Vous êtes sur la page 1sur 274

Preface

This textbook represents the Finite Element Analysis lecture course given
to students in the third year at the Department of Engineering Sciences (now
F.I.L.S.), English Stream, University Politehnica of Bucharest, since 1992.
It grew in time along with a course taught in Romanian to students in the
Faculty of Transports, helped by the emergence of microcomputer networks and
integration of the object into mechanical engineering curricula. The syllabus of the
28-hour course, supplemented by 28-hour tutorial and lab. classes, was structured
along the NAFEMS recommendations published in the October 1988 issue of
BENCHmark. The course represents only an introduction to the finite element
analysis, for which we wrote simple stand-alone single-element programs to assist
students in solving problems as homework. It is followed by an advanced course in
the fourth year at F.I.L.S., called Computational Structural Mechanics, where
students are supposed to use commercial programs.
In designing the course, our aim was to produce students capable of: (a)
understanding the theoretical background, (b) appreciating the structure of finite
element programs for potential amendment and development, (c) running packages
and assessing their limitations, (d) taking a detached view in checking output, and
(e) understanding failure messages and finding ways of rectifying the errors.
The course syllabus was restricted to 2D linear elastic structural problems.
It has been found advantageous to divide the finite element analysis into two parts.
Firstly, the assembly process without any approximations (illustrated by
frameworks) followed by the true finite element process which involves
approximations. This is achieved starting with trusses, then with beams and plane
frames, and progressively dealing with membrane and plate-bending elements.
Solid elements and shells are not treated. Our objective was to ensure that students
have achieved: (a) a familiarity in working with matrix methods and developing
stiffness matrices, (b) an understanding of global versus local coordinate systems,
(c) the abilty to use the minimum potential energy theorem and virtual work
equations, (d) the mapping from isoparametric space to real geometrics and the
need for numerical integration, (e) an insight in numerical techniques for linear
equation solving (Gauss elimination, frontal solvers etc), and (f) the use of
equilibrium, compatibility, stress/strain relations and boundary conditions.
As a course taught for non-native speakers, it has been considered useful to
reproduce as language patterns some sentences from English texts.

November 2006 Mircea Rade


Prefa
Lucrarea reprezint cursul Analiza cu elemente finite predat studenilor
anului III al Facultii de Inginerie n Limbi Strine, Filiera Englez, la
Universitatea Politehnica Bucureti, ncepnd cu anul 1992.
Coninutul cursului s-a lrgit n timp, fiind predat din 1992 i studenilor de
la facultatea de Transporturi, favorizat de apariia reelelor de calculatoare i de
includerea sa n planul de nvmnt al facultilor cu profil mecanic. Programa
cursului, care prevede 28 ore de curs i 28 ore de seminar/laborator, a fost
structurat n conformiatate cu recomandrile NAFEMS publicate n numrul din
Octombrie 1988 al revistei BENCHmark. Cursul reprezint doar o introducere n
analiza cu elemente finite, pentru care am scris programe simple, cu un singur tip
de element finit, care s fie utilizate de studeni la rezolvarea unor teme de cas. Nu
se trateaz nveliuri i elemente tridimensionale. n anul IV, planul de nvmnt
de la F.I.L.S. conine cursul Computational Structural Mechanics, la care studenii
aprofundeaz modelarea cu elemente finite i utilizeaz un program de firm.
La structurarea cursului am avut n vedere necesitatea formrii unor
studeni capabili: (a) s neleag baza teoretic, (b) s deslueasc structura
programelor cu elemente finite pentru eventuale corecii i dezvoltri, (c) s ruleze
programe i s recunoasc limitele acestora, (d) s poat verifica rezultatele i (e)
s neleag mesajele de eroare i s gseasc modaliti de corectare a erorilor.
Programa cursului a fost limitat la structuri elastice liniare
bidimensionale. S-a considerat potrivit s se prezinte analiza cu elemente finite n
dou etape: nti procesul de asamblare fr nici o aproximare (aplicat la grinzi cu
zbrele), apoi modelarea cu elemente finite, care presupune aproximarea cmpului
de deplasri, de la triunghiul cu deformaii specifice constante la elemente
patrulatere izoparametrice, incluznd integrarea numeric. S-a urmrit ca studenii
s dobndeasc: (a) familiaritate cu metodele matriciale i calculul matricelor de
rigiditate; (b) nelegerea utilitii coordonatelor locale i globale; (c) abilitatea
folosirii principiului energiei poteniale minime i a principiului lucrului mecanic
virtual; (d) trecerea de la coordonate naturale la coordonate fizice i necesitatea
integrrii numerice; (e) o vedere de ansamblu asupra rezolvrii sistemelor algebrice
liniare (eliminarea Gauss, metoda frontal etc.) i (f) utilizarea celor patru tipuri de
ecuaii echilibru, compatibilitate, constitutive i condiii la limit.
Fiind un curs predat unor studeni a cror limb matern nu este limba
englez, au fost reproduse expresii i fraze din cri scrise de vorbitori nativi ai
acestei limbi.

Noiembrie 2006 Mircea Rade


Contents

Preface i
Contents iii
1. Introduction 1
1.1 Object of FEA 1
1.2 Finite element displacement method 3
1.3 Historical view 4
1.4 Stages of FEA 5

2. Displacement Method 9
2.1 Equilibrium equations 9
2.2 Conditions for geometric compatibility 10
2.3 Force/elongation relations 11
2.4 Boundary conditions 12
2.5 Solving for displacements 12
2.6 Comparison of the force method and displacement method 13

3. Direct Stiffness Method 17


3.1 Stiffness matrix for a bar element 17
3.2 Transformation from local to global coordinates 19
3.2.1 Coordinate transformation 19
3.2.2 Force transformation 20
3.2.3 Element stiffness matrix in global coordinates 21
3.2.4 Properties of the element stiffness matrix 22
3.3 Links truss 25
3.4 Direct method 26
3.5 Compatibility of nodal displacements 28
3.6 Expanded element stiffness matrix 29
3.7 Unreduced global stiffness matrix 30
iv FINITE ELEMENT ANALYSIS

3.8 Joint force equilibrium equations 31


3.9 Reduced global stiffness matrix 33
3.10 Reactions and internal forces 35
3.11 Thermal loads and stresses 36
3.12 Node numbering 37
Exercises 41

4. Bars and shafts 47


4.1 Plane bar elements 47
4.1.1 Differential equation of equilibrium 47
4.1.2 Coordinates and shape functions 48
4.1.3 Bar not loaded between ends 49
4.1.4 Element stiffness matrix in local coordinates 51
4.1.5 Bar loaded between ends 52
4.1.6 Vector of element nodal forces 55
4.1.7 Assembly of the global stiffness matrix and load vector 56
4.1.8 Initial strain effects 59
4.2 Plane shaft elements 60
Exercises 63

5. Beams, frames and grids 79


5.1 Finite element discretization 79
5.2 Static analysis of a uniform beam 81
5.3 Uniform beam not loaded between ends 83
5.3.1 Shape functions 84
5.3.2 Stiffness matrix 86
5.3.3 Physical significance of the stiffness matrix 88
5.4 Uniform beam loaded between ends 89
5.4.1 Consistent vector of nodal forces 89
5.4.2 Higher degree interpolation functions 92
5.4.3 Bending moment and shear force 95
5.5 Basic convergence requirements 96
5.6 Frame element 97
5.6.1 Axial effects 97
5.6.2 Stiffness matrix and load vector in local coordinates 98
CONTENTS v

5.6.3 Coordinate transformation 98


5.6.4 Stiffness matrix and load vector in global coordinates 100
5.7 Assembly of the global stiffness matrix 100
5.8 Grids 111
5.9 Deep beam bending element 116
5.9.1 Static analysis of a uniform beam 117
5.9.2 Shape functions 118
5.9.3 Stiffness matrix 121

6. Linear elasticity 123


6.1 Matrix notation for loads, stresses and strain 123
6.2 Equations of equilibrium inside V 125
6.3 Equations of equilibrium on the surface S 126

6.4 Strain-displacement relations 127


6.5 Stress-strain relations 128
6.6 Temperature effects 130
6.7 Strain energy 130

7. Energy methods 131


7.1 Principle of virtual work 131
7.1.1 Virtual displacements 131
7.1.2 Virtual work of external forces 133
7.1.3 Virtual work of internal forces 133
7.1.4 Principle of virtual displacements 134
7.1.5 Proof that PDV is equivalent to equilibrium equations 137
7.2 Principle of minimum total potential energy 139
7.2.1 Strain energy 139
7.2.2 External potential energy 140
7.2.3 Total potential energy 140
7.3 The Rayleigh-Ritz method 143
7.4 FEM a localized version of the Rayleigh-Ritz method 148
7.4.1 FEM in Structural Mechanics 148
7.4.2 Discretization 149
7.4.3 Principle of virtual displacements 149
vi FINITE ELEMENT ANALYSIS

7.4.4 Approximating functions for the element 149


7.4.5 Compatibility between strains and nodal displacements 150
7.4.6 Element stiffness matrix and load vector 151
7.4.7 Assembly of the global stiffness matrix and load vector 151
7.4.8 Solution and back-substitution 152

8 Two-dimensional elements 153


8.1 The plane constant-strain triangle (CST) 153
8.1.1. Discretization of structure 153
8.1.2 Polynomial approximation of the displacement field 154
8.1.3 Nodal approximation of the displacement field 155

8.1.4 The matrix [ B ] 158

8.1.5 Element stiffness matrix and load vector 159


8.1.6 Remarks 160
8.2 Rectangular elements 176
8.2.1 The four-node rectangle (linear) 176
8.2.2 The eight-node rectangle (quadratic) 178
8.3 Triangular elements 180
8.3.1 Area coordinates 180
8.3.2 Linear strain triangle (LST) 182
8.3.3 Quadratic strain triangle 185
8.4 Equilibrium, convergence and compatibility 187
8.4.1 Equilibrium vs. compatibility 187
8.4.2 Convergence and compatibility 188

