 ^ Ja dx \ du1
e
dF
lb
+
a du
A. ^L
dx du1
<pdx (A.4.6)
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
^ ! ^ - | ^ = 0 , (.- = l , . . . , n ) (A.4.12)
dx du\ oui
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
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)
6fj = ^ 6 u i = 0 (A.5.3)
OU{
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
„. , dF
dF d dF ,. r „.
dui dx du
346
From (A.5.4), collecting the terms which have the same £u,, we get
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
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
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
+du'S^I+-
2 (A-6-1)
Equation (A.6.1) can be rewritten as
having introduced
where
rb ( dF dF \
I'Su = 6I= [ — 6u + — 1 6u')dx (A.6.6)
Ja \ OU OU J
and
(A.6.7)
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
F = F{Xi) , (i = l , . . . , n ) (A.7.1)
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)
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)
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.
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
warping constant, 31