Sei sulla pagina 1di 212

Lecture Notes in

Engineering
Edited by C. A. Brebbia and S. A. Orszag

64

A. Elzein

Plate Stability
by Boundary Element Method

i
.,
~
Springer-Veriag
Beriin Heidelberg New York
London Paris Tokyo
Hong Kong Barcelona Budapest
Series Editors
C. A. Brebbia . S. A. Orszag

Consulting Editors
J. Argyris . K -J. Bathe' A. S. Cakmak . J. Connor· R. McCrory
C. S. Desai· K-P. Holz . F. A. Leckie' G. Pinder' A. R. S. Pont
J. H. Seinfeld . P. Silvester· P. Spanos' W. Wunderlich· S. Yip

Author
Abbas Elzein
Computational Mechanics Institute
Ashurst Lodge, Ashurst
Southampton S04 2M
United Kingdom

ISBN-13:978-3-540-5371 0-6 e-ISBN-13:978-3-642-84429-4


001: 10.1007/978-3-642-84429-4

This work is subject to copyright. All rights are reserved, whether the whole or part of the material
is concerned, specifically the rights of translation, reprinting, re-use of illustrations, recitation,
broadcasting, reproduction on microfilms or in other ways, and storage in data banks. Duplication
of this publication or parts thereof is only permitted under the provisions of the German Copyright
Law of September 9, 1965, in its current version, and a copyright fee must always be paid.
Violations fall under the prosecution act of the German Copyright Law.
© Springer-Verlag Berlin, Heidelberg 1991

The use of registered names, trademarks, etc. in this publication does not imply, even in the
absence of a specific statement, that such names are exempt from the relevant protective laws
and regulations and therefore free for general use.

61/3020-543210 Printed on acid-free paper.


To my mother and father
TABLE OF CONTENTS

CHAPTER 1 INTRODUCTION

1.1 His tori cal Background................................. 1


1.2 Stability............................................. 3
1.3 Experimental and Numerical Modelling.................. 4
1.4 The Boundary Element Method........................... 6
1.5 Plate Stability by BEM................................ 8
1.6 Scope of the Present Vork............................. 9

CHAPTER 2 PLATE STABILITY THEORY

2.1 Introduction.......................................... 11
2.2 Stability of Structural Systems ....................... 11
2.3 Linear Theory......................................... 15
2.4 Large Deflections..................................... 24
2.5 Boundary Conditions................................... 26
2.5.1 Out-of-Plane Boundary Conditions ................ 27
2.5.2 In-Plane Boundary Conditions ...•................ 31
2.6 Numerical and Experimental Studies ..............•..... 35
2.7 Conclusions. . . . . . . . . • . . . . • . . • . . . • • . . . . . . . . . . . . . . . . . . . . 37

CHAPTER 3 MEMBRANE STATE OF STRESS

3.1 Introduction.......................................... 38
3.2 Boundary Integral Formulation ......................... 38
3.3 Boundary Element Solution ............................. 43
3.4 Numerical Implementation .............................. 47
3.5 Resul t s . . • . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
3.6 Conclus ions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
VI

CHAPTER 4 CRITICAL LOADS

4.1 Introduction.......................................... 54
4.2 Boundary Integral Formulation .....••...•.........•...• 54
4.3 Boundary Element Solution............................. 59
4.3.1 Modelling of Boundary Unknowns ••••.•••••..•.•••. 59
4.3.2 Domain Deflection Models .•.....••.•.••....•.•••. 60
4.3.2.1 Continuous Cells •.•••..•.•..•••..•.....• 63
4.3.2.2 Discontinuous Cells .•.•...••....••.•.... 66
4. 3. 3 Free Boundary................................... 67
4.3.4 Eigenvalue Problem ••••.....•.•...•••.•.....•.... 68
4.4 Numerical Implementation .............................. 72
4. 5 Resul t s. . • • . . . . . • . • • • . • • . . . . . . • • • . . . . . . . . . • • • . . . . . . . . . 74
4.5.1 Optimum Nodal Position in Discontinuous Elements 76
4.5.2 Performance of the Various Interpolation Models 79
4.5.3 Convergence of Results from the Linear
Discontinuous Model............................. 81
4.5.4 Comparison with Exact Solutions ...•.•........... 83
4.5.5 Comparison with the Finite Element Method ....... 86
4.6 Conclusions........................................... 88

CHAPTER 5 DUAL RECIPROCITY

5.1 Introduction.......................................... 89
5.2 Outline of the Method................................. 90
5.3 The Discrete Points Fourier Analysis ..•........•...... 93
5.3.1 The One-Dimensional Fourier Series ..........•... 93
5.3.2 The Two-Dimensional Fourier Series .......•..•... 94
5.3.3 The Discrete Points Two-Dimensional Fourier
Analysis. • . • • • • • • • . . . . . . . • • . • • . . . . . • • . . . . . . . . . . • 96
5.4 The Deflection Models................................. 99
5.4.1 The Trigonometric Deflection Model ....••........ 99
5.4.2 The Nodal Deflection Model ..•....•.•.•...•..••.. 104
5.5 Transformation of L(w) ...••••••.•....•.••............. 108
VII

5.6 Transformation of the Domain Integral ••••••••.••••.••• 109


5.7 The Problem of Singular Integrals •••••••••••.•..••..•• 112
5.8 Eigenvalue Problem ••••••••.•••••••••.••••••.••.•••••.. 113
5.9 Numerical Implementation •.•.••..•.•••••••••••.•••••••. 115
5 .10 Results............................................... 118
5.10.1 Convergence of the Fourier Transformation •••... 119
5.10.2 Convergence of the Transformed Integrals ••.••.. 124
5.10.3 Examples of Critical Loads ..••.••..••.••...•... 133
5.10.3.1 The Trigonometric Deflection Model ••....... 134
5.10.3.2 The Nodal Deflection Model .........•....... 141
5.10.3.3 The Plates Deflected Shape ...•..•........•• 145
5.11 Conclusions........................................... 149

CHAPTER 6 LARGE DEFLECTIONS

6.1 Introduction .....•.......•.•.........•..•........•..•. 151


6.2 Boundary Integral Formulation .•..•.••...••.••..•...... 152
6.3 Domain Deflection Models ...••....•••.•....•.....•..... 155
6.4 Boundary Element Solution ........•.....•..•........•.. 162
6.5 Solution of the System of Equations .•..••..•.......... 165
6.6 Numerical Implementation .....••...•••...•..•.......•.. 168
6.7 Results •.............•..••....................•..•.... 170
6.8 Conclusions..... • • • . . . • • • • . . . . . . • • . . • . . • • • . . • . • • . . • . .. 177

CHAPTER 7 CONCLUSIONS •..•.......•.•••••••••••••.•••••••••.•• 178

APPENDIX A The Green's Identities •....•••.•.••••..•...••..... 185

APPENDIX B Functions of the Fundamental Solutions .•.•.••••..• 187

APPENDIX C Trigonometric Deflection Functions •••.••••••..•.•• 191

REFERENCES •••••••••••.•••••••••••••••••••.••••••••.•••••••..••• 194


CHAPTER ONE

INTRODUCTION

1.1 Historical Background

Thin plates and shells are widely used structural elements in


numerous civil, mechanical, aeronautical and marine engineering
design applications. Floor slabs, bridge decks, concrete pavements,
sheet pile retaining walls are all, under normal lateral loading
circumstances, instances of plate bending in civil engineering. The
problem of elastic instability of plates occurs when load is
applied in a direction parallel to the plane of the plate. The deck
of a bridge subjected to a strong wind loading, the web of a girder
under the action of shear forces transmitted by the flanges, the
turbine blade of a machinery undergoing longitudinal temperature
differentials, would all eventually buckle when the applied load,
or its temperature equivalent in the last case, exceeds a certain
limit, that is the buckling load. Although the plate may exhibit a
considerable post-buckling strength, the buckling load is
considered in many design instances, especially in aeronautical and
marine engineering, as a serviceability limit because of the abrupt
and substantial change in the dimensions and shape of the buckled
plate. Nevertheless, the post-buckling region retains its
importance either as an essential safety margin or as a stage of
loading actually reached under normal loading conditions. The
design engineer will therefore need rigorous tools of analysis to
predict, in addition to the buckling load, the deflections and
stresses at both buckling and initial post-buckling stages.

The elastic stability of slender structures first became a subject


of interest when iron structures were used in construction towards
the beginning of the last century (Timoshenko, 1953). The first
2

experimental observation of the plate buckling phenomenon took


place in 1845 when tubular iron beams, specially designed for the
Conway bridge railroad, were tested and failed on the compression
rather than the tension side in what is today a classical case of
local buckling.

The first satisfactory equation describing plate bending was


derived by Navier in 1823 (Timoshenko, 1953) with a flexural
rigidity defined in terms of only one elasticity constant and,
following Poisson's earlier analysis, a requirement of three
natural boundary conditions. It was Kirchhoff (1850), however, who,
thirty years later, stated the fundamental plate theory
assumptions, derived the potential energy expression of the bent
plate and applied the virtual work method to obtain the
differential equation for plates with the flexural rigidity defined
in terms of both Young's modulus and Poisson ratio. In addition, he
realised that the three natural boundary conditions proposed by
Poisson are incompatible with the fourth order nature of the
differential equation and contracted them into two boundary
conditions. This 'Poisson-Kirchhoff boundary condition paradox', to
use Reissner expression, will only be fully apprehended a little
less than a century later (Reissner 1985). In 1883, St-Venant
presented the problem of a flat plate loaded in compression in its
own plane and derived its differential equation (Timoshenko, 1953).

Kirchhoff (1877) and Clebsh (1862) also investigated the large


deflection theory of plates but it was Foppl who, in 1907, first
used the concept of stress function in order to simplify the
originally complicated equations. Von Karman (1910) later removed
the restriction on the thinness of the plate; he derived, in 1910,
the equations called after him and later proposed an approximate
formula for the effective width.

Later in the twentieth century, thin plates and their elastic


stability gained particular interest because of the extensive use
of thin walled steel structures and the boost of the mechanical,
3

aeronautical and marine engineering industries. Although, since


then, much work has been done on the clarification, extension and
solution of the equations obtained as a result of the pioneering
work described above, the equations themselves retained their
validity. Hence, the Navier's equation as presented by Kirchhoff
and the Von Karman equations are still considered to be efficient
and accurate representations of the small and large deflection
theories of plates, respectively.

1.2 Stability

A plate subjected to in-plane loads remains in a primary state of


stable equilibrium until the buckling load is reached, provided of
course loading is applied in small enough increments so that any
deviations due to dynamic disturbances are damped out (Budiansky,
1974). Bifurcation buckling indicates the splitting of the
equilibrium path into two or more theoretically possible secondary
loading curves. An unstable equilibrium path may occur with no
lateral deflections taking place. Another equilibrium path is a
stable one producing large deflections after buckling.

The Kirchhoff's plate theory predicts the buckling load and


buckling mode but fails to depict the bifurcation equilibrium path
after buckling. On the other hand, the large deflection theory and
the von Karman equations simulate the stable equilibrium path and
predict the initial post-buckling strength. The two theories,
however, start from the same fundamental assumptions.

Owing to the small thickness of the plate compared to its other


dimensions, the theoretical analysis can be reduced from a three
dimensional elastic boundary value problem into a two dimensional
one by assuming simple distributions of displacements and stresses
along the thickness.
4

The Kirchhoff's plate theory, a two dimensional extension of the


Bernoulli beam theory, neglects the effect of shear deformations
and assumes plane sections to remain plane after bending with the
middle plane of the plate taken as neutral. In addition,
deflections are assumed to be much smaller than the thickness so
that the principles of linear elasticity can be applied. The
Kirchhoff's assumptions result in a fourth order differential
equation, called Navier's equation, with the deflection function of
the middle plane as unknown. Despite the boundary conditions
contraction vital to the Kirchhoff's plate theory, the resulting
two dimensional boundary value problem accurately models the
plate's behaviour in a wide variety of loading and geometry cases.

Ultimately, if the thickness of the plate is very small, resistance


to bending becomes negligible and the flexural rigidity of the
plate approaches zero. Deflections in that case, and after buckling
in general, would then be of the same order, if not larger, than
the thickness; stretching strains in the middle plane of the plate
cannot then be ignored. The von Karman equations take into account
this coupling between bending and membrane action due to
deflections that are large enough for the nonlinear
strain-displacements relations to apply but still small enough for
the linear curvature-displacement relations to hold.

Although sixth order theories of plates (Reissner 1944, Mindlin


1951) taking into account shear deformations are more refined and
complete than the fourth order ones, the use of the latter is
justified in most practical applications.

1.3 Experimental and Numerical Modelling

Exact solutions of the small deflections equations for a plate


compressed in its own plane exist only for a very limited class of
support and loading conditions. They do not exist at all for the
large deflexions theory. Approximate solutions must be used in most
5

cases. The most direct, but perhaps least economical, way of


predicting plates behaviour is through experimental investigation.
Only experiments can confirm the predictions of analytical or
numerical results and closely simulate the actual behaviour of the
plate. The high financial cost and the various technical
difficulties, mainly the simulation of the out of plane boundary
conditions and the measurement techniques, however reduce the
potentials of experimental modelling.

Approximate analytical solutions have been and still are widely


investigated and used. In 1891, Bryan used an energy method to find
the critical stresses of a simply supported compressed plate
(Timoshenko & Gere 1961). The Rayleigh method for the vibration of
elastic systems was later developed (Timoshenko & Gere 1961) and
applied to the stability problem of plates; results were derived
for various support and loading conditions using a Fourier series
approximation of the deflections. Alfutov and Balabukh (1967)
proposed another energy method for the determination of the
critical load that does not require the determination of the
membrane stress distribution which is replaced instead, in the
energy equation, by a statically equivalent system of stresses. As
for large deflections and the coupled von Karman equations, various
analytical approaches are also possible. Berger (1955) uncoupled
these equations by neglecting the second invariant of the strains
in the middle plane. Nowinski and Ohnabe (1958) pointed out that
this assumption, and the stemming results, can be accurate enough
in the case of immovable edges but become highly inaccurate
otherwise. Nevertheless, the method has been widely used and
developed by a large number of authors. Resort is also made to
classical energy methods and variational principles that employ
various forms of displacements and membrane stress approximations.

Although some of the analytical methods are general in scope, it is


rather difficult to find accurate representations of the
deflections in the case of non-rectangular plates and complicated
boundary conditions. The lengthy calculations required were made
6

unjustifiable, except perhaps for the purpose of research and


teaching, by the advances made in the field of computational
mechanics as numerical methods like the finite difference and
finite element developed rapidly.

The finite difference method has been successfully applied to the


problems of plate stability. The method is fairly simple and
versatile. It is not however always easy to program and boundary
conditions may be difficult to accommodate. In addition, serious
modelling problems usually arise on free boundaries and near holes.
The finite element method is nowadays the most popular tool of
analysis in engineering design. The generality of its concepts and
the accuracy of its predictions compared to previous methods have
made it particularly attractive to designers. Some disadvantages do
exist however. The sometimes poor accuracy of the results
especially in stress concentration regions, the computational cost
required by nonlinear and infinite mediums problems, the
complicated mesh when unusual shapes and stress concentrations are
involved can be serious problems in many applications of the FEM.

The recently developed boundary element method, on the other hand,


has introduced two useful concepts in the field of computational
mechanics: that the geometry of a body can primarily and fully be
defined by its boundary surface rather than its volume and that
known fundamental solutions of infinite bodies can be used to
derive highly accurate results.

1.4 The Boundary Element Method.

The boundary element method is based on a numerical solution of


integral equations, as opposed to differential ones, describing the
problem under consideration. Only values at the extreme points of
the integration interval would then appear in that equation, hence
the reduction by one dimension in the size of the problem.
Fundamental solutions of problems with infinite boundaries are used
7

as weighting functions in the integration. As the boundary


conditions are then accounted for in the numerical solution
process, no approximations usually need to be made in the domain of
the body under consideration.

A considerable amount of theoretical work on integral equations of


static and dynamic elasticity problems started in the late
nineteenth century. It was Fredholm (1903) however, who first
presented integral equations of linear elasticity problems. He
enunciated the conditions of existence and uniqueness, known as
Fredholm theorems, of the solutions to those equations. The work of
a series of Russian authors (Kupradze 1965, Muskhelishvili 1953,
Mikhlin 1957, Smirnov 1964), in the 1950's and 1960's, offered a
more rigorous understanding of integral equati.ons. They solved
singular integral equations of plane elasticity problems using
complex potentials. The method of solution was an indirect one
since the unknowns of the problems had no physical significance.
Rizzo (1968), and later Cruse (1969) applied the direct method to
problems of practical importance. More recently, Jaswon and Symm
(1977) also used the indirect method but developed more computer
oriented approaches.

Domain integrals may appear in integral equations either because


fundamental solutions of the problem under consideration cannot be
found and alternative ones are used or, more simply, as a result of
the loading functions. As part of the research effort to eliminate
domain integrals from the integral equations, the dual reciprocity
method has evolved as an efficient technique for transforming these
integrals into boundary ones. The method was first devised to solve
elastodynamic problems by Brebbia and Nardini (1982) and was later
applied to time dependent problems (Vrobel, Brebbia & Nardini 1986,
Vrobel, Telles and Brebbia 1986). Kamiya and Sawaki (1988) proposed
a dual reciprocity solution for the bending problem of plates
resting on Vinkler type elastic foundations. Tang (1987) combined
the method with Fourier series techniques to transform domain
integrals of a number of problems.
8

The computational efficiency of the boundary element method is well


established now. Because of the smaller order of the final system
of equations, the simpler mesh and the more accurate results
specially in infinite domains and stress concentration regions, the
method can be very advantageous over other numerical techniques in
computer aided design environments. It has to be mentioned,
however, that those advantages are gained at the expense of a more
complicated mathematical background and programming requirements.

1.5 Plate Stability by BEM.

General fundamental solutions of the plate buckling equations


cannot be found so that the fundamental solutions of the plate
bending problem are usually used instead. Consequently, domain
integrals containing domain curvature terms arise in the
formulations which usually result in a combined boundary-domain
solution of the problem. Nevertheless, the method retains its major
advantages with regard to the accuracy and computational efficiency
despite the difficulties introduced by the existence of those
domain integrals.

The first boundary integral solution of the problem of elastic


buckling of plates is due to Niwa, Kobayashi and Fukui (1974) who
used an indirect approach to solve for plates uniformly loaded in
their plane. Later, Bezine (1983) then Costa and Brebbia (1985)
used the direct method to obtain a general formulation of the
problem. The numerical implementation required a domain
discretization with the three curvatures as unknowns at every
domain node. Syngellakis and Kang (1987) eliminated the curvatures
from the domain integral and presented a solution which required
the modelling in the domain of the deflections only. On the other
hand, solutions of large deflections problems by BEM have mostly
considered nonlinearities induced by lateral loads (Kamiya and
Sawaki 1982, Katsikadelis and Nerantzaki 1988). Tanaka (1983)
9

presented an integral formulation of the incremental form of the


coupled von Karman equations and proposed a transformation
technique of the traction boundary conditions into stress function
boundary conditions. No numerical results were however shown.
Elzein and Syngellakis (1989) considered the problem of plates with
imperfections loaded in their own plane and used high order domain
elements to obtain approximations of the curvatures.

1.6 Scope of the Present York

The validity, efficiency and versatility of the BEM in elastic


plate stability problems is assessed by implementing various
integral formulations for various stages of loading and solution.

A general account of the plate stability theory is given in the


second chapter. Bifurcation buckling of structural systems is first
described; governing differential equations for the linear theory
of plate buckling and large deflection theory of imperfect plates
are then derived, together with in-plane and out-of-plane boundary
conditions.

In chapter three, a boundary element solution of the plane stress


problem, starting from the Rayleigh-Green theorem expressed in
terms of the stress function, is formulated. The resulting
algorithm provides an accurate estimate of the membrane stress
distribution in the domain of the plate prior to and at buckling.
The same algorithm is later modified to take into account nonlinear
terms - resulting from middle plane stretching strains, that affect
the membrane stress distribution during and after buckling.

In chapter four, boundary integral equations which contain only


domain deflections but not curvatures are derived and a numerical
implementation of the formulation is achieved. The efficiency of
various boundary and domain element models is assessed. Special
treatments of corners and free boundaries are also presented.
10

The dual reciprocity technique is applied, in chapter five,


combined with a Fourier series representation of the deflections
and curvatures, to transform domain integrals into boundary ones
and thus obtain a purely boundary type solution. Two models for
domain deflections are proposed. The domain integral is actually
transformed into a weighted sum of domain integrals containing
simple trigonometric functions. These integrals are then easily
reduced to the boundary by applying the Rayleigh-Green reciprocal
identity.

An incremental boundary integral formulation of the von Karman


equations for initially imperfect plates is presented in chapter
six. High order domain elements for the plate deflections are
introduced that result in an approximate representation of the
domain curvatures. The bifurcation path of buckling and initial
post-buckling behaviour are thus modelled. At each load step,
second-order terms in deflections, curvatures, membrane stresses
and boundary unknowns are taken into consideration in an iteration
procedure that follows the solution of the linearized primary
system.

All boundary integral formulations mentioned above are followed by


numerical implementations tested through computer programs written
for the purpose. Comparison is made with experimental, analytical
and numerical results obtained from existing literature.

Finally, general conclusions and recommendations for future


research are discussed in chapter seven.
CHAPTER TVO

PLATE STABILITY THEORY

2.1 Introduction

In this chapter, the classical theory of plate stability is


described. The buckling and initial post-buckling behaviour of
structural systems in general is first presented. The differential
equation describing the linear bifurcation (or eigenvalue) problem
of plates loaded in their own plane is then derived, starting from
the Kirchhoff's assumptions. Next, by assuming that deflections are
of the same order of the thickness, the von Karman equations are
derived for plates with initial deviations from flatness. The plate
stability problem under small and large deflections is thus
formulated as a boundary value problem.

2.2 Stability of Structural Systems

Simple structural models having a finite number of degrees of


freedom can be used to illustrate essential properties of buckling
and post-buckling behaviour. These properties are equally valid in
the case of continuous structures having an infinite number of
degrees of freedom. This approach is all the more efficient since
numerical methods such as boundary element and finite element
actually treat continuous structures such as plates, as
discontinuous ones.

Such models, also used by Budiansky (1974), will be discussed here.


Consider a rigid rod, shown in figure 2.1, having a length L and an
initial imperfection 90 , Rotation at the base of the rod is
constrained by an elastic spring providing a restoring moment f(9),
and vertical translation is restrained. A vertical compressive load
P is applied at the top of the rod.
12

L
!
Figure 2.1 Simple stability model, rigid rod.

The static equilibrium equation can be written as follows:

P L sin(9+00 ) f(0)

In addition to the fundamental state of equilibrium corresponding


to 960, a set of equilibrium states can be represented by:

f(0)
P

These states can be shown, for the case where K2 is negative, on


P-0 diagrams as the ones in figure 2.2; the two diagrams a and b
correspond to positive and negative values of 90 , respectively. The
dashed lines in figure 2.2 indicate the equilibrium paths in the
absence of initial imperfections (90=°). At the critical load
13

Pc=KI/L, the fundamental equilibrium state appears to be splitting


into a number of possible paths in what is described as
"bifurcation buckling".

p p

0 .......
.......
~",
9 9
9 >0 9 <0
(0.) (b)

Figure 2.2 Equilibrium paths of the imperfect rod model.

The parts of the curves which do not cross the origin are of course
of no physical relevance. In diagram a, a local maximum is reached
in the vicinity of the critical load. This maximum load, called
"snap load", is actually lower than the critical load P ; thus the
c
equilibrium path is an unstable one. This type of sharp buckling is
also called "snap buckling". In diagram b, a considerable increase
in deflections also occurs in the region of the critical load but a
stable equilibrium follows; some post-buckling strength is
exhibited in that case. As to which equilibrium path will actually
be followed by the structure clearly depends on the direction of
the initial imperfection. Other possible equilibrium paths can be
obtained by varying the relative signs of parameters KI , K2 , K3 .•. ;
14

however, the two types of buckling described above are somewhat


typical.

Figure 2.3 shows the different equilibrium paths obtained for


different initial imperfections for both negative and positive
values of eO.

p p

--------~--------~e --------~--------~e

a >0 9<.0
(0.) (10)

Figure 2.3 Equilibrium paths for various initial imperfections.

The buckling and pre-buckling deflections are clearly affected by


the value of the initial imperfection in both stable and unstable
equilibrium modes. The position of the snap load becomes lower as
the amount of the imperfection is increased. In the post-buckling
region however, the curves appear to be converging; the initial
post-buckling strength therefore remains essentially unaffected by
the value of the initial imperfections.

When considering the problem of plate stability, we will


essentially try to establish the critical buckling load and the
stable equilibrium path in the initial post-buckling region.
15

2.3 Linear Theory

In what follows the governing differential equations of a


homogeneous plate are derived assuming that the material is elastic
and the strain-displacement relations are linear. Load is applied
on the edges of the plate in a direction parallel to its own plane.
It is assumed that no transverse loading is applied and the effects
of body forces and thermal stresses are neglected.

A cartesian coordinate system of reference xyz is adopted. The


plate, shown in figure 2.4, has a domain Q and a boundary r. The
dimension of the plate in the z direction, that is its thickness h,
is considerably smaller than its other two dimensions. The origin
of the z-axis is taken in the undeformed middle plane of the plate.

Figure 2.4 Thin flat plate under the action of in-plane loads.

The Kirchhoff's theory of plates assumes that plane sections remain


plane after deformation and sections originally normal to the
undeformed middle plane of the plate are, after deformation, normal
16

to the deformed middle plane. Distortions due to shear deformations


are thus neglected. In addition, normal stress azz and shear
stresses azx and azy are considered negligible when compared to the
other stress components. The transverse deflection w is assumed to
be much smaller than the thickness and the stretching strains in
the middle plane of the plate are therefore neglected. Linear
strain-displacement relationships hence apply.

The deformation of a plate strip in the xz plane is shown in figure


2.5; a yz strip would have similar characteristics. The three
components of displacement ux '
uy and Uz are approximated as
follows (Timoshenko & Gere 1961, Dym & Shames 1973). Because of the
small thickness, uz is assumed to be constant along the thickness
of the plate. Hence, the deflection w of the middle plane
completely describes displacement u z :

u z (x,y,z)=w(x,y) (2.1)

0.

-- -- - - - - - - - - - - ----

z b
w -1

Figure 2.5 Deformation of an xz thin plate strip.


17

Two components of deformation Uxuy are considered, the


and
stretching displacements uxs and uys and the bending displacements
uxb and uyb :

(2.2)

The stretching action is due to the in-plane load acting on the


edges of the plate. Displacements u and u can again be assumed
xs ys
to be constant along the thickness.

As for the bending action and in accordance with the assumption


that plane sections remain plane, line ab in figure 2.5, deformed
into a'b', undergoes both a vertical translation and a rigid
rotation. Hence the following expressions of uxb and uyb :

aw(x,y)
uxb=-z
ax
(2.3)
aw(x,y)
uyb=-z
ay

The displacement field (u x ,U y ,u Z ) has now been fully described in


terms of the middle plane deformation parameters:

aw(x,y)
ux=uxs(x,y)-Z--------
ax

aw(x,y)
uy =uys (x,y)-z-------- (2.4)
ay
18

The linear strain-displacement relations are given by:

au x
£xx =
ax
au
£yy = - L (2.5)
ay

_
1 _ _ +-
( aux L
au)
Yxy =
2 ay ax

Differentiating £xx with respect to y twice, £yy with respect to x


twice and y xy with respect to x and y and combining the three
expressions we obtain the following strain compatibility relations:

+ o (2.6)

Combining (2.5) with (2.4) on the other hand:

auxs(x,y) a2w(x,y)
£xx = - z
ax ax 2

auxs(x,y) a2w(x,y)
£yy = - z (2.7)
ay a/
a2w(x,y)
Yxy=
+( auxs(x,y)
ay
+
auxs(x,y)
ax
) - z
axay

The stress-strain relationships for an elastic isotropic body is


given by the Hook's law. In the following equations azz ' azx and
azy are assumed to be zero:
19

€xx =
E ~(axx - vayy )

~[-vaxx + aYY1
€yy = (2.8)
E

l+v
Yxy = - - a
E xy

Or expressing stresses in terms of strains:

E
axx =
I-v
2 [€xx + v€YY 1
E
a (2.9)
YY 2 (V€xx + €YY)
I-v
E
a = -y.
xY l+v xy

Ve can now rewrite equation 2.6 in terms of stresses by


substituting equation 2.8. Ve obtain:

(2.10)

As in the case of beams, it is more convenient to work with stress


resultants shown in figure 2.6. These are obtained by integrating
along the thickness. The inplane forces are thus defined as:

" .J xx
h/2 axx dz
-h/2

J
"yy' h/2 a
yy dz (2.11)
-hl2

"
xy
= J h/2 a
xy dz
-hl2
20

Shear forces Ox and 0y and bending moments Hxx' Hxy and Myy are
defined as follows:

Ox, J hl2 a
xz dz
-h/2
(2.12)
Oy' J h!2 a
yz dz
-hl2

M •
xx
J h!2
zaxx dz
-hl2

M •
yy
J h!2 za
yy dz (2.13)
-hl2

M •
xy
J hl2 za
xy dz
-hl2

Y~ z
z

&,P"
0.. In-plo.ne forces b. Sheo.r forces o.nd MOMents

Figure 2.6 Stress resultants of a plate element.