9 Isoparametric elements 191


9.1 Linear quadrilateral element 191
9.1.1 Natural coordinates 192
9.1.2 Shape functions 193
9.1.3 The displacement field 194
9.1.4 Mapping from natural to Cartesian coordinates 195
9.1.5 Element stiffness matrix 198
9.1.6 Element load vectors 199
9.2 Numerical integration 200
9.2.1 One dimensional Gauss quadrature 200
CONTENTS vii

9.2.2 Two dimensional Gauss quadrature 203


9.2.3 Stiffness integration 204
9.2.4 Stress calculations 207
9.3 Eight-node quadrilateral 208
9.3.1 Shape functions 209
9.3.2 Shape function derivatives 210
9.3.3 Determinant of the Jacobian matrix 211
9.3.4 Element stiffness matrix 211
9.3.5 Stress calculation 213
9.3.6 Consistent nodal forces 214
9.4 Nine-node quadrilateral 219
9.5 Six-node triangle 221
9.6 Jacobian positiveness 223

10 Plate bending 225


10.1 Thin plate theory (Kirchhoff) 225
10.2 Thick plate theory (Reissner-Mindlin) 229
10.3 Rectangular plate-bending elements 232
10.3.1 ACM element (non-conforming) 232
10.3.2 BFS element (conforming) 238
10.3.3 HTK thick rectangular element 239
10.4 Triangular plate-bending elements 244
10.4.1 Thin triangular element (non-conforming) 245
10.4.2 Thick triangular element (conforming) 248
10.4.3 Discrete Kirchhoff triangles (DKT) 250

References 257
Index 265
1.
INTRODUCTION

Finite Element Analysis (FEA) as applied to structures is a


multidisciplinary technique, based on knowledge from three fields: (1) Structural
Mechanics, encompassing elasticity, strength of materials, dynamics, plasticity, etc,
(2) Numerical Analysis, involving approximation methods, solving linear sets of
equations, eigenproblems, etc, and (3) Applied Computer Science, dealing with the
development and maintenance of large computer codes.
FEA is used to solve large-scale analytical problems. Its task is to model
and describe the mechanical behaviour of geometrically complex structures. The
procedure is a discretized approach: the geometric shape or the internal stress-
strain-displacement field are described by a series of discrete quantities (like
coordinates) distributed through the structure. This requires a matrix notation. The
tools are the computers, able to store long lists of numbers and manipulate them.

1.1 Object of FEA

The object of FEA is to replace the infinite degree of freedom system in


continuum applications by a finite system exhibiting the same basis as discrete
analysis.
The aim is finding an approximate solution to a boundary- and initial-
value problem by dividing the domain of the system into a set of interconnected
finite-sized subdomains of different size and shape, and defining the unknown state
variable approximately, within each element, by means of a linear combination of
trial functions. The subdomains are called finite elements, the set of finite elements
is known as the mesh and the trial functions are referred to as interpolation
functions. With the individually defined functions matching each other at certain
points called nodes, the unknown function is approximated over the entire domain.
The primary difference between the FEA and other approximate methods
for the solution of boundary-value problems (finite-difference, weighted-residual,
2 FINITE ELEMENT ANALYSIS

Rayleigh-Ritz, Galerkin) is that in the FEA the approximation is confined to


relatively small subdomains.
FEA is a localized version of the Rayleigh-Ritz method. Instead of finding
an admissible function satisfying the boundary conditions for the entire domain,
which is often difficult, in the FEA the admissible functions (called shape
functions) are defined over element domains with simple geometry and pay no
attention to complications at the boundaries.
Since the entire domain is divided into numerous elements and the function
is approximated in terms of its values at the element nodes, the evaluation of such a
function will require the solution of simultaneous equations. This was possible only
at the time the computers became available. The outstanding success of the finite
element method can be attributed to a large extent to timing. While the finite
element method was being developed, so were increasingly powerful digital
computers, which led to automation. The computer is not only able to solve the
discretized equations of equilibrium, but also to carry out such diverse tasks as the
formulation of equations, by making decisions concerning the finite element mesh
and the assembly of stiffness matrices.
Perhaps more important is the fact that the finite element method can
accommodate systems with complicated geometries and parameter distributions.
The wide use of the classical Rayleigh-Ritz method has been limited by the
inability to generate suitable admissible functions for a large number of practical
problems. Indeed, systems with complex boundary conditions or complex
geometry cannot be described easily by global admissible functions, which tend to
have complicated expressions, difficult to handle on a routine basis. In turn, in the
FEA an approximate solution is constructed using local admissible functions,
defined over small subdomains of the structure. In order to match a given irregular
boundary, or to handle parameter non-uniformities, the FEA can change not only
the size of the finite elements but also their shape. This extreme versatility, coupled
with the development of powerful computer codes based on the method, some of
them made available as open source free software, has made the FEA the method
of choice for the analysis of structures.
In FEA, the equations of equilibrium are obtained from variational
principles implying the stationarity of the functional defined by the total potential
energy. While solving differential equations with complicated boundary conditions
may be difficult, integrating known polynomial functions, even approximately,
should be easier. Mathematically, solving [ A ] { x } = { b } is equivalent to
1
minimizing P (x ) = { x }T [ A ]{ x } { x }T { b }. This is the heart of the FEA when
2
applied to structures.
1. INTRODUCTION 3

1.2 Finite element displacement method

In the finite element modeling, a structure is discretized (hypothetically)


into finite elements and points named nodes are selected on the inter-element
boundaries or in the interior of the elements. Displacements at the nodes are taken
as the primary discrete variables. Displacements within the elements are expressed
in terms of these nodal displacements using interpolation functions referred to as
shape functions. Finite elements are so small that the shape of the displacement
field can be approximated without too much error, leaving only the magnitude to
be found. The shapes are polynomials, but may be trigonometric functions as
well.
All individual elements are assembled together in such a way that the
displacements are continuous in some fashion across element interfaces, the
internal stresses are in equilibrium with the applied loads, and the prescribed
boundary conditions are satisfied. Finally, the governing discrete equations are
generated by a variational approach.
The first part of the finite element modeling process involves choosing the
correct and appropriate types of elements, understanding the pedigree of elements
and spotting wrong answers due to the use of inadequate elements. The second part
of the process is the assembly of the elements and the solution of the complete
structural equations. This involves recognizing error messages when this process
breaks down or when it simply becomes inefficient because the structure has been
modeled inconveniently.
The six basic steps of FEA are the following: (1) discretize the continuum,
(2) select interpolation functions, (3) find the element properties, (4) assemble the
element properties, (5) solve the system of equations, and (6) make additional
computations if desired.
The three main sources of approximation are: (1) the definition of the
domain (physically or geometrically), (2) the discretization of the domain (cutting
the corners, making curved lines straight, and curved elements flat), and (3) the
solution algorithms. Modeling the joints and the contact between structural parts as
well as the damping in dynamic problems are the most difficult tasks. Mesh
refinements (and automatic mesh generation) do not bring necessarily increased
accuracy. Finer mesh yields also larger stiffness matrices, a larger number of
equations to be solved, hence larger computer storage space and running time.
Among the reasons why the FEA has gained such universal acceptance are:
(1) the routine choice of shape functions, (2) the easiness of producing stiffness
matrices (and load vectors), by just assembling predetermined element matrices,
and (3) the versatility. Developed originally as a method for analyzing stresses in
complex aircraft structures, FEA has evolved into a technique that can be applied
4 FINITE ELEMENT ANALYSIS

to a large variety of linear and nonlinear, static, stability and dynamic engineering
problems.

1.3 Historical view

The idea of representing a given domain as a collection of discrete


elements is not novel with the finite element method. Ancient mathematicians
estimated the value of approximating the circumference of a circle by the
perimeter of a polygon inscribed in the circle. In modern times, the idea found
application in aircraft structural analysis, where wings and fuselages are treated as
assemblages of stringers, panels, ribs, stiffeners and spars.
The use of piecewise continuous functions defined over a subdomain to
approximate the unknown function dates back to Courant (1943), who used an
assemblage of small triangular elements and the principle of minimum potential
energy to study Saint Venants torsion problem. The reason why Courants paper
did not attract more attention can be attributed to poor timing. In the early 1940s,
computers capable of solving large sets of equations of equilibrium did not exist, so
that the method was not practical then.
The theoretical background of FEA laids on the energy approach of
Structural Mechanics and on the approximation techniques. The first energy
theorems have been established by Maxwell (1864) and Castigliano (1875). The
approximation methods have been developed by Ritz (1908) and Galerkin (1915).
Ostenfeld (1926) is credited with the first book on the deformation method.
After the Second World War, the Force Method (Flexibility Method) was
sustained by Levy (1947) and Garvey (1951) and the matrix Displacement Method
(Stiffness Method) was used by Levy (1953) in the sweptback wing analysis.
Turner formulated and perfected the Direct Stiffness Method at Boeing (1959). The
development of the Force Method ended in 1969.
The development of delta wing structures revived the interest in stiffness
methods. Modeling delta wings required two-dimensional panel elements of
arbitrary geometry. After a first attempt by Levy (1953) with triangular elements,
the article series by Argyris in four issues of Aircraft Engineering (1954, 1955),
collected later in a book by Argyris and Kelsey (1960), contains the derivation of
the stiffness matrix of a flat rectangular panel using bilinear displacement
interpolation. But that geometry was inadequate to model delta wings.
The formal presentation of the finite element method is attributed to
Turner, Clough, Martin, and Topp (1956), who during 1952-1953 succeeded to
directly derive the stiffness of a triangular panel at Boeing. The term finite
element was first used by Clough (1960).
1. INTRODUCTION 5

The first book devoted to FEA was written by Zienkiewicz and Cheung
(1967), followed by books by Przemieniecki (1968) and Gallagher (1964).
Influential papers have been written by Argyris (1965), Fraeijs de Veubeke (1964)
and Irons and coworkers (1964, 1966, 1970). Research developed in the Civil
Engineering Department at Berkeley, directed by Clough, at Washington
University, under Martin, and at Swansea University, lead by Zienkiewicz. Since
1963, finite element computer programs were freely disseminated into the non-
aerospace community.
Major contributions are due to B. M. Irons, the inventor of isoparametric
models, shape functions, frontal solvers and the patch test (1964-1980), R. J.
Melosh, who systematized the variational derivation of stiffness matrices and
recognized that FEA is a Rayleigh-Ritz method applied on small size elements
(1963), J. S. Archer, who introduced the consistent mass matrix concept (1963),
and E. L. Wilson, who studied the sparse matrix assembly and solution techniques
(1963), developed the static condensation algorithm (1974) and three SAP
computer programs (the first open source FEA software). He was joined later by
K.-J. Bathe to develop the finite element codes SAP4 (1973), SAP5 and NONSAP.
Starting with 1965 the NASTRAN finite element system was developed by
COSMIC, MacNeal Schwendler, Martin Baltimore and Bell Aero Systems under
contract to NASA, completed in 1968 and first revised in 1972. Other known finite
element codes are ANSYS, developed by Swanson Analysis Systems (1970),
STRUDL - by the Civil Engineering Department at Massachusetts Institute of
Technology and McDonnell Douglas Automation Company (1967), STARDYNE -
by Mechanics Research Inc, ADINA developed by K.-J. Bathe at M.I.T. (1975),
SESAM by Det Norske Veritas, NISA by Engineering Mechanics Research
Corporation, MARC by Marc Analysis Research Corporation, ABAQUS - by
Hibbitt, Karlsson @ Sorensen, Inc. (1978), COSMOS-M by Structural Research
& Analysis Corp. (1985), SAMCEF - by SAMTECH (1965), IDEAS-MS, PATRAN,
ALGOR etc.
General purpose programs have capabilities of linear dynamic response,
including computation of natural frequencies, nonlinear static and dynamic
response, crashworthiness, static and dynamic stability, and thermal loading. After
1967 the FEA has been applied to non-structural field problems (thermal, fluids,
electromagnetics etc).

1.4 Stages of FEA

FEA involves three stages of activity: (1) preprocessing, (2) processing,


and (3) postprocessing.
Preprocessing involves the input and preparation of data, such as nodal
coordinates, element connectivity, boundary conditions, material properties and
6 FINITE ELEMENT ANALYSIS

loading. Data input can be carried out either in an interactive way, through a user-
friendly interface, or reading from a data file. Alternatively, some input data can be
imported from other F.E.A. or C.A.D. programs.
Automatic mesh generation can be used to produce nodal coordinate data
and optimal node numbering, as well as element connectivity data. Mesh plotting is
a convenient and useful way of checking the input data. Badly placed nodes or
improper blocking of boundary nodes can be easily traced.

Fig. 1.1

Three-dimensional finite element meshes, represented with hidden line


removal, are presented for a connecting rod (Fig. 1.1) and a car engine piston (Fig.
1.2), as obtained using the program SIMPAT developed by I.T.I. Italia.

Fig. 1.2
1. INTRODUCTION 7

The finite element model of a vehicle cabin frame obtained with


MSC/NASTRAN is shown in Fig. 1.3.

Fig. 1.3

In the processing stage, the finite element program processes the input data
and calculates the nodal variable quantities such as displacements and temperatures
(equation solving), and element quantities such as stresses and gradients (back-
calculation).

Fig. 1.4

In dynamic analyses, processing involves solving an eigenproblem or


determining the transient response by incremental techniques. The cost in terms of
computer resource increases with the cube of the problem size. In static analyses,
the cost of the solution of the linear set of equations increases linearly with the
problem size. It would be obviously convenient to use in dynamic analyses the
same finite element model that was built for the static analysis. Often this contains
much more detail than the dynamic analysis requires, so that condensation and
dynamic substructuring are used to reduce the size of the dynamic problem before
the processing stage.
8 FINITE ELEMENT ANALYSIS

Postprocessing deals with the presentation of results. Early programs used


tabular presentations. Most programs produce displays of the deformed
configuration, vibration mode shapes and stress distributions. Fig. 1.4 shows the
initial mesh and the deformed shape of a cooling tower under the wind action, as
obtained using ALGOR SUPERSAP.

Fig. 1.5

Scalar nodal variables such as temperatures or pressures are presented in


the form of contour plots of isotherms or isobars.
More recent finite element programs show animated displays of the
deformed configuration, as in Fig. 1.5 for a crankshaft.

a b
Fig. 1.6

Fig. 1.6, a shows the two-dimensional initial mesh for the analysis of a
gear tooth, with 1174 triangular 6-node elements. Fig. 1.6, b shows the optimized
mesh obtained with the postprocessor ESTEREF, containing only 814 elements and
a four times reduced global discretization error.
2.
DISPLACEMENT METHOD

In solving any structural problem, there are four types of equations that
should be used: equilibrium equations, geometric compatibility conditions,
constitutive relationships and boundary conditions. In order to illustrate the usual
longhand analytical procedure, a relatively simple pin-jointed framework will be
used. Variables include reaction forces at the supports and internal forces,
displacements of the bar ends and bar extensions (elongations). If forces and
elongations are eliminated and the displacements are the variables which are solved
first, the procedure is referred to as the displacement method. It works whether the
structure is statically determinate or not. Once the displacements are determined,
they are back-substituted into the compatibility equations to obtain bar extensions,
and hence strains, then stresses from the constitutive relationship.

2.1 Equilibrium equations


Consider the truss shown in Fig. 2.1 [74], assume that all members are in
tension and write the equilibrium of each node in turn. T1 , T2 , T3 are the tensions
in members and F1 , F2 , F4 , F5 are the reaction forces at the supports.

Fig. 2.1
10 FINITE ELEMENT ANALYSIS

At Node 1 (Fig. 2.2, a), resolving forces horizontally and vertically leads to

T1 + T2 2 2 + F1 = 0,
(2.1)
T2 2 / 2 + F2 = 0.
At Node 2 (Fig. 2.2, b), equilibrium gives

6 F T1 T3 2 /2 = 0,
(2.2)
T3 2 / 2 + F4 = 0.
At Node 3 (Fig. 2.2, c)

T3 2 /2 + F5 T2 2 /2 = 0,
(2.3)
9 F T2 2 / 2 T3 2 /2 = 0.
The six equations (2.1) to (2.3) have seven unknowns. The system is
statically indeterminate. The solution is not possible by using only equilibrium, and
consideration must be given to the geometry of deformation.

a b c
Fig. 2.2

2.2 Conditions for geometric compatibility

The compatibility equations relate a bar extension l to the displacements


of the ends of the bar.
Consider a typical pin-jointed element 1-2 of a frame, Fig. 2.3, inclined an
angle with respect to the X axis of the global coordinate system.
The displacements in the local coordinate frame xOy can be expressed in
terms of the displacements in the global coordinate frame as
u 1 = U 1 cos + V 1 sin ,
(2.4)
u 2 = U 2 cos + V 2 sin .
The change in length of the member is
2. DISPLACEMENT METHOD 11

( ) ( )
l = u 2 u1 = U 2 U1 cos + V 2 V1 sin . (2.5)

Fig. 2.3
Applying equation (2.5) to each member in turn leads to
l12 = U 2 - U1 , (2.6)

2 2
l13 = (U 3 U1 ) + (V3 V1 ) , (2.7)
2 2
2 2
l 23 = (U 3 U 2 ) + (V3 V2 ) . (2.8)
2 2

The three equations (2.6) to (2.8) have nine unknowns, six displacements
and three elongations.

2.3 Force/elongation relations

Truss members are in either simple tension or compression. Starting from


the Hookes law for uniaxial stress-strain conditions, the force/elongation relations
can be written

2 2
T2 l T3 l
T l 2 , 2
l12 = 1 , l 13 = l 23 =
EA 2E A 2E A
or
T1 l T l T l
l12 = , l13 = 2 , l 23 = 3 . (2.9)
EA 2E A 2E A
Three more equations have been added and so there are now 12 equations
for 16 unknowns.
12 FINITE ELEMENT ANALYSIS

2.4 Boundary conditions

The discrepancy of 4 equations is made up by adding the boundary


conditions
U1 = V1 = V2 = U 3 = 0 (2.10)
which complete the set of 22 linear algebraic equations.

2.5 Solving for displacements

Equations (2.9) can be used to convert the compatibility equations (2.6)-


(2.8) to force/displacement equations as follows
EA EA
T1 = l 12 = ( U 2 U1 ) ,
l l
2E A EA
T2 = l 13 = 2 ( U 3 U1 + V3 V1 ) ,
l l
2E A EA
T3 = l 23 = 2 ( U 3 + U 2 + V3 V2 ) .
l l
Substitution of the expressions for T1 to T3 into the equilibrium equations
(2.1) - (2.3) yields 6 equations
F1 l
2 U1 + V1 U 2 U 3 V3 = ,
EA
F2 l
U1 + V1 U 3 V3 = ,
EA
6F l
U1 + 2 U 2 V2 U 3 + V3 = ,
EA
F4 l
U 2 + V2 + U 3 V3 = ,
EA
F5 l
U1 V1 U 2 + V2 + 2 U 3 = ,
EA
9 Fl
U1 V1 + U 2 V2 + 2 V3 = .
EA
Taking into account the boundary conditions (2.10), this set of six
equations can be written in matrix form as
2. DISPLACEMENT METHOD 13

2 1 1 0 1 1 0 F1
1
1 0 0 1 1 0 F2
EA 1 0 2 1 1 1 U 2 6 F
= . (2.11)
l 0 0 1 1 1 1 0 F4
1 1 1 1 2 0 0 F5

1 1 1 1 0 2 V3 9 F

The third and the sixth equation can be decoupled


EA 2 1 U 2 6 F
l 1 2 V = 9 F
3
and solved to give
Fl Fl
U2 = , V3 = 4 . (2.12)
EA EA

Substituting U 2 and V3 into the remaining equations yields the reaction


forces
F1 = 5F , F2 = 4 F , F4 = 5 F , F5 = F . (2.13)
Substituting these forces into the equilibrium equations (2.1)-(2.3) yields
the tensions in the members, which are equal and opposite to the forces acting on
the joints. Divided by the corresponding area they give the stresses.