21

The equilibrium equations for an infinitesimal plate element are


given by:

aaxx aaxy aaxz


+ + 0 (a)
ax ay az

aa1X aa aa1Z
+ 11 + 0 (b) (2.14)
ax ay az

aa zx aa aazz
+
z1 + 0 (c)
ax ay az

Provided axz and ayz vanish at z=±h/2, that is no shear loads are
applied at the surface of the plate, and azz is small compared to
the other stresses, the following relations in terms of the
in-plane forces can be obtained by integrating equations 2.14a and
band 2.10 over the thickness of the plate:

aNxx aN
+
x1 0
ax ay
(2.15)
aN aN
X1 + 11 0
ax ay

o (2.16)

Equations 2.15 and 2.16 can be used to obtain the membrane stress
distribution in the domain of the plate. A further simplification
of these equations can be achieved by introducing the Airy's stress
function F. It can be easily proved that equations 2.15 are
identically satisfied by the following definition of F due to Foppl
(Timoshenko, 1953):
22

Nxy = N
yy
(2.17)
axay

Replacing the expressions appearing in equations 2.17 into


equations 2.16 we obtain:

(2.18)

Having obtained the equation for the membrane stresses, we will


manipulate equation 2.14c to derive the relations describing the
bending action of the plate. First, by combining equations 2.13
with equations 2.7 and 2.9, the bending moments can be expressed in
terms of the middle plane deformation parameters. The stretching
displacement terms uxs and uys disappear during the integration
process and only middle plane deflection w appears in the final
expressions:

Mxx =

M
yy
(2.19)

M -(l-'\I)D
xy
axay

where D=

Multiplying the equation of equilibrium in the x direction (2.14a)


by z and integrating over the plate thickness we obtain the
following equation, provided, again, no shear forces are applied on
the upper or lower surfaces of the plate:

aM xx aM xy
Qx = --'-=- + (2.20)
ax ay
23

Developing the equation of equilibrium in the y direction (2.14b)


in the same way:

Qy = (2.21)
ax

Thus, expressing Qx' Qy in terms of w by combining 2.20 and 2.21


with equations 2.19, we obtain:

Qx= -o[-~~
ax 3
· a3w
axay2
1
(2.22)
[ ,3.
Q = -D
Y ax 2 ay '~l ay3

Integrating equation 2.14c over the thickness and applying


equations 2.20 and 2.21 we obtain:

a2Mxx a2Mxy a2Myy a2w a2w a2w


+ 2 + + N - - + 2Nxy + Nyy -2-= 0
ax 2 axay ai xx ax2 axay ay
(2.23)

Now substituting equations 2.19 into equation 2.23 we obtain the


well known Navier's differential equation, modified to take into
account the in-plane loading:

a2w a2w a2w


4
DV w = N - - + 2N xy
xx ax2 axay
+ Nyy
--;; (2.24)

Equations 2.18 and 2.24, combined with the corresponding boundary


conditions that will be described in section 2.5 of this chapter,
can be used to solve the stability problem of thin plates. The
critical buckling load is defined as the smallest load which can be
applied at the edge of the plate to produce an unstable buckling
configuration. The buckling factor ~ is defined in the following
24

dimensionless form:

where b is the width of the plate and acr is the critical stress
resultant.

2.4 Large Deflections

As the deflections of the plate increase beyond a certain limit,


stretching strains in the middle plane cannot be neglected anymore.
For instance, in the case of a circular plate in pure bending
having a maximum deflection of 60% of the plate's thickness, the
maximum stretching stress is around 18% of the maximum bending
stress (Timoshenko & Goodier, 1970).

Considering moderately large deflections -that is of the same order


of the thickness, linear curvatures-displacements relations hold
but non-linear strains-displacements relations apply.

The expression for the strains in terms of the displacements is


modified in the following way:

au x

xx = - - + _1 [~r
ax 2 ax

au

yy = - L + _1 (~r (2.25)
ay 2 ay

Yxy=
[ au
1 _
_ au
_x_+ -L+ ~.~)
2 ay ax ax ay
25

As a result of the non-linear terms in equations 2.25, the


right-hand side of equation 2.6 becomes non-zero:

f- (2.26)

Since the right-hand side of equation 2.18 stems directly from that
of equation 2.6, the final equation in F becomes:

Eh[(-~-:-y-r- ~; <; 1 (2.27)

Equation 2.24, on the other hand, is equally valid in the case of


large deflections and will be rewritten here for the sake of
completeness using a stress function notation:

a2F a2w a2F a2w a2F a2w


----- 2------+-- (2.28)
ay2 . ax 2 axay axay ax 2 . ay2

Equations 2.27 and 2.28 are the well known von Karman equations.

It is of practical importance to consider initial deviations from


flatness in plates prior to loading. If these deviations are
represented by function wo' equation 2.28 becomes:

a2F a2wt a2F a2wt a2F a2wt


v4w -2-· - 2 + -2-· (2.29)
3y ax 2 3x3y 3xay ax 3/

where wt = w+wO.
26

When initial deviations from flatness are present, lateral


deflections keep increasing even before the applied load exceeds
the critical buckling load; however, they only become significantly
large when that load is reached. The post-buckling strength on the
other hand, as discussed in section 2.2, remains practically
unaffected by those deviations.

As equation 2.27 suggests, the membrane stress distribution


gradually changes as the applied load exceeds the buckling load and
the contribution of the middle plane stretching strains to the
total strain increases. At a certain stage of loading, experiments
show that only part of the plate width, called "effective width",
seems to be carrying the increase in stresses.

The coupled von Karman equations 2.27 and 2.29 combined with the
corresponding boundary conditions can be solved to predict the
stable path of the bifurcation buckling. Both the buckling load and
the initial post-buckling strength can be found. These equations do
not, of course, yield an eigenvalue problem in terms of the
critical load anymore. Thus, a number of methods for calculating
the buckling load from the loading curve obtained by solving the
von Karman equations have been suggested. The critical load can be
defined as that at which the extreme fiber-strain, on the convex
side of the buckle crest, stops increasing. Slightly higher values
are obtained when the load is derived by the "Top of the knee"
approach, best described by its title. Another approach, giving
still higher values, assumes the buckling load to be the one
corresponding to the inflexion point of the curve.

2.5 Boundary Conditions

The differential equations described above in both small and large


deflection situations cannot be solved without prescribing the
right boundary conditions. Two types of boundary conditions have to
be considered: the out-of-plane and the in-plane conditions.
27

s is taken as a direction tangent to the edge at which the boundary


conditions are applied and n is a direction normal to that edge.

2.5.1 Out-of-Plane Boundary Conditions

Three most common cases of out-of-plane boundary conditions are


considered here.

a. Fully Clamped Edges (figure 2.7)

Plo.n YIn 09 0. plo.te


deutped an one side

Figure 2.7 Clamped Edge.

A fully clamped edge is prevented from translating laterally and


from rotating about an axis parallel to that edge. The boundary
conditions for such a case can thus be expressed in the following
equations:

aw
w• 0 -=0 (2.30)
an
28

b. Simply Supported Edge (figure 2.8)

PlClll VieW of Cl Sec:'tlon VIew of Cl plCl'te por't!on

Sll'lply supported plCl'te straply S\.Ippor'ted on one SIde

Figure 2.8 Simply Supported Edge.

A simply supported edge cannot translate laterally but it can


freely rotate about an axis parallel to that edge; thus there can
be no bending moments along it. The boundary conditions can be
written as follows:

Since w=O then --~~-O and the conditions can be written as:
as 2

w=0
(2.31)
29

c. Free Edge (figure 2.9)

Free Edge

Sectton VIew oF a. pla.te portion


Plcan VIew of a. pla.te
WIth Free egdes

Figure 2.9 Free Edge.

The formulation of the boundary conditions on a free edge presents


some difficulties. Intuitively, one would expect the shear force
On' bending moment Hn and transverse moment Hns to be zero. These
boundary conditions were actually used by Poisson (1829).

However, since the differential equation in w is of the fourth


order, only two, not three, boundary conditions can be
accommodated. The inconsistency is due to the assumption that plane
sections remain plane. Kirchhoff (1850) first noticed the
inconsistency and proposed a contraction of the two shear and
transverse moment boundary conditions into one. The physical
interpretation of that contraction has been presented by Kelvin and
Tait (1883).

A view of a free edge, assumed parallel to the y-axis for the sake
of simplicity, is shown in figure 2.10. The assumption that plane
sections remain plane translates into line ab remaining straight
30

after deformation. The assumption that sections normal to the


undeformed middle plane remain, after deformation, normal' to the
deformed middle plane translates into line ab remaining
perpendicular to line a'b'.

~'

Figure 2.10 Contraction of Mxy boundary condition on a free edge.

Because line ab is not distorted, the axy stress distribution can


only be linear and torsional moment Myx acting on the free edge can
be replaced by a couple of horizontal forces ±P h shown in figure
2.10. Furthermore, since line ab does not rotate with respect to
line a'b', it is possible to replace ±Ph by a system of vertical
equal and opposite forces Pv acting over an infinitesimal distance
~ where P =±M I~. Thus, it is possible to replace the whole M
v ~ ~
distribution along the length Xo of the edge, in a similar way. The
difference 6P resulting from two successive couples acting at line
ab would then be:
31

Adding the AP contribution to shear a and replacing x and y by the


more general notation using nand s, we get:

If the resulting vertical force Vn is made zero along with the


bending moment Hn , the two boundary conditions, reduced to their
expression in terms of w, are written in the following form:

(2.32)

Vn -- + (2-'\1) z 0
an 3 anas 2

2.5.2 In-plane Boundary Conditions

The stress function F appearing in equations 2.18, 2.27, 2.28 and


2.29 depends on the membrane stresses of the plate. The state of
traction and/or displacement in the x and y directions has to be
specified on the edges of the plate in order to solve the above
mentioned equations.

Vhen small deflections are considered, bending and membrane actions


are uncoupled and finding the membrane stress distribution is a
classical plane stress problem. Traction boundary conditions need
to be transformed into conditions satisfied by the stress function
if equations 2.18 and 2.28 are to be solved. Consider for that
purpose a curvilinear coordinate system s over the boundary of the
plate with an arbitrary origin 0 and a normal direction n, shown

-31-
32

in figure 2.11. P is an arbitrary point at which F and dF/dn are to


be determined.

Ts
n

Figure 2.11 Curvilinear coordinate system on the boundary of the


plate.

The boundary tractions Tn and Ts' force per unit thickness, are
related to the membrane forces, shown in figure 2.12, as follows:

Tn Nxx cos(n,x) + Nxy cos(n,y)


(2.33)
Ts Nxy cos(n,x) + Nyy cos(n,y)
dy dx
where cos(n,x) and cos(n,y)
ds ds

We can express Tn and Ts in terms of the stress function F using


equations 2.17 and 2.33:
33

Figure 2.12 Boundary Tractions and Membrane Stresses.

T = a ( aF _) dy a ( aF) dx d ( aF )
n ay -;- ~ + ax -;- ~ = ds -;-

or:

d [ :: 1= Tnds
(2.34)

d ( :: ) = -Tsds

On the other hand, F can be deduced from:

F= J [~x+~y 1
aF aF
(2.35)
o
34

Constants of integration have been disregarded in the above


expression since they have no effects on the value of the membrane
stresses. Integrating by parts, we obtain:

F'" [~ +~lP- J
ax ay 0
x d[~)+
ax
y d(~)
ay
(2.36)
o

now combining 2.34 with 2.36:

F = J(X-Xp)T g ds - (2.37)
o

where xp and yP are the coordinates of point P. Equation 2.37


actually provides a physical interpretation of the stress function
as the moment about point P of the resultant of all traction forces
acting over curved segment OP.

aF/an can also be expressed in terms of edge tractions by writing:

aF aF dx aF dy
- - .. - - - - + (2.38)
an ax dn an dn

and substituting equations 2.34 into 2.38:

P p
aF
--- - -dx- JT ds + ~ J Tn ds (2.39)
an dn s dn
o o
35

3F/3n can therefore be interpreted as the projection over the


tangent to the boundary at point P of the resultant of all
tractions forces acting over curve segment OP.

The traction boundary conditions have thus been expressed in terms


of the stress function. The transformation of the displacement
boundary conditions is much more complicated. Ve will therefore
replace them by the traction conditions that most closely
approximate them. For instance, an edge that is restricted against
in-plane displacement will be assumed free to move and an average
traction will be taken along that edge. The effect of such an
approximation will only be noticeable at a late stage of loading,
well beyond the critical buckling load, when the membrane stress
distribution over the plate has greatly deviated from the original
pre-buckling distribution. The effect on the accuracy of the
computed initial post-buckling strength is expected to be minor.
This was actually confirmed by the results presented in chapter 6.

At this stage, we have completed the formulation of the boundary


value problems of small and large deflection theories. A brief
review of the most common, numerical and experimental, plate
stability modelling techniques will now be presented.

2.6 Numerical and Experimental Studies

Design engineers often resort to experimental and/or numerical


modelling to predict the buckling load and post-buckling
deflections and stresses. Among the various existing numerical
methods, the finite difference and finite element methods are the
most popular.

In solutions based on the finite difference method, the plate is


split up into a mesh of nodes at which the deflection w is taken as
unknown. Differentiation of w is obtained through direct
differences and the differential equation of the problem is thus
36
satisfied at the various nodes. A system of simultaneous equations
is hence obtained and solved to yield the nodal deflections.
Although reasonably accurate, the finite difference method is not
always easily programmable. In addition, it may become highly
inaccurate when stress concentrations and holes are involved.

The finite element method discretizes the plate in a set of


elements with a finite number of degrees of freedom. Conforming or
non-conforming linear rectangular elements with either 16 or 12
degrees of freedom, respectively, are usually used. Vith the
non-conforming element, inconsistency occurs in the modelling of
the normal slope. Vhen the conforming element is adopted, the
inconsistency is removed and accurate results can be obtained.
Because of the generality of its concepts, the accuracy of its
results and its programmable aspect, the finite element method is
the most commonly used tool of analysis in design offices.
Improvements on accuracy and overall computational efficiency are
however still largely possible. Modelling difficulties on free
boundaries and near stress concentrations still exist.

Finally, experimental investigation of plate stability problems can


be expensive; if properly designed and monitored, it is however the
most reliable research tool. In addition to obtaining the loading
curves, particular topics such as "effective width" (Hoff 1938,
Vinter 1947), secondary buckling where sudden transitions to higher
modes take place (Stein 1959, Uemura & Byon 1978), in-plane
boundary conditions effects (Uemura & Byon 1978, Yamaki 1961), load
eccentricity effects (Valker 1964, Rhodes & Harvey 1975) etc ...
have all been thoroughly studied through experimental
investigation. Experimental records have also provided a reliable
reference against which results of numerical and analytical methods
can be compared.

Results obtained by analytical methods, finite element programs and


experimental investigations will be used for comparison with and
assessment of boundary element results.
37

2.7 Conclusions

The small and large deflections theories of plate stability have


been formulated as boundary value problems. The differential
equations with the corresponding boundary conditions have been
derived.

These differential equations can now be transformed into integral


ones and the boundary element method can thus be applied.
CHAPTER TBRBB

MEHBlWIB STATE OF STRESS

3.1 Introduction

In the previous chapter, we have derived the uncoupled differential


equations (2.18) and (2.24) for the small deflection theory of
plates. Through the first equation the membrane stress distribution
for a plane stress problem can be obtained. In this chapter, a
boundary integral formulation of this problem, based on the
Rayleigh-Green theorem, will be derived. A boundary element
solution and its numerical implementation will then be presented.
Finally, results from a computer program written for the purpose
will be compared with other results from literature.

3.2 Boundary Integral Formulation

In this section, boundary integral equations of the plane stress


problem will be derived in terms of the stress function F.
Alternatively, integral equations governing displacement rather
than stresses can be obtained and fundamental solutions for the
displacement used instead. In-plane displacement boundary
conditions would then be easily accommodated in the solution. The
present formulation is however more suitable in our case because
it can be easily extended to calculate the membrane stress
distribution of the large deflections problem, by including the
non-linear terms appearing on the right hand side of the von Karman
equation 2.27.
39

The equation of the plane stress problem in terms of the stress


function F will be rewritten here with its corresponding boundary
conditions, for completeness:

(3.1)

where F and aF/an are specified on any boundary point P(xP,yp) as:
P

F - J(X-Xp)T s ds -
o

P P
aF
--= -dx-
dn
J
T
s
ds + -dy-
dn
J n ds
T
an
o o

s is a curvilinear coordinate system on the boundary r shown


in figure 2.11, 0 its arbitrary origin and n the normal direction
to the boundary at point P.

The integral equations will be derived using the following two


fundamental solutions:

r2
u1 (Q)--- lnr
8n
(3.2a)
r
u2 (Q)---(21nr+1)cos ex
8n

where r designates the distance between source point P and field


point Q and ex the angle between the r vector and an arbitrary
direction usually taken as the normal to the boundary at point P,
as shown in figure 3.1.
40

n
n

y p
b-- r

<Pc

Figure 3.1 Source and action points.

u1 and u2 have the following important property:

(3.2b)

where is the dirac delta function defined as:

:
~p

if p
~p(Q)
{ if p
" Q
= Q

Another property stems from the previous one for a continuous


function f:

(3.3)

where f(P) is the value of f at point P.

Ve will now derive the Rayleigh-Green identity containing the


biharmonic operator, using the Green's theorems.
41

The Green's second identity, described in appendix A, can be


written for two functions f and g, continuous and differentiable
twice over domain 2 and boundary r, as follows:

d2 ... J (f
ag af
- - - g - - ) dr
an an
(3.4)
r

Applying equation 3.4 and replacing f by u (u 1 or u2 ) and g by vlF:

d2 = J(u a(vlF)
an
- vlF -
au
an
) dr (3.5a)
r

Now applying the same identity with vlu for f and F for g:

J 4
(vluvlF-FV u) d2 =
J (vlu -
aF
- F
a(vlu)
) dr (3.5b)
an an
2 r
Adding equations 3.5a and 3.5b we get:

a(vlF) au aF a(vlu)
J
d2 = ( u - - - vlF- + vlu- - F
an an an an
)dr (3.6)
r
Equation 3.6 is the well known Rayleigh-Green identity for the
biharmonic operator. Substituting the left-hand terms of that
equation by their values obtained from equations 3.1 and 3.3 and
taking into account the singularity of the integral when the source
point is on the boundary, we obtain:

a(vlF) au aF a(vlu)
cZ(F(P» + J( Uanl - - - vlF- + vlu- - F
an an an
)dr = 0 (3.7)
r
42

where Z is a function of f defined as follows:

i f u=u 1
Z(f) =
{ :flan i f u=u 2

and c is a factor given by:

l. if P is inside the domain Q


c ={
.c/2n if P is on the boundary r
.c is the discontinuity angle at a corner point of the boundary
also shown in figure 3.1. Thus, on a smooth boundary.c =n so c=1/2.

Integral equation 3.7. contains only boundary integrals. Functions


u, au/an, v2u and a(v2u)/an can all be directly derived from u1 and
u 2 • Expressions of these functions are given In appendix B where u1
and u 2 are divided by the plate rigidity D in accordance with the
physical intepretation of the fundamental solutions that will be
given in chapter 4. Values of F and aF/an on the boundary can be
obtained from the boundary conditions of the problem. A physical
interpretation of these functions in terms of the boundary traction
has already been presented in section 2.5. Thus, if source point P
is only placed on the boundary, v2F and a(v2F)/an are the only
unknowns in equation 3.7. Yriting two equations in terms of u1 and
u 2 ' assuming the source point P is on a smooth boundary, and
separating known, right-hand side, from unknown, left-hand side,
functions in the equation, we obtain:
43

The evaluation of functions VZF and 3(VZF)/3n does not in itself


determine the membrane stress distribution in the domain of the
plate. Additional boundary integral equations are needed for that
purpose. Equation 3.8 is differentiated with respect to directions
~ and ~ where ~ and ~ can be either x or y. Using c=l for a domain

position of source point P, we obtain the following equation,


rearranged to have the single term a2F/3~a~ on the left-hand side:

(3.10)

where u~~ indicates differentiation of u 1 with respect to ~ and ~.


Equation 3.10 can be used with (a=y,a=y), (a=x,a=y) and (a=x,a=x),
alternately to obtain the value of membrane forces Nxx ' -Nxy and
Nyy ' respectively, at point P. Explicit expressions of the
various functions of uxx ' Uxy and uyy appearing in equation 3.10
are given in appendix B.

Equations 3.8, 3.9 ~nd 3.10 can now be solved numerically. The
formulation of the numerical problem is described in the next
section.

3.3 Boundary Element Solution

The boundary of the plate is divided into a set of Nbe boundary


elements consisting of straight segments. Linear discontinuous
elements with two internal nodes, shown in figure 3.2, are used.

The linearly discontinuous element has been actually chosen in


consistency with the plate buckling programs that will be described
in later chapters. The reason for such a choice in plate buckling
solutions will then be fully explained.
44

y 5 = -0.5 Se = 0.5
eX I X
Se= -1 I Se = 1
I
l----.
Se

Figure 3.2 Linear discontinuous boundary element.

Figure 3.2 clearly shows that a useful distinction should be made


between the nodes at the extreme points of the element only used to
define its location, and those inside the element at which the
unknowns of the system are actually taken. The first set of nodes,
marked by dots in figure 3.2, will be called mesh nodes and the
second set, marked by letter X, simply nodes or unknown nodes. Any
boundary function g can now be defined in terms of its values at
the boundary nodes using linear interpolation_functions t1 and t2:

t1 and 12 are given in terms of the local coordinate system se'


also shown in figure 3.2, as follows:

112 - s e
(3.12)

By using a Gaussian integration scheme with


Ng stations, the
following discretized form of equations 3.8 and 3.9 is obtained:
45

tIJr[:
i=l j=l
['1<.1'F)1+ 02<.1'F)2) -uh '(~F)
(i"k)

J au
-i-Fdr - u
J
a(i-F) 1
dr= -Z(f(P»+
L N
be
N

IJ ~[i-u-
L aF
- a(i-u) 1
F w. +
an an 2 an an jJ
k k i=l j=l
(i"k)

[
-'1 aF
v-u--dr -
J a(
---Fdr
i-u) (3.13)
an an
k k

where IJI is the jacobian of the transformation from global to


local coordinates, Wj the Gaussian integration weights and k the
boundary element containing source point P. Singularities in the
integrals over element k prevent numerical integration; such
integrals are evaluated analytically and are shown in a continuous
form in equation 3.13.

Equation 3.10 can be similarly transformed into an algebraic form


and the following equation is obtained:

_a(_~_F_) Jth
(3.14)
No singularities occur in the integrals of equations 3.10 and 3.14
because the source point will only be placed inside the domain.
Values of the membrane stresses on the boundary can either be
extrapolated from values inside the domain or they can be
46

calculated by using equations (2.33) giving the boundary tractions


in terms of the three membrane stresses, in addition to the
equation:

(3.15)

The value of Ng depends on the required degree of accuracy. For


most of the numerical integrations of this chapter, 4 points scheme
gave accurate results. Vhen the source point is close to the
boundary element over which the integration is being performed,
pseudo-singularity might occur and a higher value of Ng , 8 or 16
points, would be needed to obtain a reasonable accuracy.

Once the boundary has been discretized, the source point is placed
on the 2Nbe boundary nodes. Equation 3.13 is evaluated for both
u=u 1 and u=u 2 . A system of 4Nbe equations with 4Nbe unknowns -2
nodes per element and 2 unknowns iF and a(iF)/an per node, is
thus obtained:

(3.16)

where {b f } is a vector containing the boundary unknowns, [B1l


contains the coefficients resulting from the integrals on the
left-hand side of equation 3.13 and {B 2} contains the values of the
terms and integrals on the right-hand side of that equation.

The system of equations (3.16) is solved and vector {bfl is


obtained. The source point is then placed at Ndn selected domain
nodes at which the plane stress state is to be determined. Equation
3.14 is applied to each of these nodes and forces Nxx =a 2Flay, 2

Nxy =-a 2F/axay and Nyy =a 2F/ax 2 are thus evaluated. In accordance
with the matrix notation used before, the three membrane forces of
all Ndn domain nodes can be assembled in a 3Ndn vector {N«~}, and
the following matrix equation is obtained:

(3.17)
47
Vector {B3 } results from the first term in the double summation of
equation 3.14 and matrix [B4 ] from the remaining two terms.

3.4 Numerical Implementation

The boundary element solution described above has been implemented


in a computer program PLASTR, written for the purpose. The program
is written in FORTRAN-77 on a VAX/VMS microcomputer. The program
automatically discretizes the boundary of polygonal or circular
plates with the user having full control over the number of
boundary elements on each part of the plate. In addition to
concentrated forces, up to linearly varying normal and shear
loading can be accommodated by the program. Nodes at which the
membrane stresses are to be determined are also input to the
program. A flow chart of this program is shown in figure 3.3.

Read Data
I
Prepare F and dF/dn
boundary conditions
I
Assemble coefficient
matrices and vectors
L
Solve system of equ-
ations
I
Calculate membrane
stresses
I
Vrite results

Figure 3.3 Program PLASTR flowchart.


48

After reading the data, the program discretizes the boundary


automatically if required to do so by the user. The stress function
boundary conditions are then evaluated by calculating the resultant
tractions Tt acting on the portion of the boundary between
curvilinear origin 0, randomly taken as the first mesh node of the
first boundary element, and point H at which the values of F and
3F/3n are required. The projection of Tt on the tangential
direction to the boundary at point H equals 3F/3n at that point and
the moment of Tt about H equals F. To minimise the time required
for evaluating these boundary conditions and to avoid repeating
unnecessary operations, the values of Tt at all mesh nodes and
their moments about these nodes are evaluated and stored in an
array. Next, the coefficient matrices [B 1 ] and [B 4 ] and vector
matrices {B 2}, {B 3} are assembled. The system of equations is
solved and the membrane stress vector is calculated. Output data is
finally written in an output file.

3.5 Results

Three plane stress problems taken from existing literature have


been solved using PLASTR program to demonstrate its validity.

Example 1

A simple example, shown in figure 3.4a, of an axa square plate


(a~20) uniformly compressed in the x direction along its two
parallel edges was first tried. The membrane stress distribution
for this problem is uniform over the domain the plate:

~x -1

a = 0
~

a = 0
"
------==0-----=-
49

-- -
DO 0 0
Nloe=8 Nloe=16 Nloe=24 Nloe=40

-a-

EXAMPLE 1

+: BEM. 8 Eleme nts


X: BEM. 16 Eleme nts
*: BEM. 24 Eleme nts
Y: BEM. 40 Eleme nts -
-: Exact solut Ion

CD
o
-'~~--~-.--r-'--'--~-r--'--r-.
'0.00 4.00
-b- x

Figure 3.4 Membrane Stress by BEM: Example 1.


50

Meshes having 8, 16, 24 and 40 elements, of equal length and


equally distributed among the four edges, were used. These are
shown in figure 3.4a. The axx distribution obtained by the program
along the horizontal centre line of the plate is plotted in figure
3.4b. Results obtained by the program clearly converge as the
discretization mesh is refined. The smaller accuracy of the
stresses at the two points near the boundary is due to the
pseudo-singularity of the integral. When using 40 elements the
value of -0.9832 is obtained and the error is 1.68%. The
convergence of the stresses obtained at these two points
demonstrates however that accuracy can still be improved either by
refining the mesh or by increasing the order of the Gaussian
integration scheme. The accuracy can also be increased by resorting
to special techniques for dealing with pseudo-singular integrals;
such techniques will be described and used in chapter five.

Example 2

Timoshenko & Goodier (1970) propose an analytical plane stress


solution for the cantilever beam of Ie! gth L, depth 2c and a narrow
cross-section, shown in figure 3.5a. J ne beam is under the action
of a concentrated load P acting along its free end. They obtained
the following expressions for the membrane stresses:

3P
O'xx - --xy
2c 3

O'xy -~:.+ - 1
4c
Y:c

O'yy 0

Boundary element solutions of the problem were obtained using 12,


14, 20 and 32 elements discretizations shown in figure 3.5a. The
51

L I
/
It
Lx 1~ It
V
V
L=8,
c=l,
P=I,
P V

1 : ~ ~ Jt : ~ Jt~ ': ~ ::J ~:<; ~ ~ ~: ~J


, , , ,
, , } , , }
~ r , , ; r i )

Nbe=12 Nbe=14 Nbe=20 Nbe=32

-a-
EXAMPLE Z

6>ey c
CD +
c
x

c
c'"

c
~
c
+. BEM. 12 Elements *
X. BEM. 14 Elements
c
N
c
*_ EEM. 20 Elements
~ BEM. 32 Elements
c -. Analytlcal.Tlmoshenko&Goodler
~:;---.--'r--.---r--.---'--.---r--,---t--.---; Ylc:.
0.00 0.20 0.40 0.60 0.80 1,00 1.20

c
c XIL

0.32 0.48 0.64 0.80 0.96 1.12

-;-
+
c
~

+
c
OJ?
+.X.BEM. 12 Elements
BEM. 14 Elements
1
*. BEM. 20 Elements
y. BEM. 32 Elements
6xx c
~
--. Analytlcal.Tlmoshenko&Goodler

-b-

Figure 3.5 Membrane Stress by BEM: Example 2.


52

resulting profile of the normal stress axx along the horizontal


centre line and the shear stress axy along half the vertical centre
line were plotted in figure 3.Sb and compared to equivalent results
given by Timoshenko & Goodier.

'Again, there is clear convergence of the results as the number of


elements is increased for both normal and shear stresses plot. Part
of the error is due to the fact that the quadratic shear load is
modelled as a set of linear loads. Thus, if a quadratic variation
of load is accommodated in the program, higher accuracy would be
obtained at a lower number of elements.

Example 3

A more general example having a non-uniform membrane stress


distribution is shown in figure 3.6a. A rectangular plate of width
d=l. and depth 3d-3. is uniformly compressed in the y-direction
along half the width of its two horizontal edges. The problem was
solved using both PLASTR and another plane stress boundary element
computer program based on displacement fundamental solutions and
constant elements (Brebbia, 1980). Results were obtained from each
of these programs using meshes, also shown in figure 3.6a, giving
the same overall number of unknowns Nbu=120. The direct stress ayy
as obtained by both solutions is plotted in figure 3.6b along the
vertical center line of the plate. The two curves are reasonably
close although the difference is slightly higher than average at
two intermediary points.

3.6 Conclusions

The validity of the boundary element solution of the plane stress


problem in terms of the stress function has been demonstrated. This
solution will not only provide the fundamental membrane stress
distribution before buckling in linear problems but will also be
modified and extended to the non-linear solution in chapter 6.
53

!H
I
3d

1 D iti
..,.

NIM-30 Nbe=60
Llneo.r Discontinuous Ele...ents ConFtant [le"ents
Stress 'unctlon Solution Dlspl=_nt Solution

-a-

EXAMPLE 3

+ +

Stress Function Solution

Displacement Solution. Brebbla

I
1.20 1.60 2.00 2.40

-b-

Figure 3.6 Membrane Stress by BEM: Example 3.


CBAPTBR POUR

CRITICAL LOADS

4.1 Introduction

Critical buckling loads at which plates suddenly, undergo


considerable deflections can be obtained by solving the boundary
value problem in the form of differential equation 2.24 subject to
the boundary conditions described in section 2.5. The integral
formulation of this problem is not a purely boundary one as domain
integrals containing in-plane stresses cannot be easily reduced to
integrals over the boundary. However, a boundary element solution
of the problem remains efficient because of the potential accuracy
inherent to the boundary element method. In addition, a technique
for transforming the domain integrals into boundary ones has proven
successful and will be presented in the next chapter. In this
chapter however, a direct boundary element solution requiring
boundary as well as domain discretizations is presented. An
algebraic system of equations of the eigenvalue problem is
obtained. The system of equations is solved and the lowest load
factor \ -that is the buckling load factor, and fundamental
buckling mode are obtained. The accuracy of the method is tested by
comparing its predictions to results from existing literature.

4.2 Boundary Integral Formulation

The Navier's equation derived in chapter two will be rewritten here


with its boundary conditions, for completeness:

DV4w = >J.(N Ot~' w, Ot~) (4.1)


where
a2w a2w a2w
L(N Ot ~,w ,Ot ~) = N - - + 2Nxy + N
xx ax2 axay yy~
55

Nxx ' Nxy and Nyy are the membrane stresses corresponding to a
reference boundary traction Tn' Ts and A is a load factor.

The following conditions may be satisfied on the boundary:

fully clamped edge:

aw
w=O --= 0 (a)
an

simply supported edge

a2w
w=O -2-= 0 (b) (4.2)
an

free edge

(2-v)----::~ o (c)
ana/

The term w Q indicates a double differentiation of w with respect


,ex....
to a first and then to ex.

The fundamental solutions u1 and u2 defined in equation 3.2 and


used in the formulation of the plane stress problem are divided by
the plate rigidity D and used here again. So defined, u1 is
proportional to the deflection, at the field point a of an infinite
plate in bending, due to a transverse concentrated unit force
acting at the source point P; and u2 is proportional to the
deflection at point a of such a plate due to a concentrated unit
moment acting at point P about an axis having an angle ex with line
PO, as shown in figure 3.1 (Timoshenko & Gere, 1961).

Multiplying equation (4.1) by functions u denoting either u1 or


u2 ' and integrating over domain g we obtain:
56

AJ UL(N«~,W,«~) d2 (4.3)
2

The generalised Rayleigh-Green identity for plates, taking into


account transverse moments jump at corners is derived in appendix
A. It can be written in the following form:

DJ(UV4W-WV4U)d2 = J[ awau
Vn(u)w - H (u)-- + --H (w) - uVn(w) dr
1
n an ann
2 r

(4.4)

where
Vn equivalent shear force
Mn bending moment
Mns twisting moment
Nc : number of corners

+
f I = discontinuity jump of f at corner.

Combining equations (4.3) and (4.4), using the property of the


fundamental solutions expressed by (3.2b), we obtain:

(4.5)

where

Id =J uL(N«~,w,«~)d2
Q

aw au
- H (u)-- + --H (w) - uVn(w) dr
1
n an ann
57

N
c
J b= [[W(' Mns(u) ,>-U(' Mns(w) ,>1
h1

The definitions of c and Z are given in section 3.2.

Equation (4.5) includes one domain integral Id which contains the


unknown domain curvatures a2w/a«a~. Previous boundary element
solutions of the plate buckling problem have taken this equation as
a starting point for the numerical implementation of the
problem (Costa and Brebbia, 1985). The domain of the plate is
discretized and three curvature unknowns are taken at every domain
node. A solution was later proposed which eliminated integral Id
and replaced it by a domain integral which contains only unknown
deflections (Syngellakis and Kang, 1987), through a process that is
described next.

Id can be rewritten as follows:

a
(N ~+ N ~) (N ~ N ~)]
ax t xx ax xy ay
+ u
ay t xy ax
+
yy ay
dQ -

aw aw aN aw_ +
--+ --+
y__
__x..... _a_N.Jl.Y.ll..Y__a_w_l d Q
ax ay ay ax ay ax

The second integral on the right hand side vanishes by applying the
equations of equilibrium 2.14 of a plate infinitesimal element. The
first integral can also be rewritten as:

Id =J [ aax [u (Nxx ~
ax
+ N
xy ay
~)l + _a_[u (Nx
ay y ax
~ + NYY ~)lldQ-
ay

J [-auax- NIxx-axaw- + Nxy-ayaw-1 + -au- N


ay
I -aw-
xy ax
+ N -aw- dQ
yy ay
II
Q
58

By applying the first Green identity, described in appendix A, the


first integral can be transformed into a boundary one; the second
integral can be rewritten as the sum of yet another three domain
integrals:

Id J[u (N
aw aw aw
aw
- - + N --)dx - u (N - - + N --)dy
xx ax xy ay xy ax yy ay
1
r

J[- aax ((Naxx-uax- au) + -


+ N --)w
xy ay
a (
ay
au au
(N - - + N --)w )]dQ +
xy ax yy ay
Q

J w[
aN xx au
--+
aNxy au
--+
aN xy au
--+
aNyy
1
-au- dQ +

+
ax ax ax ay ay ax ay ax
Q

J ,2~ + 2.xy
a2u ,2
+ N --;- dQ
1 (4.6)
xx ax axay yy ay
Q

The first integral on the right hand side of equation (4.6) is a


boundary one that can be expressed in terms of the boundary
tractions T,
n Ts . The second integral can again be transformed by
the Green's first identity into a boundary one. Finally, the third
integral vanishes on account of equilibrium equations 2.14 and Id
is finally expressed as the sum of one boundary and one domain
integral:

(4.7)

where

J[ U[T
n an
~ + Ts ~l-
as
W[T ~ + T ~ll
n an s as
dr
r

J+xx ::~ +
2N _a_2_u_
xy axay
+ N
yy ay
a2; 1dQ
Q
59

Equation (4.5) hence becomes:

(4.8)

Equation (4.8) can be used to generate a boundary element


formulation of the plate buckling problem which would yield the
critical buckling load and the critical buckling configuration.
Explicit expressions of the functions of the fundamental solutions
u 1 and u 2 appearing in the above integrals are given in appendix B.

4.3 Boundary Element Solution

4.3.1 Modelling of Boundary Unknowns

The evaluation of integrals It and Ib requires the modelling of


unknown terms on the boundary, namely w, awl an, Vn (w), and Mn (w).
Two out of these four variables will be specified at every node
according to the boundary conditions to the problem. The other two
will be system unknowns. The term J b , on the other hand, requires
an approximation at corners of the twisting moment Mns (w) given by:

M (w) = - D(1-\) _a [~l


ns as an

The boundary is again discretized into a set of Nbe boundary


elements. Discontinuous, constant and linear, elements will be
tested and used. The reason for such a choice is that the
discontinuity of the normal slope at corners can thus be modelled
automatically. Although other methods have been proposed to model
both geometrical and physical discontinuities, this approach is one
of the most efficient because the discontinuity is inherent to it.

The constant element is implemented by simply taking one system


unknown at the mid-point of the mesh nodes. Problem variables would
then be assumed to be uniform over the element and equal to their
value at the middle node. The final system of equations would
60

contain 2Nbe unknowns. The linear discontinuous element, on the


other hand, has been already described in section 3.3; 4Nbe
unknowns are generated by this model.

Mns(W) on either sides of a corner can be approximated in the


following way. Direct finite difference is used to approximate
a2w/anas in terms of the values of the normal slope over two
consecutive boundary elements shown in figure 4.1:

a2w
anas · ~[[: L- [:: II 1 from two constant elements

L- (:: t 1
a2w
(anas II · ~[[:
from two linear elements

[ L· ~[[: L- (:: l31


a2w
anas

1--50--1 )-sl-i

L.
0.. Consto.nt EleMents b. Llneo.r EleMents

/ I eletWn'ts l'Iesh points

tit I elePlents nodes

X I I'llpoInt ~ two consecuttve nodes

Figure 4.1 Approximation of the Transverse Moment Mns(W) at the


corners.
61

All terms appearing in the above equations are defined in figure


4.1. The value of Mns(w) at the corner point can then be, either
uniformly or linearly, extrapolated depending on whether 3w/3n is
approximated with constant or linear variations, respectively.

4.3.2 Domain Deflection Models

The evaluation of integral I dw requires the approximation of the


domain deflections w. Thus, the domain is discretized into a set of
Nde triangular cells and the deflection w is defined as follows:

(4.9)

where w is the deflection at any point of the cell, wi the


deflection at the ith node, f l, the interpolation functions and Nne
the number of nodes per element. It is more convenient to work with
a local triangular coordinate system shown in figure 4.2.

3
~ 1=0')
s2=0.
s3=1.

( SI=I.).l
&
s2=0.
s3=0. 1=0.)
2 s2=1.
s3=0.
Figure 4.2 Triangular Coordinates System.
62
The transformation from the global coordinates system (x,y) to the
local one (sl,s2,s3) for a cell defined by its three vertices
(x 1 'Y1)' (x 2 'Y2) and (x 3 'Y3) is performed using the following
relations (Zienkiewicz, 1977):

sl (b 1x + a 1y + 2A1 )12A

s2 (b 2x + a 2y + 2A 2 )/2A (4.10)

s3 (b 3x + a 3y + 2A 3 )12A

(sl + s2 + s3 = 1)
where

a1 x3 - x 2 b1 Y2 - Y3
a2 xl - x3 b2 Y3 - Y1
a3 x 2 - xl b3 Y1 - Y2
A1 (x 2Y3 - x3Y2)/2

A2 (x 3Y1 - x1Y3)12

A3 (x 1Y2 - x 2Y1)/2
A Al + A2 + A3 = Area of Triangle
while the inverse transformation is given by:

x = sl x 1 + s2 x 2 +s3 x 3
(4.11)
Y sl Y1 + s2Y2 +s3Y3

Conventional continuous cells as well as discontinuous ones in


which all nodes are taken inside the cell will be tried.

The terminology that will be used subsequently is made somewhat


complicated by the fact that both domain and boundary
discretizations are required. The various types of nodes, shown in
figure 4.3, will be designated in the following way:

1. The word 'mesh nodes', already used in a similar context for


boundary elements in section 3.3, is extended to cover both
domain and boundary elements: these are the nodes that define
63

the geometry of the element namely the three vertices of a


domain cell and the two extreme points of boundary elements.

2. The word 'unknown nodes' again refers, for both domain and
boundary elements, to the nodes at which the system unknowns are
taken. They can either coincide with or be distinct from the
mesh nodes depending on whether the element is continuous or
discontinuous, respectively.

3. 'Cell boundary nodes' indicates the mesh nodes of the domain


cells that are located on the boundary.

I X X CE
I daNIn
UMrcar dllicllu'I:II lJIIUII
c:eU.
I
I

'* '*
LDE LDE LDE
I Mesh nodes I Mesh nodes

X I Unknown nodes X , Unknown nodes

D I Boundary cell nodes CE I Consto.nt elel'lent

S I Boundary cello.nd unknown LDE I UneClr discontInUOUS elel'lent


nodes colnddlng

- - - - , Boundory - I BoundClr"y

0.. DOMo.ln nodes b. Boundo.ry nodes

Figure 4.3 Discretization Nodes.


64

4.3.2.1 Continuous Cells

Linear cells defined by three unknown nodes, quadratic by six


unknown nodes and cubic by ten, are shown in figure 4.4.

0.. lIneo.r 10. quo.dro.tic C. cubic

Figure 4.4 Linear, Quadratic and Cubic Continuous Cells.

The interpolation functions for the linear cell are the triangular
coordinates themselves~

(4.12)

Those for the quadratic cell are given by:

11=s1(2s 1-1)

12=s2( 2s 2- 1 )

13.. s 3 (2s 3-1)


(4.13)
14..4s 1s 2

15=4s 2s 3
16=4s 3s 1
65

Finally, interpolation functions of the 10 nodes cubic cell are


given by:

1
11- - (3s 1-1)(3s 2-1)sl
2
9
12= - sls 2(3s 1-1)
2

9
13= - sl s 2(3s 2-1)
2

1
17= - ( 3s 3-1 )(3s 1-1)s3
2

9
IS" - s3 s 1 ( 3s 3- 1 )
2

Although unknowns in quadratic and cubic cells are defined using


non-linear interpolation functions, the geometry of the element is
still defined using the linear model, that is only the three
vertices are needed to define the cell. In finite element
terminology, this is known as a non-isoparametric element. This
66

also applies to the discontinuous linear and quadratic cells that


will be described next.

4.3.2.2 Discontinuous Cells

If the unknown nodes are taken inside the cell, discontinuity of


the modelled deflection occurs at the interface of two adjacent
elements and the cell is thus called discontinuous. Two such cell
models, linear and quadratic, are shown in figure 4.5.

0.. Llnea.r cell b. Quo.dro. tic cell

Figure 4.5 Linear and Quadratic Discontinuous Cells.

The unknown nodes in the linear cell are taken on the medians of
the triangle at the same distance sO' in triangular coordinates,
from their respective vertices. The quadratic cell is constructed
in a similar way with the addition of the three midpoints of the
segments joining the unknown nodes of the linear model. Hence, in
both cases, only one parameter So defines the distribution of the
nodes within the cell.

The interpolation functions of the linear and the quadratic cells


can be determined numerically by expressing them as linear and
67

quadratic polynomial functions of sl and s2' respectively:

I.1 for linear cells

for quadratic cells

The parameters Kij can be determined by forcing each of the


functions Ii to have the value of 1 at node i of the cell and 0 at
all other nodes. Systems of 3 equations with 3 unknown parameters
Ki1' Ki2' Ki3 for the linear cell and 6 equations with 6 unknowns
for the quadratic one are obtained for each interpolation function
ti which can thus be numerically calculated for any position So of
the internal nodes. The coefficients K.. need be calculated only
1J
once throughout the solution.

4.3.3 Free Boundary

On a free boundary both deflection wand its slope in the direction


normal to the boundary awlan are non-zero. If these two variables
are taken as unknowns on the boundary, two problems arise. First,
non-conformity would occur between values of w given by the
boundary element model and those given by the adjacent domain cell
model. The second problem is associated with the singularities of
order 1/r2 that exist in the integral containing Vn (u 2 ), shown in
appendix B, when the source point P is placed on nodes over the
free boundary. Such singularities can only be solved using special
numerical integration schemes.

Both of these problems can be avoided by evaluating wand awlan on


the free boundary using the adjacent domain cell model only.
Boundary Gaussian stations are treated as domain cell points and
equation (4.9) is used to calculate w, while awlan is obtained by
differentiating the same equation:
68

Nne
w
=L
i=l
at i
an
w.1 (4.14)

No additional unknowns would then be generated over the free


boundary element itself. Consequently, the source point need not be
placed on free boundary nodes and the singularity problem is thus
avoided. Free boundary edges will be discretized into boundary
elements only for the purpose of performing the Gaussian
integration.

Another advantage of this treatment is the reduction of the overall


number of boundary unknowns. However, accurate modelling of aw/an
may require domain mesh refinement to make up for the loss of
accuracy due to the differentiation of the interpolation functions
in equation 4.14.

4.3.4 Eigenvalue Problem

A Gaussian scheme of integration is adopted to evaluate boundary


integrals Ib and It using Ngb Gaussian stations and domain integral
I dw using Ngd Gaussian stations. The following discretized forms of
integrals I b , It and I dw are obtained:

+
69

(4.15)

au au)
[T - - + T - -
n an s as j

(4.16)

(4.17)

where notation (E)j for any expression E indicates the value of E


at Gaussian station j; (wg)j is the Gaussian weight of the jth
Gaussian station. In integrals It and I b , Nne indicates the number
of boundary nodes per element except for the free boundary case
where wand 3w/an are calculated using the domain model. In
integral I dw ' Nne denotes the number of nodes per domain cell.
Singularities occur when the source point is within the integration
element. Such integrals are evaluated analytically.

If the source point is placed on the boundary nodes, two equations


with u1 and u2 can be set at each node. A system of Neb equations
is obtained where Neb-4(Nbe-Nfe) in the case of linear
70

discontinuous boundary elements and Neb=2(Nbe-Nfe) in the case of


constant elements where Nfe is the number of elements on a free
boundary. The system of equations can be written as follows:

(4.1S)

where {bw} is a vector containing the Ne b boundary unknowns among


3w/3n, Mn(w) or Vn(w) and {w} is a vector containing domain
deflections at the Ndn domain nodes. The (NebxNeb) array [BS]
results from integral Ib and the term cZ(w) when it does not
vanish. The (NebxNeb ) array [B 6 ] results from integral It. However,
terms in Ib and It containing variables wand 3w/3n over a free
boundary contribute to the (NebxNdn ) matrices [01] and [02] instead
of [B 1 ] and [B 2 ] as these two variables are expressed in terms of
domain unknowns. [01] contains only that contribution and is
identically zero when the plate does not have any free boundaries.
[02] contains the contributions of integrals It and I dw .

Equations (4.1S) represent a system of Neb equations with Neb+Ndn


unknowns in addition to the unknown parameter A. This system,
clearly, cannot be solved without Ndn additional equations. These
equations can be obtained by placing the source point P on the Ndn
domain nodes and writing equation (4.S) with u=u 1 only. Another
system of Ndn equations is thus obtained:

(4.19)

Matrices IB 7 ], [BS ]' [03] and [04] are obtained in exactly the same
way described for matrices [B S ]' [B 6 ], [01] and [02] respectively,
except that term cZ(w)=w p contributes to matrix [03] and not [B 7 ]
because the source point is in the domain and not on the boundary.

Equations (4.1S) and (4.19) can be transformed into an eigenvalue


problem with Neb+Ndn equations with A as the unknown parameter:
71

[C]{U} clUJ (4.20)

where

c = 1IA

{b }}
{U} = { w
{w}
-1
[C) = [A] [B)

[A]

[B)

The system of equations (4.20) represents a classical form of an


eigenvalue problem. An iterative method of solution which assumes a
starting vector {Uo} is used (Spencer et aI, 1980) to obtain the
largest lowest load factor Ab .
eigenvalue c i.e. the This method
uses the following recurrence relations:

1
{Un } = -{Rn }
rn

where rn is the numerically largest element of {Rn }. It can be


shown that as n tends to infinity rn converges to the largest
eigenvalue c. Although the convergence of the method may
occasionally depend on the choice of {UO}' results showed it to be
a reliable and economical solution. Thus we have obtained the
critical factor ~ and the buckling mode represented by vector {U}.
72
4.4 Numerical Implementation

A computer program, PLABEH, has been developed to test the validity


of the formulation, the convergence of the results, the efficiency
of the various boundary and domain elements described above and the
accuracy of the solution.

Although the data input has broadly the same characteristics as the
program PLASTR described in chapter 3, additional subroutines had
to be written to accommodate and perform the domain discretization.
A fully automatic scheme for the discretization into triangular
cells of polygonal and circular plates has been developed. The user
controls the discretization by specifying the coordinates of the
cells boundary nodes; these nodes are then used to generate the
domain cells. Irregular curvilinear shapes can be dealt with by
simply considering adjacent cell boundary nodes as the ends of one
polygon side. The user has the option of using the automatic
boundary discretization scheme to generate these nodes; in fact,
the boundary mesh itself can be used as the generating mesh. Since
the consistency of domain and boundary meshes is recommended
anyway, the mesh preparation effort required by the user to
discretize both the boundary and the domain can actually be reduced
to the supply of the number of boundary elements per polygon side
or, in the case of circular plates, the total number of boundary
nodes, as in any purely boundary element solution. If mesh
refinement is required over specific areas, the user's own
distribution of boundary nodes can be supplied that would be
translated by the program into refined meshes in both boundary and
domain. Naturally, other alternatives are available such as
automatic boundary discretization with non-automatic domain one or
vice verca etc •.. The user can also specify symmetry about one or
more axes; pairs of unknown nodes, both on boundary and in domain,
that are symmetric with respect to these axes are identified and
the overall number of unknowns accordingly reduced.
73
Once the meshes have been prepared, the relative position of
elements over free boundaries, if any, and domain cells adjacent to
the boundary is established and the membrane stress distribution
read from an output file of program PLASTR. The program then
proceeds to assemble the coefficient matrices by performing the
integrations. The source point is identified with all boundary and
domain unknown nodes and the corresponding rows of the system
matrices are calculated in the fashion described in the previous
section. Finally, the eigenvalue problem matrix [C] is evaluated
and the algebraic system of equations solved. The calculated
critical load ~ and buckling shape {w} are then written into
output files.

A flow chart of PLABEM is shown in figure 4.6.

Read Data
I
Discretize domain and
boundary and locate
relative position of
boundary and domain
elements
I
Assemble boundary and
domain source equati-
ons
I
Solve system of equ-
ations
I
Vrite results

Figure 4.6 Program PLABEM flowchart.


74

4.5 Results

Results obtained from program PLABEM were used to assess the


validity and efficiency of the proposed solution. Results for the
constant boundary elements were obtained from a computer program
using the same formulation and developed by Syngellakis and Kang
(1986).Optimum positions of unknown nodes in discontinuous domain
cells and boundary elements are first established through a
systematic parametric study involving a number of plates having
simple geometries, support and loading conditions. Next the
performance of the various continuous and discontinuous, domain and
boundary elements described in section 4.3, is assessed through
comparison with closed form solutions. Finally, results for rather
more complex situations such as plates with free boundaries,
non-rectangular shapes and various loading conditions were obtained
and compared to results from other numerical and analytical
methods. Plate shapes, loading and support condition shown in
figure 4.7 will be used as examples in sections 4.5.1 to 4.5.3. The
critical load factor values quoted in figures 4.7 and 4.11 have
been obtained from Timoshenko and Gere (1961), Bulson (1970) and,
for triangular plates, Tan (1984).

The results are given in tables where the following notation has
been adopted:

NBE: number of boundary elements


NBU: number of boundary unknowns
NDE: number of domain elements
NOU: number of domain unknowns
NTU: total number of unknowns
C constant element
LC linear continuous element
LD linear discontinuous element
QC quadratic continuous element
QD quadratic discontinuous element
CC cubic continuous element
75

D
SIMple
Supports
0
Clo.Mpea
Supports
D
Free
Supports

-- D~ -0-
--- --- -0-
-- 1111
-- ---
1111

ffff
-
ffff
~=2. ~=5.3 ~=7.69
EXQl'lple 1 EXQl'lple 2 EXQl'lple 3

., 1D~ ;;>'

~ =12.28 ~=9.34
EXQl'lple 4 EXQl'lple 5

~~/
ffff

EXQl'lple 6 EXQl'lple 7

Figure 4.7 Plate Examples.


76

4.5.1 Optimum Nodal Position in Discontinuous Elements and Cells

Four square plates, shown in figure 4.7, subject to various loading


and boundary conditions were analysed to locate the optimum
position of the nodes in the discontinuous elements. The various
boundary and domain discretizations used for that purpose are shown
in figure 4.8. Example 1 represents a square plate simply supported
along its four edges and uniformly compressed in two directions.
Example 2 is identical to example 1 except that the plate's four
edges are fully clamped. Example 3 is a square plate having two
opposite simply supported edges while the other two are clamped;
the plate is uniformly compressed along its simply supported edges
and the resulting fundamental buckling shape is anti-symmetrical in
the direction of loading. Finally, example 4 is again a square
plate having the same support conditions as example 3; it is loaded
by uniform shear along its four edges and by a linearly varying
normal load along its simply supported edges.

Llneo.r Llneo.r Quo.~ro. tIc


DIscontInuous DIscontInuous DIscontInuous
Boundo.ry DOMo.ln Cells DOMo.ln Cells
EleMents
Nbe=16 Nde=24 Nde=18

Figure 4.8 Plate Discretizations.


77

Results for examples 1,2 and 3 were obtained for various values of
So defining the nodal position in the linear and quadratic domain
cells and shown in figure 4.5. So is 1 when three unknown nodes
coincide with the respective cell vertices and 0 if all unknown
nodes coincide with the centroid of the cell. The absolute value of
the error in buckling load factor vs So is plotted in figures 4.9a
and 4.9b for linear and quadratic cells, respectively. 24 domain
cells for the linear and 18 cells for the quadratic model were used
in addition to 16 linear discontinuous boundary elements. Although
various positions of the boundary nodes have been tried, only
results correponding to se=±0.5, where se is a boundary element
local coordinate defined in figure 3.2, are shown here because the
same optimum location was obtained for all values of se'

Figures 4.9a and 4.9b clearly indicate that the optimum position
occur for sO=0.5; this corresponds to local nodal triangular
coordinates of (2/3,1/6,1/6), (1/6,2/3,1/6) and (1/6,1/6,2/3). As
So decreases below that value, the unknown nodes get closer to the
centroid of the cell and to each other, the model becomes closer to
the constant model and the accuracy decreases considerably.
Therefore, this nodal position sO=0.5 is adopted in the subsequent
examples.