2.6 Comparison of the force method and displacement method

Naviers Problem. Consider the 7-bar pin-jointed framework shown in


Figure 2.4. The joint 8 is subjected to a force of components Fx and Fy .
Determine the internal bar forces and the displacement of joint 8.

Force Method
Denoting Ti ( i = 1,...,7 ) the forces applied by each bar to the end nodes,
the equilibrium equations of joint 8 can be written
7
Fx Ti cosi = 0,
i =1
(2.14)
7
Fy Ti sini = 0.
i =1
14 FINITE ELEMENT ANALYSIS

This is a set of two equations with seven unknowns, so the framework is


statically indeterminate. We choose X1 = T3 , X 2 = T4 , X 3 = T5 , X 4 = T6 , and
X 5 = T7 as redundant forces.

Fig. 2.4

The strain energy is

1 2 5
U=
2 EA
T1 l 1 + T22l 2 + X i2l i

(2.15)
i =1
where T1 and T2 are functions of X 1 to X 5 , according to (2.14).
Using Menabreas theorem, the five deformation conditions can be written
U
= 0, ( i = 1,...,5 ) (2.16)
X i
since we are assuming no support movement. They are of the form
T1 T
l1 T1 + l 2 T2 2 + l i X i = 0 . ( i = 1,...,5 ) (2.17)
X i X i
This is a set of five linear equations wherefrom the five redundant forces
X 1 to X 5 are determined.
The components of the displacement of joint 8 are given by Castiglianos
second theorem
U U
u8 = , v8 = . (2.18)
Fx Fy
2. DISPLACEMENT METHOD 15

In the force method, the larger the number of bars, the larger the number of
statically indeterminate forces, hence the number of equations (2.16).

Displacement Method

The elongation-displacement (compatibility) equations (2.5) are

l i = u 8 cosi + v 8 sin i . ( i = 1,...,7 ) (2.19)

The bar length is


a
li = . ( i = 1,...,7 ) (2.20)
sin i
Substituting (2.19) into the force-elongation equations

l i
Ti = EA , ( i = 1,...,7 ) (2.21)
li
gives the force-displacement relations

Ti =
EA
a
(
u 8 cos i + v 8 sin i sin i . ) ( i = 1,...,7 ) (2.22)

Inserting (2.22) into the equilibrium equations (2.14) of joint 8 yields

7 7
a
EA
Fx u8 cos 2 i sin i v 8 sin 2 i cos i = 0 ,
i =1 i =1
(2.23)
7 7
a
EA
Fy u8 2
cos i sin i v 8 3
sin i = 0 ,
i =1 i =1

which is solved for u8 and v 8 .

Regardless the number of concurrent bars, only two equations are obtained
for the two joint displacements. In matrix form they can be written

Fx K11 K12 u8
= (2.24)
Fy K 21 K 22 v 8

where the stiffness coefficients are


16 FINITE ELEMENT ANALYSIS

7
EA EA
K11 =
a cos2i sini = 1.409 a
,
i =1
7
EA
K12 = K 21 =
a cosi sin 2i = 0, (2.25)
i =1
7
EA EA
K 22 =
a sin 3 i = 4.567
a
.
i =1

Solving (2.24) gives


Fx a Fy a
u 8 = 0.7097 , v 8 = 0.219 . (2.26)
EA EA

The approach used in the displacement method is the same whether the
structure is statically determinate or not.
3.
DIRECT STIFFNESS METHOD

The Finite Element Method (FEM) started as an extension of the stiffness


method or displacement method. In the stiffness method for skeletal structures, the
elements of the actual structure are connected together at discrete joints. The
relationship between the end forces and end displacements of each member is
represented by an element stiffness matrix.
We may imagine that the structure is built by adding elements one by one,
with each element being placed in a preassigned location. As elements are added to
the structure, contributions are made to the structure load carrying capacity, hence
to the structure stiffness matrix, which relates all joint displacements to all joint
forces. If the members are pin-ended bars they are real distinct elements requiring
no approximation. They are natural finite elements. Assembly and solution for
displacements are of main concern.
Starting with simple planar frameworks it is possible to explain the
assembly process and to make an introduction into the matrix stiffness method. In
the following, the basic steps of the Direct Stiffness Method (DSM) are shown
using a pin-jointed plane truss.

3.1 Stiffness matrix for a bar element

In the FEM, the names joint and member are replaced by node and
element, respectively. Consider a two-noded pin-jointed element in the own local
coordinate system (Fig. 3.1). It has length l e , cross section area Ae and Youngs
modulus Ee . Nodes are conveniently numbered 1 and 2. It is acted upon by the
nodal forces f1 , f 2 . The nodal displacements are q 1 , q 2 .

Generally, bar elements are assumed to be uniform (EAe = const.) , pin-


connected at the ends, linearly elastic, axially loaded (no bending) and with no
forces between ends.
18 FINITE ELEMENT ANALYSIS

Fig. 3.1
Both the nodal displacements q 1 , q 2 and the nodal forces f1 , f 2 are
positive in the positive x direction.
The equilibrium equation for the bar element is
f1 + f 2 = 0. (3.1)
Next, the force/elongation relations are used, which incorporate
compatibility and stress/strain relations. If end 1 is fixed and end 2 is allowed to
f l
move, then for q 1 = 0 , q 2 = 2 e and
Ee Ae
E A
f1 = f 2 = e e q 2 . (3.2)
le
Similarly, if end 2 is now fixed but end 1 allowed to move,
f l
q 2 = 0 , q 1 = 1 e and
Ee Ae
Ee Ae
f1 = f 2 = q1 . (3.3)
le
Combining equations (3.2) and (3.3), the complete stiffness relationship is
obtained as
f 1 Ee Ae 1 1 q1
= q
l e 1 1
(3.4)
f 2 e 12
2 e
123 1442443 3
nodal element nodal
forces stiffness displacements
matrix
or

{ f }= [ k ] { q }
e e e
(3.5)

where the element stiffness matrix in local coordinates is

[ k ] = El A
e e e 1 1
1 1 . (3.6)
e
3. DIRECT STIFFNESS METHOD 19

3.2 Transformation from local to global coordinates

Bar elements in a truss have different orientations in space and it is


necessary to define their stiffness properties with respect to a single global
coordinate system attached to the whole structure. End forces and displacements
have two components at each node, so that nodal forces and nodal displacements
can be arranged into 4-element column vectors related by a 4 4 stiffness matrix.

3.2.1 Coordinate transformation

A typical bar element 1-2 is shown in Figure 3.2, where both the local
coordinate system xOy and the global coordinate system XOY are drawn. Nodal
displacements are denoted by lower case letters in the local coordinate system and
by upper case letters in the global coordinate system.

Fig. 3.2
Let the bar be inclined an angle e with respect to the X-axis of the global
coordinate system. In fact, the angle e is the angle between the positive X-axis
and the positive direction of the beam (defined as 1 to 2).
Displacements in the local coordinate frame xOy can be expressed in
terms of the displacements in the global coordinate frame as (2.4)
q 1 = Q x1 cos e + Q y1 sin e ,
(3.7)
q 2 = Q x 2 cos e + Q y 2 sin e .

In matrix form
20 FINITE ELEMENT ANALYSIS

Qx1
Q
q 1 cos e sin e 0 0 y1
q =
e
(3.8)
2 0 0 cos e sin Qx 2
1424 3 e
1444444 424444444 3 Qy2
local 1
424 e
3
transformation
displacements matrix global
displacements
or
{ q }= [T ] { Q },
e e e
(3.9)
where

[T ] = cos0
e e sin e 0
cos e
0
sin e
(3.10)
0

is a coordinate transformation matrix.


From nodal coordinate data, denoting ( X 1 ,Y1 ) and ( X 2 ,Y 2 ) the
coordinates of nodes 1 and 2, respectively, we obtain
X 2 X1 Y 2 Y1
cos e =
le
, sin e =
le
, le = (X 2 X 1 ) 2 + (Y 2 Y1 ) 2 .
The entries in the matrix (3.10) are calculated based on the above equations.

3.2.2 Force transformation

Consider the pin-jointed member (Fig. 3.3) subjected to forces f1 and f2


applied at the ends 1 and 2.

Fig. 3.3

The force components in the global coordinate system are


3. DIRECT STIFFNESS METHOD 21

Fx1 = f1 cos e , Fx 2 = f 2 cos e ,


(3.11)
Fy1 = f1 sin e , Fy 2 = f 2 sin e .
In matrix form
Fx1 cos e 0
F sin
y1 0 f1
F = e
, (3.12)
x2 0 cos e f 2 e
Fy
1 2 e 0 sin e
424 3 144 42444 3 123
global transformation local
force matrix forces
components
or
{ F } = [ T ] { f }.
e e T e
(3.13)
Equation (3.13) can be directly obtained from consideration of mechanical
work. Work is a scalar quantity, having the same value regardless the coordinate
system
{ f } { q }= {F } { Q } .
e T e e T e
(3.14)
Substituting (3.9) for the local displacements, equation (3.14) becomes

{ f } { q } = { f } [T ] { Q } = {F } { Q } ,
e T e e T e e e T e

hence
{ f } [T ] = { F }
e T e e T

which by transposition becomes (3.13).

3.2.3 Element stiffness matrix in global coordinates

Inserting equation (3.5) into (3.13), then equation (3.9) into the resulting
matrix product, and comparing with (3.13)

{F }= [T ] { f }= [T ] [ k ] { q }= [1T 4]44[2k 4][4T43] { Q }


e e T e e T e e e T e e e

we find
{F } = [ K ] { Q }
e e e
(3.15)
where
[ K ] = [T ] [ k ] [T ]. .
e e T e e
(3.16)
22 FINITE ELEMENT ANALYSIS

Carrying out the matrix multiplications and using (3.10) we find the
element stiffness matrix in global coordinates
c2 cs c2 cs

[K ]
e
= e e 2
E A cs
l e c
s2 cs s2
cs c2 cs
, (3.17)

c s s
2
cs s 2

where c = cos e and s = sin e .