A similar attempt was made to identify the optimal nodal position


in the linear discontinuous boundary element. Examples 1,2 and 4
were used with the linear discontinuous domain cell with sO=0.5.
Results are shown in figure 4.9c where, again, the absolute value
of the error in the buckling load factor is plotted against the
nodal position coordinate se'

Although the patterns exhibited by the three curves, corresponding


to the three examples differ, the accuracy does not vary greatly
with the position of the nodes. In fact, the maximum variation of
the error for any example does not exceed 1%. A local minimum does
exist at se=0.5 and this position will therefore be adopted in
subsequent calculations.
78
Cl

IIErr orl
Cl
+ : Example I
X : Examp le 2
"*: Example
3
(Bound ary Nodes. Se-I/Z )

0.64 0.80
SO
a. Linear Domain Cell

+ : Example
Cl
Cl
IIErr orl I

"*.
X: Examp le 2
Examp le 3
Cl
Cl (Bound ary Nodes: 5e=I/Z )
'0

0.64 0.80
50
b. Quadr atic Domain Cell

IIErr orl
Cl
N +: Examp le I

*.
t"l X : Examp le 2
Examp le 4-
(DomaI n Nodes: 50=I/Z )

~~
Cl
'0

Cl
Cl
Cl~---,----~--~----r---~-
-~----~--~--~~--~---

0.00 0.16 0.3Z 0.48 0.64 0.80


5e
c. Linear Boundary Elemen t

Figure 4.9 Optimum positi on of Nodes in Discon tinuou s Eleme


nts.
79

4.5.2 Performance of the Various Interpolation Models

The results shown in tables 4.1a,b,c for examples 1,2 and 5 of


figure 4.7 were obtained using the various interpolation models
discussed previously. Example 5 concerns a square plate simply
supported along its four edges and loaded by uniform shear forces.
The boundary and domain discretizations have been designed so that
the overall number of unknowns for all the tested models is more or
less the same.

Tables 4.1a,b, and c clearly show that, in general, results improve


as higher order elements are introduced. The first two combinations
of constant boundary elements with linearly continuous and constant
domain cells do not perform particularly well. In the first
example, the constant domain cell gives better accuracy than the
linear continuous one; this could be due to the consistency of
domain and boundary modelling in this case. There is a net
improvement in accuracy when the linear discontinuous domain cell
is introduced in model 3, except for example 1 where the
improvement is small compared to that of the other two examples.
When the interpolation models in both domain and boundary are
linear discontinuous, as in model 4, very good results are obtained
with a maximum error of 0.55%. When quadratic and cubic domain
cells are introduced the accuracy remains high but does not improve
further. In fact, unexpectedly high errors of 1.74% for model 5 in
the case of example 2 and -1.23% for model 7 in the case of example
5, were obtained. These tables thus clearly show that the linear
discontinuous model for both domain and boundary is very well
suited for the prediction of critical loads and that there is no
need for the use of higher order elements. The convergence of this
model as the number of domain and boundary unknowns is increased,
its performance in the case of more complicated plate examples and
the efficiency of its performance as compared to finite element
results will all be assessed in the next three sections.
80

Table 4.1a Models performance in the case of a simply supported


square plate uniformly compressed in two directions
(Example 1).
MODEL Boundary Domain NBE NBU NDE NDU ~ %Error
Element Cell
1 C LC 24 48 200 81 2.078 3.9
2 C C 24 48 72 72 2.068 3.4
3 C LD 24 48 24 72 2.049 2.45
4 LD LD 16 64 24 72 1.999 -0.05
5 LD QC 16 64 50 81 2.012 0.6
6 LD QD 16 64 18 108 2.004 0.2
7 LD CC 16 64 24 88 1.995 -0.25

~=2.0

Table 4.1b Models performance in the case of.a clamped square plate
uniformly compressed in two directions (Example 2).
MODEL Boundary Domain NBE NBU NDE NDU %Error
Element Cell ~

1 C LC 24 48 200 81 5.223 -1.45


2 C C 24 48 72 72 5.534 4.42
3 C LD 24 48 24 72 5.344 0.83
4 LD LD 16 64 24 72 5.329 0.55
5 LD QC 16 64 50 81 5.392 1. 74
6 LD QD 16 64 18 108 5.287 -0.23
7 LD CC 16 64 24 88 5.243 -1.08
Ab=5.3

Table 4.1c Models performance in the case of a simply supported


square plate under uniform shear (Example 5).
MODEL Boundary Domain NBE NBU NDE NDU %Error
Element Cell ~

1 C LC 24 48 200 81 9.890 5.89


2 C C 24 48 72 72 10.192 9.12
3 C LD 24 48 24 72 9.326 -0.15
4 LD LD 16 64 24 72 9.306 -0.37
5 LD QC 16 64 50 81 9.346 0.06
6 LD QD 16 64 18 108 9.290 -0.54
7 LD CC 16 64 24 88 9.225 -1.23

~=9.34
81

4.5.3 Convergence of Results from the Linear Discontinuous Hodel

Plate examples 2, 6 and 7 of figure 4.7 are analysed to test the


convergence of the solution as the number of linear dicontinuous
boundary elements and domain cells are separately increased.
Example 6 is a simply supported right-angle isocel triangular plate
of height b, uniformly compressed in the normal direction along its
three edges. The buckling factor is defined as:

Example 7 is clamped circular plate of radius R, uniformly


compressed in the radial direction. Ab is defined as follows:

a R2
cr
D

Figure 4.10a shows the effect of increasing the number of domain


cells, in the case of example 2 and 7. The percentage error of
results is plotted against the number of domain cells. Relatively
good accuracy is obtained with a small number of cells and results
improve further as this number is increased. In the case of the
circular plate of example 7, results improve dramatically as the
number of domain elements is increased from 16 to 24. Again,
convergence can be clearly observed in this case.

In figure 4.10b, the effects of the number of boundary elements on


the results of examples 2 and 6, are examined. In th~ case of
example 2, good accuracy is again achieved with a small number of
boundary elements (NBE=8). Results improve as NBE is increased to
12, but become less accurate when this number is raised to 16. As
for example 6, a steady convergence path is observed.

Vhile a small number of boundary elements can achieve good


accuracy, the number of domain elements appears to be more crucial.
82

o
o

0
-0
([
0-
([
([ + I Example 2 INbe= 16'
w
","0
X I Example 7 INbe- 91
0
CD

0
~

0
~
0
0 8 16 32 40 48
Nde

a. Percent error versus number of domain elements.

o
N
M

+ I Example 2 INde=32 I
X I Example 6 INde-2S1

o
CD
o

4 8 12 16
Nbe

b. Percent error versus number of boundary elements.

Figure 4.10 Convergence tests on linear discontinuous elements.


83

4.5.4 Comparison with exact solutions

A number of plate examples with various shapes, as well as load


and support conditions have been selected to demonstrate the
accuracy and versatility of the solution. These plates are shown
in figure 4.11 and results based on the linear discontinuous
boundary and domain elements are presented in table 4.2.

The first two examples are those of a parallelogram plate clamped


along its four edges and subject to uniform compression in one
direction in the first case, and to uniform shear forces in the
second case. In example 1, the buckling factor Ab is defined as:

where b is shown in figure 4.11 and ~ is the plate inclination


angle, taken as 45 0 in our case. In example 2, Ab is defined as:

The answers obtained by PLABEM are almost exact in both cases.

Examples 3 and 4 are simply supported plates having the shape of a


right-angle triangle. In example 3, the plate is uniformly
compressed along its three edges. In example 4, a uniform shear
load is applied. The buckling factor for these two examples, as
well as example 5, has been defined in section 4.5.2 with
reference to example 6 of figure 4.7. Although fairly accurate
results are produced, the error in example 3 is 1%, somewhat
higher than the 0.1% of example 4. This may due to the modelling
of awlan at the 45 degrees corners of the plate. The normal slope
appears in the jump term of the integral equations where it is
multiplied by the normal load Tn which vanishes when only shearing
loads are applied. The relative inaccuracy may have also been
aggravated by the proximity of the two boundary elements on each
84

side of the corners; the integration over one boundary element


when the source point is on the other becomes nearly singular
and the Gaussian integration less accurate. Example 5 is a simply
supported equilateral triangle uniformly compressed along its
three edges. The relatively high error (1.95%) again suggests that
the acute-angle corners reduce the accuracy of the solution.

Examples. 6 and 7 are circular plates uniformly compressed in the


radial direction; the first is simply supported and the second
clamped. The buckling factor has been defined in section 4.5.2
with reference to example 7 of figure 4.7. An accuracy of 1.02% is
obtained for the first plate with a relatively small number of
elements. In the case of the second plate the error is halved when
the numbers of boundary and domain elements are doubled.

Examples 8, 9 and 10 are square plates having two opposite simply


supported edges and at least one free; uniform compression is
applied along the opposite simply supported edges. A relatively
small number of elements clearly yields very good accuracy. Yhile
the error is somewhat higher in the case of examples 8, results
for examples 9 and 10 are exact. In the case of example 10 having
two free edges, a larger number of elements had to be taken in the
direction of the free edges in order to achieve a good modelling
of the boundary normal slope. Such mesh is shown in figure 4.11.

Examples 11,12 and 13 concern square plates with various support


and loading conditions. Example 11 is a plate simply supported
along three edges and clamped along the other and subjected to
uniform shear along its four edges. Example 12 is a plate simply
supported along two opposite edges and clamped along the other
two. It is uniformly compressed along its simply supported edges.
Plate of example 13 has the same support conditions as that of
example 12. However, it is loaded by uniform shear along its four
edges in addition to a moment couple along the simply supported
ones. Good accuracy is achieved with a maximum error of 1.75%.
85

~
b
-
~ ~ ff

~=10.
t
fff
'I 1~! --~
~= 13.56
.,.
:: 1'1'
f ff f
~=5.
-
J~
~=11.55
ExClJ'lple 1 Exo.rtple 2 ExCll'lple 3 ExC1J'lple -4

~ :::'0'-
ffff
~=39.478
,})J!.I"
\
,./~-\'
't'
~= 14.68
1-
n ,\\HI/

"'1\'
~=4.2 ~=L44
--- --
-- -- -0-
--0-
~=L7
ExC1l'lple 5 ExC1l'lple 6 ExI1I'Iple 7 ExCU'lple 8 ExQ.Mple 9

-.....-D--- -1D~ -0-


- .-- -
• - - - ~10~~
-
.

~=L
~= 1L ~=7.69 ~ =12.28
Ncle=24 ExclI'Iple 11 EXQl'lple 12 ExQl'lple 13
ExQl'lple 10

Figure 4.11 Various Plate Examples.

Table 4.2 Results for various plate examples.


NBE NBU NDE NDU
Example NTU
\ \BEM %Error
1 16 64 32 96 160 10. 10.002 0.00%
2 16 64 18 54 118 13.56 13.555 -0.03%
3 15 60 25 75 135 5. 5.050 1.00%
4 12 48 25 75 123 11.55 11.561 0.10%
5 15 60 36 108 168 39.478 40.247 1.95%
6 8 32 24 72 104 4.2 4.243 1.02%
7 16 64 48 144 208 14.68 14.595 -0.57%
8 16 48 8 24 72 1.44 1.461 1.46%
9 16 48 8 24 72 1.7 1.700 0.00%
10 16 32 24 72 104 1.0 1.000 0.00%
11 16 64 24 72 136 11.0 10.808 -1. 75%
12 16 64 24 72 136 7.69 7.762 0.94%
13 16 64 24 72 136 12.28 12.125 -1.26%
86
4.5.5 Comparison with the Finite Element Method

Results from a finite element program for plate stability using


conforming quadrilateral linear elements and four degrees of
freedom per node (w,aw/ax,aw/ay,a 2w/axay) are compared to the
results obtained by PLABEM. Finite element programs of plate
problems using conforming elements are usually expected to give
highly accurate results.

Oiscretizations were made that resulted in comparable overall


numbers of unknowns NTU for both programs. Both programs have been
run on the same computer system IBM-3090/1S0 mainframe, so that
CPU times could be compared. It must however be noted, concerning
the execution times, that the BEM program contains the automatic
discretization and the symmetry unknowns reduction facilities
which makes the use of the program easier at the expense of some
additional CPU time. In the FEM program, on the other hand,
discretization and symmetry must be given explicitly by the user.

Results of both programs are shown in table 4.3. NBU and NOU are
again the numbers of boundary and domain unknowns after symmetry
is applied. TNU is the final number of unknowns after symmetry,
and boundary restraints in FEM, have been applied. The number
corresponding to the FEM rows under the NOU heading corresponds to
the number of unknowns before the application of the boundary
restraints. The CPU time in seconds is shown in the last column of
the table and is meant to be an approximate measure only, because
factors, different for each program, such as data structures and
preparation, results post-processing, in addition to programming
methods, may affect the CPU time without being characteristic of
the solution itself. However, the CPU time can still be a rough
measure of the solution efficiency.

In the case of the simply supported plate of example 1, both BEM


and FEM give very accurate results.
87
Table 4.3 Comparison with the Finite Element Method.

NBU NOU TNU %Error CPU


\ sec.

----- D ---
llll
BEM 16 6 22 2.000 0.00 0.23

ffff
- FEM 36 36 2.001 0.05 0.25

\=2.0

--- --
- D --
BEM 24 12 36 1.461 1.46 0.9

FEM 48 28 1.435 -0.3 0.44

\=1.44

--
-- D~
BEM 24 12 36 1.700 0.00 2.62
FEM 112 72 1.699 -0.06 3.03

\=1.7

-- ---
-- 0
BEM 8 18 26 1.000 0.00 0.9l

- FEH 64 48 0.971 -2.9 1.12

\=1.0
88
The CPU time per system unknown appears to be higher in BEM. The
next example is that of a plate containing one free edge. The FEM
gives much better accuracy in this case. However, in the last two
examples, and the last one in particular, the BEM appears to be
much superior to the FEM in both accuracy and CPU times.

4.6 Conclusions

The boundary element formulation presented in this chapter has


been proven to be an efficient solution of the plate stability
problem, despite the required domain discretization. Accurate
results were obtained with a relatively small number of unknowns.
The method was also shown to be versatile when results were
derived for plates with various shapes, boundary and loading
conditions. Its performance in terms of accuracy and CPU times was
compared to that of a finite element program and the boundary
element solution seems to be very efficient, especially in the
case of plates containing free edges.
CHAPTER FIVE

DUAL RECIPROCITY

5.1 Introduction

The existence of domain integrals, usually due to body forces or


loading, in the integral equations of plate as well as non-linear,
time dependent and some other types of problems reduces the
efficiency of the boundary element method. Most essentially, the
boundary mesh simplicity, a characteristic of the method that is
crucial in computer aided design, is lost to the cumbersome domain
discretization required by the solution. That is why considerable
effort in BEM research has been concentrated on the transformation
of domain integrals into boundary ones.

The dual-reciprocity method has been first introduced by Brebbia


and Nardini (1982) in the context of elasto-dynamic problems. The
method has been extended later on to other fields of BEM
applications. A somewhat more general dual reciprocity technique
has been presented by Tang (1987) which could be applied to a wide
range of problems. The method uses a Fourier series of a known
function b(x) to transform into boundary integrals domain integrals
of the form:

J b(x)u dQ
Q

u being the fundamental solution of the problem. Kamiya and Sawaki


(1988) use the dual reciprocity technique and a polynomial
approximation of the deflection w to transform the domain
integral:
90

J w u dQ
Q

arising in the problem of a plate on elastic foundation. All these


attempts have been successful. ·The transformed domain integrals
contained either known functions f(x,y) or just the unknown
deflection w. The domain integral of the plate buckling problem,
however, is generated by the inplane stresses expression which
contains second order derivatives of the deflection i.e.
curvatures. The domain integral Idw of equation 4.7 containing only
deflection terms cannot be treated by dual reciprocity because of
the complexity of the weighting function appearing in that
integral.

In this chapter, a dual reciprocity multi-stage technique based on


a Fourier series analysis is applied to the stability problem of
plates. An overall description of this technique will first be
given, then each stage will be described in detail.

5.2 Outline of the Method

The domain integral appearing in equation (4.5):

Id J uL(w)dQ
Q

involves domain inplane stresses and curvatures terms. Ve will


assume that these stresses can be evaluated at any domain or
boundary point of the plate. A boundary element solution of the
plane stress problem which can be used for that purpose has been
presented in chapter 3.

Using the Rayleigh-Green identity, it will be shown that an


integral of the form:
91

10=[ (UV4p(x,y» dQ
Q

where u satisfies equation 3.2b, can be transformed into a boundary


integral. Hence, if the function L(w) is expressed in the form:

Nd
4
L(w)= [(Ud)i V Pi(x,y)
i=1
where (Ud)i are Nd unknown factors, then:

Nd
1d= [(Ud)i [ UI14 p(x,y) dQ
i=1 Q

and 1d can be transformed into a sum of boundary integrals.

The domain unknowns (Ud)i appear in the expression for the


deflection w which is assumed to have the form:

Nd
w(x,y)= [ (Ud)iGi (x,y)
i=1

where G.(x,y)
I
are suitably chosen deflection shapes. Then:

Nd
L(w)= [ (Ud)iHi (x,y)
i=1
where
a2G.I
H. (x,y)= N --2- +
I X ax
92
The function Hi(x,y) can be transformed into a Fourier sum of
simple trigonometric functions:

Ntf

Hi (x,y)= L Aj Tj (x,y)
j=l

where N tf is the number of Fourier terms, Tj (x,y) is either a


constant, a sine, a cosine or a single product of these two
functions, and A. are known coefficients.
J

Functions V.(x,y) can be found such that:


J

Hence, L(w) can be written as:


Nd Ntf
L(w)= L (Ud)i L Aj livj(x,y)
i=1 j=l
and

J(UU4V/X,y» dQ]}
Q

It follows from this description that the technique is applied in 3


major stages:

i. Approximation of the deflection.


ii. Fourier approximation of the function L(w).
iii. Transformation of the domain integral into a boundary
one using the Rayleigh-Green identity.

The various stages described above will now be presented in more


detail.
93

5.3 The Discrete Points Fourier Analysis

A quick review of the Fourier series theory is presented in this


section. A full description can be found in various mathematics and
engineering references (Scheid 1968).

5.3.1 The One-Dimensional Fourier Series

Given a periodic function f(x) of period 2a in the interval


[-a,a], f(x) can be represented by an infinite trigonometric
series:
+'"
inx inx
f(x)= KO + L(K li cos - + K2i sin - ) (5.1)
a a
i=l
where
+a

KO ---
2a
1
Jf(x)dx
-a
+a

~J
inx
Kli= f(x)cos d x
- (5.2)
a
-a
+a

~J
inx
K2i= f(x)sin d x
-
a
-a

The following two theorems describe the behaviour of the Fourier


functions and the conditions for convergence.

Theorem 1:

If a bounded periodic function f(x) with period 2a has at most a


finite number of discontinuous points in anyone period, then its
Fourier function F(x) converges to f(x) at all continuous points x,
94

and to the average of right-hand and left-hand values of f(x) at


all discontinuous points x. The conditions enunciated in this
theorem are known as the Dirichlet conditions.

Theorem 2:

If a function f(x) and its first (m_1)th derivatives satisfy the


Dirichlet conditions and are everywhere continuous, then
coefficients k1i and k2i tend to zero at least as rapidly as
(constant/i m+ 1 ). If the mth derivative is discontinuous at some
point(s) then k1i and k2i will tend to zero no faster than
(constant/i m+ 1 ).

The first theorem actually states the conditions for convergence of


a Fourier series. In our case, the function in question is the
deflection and its first and second derivatives which represent
slopes and curvatures. Therefore, simultaneous continuity of these
functions will have to be satisfied at all points of the plate in
any assumed deflection model. The second theorem describes the
speed of convergence and generally states that the smoother the
function the faster the convergence.

5.3.2 The Two-Dimensional Fourier Series

Based on the results for the one-dimensional Fourier analysis, one


can easily extend the concept of Fourier series to the case of
functions of two variables.

Given a function f(x,y) defined over the domain Q and satisfying


the Dirichlet conditions, Q can be extended to the rectangular
domain 2f or [-a,a]x[-b,b] shown in figure 5.1 and f(x,y) can be
assumed to be periodic with periods 2a and 2b in the x and y
directions respectively.
95

a ~

S2
T
T2b
b
I
~
S2
f

I
-1-

2a

Figure 5.1 Extended domain for Fourier analysis.

f(x,y) can therefore be approximated by an NflxN f2 Fourier series:

4 Nfl 4 Nfl Nf2


f(x,y)'" KO +[ [ Kl fl
n n
+[ [ [Klnm fln (5.3)
1=1 n=l 1=1 n=l m=l
where
nnx
f1 cos
n
a
nnx
f2 sin
n
a
nny
f3 cos
n
b
nny
f4 sin
n
b
nnx mny
f1 = cos - - cos
nm
a b
96

2 nnx mlty
f nm = cos - - sin
a b

3 nnx mlty
f nm = sin - - cos
a b

4 nnx mlty
f nm = sin - - sin
a b

and
+b +a
1
KO
4ab JJ
-b -a
f(x,y) dxdy

+b +a
1
1
JJ
Kl f(x,y) f n dxdy (5.4)
n
2ab
-b -a

+b +a
1
Kl =
nm
2ab JJ
-b -a
f(x,y) flnm dxdy

5.3.3 The Discrete Points Two-Dimensional Fourier Analysis

So far the described Fourier analysis required the evaluation of


the K's factors by continuous integration according to equations
5.4. However, computer applications of the analysis will have to
deal with a discrete set of points rather than a continuous field.
This is due to the fact that non uniform loading will result in a
complex inplane stress distribution that can only be obtained
numerically, through a BEM or FEM solution, at any chosen point. In
such cases, L(w) cannot be expressed as a continuous function.
Moreover, the discrete points approach is easier to program because
of the variety of boundary conditions to be encountered even in
simple cases of inplane loading.
97

2a

T
x-x-x-x-x-x-x-x-x
I I
x x x x x x x x x
I I
x x x x x x x x x
I I
x x x x x x x x x 2b
I I I
x x x x x x x x x
I I
x S2 x x x x x x x x
I f I ~
x-x-x-x-x-x-x-x-x

Figure 5.2 Nodes distribution in extended Fourier domain.

The function f(x,y) is actually given at discrete set of Npx x Npy


equidistant points equally spread over the interval [-a,a)x[-b,b),
as shown in figure 5.2. To ensure the periodicity of f(x,y),
f(-a,y) and f(a,y) should be equal, as well as f(x,-b) and f(x,b)j
the left-hand side column and bottom row of values will not be used
and their values will be implicitly taken as that of the right-hand
side column and top row, respectively. The coefficients KO' Kn and
Knm of the Nf1 xN f2 Fourier series (Nf1=Npx/2,Nf2=Npy/2) can be
approximated in the following manner:

Nfl Nf2
1
KO= C f ( Xn ,y)
m (S.Sa)
4NflN f2 C
n=-N fl +1 m=-N f2 +1

Nfl Nf2
1
Kl= Cf(Xn'Ym) f!(xn'Ym) (S.Sb)
n
2NflN f2 C
n=-N fl +1 m=-N f2 +1
(n=l, •.. ,N -1)
(1=1, •.. ,4
p
98

(5.5c)

(5.5d)

(n=l, ... ,Nfl-1)


(m=1, ... ,N f2 -1)
(1=1, •.• ,4)

(5.5e)

(5.5f)

(5.5g)

1
Kn=O. when n=N f1 and 1=2 or 1=4
2 344
KnNf2=KNf1m=KNf1m=KnNf2=O. for (n=1, ... ,N f1 ) and (m=1, ... ,N f2 )
(EQUATIONS 5.5)

The presentation of the Fourier analysis is thus completed. The


Fourier transformation will actually be used twice: first, for the
generation of the nodal deflection model; secondly for the
transformation of the loading function L(w).
99
5.4 The Deflection Models

5.4.1 The Trigonometric Deflection Model

Given a rectangular plate any edge of which can be simply supported


(SS), clamped (CL) or free (FR), the deflection function w(x,y) of
the plate can be assumed to be:

Nd
w(x,y)= [ (U i Vi (x,y» (5.6)

i=l
where

V.(x,y)=X
1 m
(x)Y n (y) (i=(m-1)N + n).
Ui :unknown coefficients.
Xm(x),Yn (y):trigonometric function satisfying the boundary
conditions of the plate in the x and y directions,
respectively.
N :total number of Yn functions.

Previous work on buckling and vibration of plates by energy methods


adopts a similar approximation of the deflection, essentially using
functions satisfying the geometric boundary conditions of the plate
(Richards 1977, Dym & Shames 1973, Jaeger 1964). Functions Xi(x)
and Yi(y) of this model should satisfy the geometric and kinematic
boundary conditions and should be differentiable at least twice
over the domain of the plate in order to obtain an accurate
representation of the curvatures. Trigonometric functions were
chosen because in simple cases of uniform loading and boundary
conditions such as SS and CL, exact reproduction of the loading
function is achieved by the Fourier analysis. These functions are
usually smooth enough for a rapid convergence to occur. In
addition, the trigonometric functions can be more easily combined
so that boundary conditions of any support case are satisfied. In
the case of rectangular plates, trigonometric functions Xi(x) or
100

Yi(y) satisfying the ss (w=0,a 2w/an 2=0) and CL (w=O,aw/an=O)


boundary conditions can be easily found. However, the FR boundary
condition involves the Poisson ratio and thus prevents a simple
transformation of the 'two-dimensional' deflection function into
the product of two 'one-dimensional' functions. To overcome this
problem, the terms containing the Poisson ratio have been
disregarded from the corresponding equations and thus only second
and third order derivatives of the deflection were required to
vanish at free boundaries. In other words, the plate has been
treated as a combination of two crossing beams as far as the
boundary conditions of the deflection function are concerned. This
assumption is not expected to affect the accuracy of the results
dramatically. In fact, other approximate methods for plate problems
have proceeded in a comparable way. The Grillage Analogy method,
for instance, used for bridge deck slabs, assumes the whole plate
to be a series of crossing beams (Hambly 1976).

If only SS, CL and FR edges are considered, 6 possible combinations


arise in the x or y direction. Functions corresponding to these 6
combinations are given explicitly in appendix C and sketches
showing their variation along a plate width or breadth are given in
figures S.3a to S.3f. These functions have been suggested and used
here. Other functions satisfying the boundary conditions can of
course be found and may prove even more effective depending on
plate geometry, support and loading.

It is more convenient to condense the equations using matrix


notation at this stage. Hence, the following matrices are defined:

X1(x)Y1(y)
X2 (x)Y 2 (y)
[
(Ttl= .

XNd(x)YNd(y)
so that w becomes:

(5.7)
101

W
0
(J)

0
0
0
0.00 0.20 0.40 0.60 0.80 1.00
5
55-55: M= I

0
'0

55-55: M=2

Figure S.3a SS-SS support conditions.

w
o
(J)

o
o
o4-~~---r--~---T--~--~~--~--~--~~~
0.00 0.20 0.40 0.60 0.80 1.00
5
CL-CL: M= I

w
o
(J)

~ 00 1.00
o, 5
CL-CL: M=2

Figure S.3b CL-CL support conditions.


102

W
CJ
CJ
~

CJ
CJ
CJ
0.00 0.20 0.40 0.60 0.80 1.00
5
FR-FR: M=O

W
CJ
CD
CJ

gJ
CJ
,
FR-FR: M=3
Figure 5.3c FR-FR support conditions.

w
CJ
CD
CJ

CJ
CJ
CJ
0.00 0.20 0.40 0.60 0.80 1.00
5
55-CL: M= 1

W
CJ
~

CJ

1.00
,
CJ 5
55-CL: M=Z

Figure 5.3d SS-CL support conditions.


103

W
0
CD
0

0
0
0
0.00 0.20 0.40 0.60 0.80 1.00
5
55-FR: M= 1

W
0
'0
0

0
~ 1.00
0
, 5
55-FR: M=3
Figure 5.3e SS-FR support conditions.

w
...,.
0

0
0
.
0
0.00 0.20 0.40 0.60 0.80 1.00
5
CL-FR: M= 1
W
0
CD
0

0
0
0
0.00 0.20 0.40 0.60 0.80 1.00
5
CL-FR: M=3
Figure 5.3f CL-FR support conditions.
104

Yhen using the trigonometric model, the accuracy can be improved by


selecting higher order trigonometric functions, that is by
increasing m and n. In addition, a selective choice of deflection
modes can be made depending on symmetry, anti-symmetry etc •..

However, this model cannot be easily extended to non-rectangular


plates because the choice of the deflection functions satisfying
the boundary conditions would become difficult if possible at all.
In addition, since there are no domain nodes but only domain
unknown coefficients, the location of the source points for domain
equations is rather arbitrary. An alternative deflection model
which eliminates these disadvantages is the nodal deflection model.