Remember that the element stiffness matrix is a proportionality factor
between the components of forces applied by the nodes to the element and the
components of nodal displacements in global coordinates

Fx1 c2 c s c 2 c s Qx1
F
y1 Ee Ae s 2 c s s 2 Qy1
= . (3.17, a)
F le c2 c s Q x 2
x2
Fy 2
e SYM s 2 Q y 2 e

3.2.4 Properties of the element stiffness matrix

The element stiffness matrix is symmetric, with positive diagonal


elements, singular (order 4, rank 1) and each column (row) sums to zero.

3.2.4.1 Symmetric matrix


Stiffness matrices are symmetrical

[ K ]= [K ]
e e T
.

Expressing displacements in terms of forces gives

{ Q }= [ K ] { F }= [ ] { F }
e e 1 e e e

so that the inverse of the stiffness matrix is the flexibility matrix

[K ]
e 1
[ ]
= e . (3.18)

According to Maxwells reciprocity theorem, the flexibility matrix e [ ]


must be symmetrical about the leading diagonal. And so must be its inverse, the
stiffness matrix.
3. DIRECT STIFFNESS METHOD 23

The same result can be obtained following strain energy arguments. For
linear structures, displacements are proportional to the applied loads. As forces are
increased from zero to their final values, the total work done by these forces is
1
We =
2
Qe { }{ }
T
Fe . (3.19)
In the absence of dynamic effects, this work is absorbed by the structure as
strain energy. Substituting (3.15) into (3.19) we obtain the strain energy which is a
scalar
Ue =
1
2
Qe
T
{ } [ ]{ }
K e Qe . (3.20)

It is equal to its transpose

Ue =
1
2
{ Q } [K ] { Q }
e T e T e
(3.20, a)
so that
[ K ]= [K ]
e e T
,
which defines the symmetry.

3.2.4.2 Singular matrix


The element stiffness matrix is of order 4 and rank 1. The rank deficiency
is 3 and this corresponds to the three possible and independent forms of rigid body
motion in plane for the unsupported bar: two translations and one rotation.
A single ungrounded bar can be moved in space as a rigid body without
straining it and hence with zero strain energy. This means that there exist a set of
rigid body displacements {Q } e
for which { } [ K ] { Q }= 0
2U e = Q e
T e e

wherefrom [ K ] { Q } = { 0 } . This can only be true if the determinant of [ K ]


e e e

vanishes. The matrix [ K ] is said to be singular. The rigid body modes are defined
e

by the eigenvectors corresponding to its zero eigenvalues.


The zero determinant implies that there are linear relationships between its
columns (rows). The rank of a matrix is the size of the largest sub-matrix with a
non-zero determinant. One can verify that the determinants of the 3 3 and 2 2
reduced sets are still zero, so that the stiffness matrix has rank 1 (or its rank
deficiency is 3).

3.2.4.3 Positive diagonal elements

[ ]
Each diagonal entry of the matrix K e is positive. If this were not so, a
force and its corresponding displacement would be oppositely directed, which is
24 FINITE ELEMENT ANALYSIS

[ ]
physically unsound. Moreover, the matrix K e is positive semidefinite. That is,
the quadratic form that represents strain energy (3.20) is either positive or zero.

3.2.4.4 Each column (row) sums to zero


( )
Consider a bar element with end 2 fixed Qx 2 = Q y 2 = 0 and end 1 having
( )
a unit displacement along the global X-axis Qx1 = 1, Q y1 = 0 as in Fig. 3.4.

Fig. 3.4

Equation (3.17, a) can be written

Fx1 k11 . . . Qx1 = 1


F
y1 k 21 . . . Q y1 = 0
= (3.21)
Fx 2 k31 . . . Qx 2 = 0
Fy 2 k 41
. . . Q y 2 = 0
which yields
Fx1 = k11 , Fy1 = k21 , Fx 2 = k31 , Fy 2 = k 41 . (3.22)

This shows that the first column of the stiffness matrix represents the
forces that must be applied to the element to preserve static equilibrium when
Qx1 = 1 and all other displacements are zero.

The displacement Qx1 = 1 produces an axial shortening Qx1 cos e = cos e ,


E A
which corresponds to a compressive force e e cos e , whose components must
le
be equilibrated by the external forces k11 , k 21 , k31 and k 41 . Equilibrium of
horizontal and vertical forces yields
k11 + k31 = 0 , k21 + k41 = 0 ,
3. DIRECT STIFFNESS METHOD 25

so that the first column sums to zero. The same applies for the other columns.
In the element stiffness matrix, each column represents an equilibrium set
of nodal forces produced by a unit displacement of one nodal degree of freedom.

3.3 Links truss

In order to illustrate the assembly of the global stiffness matrix from the
elemental stiffness matrices, consider Links truss [74] shown in Fig. 3.5, already
analyzed in Chapter 2, Fig. 2.1.

Fig. 3.5

The truss comprises 3 elements and 3 nodes. It is simply supported in 2 and


3, firmly located in 1, and acted upon by forces 6F and 9F. The global
displacements and nodal forces are shown in Fig. 3.6. A node whose global index
is i has associated with it the global displacements and forces (2 i 1) and 2 i .

a b
26 FINITE ELEMENT ANALYSIS

Fig. 3.6
Element data are given in Table 3.1 together with information useful for
the computation. The first three columns define the element connectivities, i.e.
their localization within the structure. Element numbering can be arbitrary.

Table 3.1

Nodes EAe
Member e cos e sin e c2 s2 cs le EAe
le
i j
1 1 2 0 1 0 1 0 0 l EA EA
l
2 2 1 1 1 2 EA
2 1 3 45 l 2 EA 2
2 2 2 2 2 2 l

2 2 1 1 1 2 EA
3 2 3 135 l 2 EA 2
2 2 2 2 2 2 l

3.4 Direct method

Using the data from Table 3.1, the element stiffness matrices (3.17) are
calculates as
K K
1 2 3 4 K
K
1 0 1 0 1
K K
0 0 0 2
[K ]
1
=
EA 0
l 1 0 1 0 3
T
K K
,

K K K K K K
0 0 0 0 4
K K K K K K

K K
1 2 5 6 K K
1 1 1 1 1
K K K K K K
[ ]K2 =
E A 1 1 1 1 2 T
l 1 1 1 1 5

K K K K K K
,

K K
1 1 1 1 6
K K
3. DIRECT STIFFNESS METHOD 27

and
K K K K K K
3 4 5 6 K
K K K K K
1 1 1 1 3
K K
1 1 4 T
[ ]
K3 =
E A 1 1
l 1 1 1 1 5

K K
.

K K
1 1 1 1 6
K K

At the top and on the right of the element stiffness matrices, the numbering
of coordinates in the global stiffness matrix is shown, according to Table 3.2.

Table 3.2
Node
1 2
Direction
X Y X Y
Local nodal coordinate
1 2 3 4
Element Global nodal coordinate
1 1 2 3 4
2 1 2 5 6
3 3 4 5 6

The assembly of the unreduced global stiffness matrix (3.28) is done


systematically, locating each coefficient of the element stiffness matrices into the
appropriate place in the global 6 6 matrix (for this example), as indicated by dots
above, eventually adding it to the coefficients already accumulated at that location.
This is referred to as the direct matrix method.
For instance, element 2 is located in the truss between the left end node
i = 1 and the right end node j = 3 (nodal labels i and j may be assigned arbitrarily).
In the global matrix, the two displacement and force components (along X and Y)
for node i = 1 are numbered 2 i 1 = 1 and 2 i = 2 , and those for the node j = 3 are
numbered 2 j 1 = 5 and 2 j = 6 .
The simple addition of different stiffness coefficients in a location is based
on the fact that finite element equations are in fact nodal equilibrium equations, so
that if a node is common to several elements, each member will contribute with a
force component to maintain equilibrium under an arbitrary set of nodal
displacements. An alternative algebraic explanation of the assembly of system
stiffness matrices is presented in the following.
28 FINITE ELEMENT ANALYSIS

3.5 Compatibility of nodal displacements

The compatibility of nodal displacements at element level, with the nodal


displacements at the whole truss structure level, can be expressed by equations of
the form

{ } [ ]
Q e = T e { Q },
~
(3.23)

where { Q } is the element displacement vector in global coordinates, {Q } is the


e

displacement vector of the truss structure and [T ] is referred to as a


~ e
full
connectivity or localization matrix, containing ones at the nodal displacements of
element nodes and zeros elsewhere.
For the truss from Fig. 3.5, equation (3.23) applied to the three elements
yields
Q1

Q1 1 0 0 0 0 0 Q2

0 Q3
{Q }
1 Q 0
= 2 =
1 0 0 0
[ ] [Q ] ,
= T1
~
Q3 0 0 1 0 0 0 Q4
Q4 0
0 0 1 0 0 Q5

Q6

Q1

Q1 1 0 0 0 0 0 Q2
Q 0
{ }
Q = 2 =
2 1 0 0 0 0 Q3
[ ] [ Q ],
~
= T 2
Q5 0 0 0 0 1 0 Q4
Q6 0 0 0 0 0

1 Q5

Q6

Q1

Q3 0 0 1 0 0 0 Q2
Q
0 Q3
{ } 0
Q3 = 4 =
0 0 1 0
[ ] [ Q ].
~
= T 3
Q5 0 0 0 0 1 0 Q4

Q6 0 0 0 0 0 1 Q5

Q6
3. DIRECT STIFFNESS METHOD 29

3.6 Expanded element stiffness matrix

The element strain energy in global coordinates can be written in terms of


the global displacement vector, substituting (3.23) into (3.20, a)

Ue =
1
{ Q }T [T~ e ] T [ K e ][T~ e ]{ Q }
2
or
Ue =
1
{ Q }T [ K~ e ]{ Q },
2
where the expanded element stiffness matrix

[ K~ ] = [T~ ] [ K ] [T~ ]
e e T e e
(3.24)

has the size of the system matrix.