5.4.2 The Nodal Deflection Model

Let a and b be the largest dimensions of the plate in the x and y


directions respectively. The domain Q of the plate is extended to a
rectangular domain Qf or [-a,a]x[-b,b] as shown in figure 5.4. Let
N1xN 2 equidistant points be chosen in the domain Qf and the
deflection at all points outside the actual plate domain and on the
boundaries, except free ones, are set to zero; deflection unknowns
are taken at Nd discrete set of points where deflection w does not
vanish. Let the deflection at these points be Y1 ' •.. 'YNd and define
{U d } as:

The Fourier expansion of function w can be constructed using the


discrete points analysis described in section 5.3.3. First, the
Fourier trigonometric functions 1,f n
and f nm are assembled in a
vector matrix. Hence, defining vector {T f }:
105

1
cos X
sin X
cos Y
sin Y
cos X cos Y
cos X sin Y
sin X cos Y
sin X sin Y
cos X cos 2Y
cos X sin 2Y
sin X cos 2Y
sin X sin 2Y

cos 2X
{Tfl= sin 2X

cos N1X
sin N1X
cos N1Y
sin N1Y
cos N1X cos Y
cos N1X sin Y
sin N1X cos Y
sin N1X sin Y

cos N1X cos N2Y


cos N1X sin N2Y
sin N1X cos N2Y
sin N1X sin N2Y
me ny
where X=-- and Y=--
a b

It is clear that -one constant function, 4Nl single trigonometric


functions and 4N1N2 product trigonometric functions- contribute to
the vector {Tfl. Thus, its dimension is Ntf=1+4Nl+4N1N2' Let {Kl be
the vector containing the Fourier coefficients of the respective
Fourier functions of vector {Tfl such that:

(5.8)
106

Equations 5.5 show that the coefficients {K} are directly related
to the values {Ud } of the deflection at the discrete points. In
fact, each coefficient is a linear combination of the nodal
deflections. Hence, a matrix [Cw] of dimensions (NtfxNd ) can be
constructed which relates vectors {K} and {Ud }:

Matrix [Cw ] is easily calculated from equations 5.5.

Replacing {K} in equation 5.3:

(5.9)

1---------2a---------l

i-x-x-x-X/'-Q '-'-j
x x x x

I
x x x x .
I
2b
I
x x x x x~.-.-.-!I
I
x x x x x x x x
I
x
Q I
I
x x
f
x x x x x x
I
x
I
I
x--x--x--x--x--x--x--x--x
I
x:points of zero deflection
.:points of unknown deflection

Figure 5.4 Nodes distribution in extended Fourier domain for the


nodal deflection model.
107

The elements of matrix [Cw) are known coefficients and vector {Tfl
depends solely on the coordinates x,y of the point at which w is
calculated. Hence, we have actually defined the deflection function
in terms of the unknown deflections at some specified nodes in the
domain of the plate.

It is convenient for the subsequent analysis to adopt a common


matrix notation for both trigonometric and nodal models. If we
define vector matrix {Fw l of dimension Ndx1 as equal to:

for the trigonometric model discussed in section


5.4.1
and {Fwl=[Cw)T{Tfl for the nodal model, then in both cases:

(5.10)

where {Udl contains either coefficients of the trigonometric modes


or the deflections at the nodes.

The nodal deflection model has the advantage that it can be used
for plates of any shape. The domain source point location is the
predetermined one of the actual nodes, unlike the trigonometric
model where this is somewhat arbitrarily chosen. Furthermore, even
though the boundary conditions cannot be fully satisfied by the
domain deflection function, zero-deflection can be forced on
boundaries in the SS and CL cases when plate shape is rectangular,
by taking zero-deflection nodes over these boundaries.

However, a disadvantage of this model is that it needs a larger


number of trigonometric functions to represent adequately the
deflection compared to the previous one. In simple cases of
rectangular plates with SS and CL boundary conditions and uniform
loading, the trigonometric deflection model is expected to be more
efficient with regard to computer speed and memory.
108

5.5 Transformation of L(w)

Having defined the deflection pattern, the curvatures can now be


evaluated. If, in addition, the plane stress problem is solved and
the inplane stress distribution of the plate Nxx(x,y), Nyy(x,y) and
Nxy(x,y) are determined, then the function L(w) can be found.

Using the common notation defined in equation 5.10 for both


deflection models, we can write:

where

If we set:

then

Vector matrix {F l } actually contains Nd functions of x and y. The


values of these functions can be determined at any point (x,y).
Hence, it is possible to transform each of these functions into its
equivalent Nl1 xN l2 discrete Fourier series using the analysis
described in section 5.3.3. In fact, the Fourier coefficients Kl
can be easily computed from equations 5.5. If we define (Kl)ij as
the coefficient of the ith function (Fl)i corresponding to the jth
Fourier function in the order it appears in the definition of
vector {T f } in section 5.4.2, then:
109

{F l } =[K l HT f }
T
L(w(x,y»={[Kl]{T f }} {Ud }
(5.11)
or T T
L(w(x,y»={T f } [Kl 1 {Ud }
where
[Kl ]: Nd xNfl

Nfl = 1+4Nl1+4Nl1Nl2

Equation (5.11) contains the known coefficients matrix [Kl1,


the domain unknowns vector {Ud} and a vector {T f } of simple
trigonometric functions. This expression of L(w) allows us to
transform the plate buckling domain integral into a boundary one
using the Rayleigh-Green identity.

5.6 Transformation of the Domain Integral

The Green's second identity can be used to transform domain


integrals containing the Poisson operator v2 into boundary ones.
The equivalent reciprocity equation for the biharmonic operator V4
is the Rayleigh-Green identity derived in section 3.2 and rewritten
here for completeness:

JUV4FdQ = cf(F(P») Ja(v2F) au


+ ( u - - - v2F-- +
an an
aF
v2u-- -
an
F
a(v2u)
an
)dr
S2 r
(5.12)
where

{ 0.5 if source point P is on the boundary.


c
1.0 if source point P is in the domain.

{ F(P) for u=u 1 ·


f(F(P»
(aFlan)p for u=u 2 ·

F(P) value of function F at source point P.

(aF/an)p value of aF/an at source point P.


110

Having expressed L(w) in a convenient form, it is now possible to


carry out the integral transformation using identity 5.12.

Using equation (5.11), integral Id =J uL(w)dQ can be written as:


Q

Id= UU{Tf}T dQ ]lKL]T{Ud } (5.13)

Vector matrix {T f } contains functions of one of these forms:

1,fnl or flnm which have been described in section 5.3.2.

Defining the following vector matrix {I u } such that:

(Iu)i= IQ
u Tfi dQ

where Tfi is the ith trigonometric function in the {T f } vector,


equation (5.13) becomes:

(5.14)

Hence, integral Id has been reduced to the sum of integrals Idi of


the following form:

I uflnm dQ
Q
111

Now I d1 , Id2 and Id3 will be transformed into boundary integrals.

Let us introduce functions FO' Fln and Flnm such that:

TJe find:

48

for 1=1,2
(5.15)

for 1=3,4

1
_:--_ _-=-_ _---::-- fl
n4 [(n/a)2 + (m/b)2]2 nm

where xo and Yo are arbitrary constants. Tang (1986) has used this
method to transform integrals containing the Poisson operator vl.
The approach is extended here to biharmonic operator integrals V4

All integrals I d1 , Id2 and Id3 can thus be written in the form:

where F is either FO or F! or F!m.


112

Using the Rayleigh-Green identity (5.12), Idi can be written as:

a( iF) iu)
cf(F(P») + J(u-- -
an
au

an
aF
iF-- + iu-- - F
an
a(

an
)dr
r (5.16)

Idi can therefore be evaluated as a boundary integral and integral


Id has now been expressed as the sum of boundary integrals.

Equation (4.5) rather than (4.8) can now be used to obtain a


boundary element solution of the problem in exactly the same way
described in chapter 4 except that, of course, no domain
discretization is required.

5.7 The Problem of Singular Integrals

Vhen the distance r between the source point and the field point
approaches zero, certain functions of 1/r, 1/r2 and In r, appearing
in integrals Ib and I d , tend to indeterminate values. Consequently,
the standard Gaussian quadrature for numerical integration becomes
less accurate and a larger number of integration points is needed.
It was found that when the shortest distance from the source point
to the field boundary element is less than 0.5 times the length of
this element, the 32 points quadrature gives accurate results.
However, when this ratio is reduced beyond 0.05, the Gaussian
quadrature fails and other techniques, either analytical or
numerical, need to be used instead. This situation actually occurs
when integration is performed over the boundary element containing
the source point P.
113

Vhen integrals Ib are calculated, singularities can be avoided by


analytically evaluating the values of these integrals as mentioned
in chapter 3. The singularities due to integral I d , however, cannot
be always evaluated analytically because they contain the
trigonometric functions of the Fourier series. Two numerical
techniques have thus been used.

The first technique has been presented by Telles (1987) as a


self-adaptive coordinate transformation for the evaluation of
singular and quasi-singular integrals. This method uses a special
purpose variable transformation that would automatically consider a
higher number of integration points near singularities, hence
increasing the accuracy of the Gaussian quadrature.

The second technique was used for the functions of (1/r2) where
singularity is too strong to be treated by the first method. Kutt
(1975) presented a general technique for the evaluation of the
principal values of integrals involving functions of the form
f(x)/x n by either equispaced stations numerical integration, or
Gaussian-type numerical integration which may involve complex
numbers stations. Using these numerical techniques, a sufficient
accuracy has been achieved and hence, the problem of singularities
has been solved.

5.8 Eigenvalue Problem

The system of equations obtained after integration over elements is


performed, is of the form:

Boundary equations [A 1 ]{Ub} ([A 2 ] + ~[A3]) {Ud} (5.17a)


Domain equations [A4] {U b } ([AS] + ~[A6]) {Ud } (5.17b)
114

where

[A l ] corresponds to boundary positions of the source point and


contains coefficients of boundary unknowns arising from
the integration of Ib over simply supported and clamped
edges in addition to jump terms J b at simply supported
edges.

[A Z] corresponds to boundary positions of the source point and


contains coefficients of domain unknowns arising from the
integration of Ib over free edges in addition to jump
terms J b at free edges.

[A 3 ] corresponds to boundary positions of the source point and


contains coefficients of domain unknowns arising from the
integration of I d .

[A 4 ] corresponds to domain positions of the source point and


contains coefficients of boundary unknowns arising from
the integration of Ib over simply supported and clamped
edges in addition to jump terms J b at simply supported
edges.

[AS] corresponds to domain positions of the source point and


contains coefficients of domain unknowns arising from
the integration of Ib over free edges in addition to jump
terms J b at free edges and single values of deflection at
the position of the source point.

[A6 ] corresponds to domain positions of the source point and


contains coefficients of domain unknowns arising from
the integration of I d .
115

Inverting [A1 ] in equation (S.17a):

Replacing {Ub} in equation (S.17b) and rearranging:

([A]-A[B]){Ud}=O. (S.18)

where

Hence, the problem can now be solved using the standard solution
technique of the eingenvalue problem described in section 4.3.4.
Vector {Ud } will either contain coefficients of trigonometric modes
or values of deflection at domain nodes; in both cases the buckling
mode of the plate can be found without using additional boundary
integral equations.

S.9 Numerical Implementation

New computer program versions, called PLADRE, have again been


developed to assess the validity and efficiency of the described
solution. Vhen the trigonometric deflection model is used, the
program can deal with rectangular plates only. Vhen the nodal
deflection model is used, any closed polygonal plate shape can be
analysed. In both cases, any inplane loading can be dealt with
since the discrete values of the membrane stresses at some specific
points of the domain or boundary of the plate are input to the
program.
116

The data structure is rather simple as expected of a BEM program in


which no domain discretization is needed. The required input data
are: the main properties of the plate, namely its shape, thickness,
Young's modulus and Poisson ratio, the number of boundary elements
on each edge of the plate and the deflection modes in the
trigonometric model or the nodal position in the nodal model.

The program can also make automatic use of any x,y or diagonal axis
symmetry or even any combination of these. This would lead to an
overall reduction in system unknowns and equations.

A flow chart summarising the program operations is given in figure


5.5.

I
INPUT I
Input data and transform L(w)1
into a Fourier series

BINE
Set boundary equations and
modify them for symmetry

DINE
Set domain equations and
modify them for symmetry

BUKLE
Solve the system equations

Figure 5.5 Flowchart of Program PLADRE.


117

The INPUT subroutine reads all data, but also prepares the symmetry
vectors, calculates the matrix [C 1 for the nodal deflection model
w
and calculates matrices [Fw 1 and [F11 thus transforming the L(w)
function into a Fourier series.

The BINE subroutine moves the source point over the boundary
elements nodes and derives two equations, corresponding to each
fundamental solution, at each node. Hence, a total of 4Nbe boundary
equations are obtained that is 2 equations per node and 2 nodes per
element. If any symmetry is taken into account, half or even a
quarter of this number of equations is actually formulated. BINE
also reduces the number of unknowns in accordance with such a
symmetry.

The DINE subroutine moves the source point over nodes in the domain
of the plate. In the case of nodal deflection model, the source
points are placed at the deflection unknowns nodes. Some of these
nodes may be on the free boundary if any. Hence, even though called
domain equations, some of them may well be due to a boundary
position of the source point. In the case of the trigonometric
deflection model, the source points are usually taken at peak
values of the deflection shapes considered. Hence, Nd domain
equations are obtained where Nd is either the number of domain
nodes in the first case or the number of trigonometric modes in the
second. Again, this number is reduced should any symmetry be
considered.

Finally, the BUKLE subroutine transforms the system of equations


into a standard eigenvalue problem with {Ud } as unknowns and A as a
parameter, solves this system and writes all results into an output
file.
118

5.10 Results

This chapter describes the results obtained at each stage of the


analysis. The convergence of the Fourier transformation is first
assessed separately. Next, the accuracy of the transformed domain
integral studied by comparing results from both boundary and domain
integrations. Finally, critical load factors and buckling modes
obtained by both trigonometric and nodal deflection models are
compared to analytical results for plates with various shapes,
support and loading conditions.

Unless otherwise stated, the plates used as examples in this


chapter have the dimensions and material properties described in
figure 5.6.

10

b=1.0
t=0.05
E=87360.
\1=0.3
D=1.0

Figure 5.6 Plate geometry and physical properties.


119

5.10.1 Convergence of the Fourier Transformation

Vhen the inplane stress distribution is uniform and the nodal


deflection model is used, the L(w) function is the sum of simple
trigonometric functions that are exactly reproduced by Fourier
analysis. However, when inplane stresses are not uniform and/or if
the trigonometric deflection model is used, L(w) is not always a
simple function. Particular modes Vi(x,y) of the trigonometric
deflection model are applied to four plate examples and the L(V i )
function is calculated and transformed into a Fourier series. The
exact values and the Fourier approximation values of the L(V.1 )
function are calculated at 441 points in the domain of the plate
and the average error t computed:

t= -Lt
1 ~
i (n=441)
n
i=1

Identical values of Fourier expansion order was taken in both x and


y directions, that is NI=Nl1=NI2' Tables S.la to S.ld show the
results for various values of NI , thus various numbers of discrete
points Np=4NI2, and different trigonometric modes mx,m y in the x
and y directions. Figures S.7a to S.7d show the graphic variation
of the error with N .
P

Example 1

A plate, simply supported and uniformly compressed on all four


edges, has an L(w) function of the form Ksin(m x nx/a)sin(my ny/b).
This function is exactly reproduced by a Fourier analysis and the
error is hence 0.00% for any mode order mx ,my . Similarly, if a
simply supported plate is shear-loaded on all four edges, its L(w)
function is of the form Kcos(mx nx/a)cos(my ny/b). Again, L(w) is
exactly reproduced and the error is 0.00%
120

Example 2

A plate, clamped along one edge and simply supported along the
other three, is uniformly compressed in both directions.

Table 5.1a Average percent error of Fourier analysis, Example 2.


N 36 64 100 144 196 256 400
P
NI 3 4 5 6 7 8 10
mx,mX
1,1 26.11 12.51 11.78 9.73 7.01 4.68 3.69
2,2 53.79 25.12 17.48 12.25 9.16 7.66 4.67
3,3 **.** 56.76 30.61 18.29 15.29 11.10 6.06

(**.**: error>100%)

1: Mode 11.1 1
o
o 2: Mode 12.21
fError 0 3: Mode 13.31
CD

o
o
o
...r

80 160 240 320 400


Np

Figure 5.7a Fourier approximation error vs N : Example 2.


p

Figure 5.7a shows that the Fourier approximation converges as the


Fourier number NI increases. In the y direction (SS-SS) the
function is an ordinary sine function which is exactly
reproduced. In the x direction (CL-SS), however, some error is
induced. Notice that as the mode order increases, the accuracy
decreases because a higher number of half waves is involved. The
(3,3) mode cannot be approximated at all with a 3x3 Fourier series.
121

Furthermore, the slopes of the curves show clearly that the


convergence becomes faster with higher mode orders.

Example 3

A plate, simply supported along two adjacent edges and clamped


along two others, is uniformly compressed in the x-direction.

Table 5.1b Average percent error of Fourier analysis, Example 3.


Np 36 64 100 144 196 256 400

Nl 3 4 5 6 7 8 10
mx,m;z:
1,1 38.16 27.95 25.24 17.71 14.38 15.86 7.72

2,2 **.** 25.49 17.34 14.99 12.01 12.01 9.53


3,3 **.** **.** 28.73 11.25 10.32 9.20 5.48

(**.**: error>100%)

1: Mode (1.1)
Z: Mode (Z.Z)
o
o 3: Mode (3.3)
lError 0
q-

o
o
N

o
o
o~---.---.----.---.---.----r--~--~r---T---'
o 80 160 240 320 400
Np

Figure 5.7b Fourier approximation error vs N Example 3.


p

Figure 5.7b shows again that the Fourier series converges towards
the L(w) function as Nl increases. Error estimates are higher than
122

in example 1 because the error is induced in both directions in


this case. As from 6x6 Fourier term order, the accuracy seems to
improve as mx,my increase.

Example 4

A plate, simply supported along two opposite edges, clamped along


another and with the last free, is uniformly compressed in the x
direction along the two simply supported edges.

Table 5.1c Average percent error of Fourier analysis, Example 4.


N 36 64 100 144 196 256 400
P
Nl 3 4 5 6 7 8 10
mx,mx:
1,1 15.15 7.38 4.44 3.42 2.85 2.41 1.69
2,2 17.57 5.54 4.26 3.58 2.97 2.53 1.63

3,3 **.** 42.22 17.29 14.62 7.95 5.85 4.60

(**.**: error)100%)

1: Mode (I. I )
o 2: Mode (2.2)
o
lError 0
CD
3: Mode (3.3)

o
o
o.q-

80 160 240 320 400


Np

Figure 5.7c Fourier approximation error vs N Example 4.


p
123

This example follows the same general pattern as the previous two.
Steady convergence and higher error with higher modes order, again
characterises the results.

Example 5

A plate, clamped along three edges and simply supported along the
last, is loaded by a uniform shear along the four edges.

Table 5.1d Average percent error of Fourier analysis, Example 5.


N 36 64 100 144 196 256 400
P
Nl 3 4 5 6 7 8 10
m ,m
X Y..
1,1 24.80 18.01 9.21 9.44 6.82 5.47 2.89
2,2 **.** 21.38 16.00 10.89 8.11 6.81 3.64
3,3 **.** **.** 17.89 14.70 12.49 9.06 4.08
(**.**: error>100%)

lError
Cl
Cl
1: Mode (1.1)
2: Mode (2.2)
3: Mode (3.3)
Tb~l
b

Cl
Cl
Cl1r---r---'---'r---~--'----r--~--~~--~--~
o 160 240 320 400
Np

Figure 5.7d Fourier approximation error vs Np Example 5.


124

This example considers a case of shear loading in which first


derivatives of the deflection functions appear in the expression of
L(w) function. Nevertheless, the same degree of accuracy is
observed.

These five examples have proved the validity of the Fourie~

analysis as applied to the representation of the L(w) function.


They have shown that the accuracy usually decreases as the modal
order increases; but a Fourier expansion of order 6x6 or higher has
consistently given a reasonable accuracy. It is important to note
that a much faster convergence is expected anyway when the
integration of uL(w) is performed. Conclusions about the efficiency
of the whole transformation technique cannot be derived before
examining the accuracy of the integral transformation through the
Rayleigh-Green identity.

5.10.2 Convergence of the Transformed Integrals

An overall assessment of the accuracy of the technique can be


achieved by comparing the value of integral Id obtained by direct
domain integration on the one hand, and that value obtained as a
result of the transformation, that is by Fourier analysis and
boundary integration, on the other.

Three plate examples already used in section 5.10.1 are considered


in this section. The integration of (uL(w» will be performed in
this section, first on the boundary using the transformed integral
expression and then in the domain using the original integral
expression. The 4 points Gaussian quadrature on the boundary and 12
points quadrature for triangular cells in the domain are used. The
boundary integration points are raised to 32 in the case of
near-singular integrands. The boundary has been discretized into 16
boundary elements with 4 elements on each side. The domain has been
discretized into 200 triangular elements. The integration has been
125

performed for various values of mx and my of the arguments mxnx/a


and myHy/b. An arbitrary position of the source point on the
boundary has been chosen. This is the first node of the second
boundary element.

The symbol FA%Err appearing in tables 5.2a and 5.2b, represents the
Fourier analysis percent error already defined in section 5.10.1.
Np is the number of discrete points in the Fourier domain. I d (u 1 )
and I d (u 2 ) are the values of the integrals obtained by domain
integration.

Example 1

A plate simply supported along its four edges is loaded by uniform


shear. This example was treated in section 5.10.1 under example 1.
The deflection function is of the form:

w=sin (mxnx/a)sin (myHy/b)

The resulting expression for L(w) is:

a2w mn mn my ny
L(w)=2N --- _x-~cos cos
xy axay a b a b

The Fourier analysis reproduces the exact function and very good
accuracy (0.00% error) is achieved regardless of the value of N .
P

Example 2

This is the same as example 3 of section 5.10.1. Table 5.2 shows


the results.
126

Table 5.2a Integral Transformation, example 2, mx =l,my =1.

Ne FA%Err I b (u 1 ) %Error I b (u 2 ) %Error


36 38.16 -0.707365E-2 -16.11 0.310055E-2 -34.66
64 27.95 -0. 773650E-2 -8.24 0.385126E-2 -18.84
100 25.24 -0.800561E-2 -5.05 0.407389E-2 -14.15
144 17.71 -0. 814355E-2 -3.42 0.426170E-2 -10.19
196 14.38 -0.822493E-2 -2.45 0.437352E-2 -7.83
256 15.86 -0.827548E-2 -1.85 0.445240E-2 -6.17
400 7.72 -0.833383E-2 -1.16 0.454995E-2 -4.12

I d (u 1 )=-0.843160E-2
I d (u 2 )= 0.474522E-2

CJ
CJ
lError

CJ Mode (1.1)
CJ
CJ
N

CJ
CJ
CJo~o~o~--~~~==~;;;F==~~~==~~~~~10.00
CJ Np
CJ
CJ

1: I (u 1 I
CJ
2: I (u2 )
CJ F: FAlErr
CJ
...,.
I

Figure 5.8a Integral transformation, %Error vs Np ' example 2, mode


(1,1).
127

Table 5.2b Integral Transformation, example 2, mx =2,my =2.


N FA%Err I b (u1 ) %Error I b (u 2 ) %Error
I!
36 **.** 0.965205E-2 -40.18 0.530705E-2 35.57
64 25.49 0.134344E-1 -16.74 0.199243E-2 -49.10
100 17.34 0.142961E-1 -11.40 0.275090E-2 -29.73
144 14.99 0.149042E-1 -7.63 0.294880E-2 -24.67
196 12.01 0.152207E-1 -5.67 0.315039E-2 -19.52
256 12.01 o. 15441OE-1 -4.30 0.330687E-2 -15.52
400 9.53 0.156897E-1 -2.76 0.349675E-2 -10.67

(**.**:error>100%)
I d (u 1 )= 0.161354E-1
I d (u 2 )= 0.391449E-2

o
o
lError

Mode (2.21
"<t

0
0
0 1
0.00 320.00 4 0.00
0
Np
0
0
"<t
I

I: I (u I J
0
2: I (u2 J
0
F: FA/Err
0

Figure 5.8b Integral transformation, %Error vs N : example 2, mode


p
(2,2).
128

Table 5.2c Integral Transformation, example 2, mx=3,my =3.


N FA%Err I b (u 1 ) %Error I b (u 2 ) %Error
12
36 **.** O.OOOOOOEOO **.** O.OOOOOOEOO **.**
64 **.** -0.217696E-2 -68.24 0.835960E-2 **.**
100 28.73 -0.448966E-2 -34.49 0.361723E-2 35.03

144 11.25 -0.530321E-2 -22.62 0.314545E-2 17.41

196 10.32 -0. 578089E-2 -15.65 0.290121E-2 8.30


256 9.20 -0.604980E-2 -11.73 0.272259E-2 1.63
400 5.48 -0.636091E-2 -7.19 0.264952E-2 -1.10

(**.**:error>100%)
I d (u 1 )=-O.685340E-2
I d (u 2 )= O.267892E-2

o
o
lError 0
'0
Mode (3.3)
o
o
o
N

D~~~~~--r-~--~---T~Ir~9===~==i
9:1.00
o
N
I

o
o 1: I (u 1 )
o
'0
I
2: I (u2 )
F: FAlErr
o
o
o
o

Figure 5.8c Integral transformation, %Error vs Np , example 2, mode


(3,3).
129

The variation of the error with Np shown in figures 5.8a,b,c,


indicates a clear convergence of the transformed integral to its
value obtained by domain integration; hence it proves the validity
of the transformation technique. It is important before discussing
these results to note once more that the FA%Err is the average of
errors of the Fourier analysis measured at 441 points of the plate
in section 5.10.1. The FA%Err is only used as a general measure of
convergence but is not an exact analytical estimate .

In the case of mode (1,1) of figure 5.8a, the integral convergence


is faster than the Fourier analysis convergence for both
f un d amen t a I so I u t 10ns
· d D h
u1 an u2 . ue to t e 1/ 2 . I ' . u2 '
r slngu arlty In
accuracy associated with u 1 is better. Figure 5.8b for mode (2,2)
shows that the accuracy of I(u 1 ) is still better than the accuracy
I(u 2 ) and that of the Fourier analysis. However, the I(u 2 )
convergence seems slower than the Fourier convergence. Figure 5.8c
for mode 3,3 shows an unexpected improvement in I(u 2 ) accuracy that
makes it once more better than the Fourier analysis accuracy. Note
also that, except for I(u 2 ) in mode (3,3), the error increases with
higher deflection modes. Finally, for Np=400 (N l1 =10,N I2 =10), the
error is always less than 11%.

Example 3

This is the same as example 4 of section 5.10.1. Tables 5.3 show


the values of integrals I(u 1 ) and I(u 2 ) and the error induced, and
figures 5.9 show the variation of the error with the number of
discrete points Np
130

Table 5.3a Integral Transformation, example 3, mx =l,my =1.


N FA%Err I b (u 1 ) %Error I b(u 2 ) %Error
l!
36 15.15 -0.431367E-2 0.69 0.178027E-1 3.15
64 7.38 -0. 428311E-2 -0.02 0.173753E-1 0.68
100 4.44 -0.428780E-2 0.09 0.171833E-1 -0.44
144 3.42 -0.428557E-2 0.04 0.171100E-l -0.86
196 2.85 -0. 428872E-2 0.11 0.170717E-1 -1.08
256 2.41 -0.428819E-2 0.10 0.170591E-1 -1.16
400 1.69 -0.428926E-2 0.12 0.170560E-1 -1.17

Cl
Cl
lError N....

Cl
Cl
'0

Mode (1.1)

Cl
Cl
CD

1: I (u 1
Z: I (uZ )
F: FAlErr
Figure 5.9a Integral transformation, %Error vs Np , example 3, mode
(1,1).
131

Table 5.3b Integral Transformation, example 3, mx=2,my =2.


N FA%Err I b (u 1 ) %Error I b (u 2 ) %Error
12
36 17.57 -0. 448004E-2 -22.19 -0. 717303E-1 -0.93
64 5.54 -0.505520E-2 -12.20 -0. 710995E-1 -1.80
100 4.26 -0.525628E-2 -8.70 -0. 715980E-1 -1.12
144 3.58 -0. 539403E-2 -6.31 -0. 717722E-1 -0.88
196 2.97 -0.546748E-2 -5.03 -0.719361E-1 -0.64
256 2.53 -0.552648E-2 -4.01 -0. 720157E-1 -0.54
400 1.63 -0.559338E-2 -2.85 -0. 721334E-1 -0.38

I d (u 1 )= -0.575735E-2
I d (u 2 )= -0.391449E-1

o
o
lError N
M

o
o Mode (Z.ZI

8 ~~F ~F F ---E-F
o·i---~~~Z=F=z,==~===r~9F==,===~==i
o 00 0.00
-----EF

o Np
o
'()

1: I (u 1I
o Z: I (uZ I
o F: FAlErr
N

Figure 5.9b Integral transformation, %Error vs N , example 3, mode


p
(2,2).
132

Table 5.3c Integral Transformation, example 3, mx =3,my =3.


N FA%Err I b (u 1 ) %Error I b (u 2 ) %Error
E
36 **.** O.OOOOOOEOO **.** O.OOOOOOEOO **.**
64 42.22 -0.444778E-1 -3.95 0.650159E-1 14.68
100 17.29 -0.454013E-1 -1.95 0.630092E-1 11.14

144 14.62 -0.456926E-1 -1.32 0.628694E-1 10.89


196 7.95 -0.458872E-1 -0.90 0.609025E-1 7.42
256 5.85 -0.459917E-1 -0.68 0.596802E-l 5.27
400 4.60 -0.461291E-1 -0.38 0.583222E-1 2.87

(**.**:error>100%)
I d (u 1 )=-0.463056E-l
I d (u 2 )= 0.566933E-l

0
0

~ D~
lError '0
lfl
1: I (u 1 )
2: I (u2 )
0
0 F: FA/Err
0
....
Mode (:3.:3 )
0
0
....
N

0
0
CD

1
EP .00 160.00 240.00 320.00 400.00
,
CD Np

Figure 5.9c Integral transformation, Error vs N , example 3, mode


p
(3,3).
133

It is clear from figures 5.9 that error decreases as Np increases.


The particularity of mode (3,3) of the previous example, where
I(u 2 ) gives better accuracy than I(u 1 ) and FA%Err, is repeated here
with mode (2,2). Apart from this mode, the characteristics of the
previous examples are observed: accuracy for I(u 1 ) is better than
that for I(u 2 ), integral convergence is faster than that of the
Fourier analysis, the error increases as higher modes are used.

The three examples considered in this section have verified the


numerical validity of the transformation technique for the
trigonometric deflection model. It has been shown that the
integrals rate of convergence depends on the particular form of
L(w) under consideration, that is the support and loading
conditions, the number of Fourier discrete points N and the degree
p
of singularity in the fundamental solution u.

5.10.3 Examples of Critical Loads.

In what follows, various examples of plates, loaded in their plane,


are analysed and their critical loads assessed using both the
trigonometric and nodal deflection models. The boundary of
rectangular plates is discretized into 16 boundary elements with 4
elements on each edge. Symmetric properties are used to reduce the
overall number of unknowns.

The critical buckling stresscr of the plate is related


a to the
buckling coefficient Ab through the following relation:

n20
acr = ---A
bb b
1

where b is the plate


width and bl the strip width of the
distributed 2
load. The n term is omitted for the triangular plate
treated under example 10 using the nodal deflection model. If the
plate is loaded by concentrated forces as in examples 18 to 20 for
the trigonometric model and 8 for the nodal model, Ab is given by:
134

acr =

5.10.3.1 The Trigonometric Deflection Model

Results for 20 plate examples are shown in tables 5.4. Various


combinations of SS, CL, FR support conditions, loading conditions
and plate dimensions are used. Vhenever the plane stress
distribution is non-uniform as in examples 15 to 20, the boundary
element program for plane stress problems, described in chapter 3,
has been used to find the stress distribution in the domain of the
plate. Symmetric properties of the plate, if any, are taken into
consideration when choosing the deflection modes.

The column'Nmo des ' in tables 5.4 indicates the number of modes
used, that is the number of domain unknowns. Each mode shape is
indicated by a pair of integers (m,n) representing the modal
parameters of the modes defined in appendix C where m corresponds
to the x-direction and n to the y-direction. The column 'Modes'
states these modes either explicitly (e.g. (1,1)(1,2) ..• ) or
implicitly, that is the notation (1~2)x(1~3) means the collection
of modes (1,1),(1,2),(1,3),(2,1),(2,2),(2,3). Nl=Nl2=Nl1 indicates
the order of the Fourier approximation in both x and y directions.
The computed buckling coefficients are shown in column 'Ab ' and the
errors with respect to analytical or numerical values of Ab are
also shown in column '%Err'. Finally, the reference numbers of the
analytical buckling loads correspond to the following publications:

1. Timoshenko & Gere (1961)


2. Bulson (1970)
3. Khan & Valker (1972)
4. Khan, Kenneth & Hayman (1977)
5. Alfutov & Balabukh (1967)
6. Shuleshko (1957)
135

Table 5.4 Examples of critical loads, Trigonometric model.


Plates N
modes Modes
\ %Err
1
~ ,J,.Jxb.Jxb

~DI1o/I'o/f~
1 (1,1) 4 1.998 -0.10%

\=2.0 (Ref 1,2)


2

~DI 1 (1,1) 4 3.997 -0.08%

Ab=4.0 (Ref 1,2)


3
~xJx1dxW;.

~DI
1 (1,1) 4 5.346 0.87%
2 (1,1),(3,3) 4 5.400 1.89%
3 (1,1)(3,3)(5,5) 4 5.340 0.75%
1"J'1'1YP'1'
Ab=5.3 (Ref 1,2)
4

~DI
1 (1,1) 4 10.690 6.16%
2 (1,1),(3,3) 4 10.800 7.25%
3 (1,1)(3,3)(5,5) 4 10.140 -0.70%
Ab=10.07 (Ref 1,2)

101
1 (1,1) 4 8.917 15.96%
4 (1~2)(H2) 4 8.094 5.25%
9 (H3)x(1~3) 4 7.669 -0.27%
16 (1~4)x(1~) 4 7.585 -1. 37%
3 (2,1)(2,2)(2,3) 4 7.662 -0.36%
\=7.69 (Ref 1,2)
136

Table 5.4 (cont'd)

Plates N
modes Hodes
\ %Err

IDI
6
1 (1,1) 4 6.750 -0.74%

4 (H2)x(H2) 4 6.750 -0.74%

9 (1-+3)x(1-+3) 4 6.746 -0.79%

Ab =6.8 (Ref 2) 16 (1-+4 )x( 1-+4) 4 6.746 -0.79%

7 1 (1,1) 4 5.295 -2.11%

IDI
2 (1,1)(1,2) 4 5.537 2.35%
3 (1,1)(1,2)(1,3) 4 5.299 -2.05%
4 (1,1)(1,2)(1,3) 4 5.650 4.44%
(1,4)
1 (1,1) 8 5.271 -2.56%
2 (1,1)(1,2) 8 5.514 1.92%
3 (1,1)(1,2)(1,3) 8 5.272 -2.55%
(a=O. 746, b=1.0) 4 (1,1)(1,2)(1,3) 8 5.622 3.92%
Ab ",5.41 (Ref 2) (1,4)

101
8

1 (1,0) 4 0.997 -0.30%

\=1.0 (v=O,Ref 1)

101
9 ~2;l2SlAbd2Sl7
1 (1,0) 4 0.995 6.99%

3 (1,1),(1,3) 4 0.912 -1. 94%


1"1"'1"'1"]" F'
Ab =0.93 (Ref 6)

10 1 (1,1)

101
4 1.391 -3.40%

2 (1,1)(1,2) 4 1.439 -0.07%

1 (1,1) 8 1.412 -1. 94%

2 (1,1)(1,2) 8 1.456 1.11%


Ab =1. 44 (Ref 1)
137

Table 5.4 (cont'd)


Plates \ %Err
- - - - - - - Nmodes Modes
11 1 (1,1) 1. 747 2.76%

4 (1,1)(1,2) 4 1.708 0.47%

1 (1,1) 8 1.738 2.24%


4 (1,1)(1,2) 8 1.698 -0.12%

12
5 (1,1)(1,3)(2,2) 4 9.920 6.21%
(3,1)(3,3)
9 (H3)x(H3) 4 9.009 -3.54%
(1,1)(2,2)(3,3)
9 (4,4)(5,5)(1,3) 4 9.292 -0.51%
A.b=9.34 (Ref 2) (3,1)(3,5)(5,3)
13
2 (1,1)(2,2) 4 20.360 39.64%
3 (1,1)(2,2)(3,3) 4 16.800 15.23%
4 (1,1)(2,2)(3,3) 4 14.490 -0.62%
(4,4)
A.b=14.58 (Ref 2)

14
4 (1-+2)x( 1-+2) 4 11.476 4.33%

~b~ ~
9

9
(1-+3 )x( 1-+3)
(H2)x(H2)
(H3)x(1-+3)
4

8
10.421
11.978
10.600
-5.26%
8.89%
-3.64%
\=11.0 (Ref 2)
15

1 (1,1) 8 3.108 1.90%

(load width=0.5)
\=3.05 (Ref 3)
138

Table 5.4 (cont'd)


Plates Nmodes Modes %Err
>0
16
~xWxI:xlxb
4 (1,1)(1,2)(3,1) 8 5.919 5.70%

~D[
(3,2)
(1,1)(1,2)(1,3)
9 (3,1)(3,2)(3,3) 8 5.559 -0.73%
(5,1)(5,2)(5,3)
>0=5.6 (Ref 3,4)
17
~

~o[
4 (1,1)(1,2)(3,1) 8 4.343 11.36%
(3,2)
(1,1)(1,2)(1,3)
9 (3,1)(3,2)(3,3) 8 4.064 4.21%
(load width=0.5) (5,1)(5,2)(5,3)
>0=3.9 (Ref 4)
18
1 (1,1) 8 0.683 1.94%

-1Dr-
(load:P)
2

3
(1,1)(3,3)

(1,1)(3,3)(5,5)
8

8
0.681

0.680
1.64%

1.49%
>0=0.67 (Ref 1,3,5)
J, J,

JRt
19
1 (1,1) 8 0.430 -14%

2 (1,1)(3,3) 8 0.379 -24.2%

(load:P) 3 (1,1)(3,3)(5,5) 8 0.381 -23.8%


>0=0.5 (Ref 5)
20

Jot
1 (1,1) 8 1. 723 -13.85%

2 (1,1)(3,3) 8 1.528 -23.6%

(load:P/2) 3 (1,1)(3,3)(5,5) 8 1.536 -23.2%


>0=2.0 (Ref 5)
139

As expected, nearly exact results have been obtained for examples 1


and 2 since the assumed deflection mode (1,1) coincides with the
exact buckling mode of the problem. Accurate results have been
obtained in examples 3 and 4 involving a uniformly compressed
clamped plate, with 3 modes including mode (5,5). Note that there
is a loss of accuracy when mode (3,3) is introduced in the second
run. This shows that mode (5,5) is crucial in this case.

The critical buckling factor of example 5 corresponds to an


antisymmetric mode in the x-direction. The accuracy of the result
improves as the number of modes is increased except when 16 modes
are used. In this latter case, the higher number of domain unknowns
proves to be disadvantageous as the modes associated with m=4 and
n=4 have negligible contribution. In fact, the modes with major
contribution can be found by comparing the values of the modal
coefficients, that is the domain unknowns, of the first four runs.
Accordingly, modes (2,1),(2,2) and (2,3) are selected alone in the
fifth run and good accuracy is obtained.

In example 6, mode (1,1) suffices for good accuracy. No improvement


is observed as other modes are introduced. Example 7 is a
rectangular non-square plate. The error in this case is around 2%,
slightly higher than the previous examples. Again, the introduction
of a fourth unknown is disadvantageous. Example 8 is a plate
containing 2 opposite free boundaries i.e. it has a FR-FR mode in
the y-direction. The use of the mode m=O, that is the constant
function simulating rigid body motion in the FR-FR direction,
results in good accuracy.

The results for example 9 which is similar to 8 except that


compression is in both directions, are not as accurate as in
example 8. The search for and use of more appropriate modes may
improve the accuracy.
140

Examples 10 and 11 both contain only one free boundary. Convergence


is observed in both examples as the order of the Fourier
approximation is increased and as higher modes are introduced. The
increase in Fourier order gives very good results in example 10,
while the addition of mode (1,2) in example 11 seems to have the
decisive effect that leads to good accuracies.

Examples 12, 13 and 14 are plates subjected to a uniform shear


loading. In examples 12 and 13 diagonal symmetry has been taken
into account when choosing the modes. Timoshenko & Gere (1961)
showed that for the simply supported plate of example 12 the sum
(m+n) is even. Convergence is observed as the number of modes is
increased despite some irregularities in example 14. Again, the
trial of different modal functions may lead to better results.

Examples 15 to 20 are all plates having a non-uniform plane stress


distribution. Unlike the previous examples, the plane stress
distribution is obtained numerically and, moreover, the buckling
coefficients to which the computed values are compared are
approximate.

In example 15 an error estimate of 1.9% is achieved. In example 16,


accuracy is considerably improved as 9 modes are used. The same
applies to example 17; however the error there remains higher.

Examples 18 to 20 are plates subjected to concentrated inplane


compressive forces. In these examples, the load has been actually
considered as a small strip of width wq=0.01 of uniformly
distributed load q and has been transformed to its concentrated
force equivalent P that is P=wqq.

In example 18, mode (1,1) is the major mode; the introduction of


modes (3,3) and (5,5) improves the result only slightly. In
examples 19 and 20, the error is noticeably high and does not
converge as higher modes are introduced. However, it must be said
that the only source of the approximate analytical results, Ab=O.S
141

for example 19 and ~=2.0 for example 20, is the paper by Alfutov
and Balabukh (1967). An alternative source is needed to confirm
these results.

To conclude, the above listed results demonstrate the overall


validity of the formulation based on the trigonometric deflection
model. The results accuracy clearly depends on the choice of the
deflection modes and the order of the Fourier approximation.

5.10.3.2 Nodal Deflection Model

Tables 5.5 show the results obtained. Pair of numbers (m,n)


indicates the order of the Fourier approximation of the deflection.
In fact, it represents the number of intervals to which the plate
is divided in each direction. The Fourier approximation of the
inplane stress function L(w) is also (m,n) except for the cases of
non-uniform membrane stress, as in examples 8 and 9, where it is
taken as (8,8) regardless of the Fourier order for the deflection.
Nd is the number of nodes where the deflection unknowns are taken.
Hence, the relation between (m,n) and Nd is:

where Nz is the number of deflection nodes located on simply


supported or clamped boundaries in addition to, in the case of
non-rectangular plates, nodes located outside the actual domain. No
deflection unknowns are taken at such nodes.

Tables 5.5 and figures 5.10 show clearly that results given by the
nodal deflection model converge towards the actual buckling factor
as the number of domain nodes is increased. Indeed, very accurate
results are obtained with (m,n)=(5,5) and (m,n)=(6,6). The results
of the triangular plate example 10 also converge; the error,
however, is very high.
142

Table 5.5a Examples of critical loads, nodal deflection model,


uniformly compressed plates.
Plates (m,n) Nd Ab %Error

1 J?JIJA!AW, 3,3 4 2.030 1.50%

~DI
4,4 9 2.003 0.15%
5,5 16 2.002 0.10%
\=2.0 1PM'1"1'
,
6,6 25 1.998 -0.10%
2 JJ$IISk\kl;. 3,3 4 5.856 10.49%

~DI
4,4 9 5.443 2.69%

5,5 16 5.388 1.66%


Ab=5.3, 'P'JYMYP' P' 6,6 25 5.348 0.91%

IDI
3 3,3 4 7.890 2.60%
4,4 9 7.772 1.07%
5,5 16 7.708 0.23%
Ab=7.69 6,6 25 7.707 0.22%
.JMMJ7

IDI
4 3,3 8 0.890 -4.3%
4,4 15 0.931 0.11%
5,5 24 0.905 -2.69%
Ab=0.93 1'o/P't"~ ~ 6,6 35 0.923 -0.75%

lError 1: Example 1
a 2: Example 2
a
(D
3: Example 3
4: Example 4

a
~~--~~~~~~~~~---F~-'---'--Tr--~
0.00

a
a

Figure 5.10a Nodal model, %Error vs Nd uniformly compressed plates.


143

Table S.Sb Examples of critical loads, nodal deflection model,


plates loaded by shear.
Plates (m,n) Nd >0 %Error
5

]tJ~
3,3 4 11. 725 25.54%
4,4 9 9.507 1. 79%
5,5 16 9.446 0.96%
~ 6,6 25 9.343 0.03%
Ab=9.34

itJJ
3,3 4 23.776 63.07%
4,4 9 15.283 4.82%
5,5 16 14.677 0.67%
6,6 25 14.697 0.80%
Ab=14.58
7

]tJ~
3,3 4 14.777 34.34%
4,4 9 10.985 -0.14%
5,5 16 10.856 -1.31%
~ 6,6 25 10.742 -2.35%
Ab=11.

lError 5: Example 5
C)
C) 6: Example 6
C)
'0
7: Example 7

C)
C)

C)
N

C)

c:n
C)
00 8.00 16.00 24.00 32.00 40.00
Nd
Figure 5.10b Nodal model, %Error vs Nd , Plates loaded by shear.
144

Table 5.4c Examples of critical loads, nodal deflection model,


non-rectangular plates and non-uniform loading.
Plates (m,n) Nd Ab %Error

8
3,3 4 0.708 5.67%

1D~
4,4 9 0.690 2.99%

5,5 16 0.680 1.49%


6,6 25 0.682 1.79%
Ab=0.67

9
3,3 4 3.213 5.34%

eD~ 4,4

5,5
9
16
3.153

3.117
3.38%
2.20%

6,6 25 3.107 1.87%


\=3.05

1&\
10
6,6 13 44.555 12.86%

7,7 18 44.221 12.01%

1?I'1'1vp 8,8 25 43.842 11.05%


(equilateral,dimension:1) 9,9 32 43.671 10.62%
Ab=39.478

lError 8: Example 8
o
o 9: Example 9
'0 0: Example 10

o
o
CD

o
o
o4----r---.---.---,---..---r---r---~--.---~

0.00 8.00 16.00 24.00 32.00 40.00


Nd
Figure 5.10c Nodal model, %Error vs Nd , non rectangular plates and
non uniform loading.
145

fictitious domain

Figure 5.11 Fourier domain of triangular plate.

The Fourier domain used for the triangular plate is shown in figure
5.11. The reason for the inaccuracy is that, unlike the rectangular
plates case, w=O boundary condition cannot be satisfied by the
deflection model. Moreover, the severe near-singularity of
integration around the corner may have increased the error since
results of the same plate example from program PLABEM in chapter 4
were also less accurate than the other examples. The trial of
different Fourier domains may lead to better results. The partially
satisfied boundary conditions and the near-singularity around free
boundaries, may also be the reason for the erratic behaviour of the
error curve in the free boundary case of example 4.

5.10.3.3 The Plates Deflected Shape

The buckling mode is directly obtained when the system of equations


is solved since deflections, rather than curvatures, are modelled
in the domain of the plate. Figures 5.12 show the deflected shape
given by both trigonometric and nodal models compared to analytical
results, for 3 square plate examples. Analytical deflection curves
are taken from the book by Timoshenko & Gere (1961). The deflection
is sketched along the half diagonal of the square plates. The
normalised distance from the lower left corner of the plate to the
point on the diagonal is designated by s with s=O at the lower left
corner and s=l at the centroid.
146

o
~U1

0
0-
0

0
..J CD
W 0
0
0
~

u 0
r-
a:..J
..JI-W 0
<WO
u~o
+ ;:~~
>- ..J 0
'0
..Jl:J<
<-8
Z([
0

<I-Z
0
lfl
I X +
0

+ ..,.
0

0
I"')

0
N
0

-
0

0
0

00' I 08'0 09'0 00'00


~

Figure 5.l2a Deflection profile of a simply supported plate


uniformly compressed in both directions.
147
o
1:?U1

o
(),

o
ID
o

o
r-
o

o
'0
o

o
lfl
o

-' .,.
o

~ o
o
::?:
U o
1"1
([-'
-ll-W o
«wo
u::?:o
;:~::?:
o
>-0--3 N
-lLJ«
o
~o:8
«I-Z
o
I X ...
o

00' I 08'0 09'0 01"0 oz·o


==

Figure 5.12b Deflection profile of a fully clamped plate uniformly


compressed in both directions.
148

0
0..
0

0
~
+ 0

0
I"-

0
'(I

...1
~ W
0
0
~

-
u
0 :...1
--1f-W
<WO
u~o
i=~~
>-0--1
--1LJ<
<
zo:-8
<f-Z

I x +

Figure 5.12c Deflection profile of a simply supported plate loaded


in uniform shear along its four edges.
149

Although all three curves are reasonably close, the trigonometric


model gives more accurate deflections than the nodal model. In
fact, the trigonometric model simulates the exact deflection mode
in the case of a simply supported plate uniformly compressed in
both directions and an almost exact mode when the plate is clamped.

5.11 Conclusions

The validity of the dual-reciprocity technique which transforms the


domain integral into a boundary one has been demonstrated in this
chapter.

The trigonometric deflection model allows a selective choice of the


trigonometric modes based on the knowledge of the plate geometric
properties and boundary conditions. It is therefore a more
realistic continuous deflection model if compared to the more
conventional piecewise finite elements based on polynomial
approximations. In simple cases of boundary conditions and loading,
a small number of modes is needed and very efficient solutions are
obtained in terms of accuracy and computer memory and speed. The
trigonometric model remains valid for more complicated examples
containing free boundaries and non-uniform membrane stress
distribution as has been shown in section 5.10.3.1; however, the
mode selection becomes more difficult and more domain unknowns are
hence required. The advantageous trigonometric modes may therefore
become a disadvantage. Moreover, the position of the source point
for the domain equations starts having a strong effect on the
results. Finally extension of the model to non-rectangular plates
is difficult.

The nodal deflection model represents the deflection in terms of


unknowns at some specified nodes of the plate domain. Any plate
shape can therefore be considered. Results for triangular plates
were however somewhat inaccurate. The data input is more automatic
then for the trigonometric model where mode selection is required.
150

There is a clear convergence as the number of domain nodes is


increased as has been shown in section 5.10.3.2. However, more
computing time and memory is needed.

The transformation of the domain integrals into boundary ones has


produced a numerical implementation of the problem having many
advantages over the ones using domain discretization. Accuracies
have been improved. The data structure of the problem has been
further simplified. The data needed is actually restricted to the
physical properties of the plate and the number of boundary
elements along each edge of the plate in addition to the domain
data which is either modes specifications or nodes number depending
on the model. Further investigations are needed however in the case
of non-rectangular plates. Saving in computer memory is produced by
the reduced data structure and by the smaller system of domain
equations. The computing time taken by the solution of the system
of equations is also reduced; however, in the case of the nodal
deflection model, this is sometimes outweighed by the time needed
for the evaluation of the unknown coefficients.
CHAPTER SIX

LARGE DEFLECTIONS

6.1 Introduction

Vhen deflections are not small compared to the thickness of the


plate, coupling between bending action and midplane stretching
results in nonlinear differential equations that can only be solved
by approximate methods. The governing differential equations of
this problem have been already derived in chapter 2. Any realistic
approach to the stability problem of flat plates has to take into
account the initial deviation from flatness existing in all
practical applications. Although the buckling load and
post-buckling strength remain essentially unaffected by the initial
deviations, the load-deflection curve in the pre-buckling and
buckling regions does depend on these deviations.

The extension of the boundary element method to the large


deflection of plates as well as non-linear problems in general has
not always been a straightforward matter. Fundamental solutions for
this class of problems are either extremely difficult or impossible
to generate; instead the fundamental solutions for the respective
linear problems are often used and, as a consequence, domain
integrals occur in the resulting boundary integral equations.

Nevertheless the overall computational performance of the method


has allowed its successful extension to the solution of a wide
range of problems involving geometrical and material
nonlinearities. Moreover, progress has been recently made in
transforming the domain integrals into boundary ones for a large
number of problems. An application of such a process to plate
buckling has been presented in chapter 5.
152

The present approach proposes a boundary element solution of the


non-linear plate response to in-plane edge loading. Proceeding
along the lines followed in chapter 4 for the solution of the
critical loads, the domain curvatures are eliminated from the
integral equations governing the deflection of the plate and its
slope within the domain. An incremental approach is thus adopted
with the deflection as the only domain unknown requiring modelling.
The membrane stresses are expressed in terms of the stress function
and their variation monitored throughout the load incrementation.
This is achieved by extending the formulation presented in chapter
3 in order to include nonlinear terms. These terms contain
quadratic expressions of the domain curvatures. Thus, the
dependence of curvatures on the domain unknowns is established
through higher-order models for the deflection. This increases the
efficiency of the method by reducing the overall number of unknowns
at no significant cost to accuracy. Initially, the linearized
system of equations is solved and the results are further improved
by subsequently including the second order terms using an iterative
procedure.

The accuracy of the boundary element solution is tested by


comparing its predictions with a selection of experimental and
numerical results. The versatility of the method is demonstrated by
considering a variety of geometries, loading and support
condi tions.

6.2 Boundary Integral Formulation

The incremental form of the von Karman equations (2.27,2.28),


including the effect of initial deviation from flatness, can be
written as follows:

4 (6.1)
DV w= NCIt",
QW Q + n QVl t
,CIt", CIt",
Q
,CIt",
+ n QW Q
CIt", ,CIt",

(6.2)
153

where

C(V,w) 2w ,xyVt,xy -w ,xxVt,yy -w ,yyVt,xx +

2
+(w
,xy ) -w ,xxw,yy (6.3)

where V is the plate deflection, N~~ are the membrane stresses, w


and n~~ their respective increments, Vo the initial plate
imperfection, f the increment of the stress function, E the Young's
modulus, h the plate thickness and 0 the rigidity of the plate.
Greek indices indicate mid-plane coordinates and a comma denotes
differentiation with respect to subscripts, with summation implied
over repeated indices. It should be noted that the second-order
terms in the incremental quantities have been retained in (6.1) and
(6.2). The incremental membrane stresses are obtained from the
incremental stress function through:

a2 f
n xx =--2-' (6.4)
ay axay
In-plane traction boundary conditions have been expressed in terms
of the stress function in chapter 2. In-plane displacement boundary
conditions, as mentioned earlier, will be replaced by the traction
conditions that most closely approximate them.

A boundary integral equation can be derived from equation (6.1) by


following the same treatment described in chapter 4. The final
equation will be similar to equation (4.8) except that incremental
deflection w replaces the buckling deflection and domain integrals
d
It and I dw become I jt and I 1j , both of which include the effect of
all terms on the right hand side of equation (6.1).

Similarly, another boundary integral equation can be derived from


equation (6.2) along the lines presented in chapter 3. The
154

non-linear C(V,w) term on the right-hand side of this equation will


cause a domain integral 1 d2j to appear in the equation.

Hence, the following two integral equations are obtained:

b d
cZ(w) = I lj + J. + I lj + I~ (6.5)
J J

cZ(f) b
1 2j + I 2j
d (6.6)

where

I bl ·= =-J [vn (w)u.J - M (W)~ + aw M (u.) - wv (u.)]dr


J D n an an n J n J
r

J(-~.
_7 au.
+ V-f J -
an J an
r

c[
N

JJ'
D
1
=- \
L u. (M
J ns
(w) t )-w(

i=l

I~j ~ J(WNaf3+Vtnaf3+wnaf3)Uj ,af3dQ


Q

I~j EhJC(V,W)UjdQ
Q

I~J =- J[(T aw,a+t aVt ,a+t aw,a)u.-


D J
r
- (wT a +Vtt a +wt a )u.J,a ]dr

and c and Z are as defined with reference to equation 4.5. T is the


current value of the traction and t its increment. The index j=1,2
refers to the fundamental solution u j ' that is either u l or u2 .
155

The integral equations (6.5) and (6.6) are used to generate a


boundary element solution of the large deflection problem. The
boundary integrals I? contain eight variables namely w, awlan,
1J
Mn(w), Vn(w), f, af/an, vlf, a(vlf)/an), four of which are known
from the specified boundary conditions.

6.3 Domain Deflection Models

Vhile curvatures have been eliminated from equation (6.5), the


domain integral I~j appearing in equation (6.6) still contains a
quadratic expression of both incremental and current state
curvatures w, ex...Q and V, ex...Q. The inclusion of w, ex...Q as domain unknowns
in the boundary element solution would incur a high computational
cost since three unknowns w,xx ,wand ,xy w,yy would have to be
considered at every node. In order to overcome this difficulty, the
use of high order elements for the domain deflection is
investigated here so that expressions of the curvatures in terms of
the nodal deflections could be obtained.

The domain of the plate is discretized into Nd triangular cells


over which the deflection w is defined in terms of its values at
the Nn nodes per cell:

(6.7)

Polynomial interpolation functions, already used for the solution


of the critical loads in chapter 4, is the most direct choice. The
curvatures are then found from:

(6.8)
1~

so that the relation between cell curvatures and nodal deflections


is directly derived. Quadratic interpolation functions with Nn =6,
as shown in figure 4.4, result in modelling the curvatures as
constant over a cell. Consequently, this model is not expected to
give accurate results.

Cubic interpolation functions resulting in a linear variation of


the curvatures over a cell, are also used. The expression of the
cubic polynomials depends on the position of the intermediate nodes
on the edges of the triangular cell shown in figure 6.1. The
interpolation functions will be calculated numerically in a fashion
similar to that described for discontinuous elements in chapter 4.
The distance d of these nodes from the apices becomes a key
parameter in the approximate curvature-deflection relationship
which needs to be thoroughly investigated.