For the truss from Fig. 3.5, equation (3.24) yields

1 0 0 0
0
1 0 0 1 0 1 0 1 0 0 0 0 0
0 0 0 0

0 1 0 EA 0
[K~ ] = [T~ ] [ K ][T~ ]
1 1 T 1 1 0
=


0 0 1 l 1 0 1 0 0

1 0 0 0 0
0 1 0 0 0
,
0

0 0 0 0 0 0 0 0 0 0 0 1 0 0

0 0 0 0

or
1 0 1 0 0 0
0
0 0 0 0 0

[ ]
~ 1 EA 1
K =
l 0
0 1 0 0 0
0 0 0 0 0
.
0 0 0 0 0 0

0 0 0 0 0 0

1 0 0 0
0
1 0 0 1 1 1 1 1 0 0 0 0 0
1 1 1 0 1 0 0 0 0
[K~ ] = [T~ ] [K ][T~ ]
2 2 T 2 2 0
=
0 0 0 EA 1

0 0 0 l 1 1 1

1 0 0 0 0 1 0
,
0

0 0 1 0 1 1 1 1 0 0 0 0 0 1

0 0 0 1
30 FINITE ELEMENT ANALYSIS

or
1 1 0 0 1 1
1
1 0 0 1 1

[K~ ]
2
=
EA 0

l 0
0
0
0 0 0
0 0 0
0
0

1 1 0 0 1 1

1 1 0 0 1 1
and
0 0 0 0
0
0 0 0 1 1 1 1 0 0 1 0 0 0
1 1 0

0 0 0 EA 1 1
[K~ ] = [T~ ] [ K ][T~ ]
3 3 T 3 3 1
=
1 0 0 l 1 1 1 1 0

0 0 1 0 0
0 0 0 1 0
,
0

0 0 1 0 1 1 1 1 0 0 0 0 0 1

0 0 0 1
or
0 0 0 0
0 0
0
0 0 0
0 0

[K~ ]
3
=
EA 0

l 0
0 1 1 1 1
0 1 1 1 1
.
0 0 1 1 1 1

0 0 1 1 1 1

3.7 Unreduced global stiffness matrix

The strain energy for the complete truss structure

{ Q } [ K ] { Q },
1 T
U= (3.25)
2
can be calculated by simply adding the element strain energies

U = 2 { Q } [ K ]{ Q } = 2 { Q } [ K ] { Q }.
1 ~ T 1 ~ T
U= e
e e
(3.26)
e e e

Comparing (3.25) and (3.26) we get

[ K ] = [ K~ e ] . (3.27)
e
The global stiffness matrix is equal to the sum of the expanded element
stiffness matrices.
3. DIRECT STIFFNESS METHOD 31

[ ]
The unreduced stiffness matrix K is symmetric, singular (for a plane
truss, the deficiency is 3), has positive elements along the main diagonal and each
column (row) sums to zero. It corresponds to the free-free system. For a grounded
system, this matrix is condensed using the boundary conditions.
The effect of elastic supports modelled as lumped springs can be accounted
for by adding their stiffnesses along the main diagonal at the appropriate locations
in the stiffness matrix.
For the truss from Fig. 3.5, equation (3.27) yields

2 1 1 0 1 1
1 1
0 0 1 1

[ K ] = [ K~ 1 ]+ [ K~ 2 ]+ [ K~ 3 ] =
EA 1 0

l 0
2 1 1 1
0 1 1 1 1
. (3.28)
1 1 1 1 2 0

1 1 1 1 0 2

The expanded element stiffness matrices have been used above only to
show algebraically how to assemble a global stiffness matrix; they are never used
in practice. The expensive product (3.24) is never formed. The global stiffness
assembly is a simple book-keeping exercise and is done by directly placing the
nonzero coefficients of element stiffness matrices in the right locations of the
global stiffness matrix based on element connectivity.

3.8 Joint force equilibrium equations

The assembly of the global stiffness matrix has been based on strain energy
considerations. An alternative presentation is given below, based on joint
equilibrium equations, using for convenience the truss from Fig. 3.5.
Note that element equilibrium equations (3.1) used so far involved only
forces applied by nodes to the elements. The joint equilibrium equations, involving
forces applied by elements to nodes, are used in the following.
An exploded layout of the truss is shown in Fig. 3.7. Apart from external
forces and support reactions, nodes are acted upon by forces equal and in opposite
direction to those applied to elements. Equal forces are labelled only once for
clarity.
Resolving nodal forces horizontally and vertically, we obtain 6 equilibrium
equations
32 FINITE ELEMENT ANALYSIS

F1 = Fx11 + Fx21 , F3 = Fx12 + Fx31 , F5 = Fx22 + Fx32 ,


F2 = Fy11 + Fy21 , F4 = Fy12 + Fy31 , F6 = Fy22 + Fy32 .

Fig. 3.7

In matrix form
Fx11
1
Fy1
1
Fx 2
Fy12
F1 1 0 0 0 1 0 0 0 0 0 0 0
F
2 0 1 0 0 0 1 0 0 0 0 0 0 Fx21
F3 0 0 1 0 0 0 0 0 1 0 0 0 Fy21
=
F4 0 0 0 1 0 0 0 0 0 1 0 0 Fx22
F5 0 0 0 0 0 0 1 0 0 0 1 0 Fy22

F6 0 0 0 0 0 0 0 1 0 0 0 1
14 42443 144244 3 14
4244 3 F3
x1
[ T~ ]
1 T [ ]
~2 T
T
~3 T
T [ ]
F3
y1
Fx32
3
Fy 2

whose partitioned form can be written in scalar product form as

{ F } = [ T~1 ] T {F 1 }+ [ T~ 2 ] T {F 2 }+ [ T~ 3 ] T {F 3 }
or generally
{ F } = [ T~ e ] T {F e }. (3.29)
e
3. DIRECT STIFFNESS METHOD 33

Substituting (3.15) and (3.23), equation (3.29) becomes

{ F } = [ T~ e ] T [ K e ]{Q e }= [1T~4]44[ K24][4T~43] {Q },


e T e e
(3.30)
e e

then, using equations (3.24) and (3.27), it can be written

{ F } = [ K~ e ] {Q } = [ K ] {Q }. (3.31)
e
[ ]
The unreduced global stiffness matrix K relates the unreduced vector of
nodal forces { F } to the unreduced vector of nodal displacements { Q }.

3.9 Reduced global stiffness matrix

The global truss should be supported adequately and there should be no


internal mechanism. Boundary conditions eliminate the possibility of the structure
to move as a rigid body.
A general type of boundary conditions include specified displacements of
the support nodes, of the type Qi = ai ( i = 1 to ns ), where ns is the number of
supports. Most often the support nodal displacements are zero, Qi = 0 . Another
type are the multipoint constraints, encountered in inclined roller supports and rigid
connections, of the type bi Qi + b j Q j = b0 , where bi , b j and b0 are known
constants.
For an N-degree-of-freedom structure, the finite element equations (3.31)
are of the form
K11 K12 L K1N Q1 F1
K
21 K 22 L K 2 N Q2 = F2 .
L L L L L L

K N 1 K N 2 L K NN QN FN
Consider a single boundary condition, Q1 = a1 . As Q1 is known, the finite
element equations in the remaining N 1 unknowns are
K 22 K 23 L K 2 N Q2 F2 K 21 a1
K K 33 L K 3 N Q3 F3 K 31 a1
32 = . (3.32)
L L L L L L

K N 2 KN3 L K NN QN FN K N 1 a1
34 FINITE ELEMENT ANALYSIS

The ( N 1 ) (N 1) stiffness matrix above is obtained simply by deleting


or eliminating the first row and column from the original ( N N ) stiffness matrix.
Equation (3.32) may be written in condensed form
[ K ] { Q } = { F }, (3.33)

where [K ] is a reduced stiffness matrix, obtained by eliminating the row and


column corresponding to the specified or support degrees of freedom, { Q } and
{ F } are the reduced vectors of global displacements and forces, respectively.
For a truss supported on many supports, the reduced global stiffness matrix
is obtained deleting as many rows and columns as the number of specified
displacements. The final equations (3.33) can be solved for the displacement vector
{ Q } using Gauss elimination. The matrix [ K ] is not singular.
It is instructive now to consider the truss from Fig. 3.5. Having assembled
the unreduced global stiffness matrix, after substitution of the boundary conditions
Q1 = Q2 = Q4 = Q5 = 0 and the values of the applied forces F3 = 6 F and F6 = 9 F ,
the following set of linear equations is obtained
2 1 1 0 1 1 0 F1
1
1 0 0 1 1 0 F2
EA 1 0 2 1 1 1 Q3 6 F
= . (3.34)
l 0 0 1 1 1 1 0 F4
1 1 1 1 2 0 0 F5

1 1 1 1 0 2 Q6 9 F

It can be seen that where the displacements are specified, the nodal forces
are unknown, and where the forces are specified, the nodal displacements are
unknown. Equations (3.34) can be rearranged as follows
2 1 1 0 1 1 Q3 6 F

1 2 1 1 1 0 Q6 9 F


EA 1 1 2 1 0 1 0 = F1 .
(3.35)
l 0 1 1 1 0 1 0 F2

1 1 0 0 1 1 0 F4

1 0 1 1 1 2 0 F5

The first two equations can be written


EA 2 1 Q3 6 F
=
l 1 2 Q6 9 F
3. DIRECT STIFFNESS METHOD 35

and solved to give


Fl Fl
Q3 =, Q6 = 4 .
EA EA
In general, for non-zero boundary conditions, equations (3.33) can be
written in partitioned form
[ K aa ] [ K ab ] { Qa } { Fa }
=
[ K bb ] { Qb } { Fb }
[K ] , (3.36)
ba

where { Fa } is the vector of known forces, { Qb } is the vector of known


displacements and [ K ba ] = [ K ab ] . T

The first set of equations can be written

{ Fa } = [ K aa ] { Qa }+ [ K ab ] { Qb } , (3.37)

then multiplied to the left by [ K aa ] 1 to yield the unknown displacements

{ Qa } = [ K aa ] 1 ( { Fa } [ K ab ] { Qb } ) , (3.38)

3.10 Reactions and internal forces

The unknown forces, which in this case are the external reactions, are
given by
{ Fb } = [ K ab ] T { Qa }+ [ K bb ] { Qb } . (3.39)
For the truss from Fig. 3.5, the support displacements are zero, so that
F1 1 1 5F
F 0 1 Q
2 EA 3 4 F
= = .
F4 l 1 1 Q6 5 F
F5 F
1 0
The elongation of an element (2.5) is
l e = q2 q1 = ( Qx 2 Qx1 ) cos e + ( Q y 2 Q y1 ) sin e (3.40)

which can be computed when all displacements have been determined.