Figure 6.1 Cubic domain cell with variable position of intermediate


nodes.
157

A one-dimensional equivalent of the cubic cell can be built by


considering one edge of the cell with 4 unknown nodes as shown in
figure 6.2.

[ B(s=d) C(s=l-d) D(s=1>


A
(s=O) 1 ¥ ¥ ¥
I 1 1
I I J
~ d -..,,,~"---- i-2d ----+,."-d--,r

Figure 6.2 Details of one edge of a cubic cell.

The four interpolation functions corresponding to this model are as


follows:

1
t 1= (l-s)(s-d)(s-l+d)
d(1-d)
1
t 2= s(l-s)(s-l+d)
d(1-d)(2d-l)
(6.9)
1
t3= s(s-d)(l-s)
d(1-d)(2d-l)
1
t 4= s(s-d)(s-l+d)
d(l-d)

with the local coordinate s as shown in figure 6.2.


158

The second derivatives of these functions are given by:

a2 t 1 1
(-6s+4)
as 2 d(l-d)

a2 t 2 1
(6s+2d-4)
as 2 d(1-d)(1-2d)
(6.10)
a2 t 1
3
(-6s+2d+2)
a/ d(1-d)(1-2d)

a2 t 4 1
(6s-2)
as 2 d(l-d)

It can be easily verified that this modelling of the curvatures is


equivalent to a finite difference scheme. Such a scheme would
calculate a first estimate c 1 of the curvature using a double
direct difference approximation applied to the nodal deflections at
points A, Band C. Another curvature value c 2 is then evaluated by
applying the same process to points B, C, and D. The following
expressions of c 1 and c 2 are thus obtained:

2 2(d-1) 2d
c 1= w1 + w2 + w3
d(l-d) d(1-d)(1-2d) d(1-d)(1-2d)

(6.11)
2 2(d-1) 2d
c 2= w4 + w3 + w2
d(l-d) d(l-d)(1-2d) d(l-d)(1-2d)

If c 1 and c 2 are assumed to be the values of curvatures at points


s=1/3 and s=2/3 respectively, not at midpoints of segments [A,B]
and [B,C] as in conventional finite difference, then the linear
equation of the curvatures obtained by joining these two points is
identical to equations (6.10). Clearly, the values of c 1 are c 2 are
key parameters in the accuracy of the solution; their dependence on
159

the value of d can be easily seen from equation 6.11. If d=1/3


then:

c1 9w 1 - 18w2 + 9w 3
c2 9w4 - 18w3 + 9w2

ifd=1I4 then:

32 16
--'WI - 16w Z + --'W3
3 3

32 16
--'W 4 - 16w3 + --'W 2
3 3

The estimates of the nodal curvatures predicted by both quadratic


and cubic models can be further improved by adopting an averaging
scheme. Average curvatures at every domain node are obtained from
the contribution of all domain cells adjacent to that node. A model
having the same order as that of the deflection model is then
adopted.

Figure 6.3 shows the w,xx curvature profile of the various models
along the width of a simply supported square plate uniformly loaded
in one direction. The curvature profile obtained from a single sine
wave deflection in both directions, is also shown. The improvement
in the curvature estimate by going to a higher order model and by
applying the averaging scheme can be clearly seen from these plots.

Another way of approximating the deflection is through simple


trigonometric functions. A smoother variation of the deflection
over the cell, and the same order of nonlinearity between
curvatures and deflections are thus achieved. Such models can be
built up using combinations of products Ii of the following
160

:Slne function
X :Ouadrallc model
+ :Ouadrallc model with averaging
*: Cubic model with averaging

Wxx '0

a x x
"" ""
CD
a + +
a
x x x

a
a
a~---.----r---'----.---'r---'----r---.'---.---~
0.00 0.20 0.40 0.60 0.80 1.00

Figure 6.3 Curvature variation for quadratic and cubic domain


models.

functions, defined in terms of a cartesian coordinate system X,Y


having its origin at a point on the boundary or in the domain of
the cell:

J1X

J1X

nY
F4 =sin
b

nY
FS=cos
b
161

where a and b are arbitrary constants that can be taken as the


plate width and breadth, respectively. The plate deflection would
be given by

Nn
w = LAi ti =<t>{A} (6.12)

i=l

where A.1 are the generalised coordinates of the model with {A}
their vector representation and <t> a single row matrix.

Vector {A} can thus be related to the nodal deflections vector {wi}
by an Nn xNn matrix equation:

from which

{A}=[Mj -1 {w.}
1

Equation (6.12) now becomes:

w=<t>[Mj -1 {w.}
1

and
-1
w Q=<t "'>[Mj {w.} (6.13)
,a.... ,aJ:S 1

where <t ...> contains the second derivatives of the functions t ..


,aJ:S 1

Rectangular plates in particular can be considered as a single cell


over which the deflection w is again given by equation (6.12) with
the functions ti chosen so as to satisfy the out-of-plane boundary
conditions of the plate. In other words, this is identical to the
162

trigonometric deflection model described in chapter 5. Accurate


representation of the deflections can be expected from this model
with a minimal number of domain unknowns A1..

Finally, whatever deflection model described above is adopted, a


linear matrix relationship can be obtained between curvatures and
deflections:

{w ,~~Q}=[Cw]{w} (6.14)

where Q} is a vector containing three curvatures at every


{w
,~~

domain node, {w} a vector containing domain nodes deflections and


[Cw] a known matrix that depends on the adopted deflection model.

6.4 Boundary Element Solution

The incremental boundary element solution described in this section


is derived along the same lines as the plane stress solution of
chapter 3 and the critical load solution in chapter 4. Linear
discontinuous boundary elements are used to model any boundary
unknown quantity. Treatment of corners and free boundaries is also
identical to that described in chapter 4. If a Gaussian integration
scheme is adopted with 4 points for boundary integration and 12
points for domain integration, then all integrals appearing in
equations (6.5) and (6.6) can be expressed in a matrix form
containing five one-dimensional arrays {b }, {b f }, {w}, {w Q} and
w ,~~

{n} representing respectively the boundary unknowns in the


deflection and stress integral equations, the domain deflections,
the domain curvatures and the membrane stresses.

4xN bn algebraic equations can thus be obtained by placing the


source point on all Nbn boundary nodes and evaluating the integrals
of equations (6.5) and (6.6) for both fundamental solutions u1 and
u 2 . The general form of these equations is:
163

(6.16)

where A is the load factor generating the current load from a


reference stage and 6A its increment. The coefficient arrays on the
left-hand side of equations (6.15) and (6.16) are all fixed and
need to be calculated only once throughout the incremental solution
process. These matrices constitute the linear part of the problem.
Matrices {E d1 } and {E d2 } contain the second-order incremental
deflection and curvature terms, respectively.

Nd algebraic equations are also obtained by placing the source


point on the Ndn domain nodes and applying equation (6.5) for the
fundamental solution u 1 with c=1. This operation results in the
following system of equations:

(6.17)

In the case of the single-cell domain model, the source point is


placed on Nn peak points of the respective trigonometric functions
~ .. Again, matrices on the left-hand side of equation (6.17) are
1
fixed and matrix {E d1 } contains second-order deflection terms.

Integral expressions for the membrane stresses are obtained by


applying relations (6.4) and equation (6.6) with c=1. The number of
nodes at which the membrane stresses are determined depends on the
desired accuracy. The following matrix equation can be obtained:

(6.18)

where [BZ1 and {Eb} are fixed, {E d2 } contains second-order terms of

curvatures.
164

The following correspondence exists between integrals of equations


(6.5) and (6.6) and matrices of equations (6.15) to (6.18):

[Olw] and [Oiw]: from term wN~~ in I~j and free boundary terms of

I~j and Il.

{E t } and {E t'}: from terms t Vt u. and Vtt u. in I~,


~ ,~J ~J,~ J

[B 2 ] and [B 2]: from the first two terms of I b2j ,

{E b } and {Eb}: from the last two terms of I b2j ,

d 2
[02f] and [02'f]: from integral I 2J. excluding (w ,xy ) -w ,xxw,yy

2 d
(w ,xy ) -w ,xxw,yy in I 2J.,
165

6.5 Solution of the Systems of Equations

The systems of equations (6.15), (6.16), (6.17) and (6.18) are


solved for every load increment. While matrices [B 1 ], [B t ], [B 2 ],
[E b ], [Bi]' [B~], [B Z] and {E b} are evaluated once throughout the

solution, new values of matrices [Olw]' [Oln]'


[Oiw]' [Din]' {E~} and [OZf] are found at every new load step.

The four systems of equations are first solved without including


the second-order terms represented by matrices {E d1 }, {E d2 }, {E d1 }

and {E d2 }. The calculated {w} and {n} are then used to evaluate

these matrices and solve the four systems of equations iteratively.


Writing the four systems without including second-order terms, and
replacing {w ,~~Q} by its expression in equation (6.14), we obtain
the following systems of equations:

([B 1 ]+(A+6A)[B t ]){bw} [Olw]{w}+[Oln]{n}+(E t } (a)

[B 2 ]{b f } - 6A(E b} [02f][Cw]{w} (b)


(6.19)
([Bi]+(A+6A)[B~]){bw} [Oiw]{w}+[Oin]{n}+(E~} (c)

{n} [BZ]{bf}+[OZf] [C w] {w}+6A(E b} (d)

Inverting [B 2 ] in equation (6.19b) we obtain:

(6.20)

Now replacing the obtained expression of {b f } in equation (6.19d),


we get:

(6.21)
166

where

Combining equations (6.19a) and (6.21), we obtain:


(6.22)
where

Finally combining (6.19c), (6.21) and (6.22) yields:

(6.23)

from which:

where
167

Incremental deflections {w} are thus obtained from equation (6.23).


Incremental curvatures {w ,at",Q}' stresses {n} and boundary unknowns
{bw} can then be found by substituting into equations (6.14),
(6.21) and (6.22), respectively. While matrix [B 2 ] is fixed and is
only inverted once throughout the solution, matrices [MS] (NbxN b )
and [M 6 ] (NdxNd ) vary from one load step to another and have to be
inverted once at each load increment. Once a first estimate of the
unknowns vectors had been found, the matrices {E d1 }, {Ed2 }, {E d1 },

and {E d2 } are assembled. The systems of equations are then solved

again, taking into consideration these second-order terms; the


expressions of incremental vectors {n}, {b } and {w} thus become:
w

{n} [M 1 ]{w} + 6:>"{M 2} + {Sl} (a)

{bw} = [M 3 ] {w} + {M 4 } + {S2} (b) (6.24)

[M 6 ] {w} {M 7} + {S3} (c)

where

{Sl} [B ZHB 2 ] -1 {E d2 } + {Edz}

{S2} [MS] ([D 1n ] {S2} + {Ed1 })

{S3} {E t } + [Din]{Sl} - [MaJ{S2}

Vectors {Sl}' {S2} and {S3} are evaluated and a new deflections
vector is calculated from equation (6.24c); the iteration continues
until the required accuracy is attained. Notice that no matrix
inversion is required in this iteration process since the inclusion
of the second order terms only affects the right hand side of
equation (6.23).
168

6.6 Numerical Implementation

The proposed solution has been implemented in a computer program


LARDEF developed as a combination and extension of programs PLASTR
and PLABEM described in chapters 3 and 4. The program performs
essentially three tasks: the data input and processing, the
assembly of the linear (fixed) matrices, the incremental solution.
In addition to the usual material and geometrical plate properties,
as well as loading and discretization data, the program requires
control data for the incremental solution: initial deviation from
flatness, size of incremental load and error tolerance when
second-order terms are included. Once the domain mesh has been
generated, the dependence of curvatures on deflections is
established as described in section 6.3. As a result, a matrix
relating all nodal curvatures to all nodal deflections is obtained.
Matrices that do not change as the load is incremented are then
evaluated in the second part of the program. Finally, the nonlinear
incremental solution is started. At each load increment, boundary
integration is first performed and vectors {E t }, {E'} are obtained.
t

Next, the integration over the domain is carried out and matrices
[D lw ]' [D ln ], [D 2f ], [Diw]' [Din] and [Dif] are evaluated. The

system of equations is solved and iteration matrices {E dl }, {E d2 },


{E dl }, and {E d2 } assembled. Once the error of iteration is small

enough, results are written into an output file and current state
matrices of deflection W, membrane stresses N and boundary unknowns
are incremented by the new solution and the next load step started.
In the first load increment, current state membrane stresses, and
consequently matrices [D lw ] and [Diw] , are identically zero. Non-

are caused by the initial deviation which initiates the solution. A


flow chart summarising these steps is shown in figure 6.4.
169

Read data and find curva-


tures matrix
~
Assemble fixed matrices of
boundary and domain source
equations

'"
Perform matrix operations
with fixed matrices
-.v
, Start load increments I
< t
Assemble boundary and do-
main source equations

Solve primary systems of


equations

Start iterative procedure

rI r :,sSeOclVoned-orde~terms
I I' : Assemble matrices .ith I
I
!

systems of equations'

~ I
I
Next i~era tion I

Figure 6.4 Flow chart for large deflections computer


program, LARDEF.
170

6.7 Results

A set of experimental, numerical and analytical results has been


gathered from existing literature on the subject in order to
compare them to results obtained from the program described above.

Table 6.1 summarises the properties of the various plates selected


for comparison. In some cases the initial imperfection was not
given in the papers; it had therefore to be estimated by fitting
the initial experimental and theoretical data.

Table 6.1 Plate dimensions, material properties and initial


deviations.

Plate dimensions(mm)
Reference Fig. Young's Poisson Initial
Length Width Thick. Modul~s Ratio Deviation
(GN/m ) (mm)

Hoff et al 6.5 508 508 6.35 18 0.25 0.349


(1948)
Uemura & Byon 6.6 300 300 2.00 68.7 0.30 0.100
(1978)
Uemura & Byon 6.7 300 300 2.00 68.7 0.30 0.160
(1978)

Stein 6.8 644 120 1.83 73. 0.33 0.055


(1959)
Walker 6.9 508 254 1.2 227. 0.25 0.088
(1964)
Rhodes et al 6.10 508 254 1. 70 205. 0.30 1.700
(1975)
Yamaki (II) 6.11 300 300 0.99 68. 0.31 0.099
(1961)

Yamaki (III) 6.11 300 300 1.24 68.2 0.33 0.124


(1961)
171

In all the above examples involving square plates, the boundary is


discretized into 16 elements and the domain into 32 cells. These
numbers are, respectively, 20 and 24, for the rectangular plates of
figure 6.8 and 6.9, 24 and 36 for that in figure 6.10. The number
of load steps considered varied from 15 to 25 depending on the
length of the experimental and analytical records. YO indicates the
initial deviation from flatness at the crest of a buckle, usually
at the centre and t the thickness of the plate. £b is the bending
strain and £c and Pc are the critical bending strain and buckling
load, respectively.

The various deflection models were first tested by comparing their


predictions with experimental results for a simply supported square
plate uniformly loaded in one direction and having uniform membrane
stress distribution (Hoff et aI, 1948). Results are plotted in
figure 6.5. As expected, the quadratic deflection model does not
perform particularly well. Its predictions deviate from the
experimental curve in the region of the buckling load. A slight
improvement is however achieved by the use of the averaging scheme.
The deflections obtained by the trigonometric deflection model are
smaller than the experimental results up to buckling, and larger
thereafter. Deviations from experimental curve also occur as the
cubic model is introduced with d=1/3. Highly accurate results are
however obtained when d=1/4. This is due to the fact that
curvature-deflection relations corresponding to the latter value of
d reproduce more accurately the actual curvatures of the plate. The
single-cell model, on the other hand, clearly gives accurate
results with only one deflection term, until a loading stage is
reached where two terms are needed to achieve the same accuracy.

Experimental and finite element results were given by Uemura and


Byon (1978) for duraluminium clamped square plates uniformly loaded
in one direction with zero in-plane displacement along the edges.
Results for two plates with different initial deviations are shown
172

in figures 6.6 and 6.7. The BEM clearly gives accurate results in
both cases. In figure 6.6, the BEM appears to be slightly more
accurate than the FEM in the region of the buckling load while the
opposite is true in the post-buckling region. The cubic deflection
model proves accurate for d=O.2 compared to d=O.25 in the previous
example. The deviation of the BEM curve from the experimental
results towards the end of the record is most probably due to the
difference in in-plane boundary conditions.

In all the following examples, results involving rectangular plates


using the single-cell model are shown along with other experimental
and numerical results.

Experimental and analytical results for a simply supported


rectangular plate with an aspect ratio of a/b=5.38 uniformly
compressed in one direction along its short edges were given by
Stein (1959). The analysis proceeded from a power series
approximation of the deflection to derive a loading profile
without, however, taking into account the initial deviations from
flatness. These results along with two BEM loading profiles
corresponding to the 5 and 6 half-wave modes are shown in figure
6.8. Very good agreement between BEM and experimental results is
clearly observed for both deflection modes.

Experiments presented by Yalker (1964) and Rhodes et al (1975)


provided the opportunity to test the BEM program for non-uniform
membrane stress distribution and free boundary. Tests on
eccentrically loaded plates were presented by Yalker that were
subsequently analysed in the paper by Rhodes
et al by an energy
method using a functional deflection representation. Experimental,
analytical and BEM results for one such plate loaded with an
eccentricity e=b/18 are shown in figure 6.9. The loading profile
obtained by the BEM is again in good agreement with both analytical
and experimental curves.
173

o
o
'0

o
o
..t

o
o
N

o
o

o
o
ID

o
o
'0

Cubic Model. d=1/4


o
o Cubic Model. d=1/3
..t Trigonometric. SIngle Cell.1 term
Trigonometric. Single Cell.2 terms
+: QuadratIc Model
o X : QuadratIc Averaged Model
o
N y : TrIgonometric. TrIangular Cell
*: Experiments. Hoff et al
o
o
o~--~---'----r---~--~---r--~--~r---r---'----r--~--~
0.00 0.16 0.32 0.48 0.64 0.80 0.96

W-Wo (In)

Figure 6.5 Applied load P versus net centre plate deflection for
various deflection models compared to experimental
results of Hoff et al.
174

o
o

_ ,BEM. SIngle Cell. 2 deflectIon terms


o -+- ,FEM. Uemuro. & Byon
o
X ,ExperIment. Uemuro. & Byon

o
o
o~--'---'---'---'---'---~--~--T---~~
0.00 0.40 0.80 1.20 1.60 2.00

CW-Wo ]/h

Figure 6.6 Axial compressive stress versus ratio of net centre


deflection to thickness of plate as given by BEM and,
FEM and experiment of Uemura & Byon.

~D~
0
p ~
r<1
(MgJ

0
"': >+- /'
N
/'"
./
~
0
~

0
0: :BEM. SIngle Ce II. 2 deflectIon terms
0
:BEM. SIngle Ce II. 1 deflectIon term
+ :BEM. CubIc Mode I. d=O.2
0 - ,ExperIment. Uemura & Byon
C!
0
0.00 0.80 1.60 2.40 3.20 4.00

W-Wo (mm)

Figure 6.7 Applied load P versus net centre plate deflection as


given by BEM and experiment presented by Uemura & Byon.
175

0
~
P N

Pc
0
"?

o
N

o
CD
o

o --:BEM. 6 half-wave mode shape


..... _ :BEM. 5 half-wave mode shape
o
-+-:Analytlcal. 5 buckles. Stein
X :Experlment. 5 buckles. Stein
o ~ ,Experiment. 6 buckles. Stein
o
o1---'---'---'---'---'---~--~--~--'---'---'---'--'
0.00 0.80 1.60 2.40 3.20 4.00 4.80
€b/€c
Figure 6.8 Maximum bending strain as given by BEM and, analysis and
experiment presented by Stein.
o
o
Pia

o
o
CD

o
o
.....
-*-..BEM. Single eel I. 2 deflection
Ana I y II cal. Rhodes e al
~,Experlment.h=O.049In.
l
Walker
terms

X ,Experlment.h-O.042In. Walker
y ,Experlment.h=O.078In. Walker
o ~ .Experlment.h=O.059In. Walker
o
o~--,---,---,---,---,---~--.---.---~--~--~--,
0.00 0.40 0.80 1.20 1.60 2.00 2.40

[W-Wo]/h

Figure 6.9 Axial compressive stress versus ratio of net centre


deflection to thickness as given by BEM, experiment
presented by Valker and analysis given by Rhodes et al.
176

o
N

a
x
x
o x
(D

o
....
o
- 'BEM. Single Cell. 4 terms
-+-, Ana I yt I co. I. Rhodes et a I
o X ,Experiment. Rhodes et al
o
o·~--.---.---,---,---,---,---.---.---.---.
0.00 1.00 2.00 3.00 4.00 5.00

[W-Wo J/h

Figure 6.10 Axial compressive stress versus ratio of net centre


deflection to thickness of plate as given by BEM and,
experiment and analysis presented by Rhodes et al.

o
P

~ 101 ~
'0

(xI0 6 1bs)- x

o
N
Case I I
/
/
/ +
o +
(D
+
o ~ ++
+
+ Case I I I
/ ++
o
.... xy +
++ --:BEM. Single Cell. Case I I
o + _ ,BEM. Single Cel I. Case I II
-r :Experlments. Yamakl. Case II
o
X ,Experiments. Yamakl. Case II I
o.~__~__~__, -__, -__, -__, -__. -__. -__. -__. -__. - - - .

0 0 . 00 0.04 0.08 O. 12 0.16 0.20 0.24


W-Wo (1 n )

Figure 6.11 Comparison of load-maximum deflection curves obtained


by BEM and the experiments presented by Yamaki.
177

Another set of tests was presented by Rhodes et al where steel


plates with one boundary edge free to move out of plane were
eccentrically compressed until failure. Results for a plate loaded
with an eccentricity e=b/10 are shown in figure 6.10 along with the
analytical and BEM predictions. While good agreement is observed
for the first part of the curve, both analytical and BEM curves
clearly deviate from the experimental one at the final stage of
loading. Non-elastic behaviour of the tested plate prior to
collapse, is a possible explanation of the observed discrepancy.
The stiffer response of the BEM model as compared to the analytical
results towards the end of the loading curve may be due either to
the difference in in-plane boundary conditions or to inadequate
representation of the deflection. In the latter case, the
introduction of different functions ~i may improve the results.

The versatility of the BEM formulation to deal with a variety of


out-of-plane boundary conditions was tested through a comparison
with results presented by Yamaki (1961). The results of two cases
from the four originally considered by Yamaki are reproduced in
figure 6.11 together with the BEM predictions. The two plate
examples have two simply supported and two clamped edges. Uniform
compression is applied along the simply supported edges in case II
and along the clamped edges in case III. The comparison of the
results shows that the proposed solution is not only effective in
this respect, but is also very accurate, particularly in the
support case III. The discrepancy between theory and test case II
has been attributed by Yamaki himself to the condition of free
rotation along the loaded edges not fully satisfied.

6.8 Conclusions

The validity and efficiency of the boundary element method as


applied to the problem of imperfect plates loaded in their own
178

plane has been demonstrated. Various high-order elements for the


domain deflections were tried. The deflections obtained by the
quadratic and trigonometric models are inaccurate in the
post-buckling region while very accurate results were obtained when
the cubic and single-cell trigonometric models were introduced.
However, an optimum value of distance d of internal nodes over the
edge of a cubic cell remains to be established. The single cell
deflection model, for rectangular plates, gives very accurate
results with a small number of domain unknowns. The versatility of
the proposed solution have also been demonstrated as various
support and loading conditions have been tried. In addition, the
method proved efficient in simulating higher deflection modes
resulting from secondary buckling.
CHAPTER SEVEN

CONCLUSIONS

Since the boundary element method has emerged, in the late


seventies, as a powerful numerical technique endowed with the
capability of solving a wide range of engineering problems, it has
been the focus of interest among researchers, design engineers and
software developers alike. The 'promise' of a by one degree
reduction in the size of a problem with all the consequent
economies in computation, time and money is perhaps behind all the
energy invested in developing the method.

The ever-increasing demand in computer aided design environments


for yet faster, more accurate and user friendly computer codes
renders the development and introduction of new numerical methods
of prime importance. As the design engineer may require a multitude
of identical, repetitive analyses before the final optimum product
is reached, a seemingly slight amelioration in the performance and
data simplicity of a single analysis may well be amplified into
substantial economies and quality improvements.

In the previous chapters, the focus has been on four main problems
of plates loaded in their own plane: the calculation of the
membrane stress distribution, the prediction of the critical
buckling factor, the transformation of domain integrals into
boundary ones and the solution of the large deflection problem of
imperfect plates.

A boundary integral equation in terms of the stress function based


on the Rayleigh-Green identity, has been used to obtain the
membrane stress distribution of the problem. The proposed solution
can be used for any plate geometry and loading. Vhile in-plane
180

displacement boundary conditions could not be easily expressed in


terms of the stress function, the solution nevertheless gave
accurate results even when extended to the non-linear problem where
displacement boundary conditions were replaced by the stress
conditions that most closely approximate them.

The elimination of the curvatures terms from the boundary integral


equation containing membrane stresses resulted in an efficient
implementation of the boundary element solution for the prediction
of the critical buckling load. A systematic parametric study was
performed to establish the optimum boundary and domain element
models. An automatic domain mesh generation scheme reduced the
effort required by the program user to discretize the plate.
Moreover, no additional equations were required to determine the
buckled shape of the plate as the domain deflections are directly
obtained when the system of equations is solved. Accurate results
were obtained with a relatively small number of system unknowns for
various plates geometries, support and loading conditions. The
performance of the BEM also compared favourably with that of a
finite element program.

An alternative dual-reciprocity method was also proposed where


expressions containing curvatures terms are transformed into a
Fourier series and the resulting domain integrals transformed to
the boundary using the Rayleigh-Green reciprocity equation. Two
deflection models were tested and both gave accurate results in the
case of rectangular plates with various support and loading
conditions. Although the solution is based on the integral equation
containing curvatures terms, the final system of domain unknowns
only contained either deflections or unknown coefficients
representing deflections.

An incremental boundary element solution of the large deflections


problem of imperfect plates proceeded along lines similar to the
181

critical buckling load solutions. In addition, curvatures terms


arising from the non-linear strain-displacement relations were
expressed in terms of the deflections, using high order deflection
models and thus avoiding the costly inclusion of these curvatures
in the final system of unknowns. The results accuracy was improved
further by including the second-order terms of deflections,
stresses and curvatures, iteratively, in the solution. Various
deflection models were tested and accurate results were obtained
when the cubic and single cell trigonometric models were used. The
proposed solution was actually found to be highly accurate in
simulating the non-linear behaviour of the plate when its
predictions were compared with experimental results obtained from
existing literature. Various cases of loading, boundary conditions
and geometries were tested with equal success.

In all of these problems, the modelling of the domain deflections


appears to be a key factor in the accuracy and numerical efficiency
of the solutions. Various domain deflection models were described
and tested namely triangular finite elements using polynomial
interpolation functions or trigonometric ones, Fourier series
models transforming a distribution of nodal deflections into a
continuous trigonometric series, and generalised functions models
applied to rectangular plates. The latter appears to be very
efficient, being understood that it only applies to rectangular
plates. Accurate critical buckling loads were obtained when linear
discontinuous domain cells and boundary elements were used; the
latter being particularly efficient in modelling corner
discontinuities both geometrical due to the duality of the normal
to the boundary, and physical resulting from the jump of the
transverse moment.

A special treatment of the free boundary unknowns expressing them


in terms of domain deflection models also yields to good results
with the additional advantage of avoiding cumbersome numerical
182

singularities. Good results were in fact obtained, sometimes with a


remarkably small number of unknowns.

In addition to the highly accurate results obtained by the BEM


through the various formulations described in the previous
chapters, the method was shown to be highly versatile in its
ability to solve plate problems with various shapes, support
conditions and loading configurations with, more or less, equal
accuracy. Free boundaries, concentrated loads and non-rectangular
plates were all successfully analysed by the method.

A question immediately arises when one attempts to solve the plate


stability problem by the BEM: why using such a solution when the
problem cannot be formulated with purely boundary integral
equations? Or, more to the point, why would a design engineer be
invited to use a BEM package for plate problems when he has to
discretize both the boundary and the domain; is it not the case, as
some might say, that the BEM has 'increased' in a way the dimension
of the problem instead of 'reducing it by one'?

The case for solving plate buckling problems by the boundary


element method, despite the domain discretization usually required,
is partially justified by the accurate results obtained using the
dual-reciprocity method and by the automatic domain mesh
discretization. Such a scheme is all the more efficient since
computer hardware industry is continuously heading towards a
cheaper memory cost, a faster execution and an easier data
handling. Only partially because the real advantage of the boundary
element method, as I hope has been demonstrated by the results
described in the previous chapters, resides in its potential for
accuracy. Most of the functional approximations of physical
quantities are made on the boundary rather than the domain because
exact fundamental solutions to the problem are being used; only one
degree of freedom, the deflection w, per domain node had to be
183

taken compared to six used in conforming shell elements of finite


elements solutions. On the other hand, although not considered
here, infinite and semi-infinite plates as well as plates with
stress concentrations around holes can be very accurately modelled
by the method. The potential accuracy and versatility of the BEM
are simply too powerful to be overshadowed by the mere existence of
the domain integrals. In addition, results from any numerical or
analytical technique have to be confirmed by comparison with
results from other methods and thus justifies the introduction of
new techniques having the data simplicity of the BEM.

For the reasons aforementioned, I believe that the real challenges


facing the boundary element method applied to plate problems, do
not concern its validity or its efficiency as much as the numerical
problems associated with singular and near-singular integrals, the
treatment of domain integrals, the development of possibly more
efficient fundamental solutions and/or boundary integral equations
and, finally, the extension of the method to other fields of plate
stability problems.

Yhile highly accurate physical modelling is obtained by the BEM,


some of the accuracy is however lost to the numerical problems
associated with singular or near-singular integrals. Efficient
analytical methods have been proposed to deal with singular
integrals that occur when the source point is in the field element.
Near-singular integrals, however, still require substantial
computational effort, and, even though good accuracy is achieved,
large amounts of CPU times are consumed.

The dual-reciprocity technique proposed in chapter 5 did result in


a boundary element solution that required no domain discretization
at all. Further developments can however be made by using more
general deflection models or improving the efficiency of the nodal
deflection model by reducing the number of fictitious nodes of the
184

plate that is the nodes outside the actual domain of the plate,
and by devising a more rigorous way of applying boundary conditions
to that model.

The treatment of the domain integral which eliminated the


curvatures and replaced them by deflections in chapter 4
(Syngellakis and Kang 1987), demonstrates how the manipulation of
the integral equations can considerably increase the efficiency of
BEM. Likewise, the use of more efficient fundamental solutions or
different forms of integral equations may do away with domain
integrals altogether.

The non-linear solution presented in chapter 6 can also be further


improved by formulating displacement in-plane boundary conditions.
The introduction of material non-linearities would also widen the
scope the method. Finally, the extension of the formulation to deal
with plates with holes and stiffened plates is of considerable
practical importance.
Appendix A

THE GREEN'S IDENTITIES

A.l The two-dimensional Green's identity

Given two continuous and twice differentiable functions P(x,y) and


Q(x,y) defined in terms of a cartesian coordinates sytem over
domain Q and its boundary r, with continuous first and second
derivatives, the two-dimensional Green's theorem can be written as
follows:

Ip(X,Y)dX + Q(x,y)dy I (_a_Q_:_:'-y-)-


ap(x,y) )
--dQ
ay
r Q

A.2 The Generalised Rayleigh-Green identity for plates

Given a plate with domain Q and boundary r defined in terms of a


cartesian coordinates system, the strain energy density can be
written as follows:

a2w a2u
E(w,u)= D 2----+
axay axay

where w is the plate deflection, u a weighting function, D the


plate rigidity and ~ the Poisson ratio of the plates's material.
Integrating the above expression by parts twice, using equations
that express the equivalent shear Vn , bending moment Mn and
twisting moment Mns in terms of w or u, and accounting for corner
points of the plate we obtain:
186

- a: LU
+
E(w,u)= D J uV4w d!i2 - J (UVn(W)
a ]
ttn(w) dr -
c
[Mns(W)
!i2 !i2 i=l

- a:
+
E(u,w)= D J wv4u d!i2 - J (wVn(U)
a ] c
ttn(u) dr - LW (Mns(U)
!i2 !i2 i=1

where Nc is the number of corners, n the direction normal to


boundary and s the boundary tangential direction. The term
+
(M is the corner jump of the twisting moment. Substracting the
~ ns
the last two equations from each other, we obtain the generalised
Rayleigh-Green identity for plates:

o J[uv"v -
!i2
vv"u ]dO = Lh (v) - :: Hn (v) - vVn (u) + :: Hn (u) ]dr
Appendix H

FUNCTIONS OF THE FUNDAMENTAL SOLUTIONS

In what follows:

r is the magnitude of vector PO where P is the source point and 0


the field point.

a is the counter-clockwise angle from vector PO to the normal to


the boundary at P.

e is the counter-clockwise angle from the x-axis to PO.

~ is the counter-clockwise angle from vector PO to the normal to


the boundary at O.

~ is the counter clockwise angle from the x-axis to the normal to


the boundary at O.

(u 2 is only used when source point P is on the boundary).

H.1 Fundamental Solution u1


r2
ln r
8nD

aU
1 r
--- (2ln r + 1) cos~
an 8nD

aU 1 r
--= (21n r + 1) sin~
as 8nD
188

1 1-v
Vn (u 1 )= - [2+ (l-v)cos2~1 cos~ + -- cos2~
4nr 4nr

l+v I-v
Mn (u 1 )= - (In r + 1) - - - cos2/3
4n 8n

I-v
M (u 1 )= sin2/3
ns
an

1
iU 1 (1 + In r)
2nD

an 2nDr

B.2 Fundamental Solution u 2

r
- - - (21n r + l)cos 0:
8nD

1 1
--= - - - (21n r + 1)cos(j3-0:) - - - cosl3 coso:
an 8nD 4nD

1
(2ln r + l)sin(j3-o:) + - - coso: sinl3
as anD 4nD

1 1-v I-v
2 sino:cos/3sin 213
Vn (u 2 )=---2 [2+( 1-v)cos2131 cos (13+0:)- -----zsin2/3sino:+--
4~ 2~ 2~

l+v I-v
Mn (u 2 )= - - coso: + - - sin213 sino:
4nr 4nr
189

1-'\)
Mns (u 2 )= - - sinet cos2/3
4nr

COSet
v'lu 2
2nDr

av'lu 2 cos( /3+et)


an 2nD/

B.3 Fundamental Solution Uxx

cos 2 9
4nD

au xx 1
[cos /3 - sin /3 sin 29)
an 4nDr

cos 29
v'lu xx
2nDr2

av'lu xx cos (29-/3)


an nDr 3

B.4 Fundamental Solution uxy

sin 29

anD

sin /3 cos 29
4nDr
190

sin 28
v'luxy
2nDr2

av'lu sin (2EHn


x;l
an nDr 3

8.5 Fundamental Solution uyy

1 ( ) sin 2 e
8 nD 2ln r + 1 + -4-nD--

1 (cos 13 + sin 13 sin 29)


4nDr

cos 28
2nDr2

cos (29-13)
nDr 3
Appendix C
TRIGONOMETRIC DEFLECTION FUNCTION5

In what follows:

s is the coordinate axis (either x or y)

L is the dimension of plate in direction s (either a or b)

k=2m-l

z=(-l) m-l (z=-l for even m, z=1 for odd m)

1. 55-55 :

Boundary Conditions: f(O)=O.,f"(O)=O.,f(L)=O.,f"(L):O.

mns
f(s)= sin (m=1,2,3, ... )
L

2. 55-CL :
Boundary Conditions: f(O)=O.,f"(O)=O.,f(L)=O.,f'(L)=O.

ns mns
f(s)= cos - - sin (m=1,2,3, ... )
2L L

3. 55-FR :

Boundary Conditions: f(O)=O.,f"(O)=O.,f"(L)=O.,f"'(L)=O.

2 [kns kns kns


f(s)=- 3 + cos-- + O.5z sin-- - 4 cos-- + 4z sin
17 L L 2L
~]
2L
(m=1,2,3, •.• )
(k=1,3,5, •.• )
192

4. CL-SS :
Boundary Conditions: f(O)=O.,f'(O)=O.,f(L)=O.,f"(L)=O.

ns mns
f(s)= sin - - sin
L L (m=1,2,3, ... )

5. CL-CL :
Boundary Conditions: f(O)=O.,f'(O)=O.,f(L)=O.,f'(L)=O.

f(s)=
1 (m-1)ns (m+1)ns
-cos-
1
2 L L (m=1,2,3, ... )

6. CL-FR :
Boundary Conditions: f(O)=O.,f'(O)=O.,f"(L)=O.,f"'(L)=O.

2 [17
f(s)=- -
1 kns z kns kns kns
- - c o s - - + --sin-- - 4cos-- - z sin--
1
17 4 4 L 2 L 2L 2L
(m=1,2,3, ... )
(k=1,2,3, ... )

7. FR-SS :
Boundary Conditions: f"(O)=O.,f"'(O)=O.,f(L)=O.,f"(L)=O.

f(S)=-=-[3 _ cos kns + ~in kns + 4cos


kns
kns
4zsin - - 1
17 L 2 L 2L 2L
(m=1,2,3, ... )
(k=1,3,S, ..• )
193

8. FR-CL :

Boundary Conditions: f"(O)=O.,f"'(O)=O.,f(L)=O.,f'(L)=O.

1 kns _kns ]
f(s)= [ 1 _ z (13 _ cos ) sin
2 14 L 2L (m=1,2,3, ... )
(k=1, 3,5, ... )

9. FR-FR :

Boundary Conditions: f"(O)=O.,f"'(O)=O.,f"(L)=O.,f"'(L)=O.

For m=O

f(s)=1.

For m=1,3,5, ... i.e. k=1,5,9, ...

1 [1 kns kns kns 1 kns knS]


f(s)=- -(2cos- - sin-) + sin- - -(4 + cos-)cos-
3 8 L L 2L 9 L 2L

For m=2,4,6, ... i.e. k=3,7,11, ...

1 [1 kns kns kns 1 kns knS]


f(s)=- -(2cos- - sin-) + sin- - - ( 9 - cos-)cos-
3 8 L L 2L 4 L 2L
REFERENCES

Alfutov, N.A. & Balabukh, L.I. (1967); "On the Possibility of


Solving Plate Stability Problems without a Preliminary
Determination of the Initial State of Stress", PMM Vol. 31, No 4,
pp.716-722

Berger, H.M. (1955); "A New Approach to an Analysis of Large


Deflections of Plates", J. appl. mech., ASME Vol. 22, pp.465.

Bezine, G. (1983); "Etude du Flambage des Plaques Minces par la


Methode des Equations Integrales de Frontiere", IIIe Congres
International sur les Methodes Numeriques, Pluralis, Paris,
pp.723-729.

Brebbia, C.A. (1980); "The Boundary Element Method for Engineers",


Pentech Press, 2nd edition.

Brebbia, C.A. and Nardini, D. (1982); "Dynamic Analysis in Solid


Mechanics by an Alternative Boundary Element Procedure", Int. J.
Soil Dynamics and Earthquake Eng'g, Vol. 2, pp.228.

Budiansky, B. (1974); "Theory of Buckling and Post-Buckling


Behaviour of Elastic Structures", Advances in Applied Mech., C.S.
Yih Ed., Vol.14, Academic Press, NY, 1974, pp. 1-65.

Bulson, P.S. (1970); "The Stability of Flat Plates", Chatto &


Vindus, London.

Clebsh, A. (1862); "Theorie de Elasticitat fester Korper", Leipzig,


B.G. Teubner.
195

Costa Jr., J.A. and Brebbia C.A. (1985); "Elastic Buckling of


Plates using the Boundary Element Method", Proceedings of the 7th
International Conference on Boundary Element Methods, Villa Olmo,
Italy, pp. 429-442, SpringIer Verlag.

Cruse, T.A. (1969); "Numerical Solutions in Three Dimensional


Elastostatics", Int. J. of Solids Struc., Vol. 5, pp.1259-1274.

Elzein, A. and Syngellakis, S. (1989); "High-Order Elements for the


boundary element method stability analysis of imperfect plates",
Advances in Boundary Elements, Proc. of the 11th Int. Conf. on
Boundary Element Methods, Vol. 3, Ed. Brebbia C.A. and Connor J.J.,
pp.269-284.

Dym, C.L. and Shames, LH. (1973); "Solid Mechanics. A Variational


Approach", McGraw Hill.

Fredholm, I. (1903); "Sur une Classe d'Equations Fonctionelles",


Acta Math., Vol. 27, pp.365-390.

Hambly, E.C. (1976); "Bridge Deck Behaviour", Chapman & Hall.

Hoff, N.J. (1938); "Instability of Monocoque Structures in Pure


Bending", Journal of the Royal Aeronautical Society, Vol. 42,
pp.291-346.

Hoff, N.J., Boley, B.A. and Coan, J.M. (1948); "The Development of
a Technique for Testing Stiff Panels in Edgewise Compression",
Proc. Soc. Exper. Stress Analysis, Vol. 5, pp.68-74, 1948.

Jagger, L.G. (1980); "Elementary Theory of Elastic Plates",


Pergamon Press.

Jaswon, M.A. and Symm, G.T. (1977); "Integral Equation Methods in


Potential Theory and Elastostatics", Academic Press, London.
196

Kamiya, N. and Sawaki, Y. (1982); "An Integral Equation Approach to


Finite Deflection of Elastic Plates", J. Non-Linear Mechanics, Vol.
17, pp.187-194.

Kamiya, N. and Sawaki, Y. (1988); "The Plate Bending Analysis by


the Dual-Reciprocity Boundary Elements", Engineering Analysis, Vol.
5, No 1, pp.36-40.

Karman, T. (1910); "Festigkei t im Maschinenbau", Encykl. Math.


Wiss., Vol. 4' pp.311.

Katsikadelis, J.T. and Nerantzaki, M.S. (1988); "Large Deflections


of Thin Plates by the Boundary Element Method", Boundary Elements
X, Ed. Brebbia C.A., pp.435-456, SpringIer Verlag.

Lord Kelvin and Tait, P.G. (1883); "Treatise on Natural


Philosophy", Vol. 1, part 2, pp. 203.

Khan, M.Z. and Walker, A.C. (1972); "Buckling of Plates Subjected


to Localised Edge Loading", The Structural Engineer, No 6, Vol. 50,
pp.225-232.

Khan, M. Z., Kenneth, C.J. and Hayman, B. (1977); "Buckling of


Plates with Partially Loaded Edges", J. of Struc. Division,
pp.547-557.

Kirchhoff, G.R. (1850); "Uber das Gleichgewicht und die Bewegung


einer elastischen Scheibe, J. Math., Crelle, Vol. 40, pp. 51-58.

Kirchhoff, G.R. (1877), "Mechanik", 2nd edition.

Kupradze, O.D. (1965); "Potential Methods in the Theory of


Elasticity", Daniel Davey & Co., NY.
197

Kutt, H.R. (1975); "Quadrature Formulae for Finite-Part Integrals",


National Research Institute for Mathematical Sciences. CSIR Special
Report YISK 178, Pretoria, September.

Mikhlin, S.G. (1957); "Integral Equations", Pergamon Press, NY.

Mindlin, R.D. (1951); "Influence of Rotary Inertia and Shear on


Flexural Motions of Isotropic Elastic Plates", J. Appl. Mech. 18,
pp.31-38.

Muskhelishvili, N.I (1953); "Some Basic Problems of the


Mathematical Theory of Elasticity", P. Nordhoff Ltd, Groningen.

Niwa, Y. Kobayashi, S. and Fukui, T. (1974); "An Application of the


Integral Equation Method to Plate Bending", Memo, Faculty of
Engineering, Kyoto University, Japan, Vol. 36, Part 2, pp.140-158.

Nowinski, J.L. and Ohnabe, H. (1958); "On Certain Inconsistencies


in Berger Equations for Large Deflections of Elastic Plates", Int.
J. Mech. Sc., Vol. 14, pp.165.

Poisson, S.D. (1829); "Memoire sur l' Equilibre et Ie Mouvement des


Corps Solides", Vol. 12, part 8.

Reissner, E. (1944); "On the Theory of Bending of Elastic Plates",


J. Math. and Phys., Vol 23, pp. 184.

Reissner, E. (1985); "Reflections on the Theory of Elastic Plates",


Appl. Mech. Rev., Vol. 38, no 11, pp.1453-1464.

Richards, T.H. (1977); "Energy Methods in Stress Analysis", Ellis


Horwood Series in Eng'g Science.
198

Rhodes, J., Harvey, J.M. and Fok, V.C. (1975); "The Load-Carrying
Capacity of Initially Imperfect Eccentrically Loaded Plates", Int.
J. Mech. Sc., Vol. 17, pp.161-175.

Rizzo, F.J. (1968); "An Integral Equation Approach to Boundary


Value Problems of Classical Elastostatics", Quart. Appl. Math.,
Vol. 25, pp.41-52.

Sheid, F. (1968); "Theory and Problems of Numerical Analysis",


Schaum's Outline Series, McGraw Hill.

Shulesko, P. (1957); "Buckling of Rectangular Plates with Two


Unsupported Edges", J. of Applied Mechanics, December.

Smirnov, V.J. (1964); 'Integral Equations and Partial Differential


Equations', Course in Higher Mathematics, Vol. 4, Addison-Vesley,
London.

Spencer A.J.M., Parker D.F., Berry D.S., England A.H., Faulkner


T.R., Green V.A., Holden J.T., Middleton D., Rogers T.G. (1980);
"Engineering Mathematics- Volume 1", von Nostrnd Reinhold Company.

Stein, M. (1959); "Loads and Deformations of Buckled Rectangular


Plates", Technical Report R-40, NASA, Langley Research Center.

Syngellakis, S. and Kang, M. (1987); "A Boundary Element Solution


of the Plate Buckling Problem", Engineering Analysis, Vol. 4, No 2,
pp.75-81.

Tan, H.K.V. (1984); "Elastic Buckling of Isotropic Flat Plates by


Finite Elements", Proc. Instn. Civ. Engnrs, Part 2, Vol. 77, pp.
13-21.
199

Tanaka, M. (1983); "Integral Equation Approach to Small and Large


Displacement of Thin Elastic Plates", Boundary Element Methods in
Engineering: Proc. of the 4th Int. Seminar, ed. Brebbia, C.A.,
pp.526-539, SpringIer Verlag, Berlin.

Tang, V. (1987); "A Generalised Approach for Transforming Domain


Integrals Into Boundary Integrals in Boundary Element Methods", PhD
Thesis, Computational Mechanics Institute.

Telles, J.C.F. (1987); "A Self Adaptive Coordinate Transformation


for Efficient Numerical Evaluation of General Boundary Element
Integrals", Intern. J. Numer. Methods in Eng'g, Vol. 24, pp. 1-15.

Timoshenko, S.P. (1953); "History of Strength of Materials",


McGraw-Hill, London.

Timoshenko, S.P. and Gere, J.M. (1961); "Theory of Elastic


Stability", Second Edition, McGraw Hill.

Timoshenko, S.P. and Goodier, J.N. (1970); "Theory of Elasticity",


Third edition, McGraw Hill.

Timoshenko, S. P. and Voinowsky-Kreiger, S. (1970); "Theory of


Plates and Shells", Second Edition, McGraw Hill.

Uemura, M. and Byon, 0-1 (1978); "Secondary Buckling of a Flat


Plate under Uniaxial Compression Part 2. Analysis of Clamped Plate
by F.E.M. and Comparison with Experiments", Int. J. Non-Linear
Mechanics, Vol.13, pp.1-14.

Valker, A.C. (1964); "Thin Valled Structural Forms under Eccentric


Compressive Load Actions", Ph.D. Thesis, University of Glasgow.

Vinter, G. (1947); "Strength of Thin Steel Compression Flanges",


Trans. ASCE, Vol. 112, pp.527-554.
200

Vrobel, L.C. Brebbia C.A. and Nardini, D. (1986); "The


Dual-Reciprocity Boundary Element Formulation for Transient Heat
Conduction", Finite Element in Vater Resources VI, Brebbia et al
eds, SpringIer Verlag.

Vrobel, L.C.Telles, J.C.F. and Brebbia, C.A. (1986); "A Dual


-Reciprocity Boundary Element Formulation for Axisymmetric
Ditiusion Problems", Proc. 8th Int. Cont. on BEM, Tanaka and
Brebbia eds, SpringIer Verlag, pp.59.

Yamaki, N. (1961); "Experiments on the Postbuckling Behavior of


Square Plates Loaded in Edge Compression", Trans. ASME, Journal of
Applied Mechanics, Vol. 28, pp.238-244.

Zienkiewics, o.c. (1977); "The Finite Element Method", Third


Edition, McGraw-Hill.
SUBJECT INDEX

Automatic discretization 47-48,72-73,76,86,


116,180,182 Hook's law 18-19

Bifurcation 3,9-11,13,26,118,133-134,139-141 In-plane boundary conditions 9,26,31-36,


Buckling load 1,3,5,12-14,23-24,26,35,54, 38,42,153,171-172,177,184
59,73,77,81,83-84,151,171-172,179- 181
Mesh nodes 35-36,44,48,59,63,72
Cell boundary nodes 63,72
Circular plates 24,47,72,75,81-82,84-85 Navier's equation 2-4,23,54
Concentrated forces 47,50-51,55,182
Corners 9,42,56,59,60-61,83-84,145,162,181 Parallelogram plates 83-85
Plate bending 1,2,4,7-8,17,22,24,31,55,151
Dirac delta function 40 Post-buckling 1,3,10-11 ,13-14,26,35, 151,
Dirichlet conditions 94 172,178
Discontinuous elements 43-44,59-60,62-63,
66-67,70,74,76-83,156,162,181 Rayleigh method 5
Rayleigh-Green equation 9-10,38,40-41,
Eccentric loads 36,172,175-177 56,90,92,109,112,124,179-180
Effective width 2,26,36
Elastic foundations 7 Secondary buckling 36,172,175,178
Second-order terms 10,152,167
Field point 39-40,55,112 Shear deformations 4,16
Finite difference method 6,35-36,158 Snap load 13-14
Finite element method 6,11,35-36,65,79,86- Source point 39-43,45-46,55,67-70,73,84,
88,171,174,180,813 104,107,109,112,114,117,125,149,162- 163
Flexural rigidity 2,4,22,42,55,153 Stress function 2,9,21,25,31-35,38-39,48,
Free boundary 6,9,29-31,36,50-51,55,67-70,73- 52,152-153,179-180
75,84-85,87-88,99-104,114,117,122,129- Stretching strains 4,9,16-17,22,24,26
132,136-140,142,145,149,152,164,172,176-
177,181-182 Top-of-the-knee approach 26
Fundamental solution 6-8,38-42,52,55-56,59, Traction boundary conditions 9,31-35,46,153
89,116,129,133,151,154,162-163,182- 184 Triangular plates 74-75,81-83,85,144-145

Green's theorems 40-41,58


Grillage analogy method 100 Von Karman's equations 2-5,9-11,25-26,38,152
LIST OF SYHBOLS

SYMBOLS

a depth of plate

width of plate

width of load strip

discontinuity jump of a fUnction f at corners

h thickness of plate

n direction normal to the boundary

incremental membrane stress resultants


distance between source and field points

s direction tangential to the boundary

s1' s2' s3 local coordinates of a domain cell

u1,u2'Uxx'Uxy'Uyy fundamental solutions

ux,uy,u z displacements in the x, y and z directions

uxs,uys stretching components of Ux and uy

uxb,u yb bending components of Ux and uy

w Uz displacement of the mid-plane of the plate

initial w pr~or to loading

gaussian weight

total w

derivative of w with respect to a then ~

normal slope of deflection

aw
tangential slope of deflection
as
203

x,y,z cartesian coordinates system


BEM Boundary Element Method
D plate flexural rigidity
E Young's modulus
F Airy's stress function
FI FI Fourier trigonometric functions
FO' n' nm

FEM Finite Element Method

IJ I jacobian
KI KI coefficients of Fourier transforms
KO' n' nm

Mxx,Mxy,Myy bending moments


Mn normal bending moment
Mns transverse bending moment
Nc number of corners

Nbe number of boundary elements

Nbu number of boundary unknowns

Nde number of domain cells


Ndn number of domain nodes
Nd , Ndu number of domain unknowns

Neb number of boundary equations for critical loads

Nfl' Nf2 order of Fourier analysis in the x and y


directions respectively
Nfe number of elements on free boundaries
Ng number of Gaussian points

Ngb number of Gaussian points for boundary integrals

Ngd number of Gaussian points for domain integrals


Nne number of nodes per element
Nz number of fictitious domain nodes
204

P source point

Q field point

shear force

radius of circular plate

normal and tangential boundary tractions

equivalent shear

Ot angle between vector r and normal n to boundary

initial rotation of a rigid bar

direct strains in the x, y and Z directions

corner angle

t.1 ith interpolation function

shear strains

load factor

critical load factor

Poisson ratio
"
domain of plate

extended Fourier domain

critical stress resultant

C1XX ' C1yy ' C1ZZ direct stresses in the x, y and z directions

'xy' 'XZ ' 'YZ shear stresses

r boundary of plate

Poisson operator

biharmonic operator

INTEGRALS

boundary integral of equation for critical loads

boundary integral of bending equation for large


deflections
205

boundary integral of stretching equation for


large deflections
domain integral of equation for critical loads,
with curvatures terms
domain integral of bending equation for large
deflections

domain integral of stretching equation for large


deflections
domain integral of equation for critical loads,
with deflections
traction boundary integral of equation for
cri tical loads

I~ traction boundary integral of bending equation


J for large deflections

corner jump term of transverse moment in


equation for critical loads
corner jump term of transverse moment in bending
equation for large deflections

ARRAYS

{b f } vector of stress function unknowns


{bw} vector of bending boundary unknowns

lew] matrix relating curvatures to deflections

{Fw} vector relating deflections to domain unknowns


{n} vector of incremental membrane stresses

{Nail} vector of membrane stresses


{T f } vector containing Fourier functions
{Ub } vector of boundary unknowns in dual-reciprocity
{U d } vector of domain unknowns in dual-reciprocity
{w} vector of domain deflections

{w Q}
,a....
vector of curvatures
Lecture Notes in Engineering
Edited by CA Brebbia and SA Orszag

Vol. 40: R. Borghi, S. N. B. Murhty (Eds.) Vol. 49: J. P. Boyd


Turbulent Reactive Flows Chebyshev & Fourier Spectral Methods
VIII, 950 pages. 1989 XVI, 798 pages. 1989

Vol. 41: W. J. Lick Vol. 50: L. Chibani


Difference Equations Optimum Design of Structures
from Differential Equations VIII, 154 pages. 1989
X, 282 pages. 1989
Vol. 51: G. Karami
Vol. 42: H. A. Eschenauer, G. Thierauf (Eds.) A Boundary Element Method for
Discretization Methods Two-Dimensional Contact Problems
and Structural Optimization - VII, 243 pages. 1989
Procedures and Applications
Proceedings of a GAMM-Seminar Vol. 52: Y. S. Jiang
October 5-7,1988, Siegen, FRG Slope Analysis Using
XV, 360 pages. 1989 Boundary Elements
IV, 176 pages. 1989
Vol. 43: C. C. Chao, S. A. Orszag, W. Shyy (Eds.)
Recent Advances in Computational Vol. 53: A. S. Jovanovic,
Fluid Dynamics K. F. Kussmaul, A. C. Lucia,
Proceedings of the US/ROC (Taiwan) Joint P. P. Bonissone (Eds.)
Workshop in Recent Advances in Expert Systems in Structural
Computational Fluid Dynamics Safety Assessment
V, 529 pages. 1989 X, 493 pages. 1989

Vol. 44: R. S. Edgar Vol. 54: T. J. Mueller (Ed.)


Field Analysis and Low Reynolds Number
Potential Theory Aerodynamics
XII, 696 pages. 1989 V, 446 pages. 1989

Vol. 45: M. Gad-el-Hak (Ed.) Vol. 55: K. Kitagawa


Advances in Fluid Mechanics Boundary Element Analysis
Measurements of Viscous Flow
VII, 606 pages. 1989 VII, 136 pages. 1990

Vol. 46: M. Gad-el-Hak (Ed.) Vol. 56: A. A. Aldama


Frontiers in Experimental Filtering Techniques for
Fluid Mechanics Turbulent Flow Simulation
VI, 532 pages. 1989 VIII, 397 pages. 1990

Vol. 47: H. W. Bergmann (Ed.) Vol. 57: M. G. Donley, P. D. Spanos


Optimization: Methods and Applications, Dynamic Analysis of Non-Linear
Possibilities and Limitations Structures by the Method of
Proceedings of an International Seminar Statistical Quadratization
Organized by Deutsche Forschungsanstalt fUr VII, 186 pages. 1990
Luft- und Raumfahrt (DLR), Bonn, June 1989
IV, 155 pages. 1989 Vol. 58: S. Naomis, P. C. M. Lau
Computational Tensor Analysis
Vol. 48: P. Thoft-Christensen (Ed.) of Shell Structures
Reliability and Optimization XII, 304 pages. 1990
of Structural Systems '88
Proceedings of the 2nd IFIP WG 7.5 Conference
London, UK, September 26-28, 1988
VII, 434 pages. 1989

For information about Vols. 1·39 please contact your bookseller or Springer-Verlag.

Potrebbero piacerti anche