The axial force in a truss element is
36 FINITE ELEMENT ANALYSIS

Qx1
Q
E A E A y1
Te = e e l e = e e b c s c s c . (3.41)
le le Q x 2
Q y 2
e
For the truss from Fig. 3.5, the member forces are
Q1
Q
EA EA
T1 = b1 0 1 0c 2 = Q3 = F ,
l Q
3 l
Q4

Q1
Q
2 EA 2 2 2 EA
T2 = b 1 1 1 1c = Q6 = 4 2 F ,
l 2 Q5 l
Q6

Q3
Q
4
T3 =
2 EA 2
b1 1 1 1 c =
2 EA
( Q3 + Q6 ) = 5 2 F .
l 2 Q5 l
Q6

Stresses can be determined dividing these forces by the element cross-


section areas.

3.11 Thermal loads and stresses

Thermal stresses are calculated using the restraining method suggested


by J. M. C. Duhamel (1838).
Suppose the node displacements are completely restrained (blocked). This
produces thermal strains T = T , where is the coefficient of thermal
expansion and T is the amount of uniform heating (temperature difference). The
restrained state is equivalent to a pre-stressing with compressive stresses
T = ET , where E is Youngs modulus.
Restraining produces a compressive axial force EAT in the element,
where A is the cross-section area. Accordingly, the element acts upon its nodes
with equal and opposite forces

{ FT } = EAT b c s c s cT , (3.42)
3. DIRECT STIFFNESS METHOD 37

or, in local coordinates,

{ fT } = EAT b 1 1cT . (3.42, a)


These forces should be included in the vector of nodal forces (added to the
external forces, if the case). After determining the displacements produced by these
forces, the element elongations are determined from (3.37) and the stresses are
calculated as
l
e = Ee e T , (3.43)
le
i.e. adding to the stresses produced by the thermal (and external) loads, the initial
stresses produced by restraining. It is as if unrestraining forces are applied at the
ends of the element to free it from the initial restraining.

3.12 Node numbering

Stiffness matrices are symmetric and sparse. They are also banded, i.e.
with the nonzero elements clustered in a band along the main diagonal. This is
obvious for one-directional structures and can be achieved by rational node
numbering in other structures.
Consider the 12 12 unreduced global stiffness matrix of the truss from
Fig. 3.8, a, with nonzero elements in the upper triangle identified by the symbol X

B

0







[ ]
K =

. (3.44)







s y m m
e t r i c

Because of symmetry, only the diagonal elements and those from one side
of the main diagonal are retained. Even so, due to the sparseness, there are many
zero elements in the upper triangle. The half-bandwidth is denoted B.
38 FINITE ELEMENT ANALYSIS

Since only nonzero elements need to be stored, the information in the


above matrix can be compactly stored in the 12 6 matrix below
1 2 B








.

[ Kb ] =









0


The first column contains the diagonal elements of the full matrix (3.44).
The second column contains the elements from the second diagonal. In general, the
mth diagonal of the original matrix is stored as the mth column. The number of
columns in the banded-form storage is equal to B - the half-bandwidth of the
original matrix. The efficiency of band-storage increases with the order of the
matrix.

a b
Fig. 3.8

For plane trusses, B is equal to 2 plus twice the maximum node number
difference in an element. For the numbering scheme of Fig. 3.8, a , B = 6 .
However, for the numbering from Fig. 3.8, b, B = 12 , i.e. equal to the size of the
original matrix.
As a general rule, a small bandwidth can be obtained by numbering nodes
along the shorter dimension of a structure, then progressing along the longer
dimension.
3. DIRECT STIFFNESS METHOD 39

The half-bandwidth is automatically determined within the finite element


program from the node numbering. Apart from storage savings, Gaussian
elimination algorithms are used for symmetric banded matrices which enable also
the reduction of computer time.
For large sparse stiffness matrices, efficient reduction of both storage and
computing time can be achieved using the skyline storage and a skyline equation
solver. In this case, the columns of the matrix upper triangle are stored serially and
concatenated in a column vector { K s } . If there are zeros at the top of a column,
only the elements between the diagonal term and the first nonzero term need be
stored. The line separating the top zeroes from the first nonzero element is called
the skyline.
Consider the following matrix
Column
1 2 2 4 3 1 4 5
height
k11 k12 0 k14 0 0 0 0

k 22 k 23 k 24 0 0 0 0
k33 k34 k35 0 0 0
Skyline
k 44 0 0 k 47 k 48
k55 0 0 0

k 66 k67 0
k77 k 78

k88

The active columns are stored in the column vector { K s } and a diagonal
pointer vector { ID } is built up with the indices of diagonal elements in { K s }

k11
k 1
12
k22 1
3 3
k23
k33 5
5
k14
{ ID } = .
9
{Ks } =
k24 12
k34 13

k44 17
L 9 22

L
k
88
40 FINITE ELEMENT ANALYSIS

The height of the jth column is given by ID ( j ) ID ( j 1) . For the


solution of the finite element equations, Gaussian elimination can be applied using
a skyline solver program.

A generic flow chart of the Matrix Displacement Method is given in the


following.

Matrix Displacement Method

Input Data
Geometric data of truss (Nodal coordinates)
Material properties + Cross-section area of members 1
Connectivity table of elements
Boundary conditions

Element stiffness matrix in local coordinates 2
Transformation to global coordinates

Assembly of the global stiffness matrix 3


Introduction of boundary conditions 4

Solving the linear set of equations 5

Back-calculation
6
External reactions
Member forces (and stresses)
3. DIRECT STIFFNESS METHOD 41

Exercises
E3.1. For the truss in Fig. E3.1, a, determine: a) the maximum nodal
displacement and its location b) the maximum stress and its location, and c) the
support reactions. Plot the deformed shape.

Fig. E3.1, a

Answer. a) The finite element model consists of 8 nodes and 13 elements.


Taking l = 1 , A = 1 , E = 1 and F = 1 , we obtain the following:

Nodal data
Node Restr Restr Coord Coord Displ Displ
nr X Y X Y X Y
1 1 1 0 0 0 0
2 0 0 4 0 64 -361.1
3 0 0 4 3 208.3 -361.1
4 0 0 8 0 128 -391.8
5 0 0 8 6 127.6 -343.7
6 0 0 12 0 176 -339.7
7 0 0 12 3 62.9 -339.7
8 0 1 16 0 224 0

Element data
Element Axial Element Axial
nr Nodes stress nr Nodes stress
1 1, 2 16 8 4, 6 12
2 1, 3 -10 9 4, 7 -4.16
3 2, 3 0 10 5, 7 -10.83
4 2, 4 16 11 6, 7 0
5 3, 4 -9.16 12 6, 8 12
6 3, 6 -10.83 13 7, 8 -15
7 4, 5 8
42 FINITE ELEMENT ANALYSIS

The support reactions are R1 = 8 , R2 = 6 , and R3 = 9 . The deformed


shape is shown in Fig. E3.1, b.

Fig. E3.1, b

E3.2. Consider the pin-jointed framework shown in Fig. E3.2, a.


Determine: a) the maximum nodal displacement, b) the maximum stress, and c) the
support reactions. Plot the deformed shape.

Fig. E3.2, a

Answer. The finite element model consists of 7 nodes and 11 elements. a)


Fl
The vertical displacement of point 7 is v7 = 219.787 . b) The axial stress in
EA
element 7 is N 7 = 6 F A . c) The support reactions are R1 = R3 = 5.4 F , and
R2 = 6 F .
The deformed shape is presented in Fig. E3.2, b.
3. DIRECT STIFFNESS METHOD 43

Fig. E3.2, b

E3.3. Consider the truss shown in Fig. E3.3, a. Determine the location and
the value of: a) the maximum nodal displacement, b) the maximum stress.
Calculate c) the support reactions. Plot the deformed shape.

Fig. E3.3, a

Answer. The finite element model consists of 6 nodes and 9 elements. a)


Fl
The vertical displacement of point 3 is v3 = 137.83 . b) The axial stress in
EA
element 1 is N1 = 7.25 F A . c) The support reactions are R1 = 0 , R2 = 3.4 F and
R3 = 2.6 F .
The deformed shape is presented in Fig. E3.3, b.
44 FINITE ELEMENT ANALYSIS

Fig. E3.3, b

E3.4. Consider the truss shown in Fig. E3.4, a. Determine: a) the


maximum nodal displacement, b) the maximum stress, and c) the support reactions.
Plot the deformed shape.

Fig. E3.4, a

Answer. The finite element model consists of 4 nodes and 5 elements.


Taking l = 1 , A = 1 , E = 1 and F = 1 , we obtain the following:

Nodal data
Node Restr Restr Coord Coord Displ Displ
nr X Y X Y X Y
1 1 1 0 0 0 0
2 1 1 0 1 0 0
3 0 0 1 0 -1.3294 -3.2095
4 0 0 2 1 2.6705 -9.0896
3. DIRECT STIFFNESS METHOD 45

Element data
Element Axial
nr Nodes stress
1 3, 4 - 0.940
2 1, 3 - 1.329
3 2, 3 0.940
4 2, 4 1.335
5 1, 4 - 0.749

a) The vertical displacement of point 4 is v4 = 9.09 F l E A . b) The axial


stress in element 4 is N 4 = 1.335 F A . c) The support reactions are
R1 = R3 = 2 F , R2 = 0.335F and R4 = 0.665F . The deformed shape is presented
in Fig. E3.4, b.

Fig. E3.4, b

E3.5. Consider the pin-jointed framework shown in Fig. E3.5, a where


point 6 is displaced v6 = 5 . Determine the location and the value of: a) the
maximum nodal displacement; b) the maximum stress. Plot the deformed shape.

Fig. E3.5, a
46 FINITE ELEMENT ANALYSIS

Answer. The finite element model consists of 14 nodes and 25 elements.


Taking l = 1 , A = 1 and E = 1 , the vertical displacements of points 7 and 8 are
v7 = v8 = 4.538 , and the axial stress in element 15 is N15 = 0.366 . The deformed
shape is presented in Fig. E3.5, b.

Fig. E3.5, b
4.
BARS AND SHAFTS

This chapter deals with simple one-dimensional structural elements, having


one degree of freedom per node. The displacement within the element is expressed
in terms of the nodal displacements using shape functions. The unknown
displacement field within an element is usually interpolated by a linear distribution.
This approximation becomes increasingly accurate as more elements are
considered in the model. For a bar without loads between ends the linear
interpolation is exact. The compatibility of adjacent elements requires only C 0
continuity. Displacements must be continuous across the element boundary.
For bars with distributed loads, true displacements are described by higher
order polynomials. It is shown that their use is tantamount to adding internal
nodes. However, it is common practice to use linear shape functions and two-node
elements without loads between ends. This implies replacement of the distributed
loads by equivalent forces applied to nodes. These kinematically equivalent forces
are determined using the appropriate shape functions from the condition to perform
the same mechanical work as the actual loading. In this section, the corresponding
element stiffness matrix and load vectors will be derived.

4.1 Plane bar elements


Bars are structural elements used to model truss elements, cables, chains
and ropes. Their longitudinal dimension is much larger than the transverse
dimensions. Bars are loaded only by axial forces. They are modeled by elements
having one-degree-of-freedom per node.

4.1.1 Differential equation of equilibrium

In a thin uniform rod of cross-section area A and Youngs modulus E, there


are axial displacements u = u (x ) due to axial loads p(x ) . The dimensions of p are
force/length. The displacement at x + d x will be u + du .
48 FINITE ELEMENT ANALYSIS

The axial strain is given by the strain-displacement relation


x = du d x . (4.1)
The normal stresses result from Hookes law
x = E du d x . (4.2)

Fig. 4.1
The internal axial force N is
du
N = A x = E A . (4.3)
dx
The equation of equilibrium of an infinitesimal element of length dx of the
rod (Fig. 4.1) is d N + p d x = 0 ,

dN
+ p (x ) = 0 ,
dx
or, using (4.3),
d2 u
EA = p (x ) . (4.4)
d x2

4.1.2 Coordinates and shape functions

Consider a two-node pin-jointed element in the own or local coordinate


system. Nodes are conveniently numbered 1 and 2, their coordinates in the physical
(Cartesian) reference system being x 1 and x 2 respectively (Fig. 4.2, a).

We define a natural or intrinsic reference system which permits the


specification of a point within the element by a dimensionless number

2 x + x2
r= x 1 (4.5)
x 2 x 1 2

so that r = 1 at node 1 and r = +1 at node 2 (Fig. 4.2, b).


4. BARS AND SHAFTS 49

Fig. 4.2

Expressing the physical coordinate in terms of the natural coordinate yields


x = N1 (r ) x 1 + N 2 (r ) x 2 , (4.6)
where
1
N1 (r ) = ( 1 r ) and N 2 (r ) = 1 ( 1 + r ) (4.7)
2 2
can be considered as geometric interpolation functions. The graphs of these
functions are shown in Figs. 4.3, a,b. They have a unit value at the node of the
same index and zero at the other node.

Fig. 4.3

4.1.3 Bar not loaded between ends

For a prismatic bar not loaded between ends, p = 0 , d 2u d x 2 = 0 ,


d u d x = const . , so that the displacement field within the element may be expressed
as a linear polynomial
50 FINITE ELEMENT ANALYSIS

u (x ) = a + b x . (4.8)
The two integration constants above may be determined from the nodal
displacements and the geometry of the element. With q1 = a + b x1 at node 1 and
q2 = a + b x2 at node 2, equation (4.8) becomes
q2 q1 q x q2 x1
u= x+ 1 2 . (4.9)
x2 x1 x2 x1
Equation (4.9) can also be written
x + x2 x x1
u= q1 + q2 ,
x2 x1 x2 x1
or, using (4.5),
1 r 1+ r
u= q1 + q2 .
2 2
The displacement of an arbitrary point within the element can be expressed
in terms of the nodal displacements q 1 and q 2 as

u = N1 (r ) q1 + N 2 (r ) q2 , (4.10)

1
where N1 (r ) = ( 1 r ) and N 2 (r ) = 1 ( 1 + r ) are the shape functions of the
2 2
element.
The polynomial form (4.6) is simpler, but the integration constants, a and
b, have no simple physical meaning. The nodal expansion (4.10) is more
complicated, but the integration constants, q 1 and q 2 , are the nodal
displacements.
In matrix form
2
u= N q = N { q },
i =1
i i
e
(4.11)

where
N = N1 N 2 and {q }= { q
e
1 q2 }T . (4.12)

Thus, the displacement at any point within an element can be found by


multiplying the matrix of shape functions by the vector of nodal displacements.

{ }
In (4.11), q e is the column vector of element nodal displacements and
N is the row vector of displacement interpolation functions also named shape
4. BARS AND SHAFTS 51

functions. It is easy to check that u = q 1 at node 1 and u = q 2 at node 2, and that u


varies linearly (Fig. 4.3, c).
Equations (4.6) and (4.10) show that both the element geometry and the
displacement field are interpolated using the same shape functions, which is
referred to as the isoparametric formulation.

4.1.4 Element stiffness matrix in local coordinates

Strains can be expressed in terms of the shape functions as

x =
du d
=
dx dx
N q = B q
e e
{ } { } (4.13)

where
d
B = d x N (4.14)

is the row vector of the derivatives of shape functions, generally called the element
strain-displacement matrix. It gives the strain at any point due to unit nodal
displacement.
The transformation from x to r in equation (4.5) yields
x 2 x1 le
dx = dr = dr , (4.15)
2 2

where 1 r +1 and the length of the element is l e = x 2 x 1 .

The element strain energy U e is

1 1
E
2
Ue = x x dV = e x dV . (4.16)
2 2
e e

Substituting (4.13) in (4.16) we obtain

Ue =
1
2
{ q } B
e T T
Ee B dV { q }.
e
(4.17)
Ve
144
42444
3

The above equation is of the form

Ue =
1
2
{ q } [ k ] { q },
e T e e
(4.18)

[ ]
where the element stiffness matrix k e is given by
52 FINITE ELEMENT ANALYSIS

[ k ] = B
e T
Ee B dV . (4.19)
Ve

[ ]
This form guarantees that k e will be a symmetric matrix.
In terms of the shape functions

[k ]= E A
T
e dN dN
e e dx . (4.20)
le
dx dx

dN dN d r 2 dN
Because = = , we can write
dx dr d x l e dr
+1 T
dN dN
[ ke ] = 2 Ee Ae dr . (4.21)
le
1
dr dr

d N1 1 d N2 1
Substituting = and = + , equation (4.21) yields
dr 2 dr 2
+1
[k ]
e 2E A
= e e
le
1 2 1
1 2 2
1
2
dr
1
or

[ k ] = El A 11
e e e 1
1
. (4.22)
e
which is the same as equation (3.6).

4.1.5 Bar loaded between ends

For a prismatic bar acted upon by a uniformly distributed axial load,


p = const. , d 2u d x 2 = const . , so that the displacement field within the element
may be expressed as a quadratic polynomial

u (x ) = a + b x + c x 2 . (4.23)
The three integration constants, a, b and c, have to be determined from
three boundary conditions. This can be done if we use a three-node one-
dimensional element. An internal node is added at the midpoint to comply with the
requirement of a quadratic fit (Fig. 4.4, a).
4. BARS AND SHAFTS 53

For a linearly distributed axial load (as in a bar rotating at constant angular
speed around an end, acted upon by a distributed centrifugal load proportional to
the distance to the rotation centre), the true displacement field is given by a cubic
polynomial, involving four integration constants (see Example 4.8).
For an exact solution, this implies using a four-node element (adding two
internal nodes). The usual practice is to assume an approximate lower order linear
displacement field, i.e. a two-node element and to replace the actual linearly
distributed load by equivalent nodal forces, having thus an element not loaded
between ends describable by linear shape functions.

Fig. 4.4

Consider the three-node quadratic element from Fig. 4.4, a, with node 3 at
the midpoint. Nodal coordinates are x1 , x2 , x3 , and the vector of element nodal
{ }
displacements is q e = { q1 q2 q3 }T . The x-coordinate system is mapped
T

onto an intrinsic r-coordinate system, given by the transformation


2 ( x x3 )
r= . (4.24)
x2 x1
It comes out that r = 1 , 0, and + 1 at nodes 1, 3, and 2 (Fig. 4.4, b).
The displacements within the element can be written in terms of the three
nodal displacements q 1 , q2 , and q 3 as

u = N1 (r ) q 1 + N 2 (r ) q 2 + N 3 (r ) q 3 , (4.25)

where the quadratic shape functions N i ( i = 1, 2, 3 ) are


1 1
N1 (r ) = r ( 1 r ) , N 2 (r ) = r ( 1 + r ) ,
2 2 (4.26)
N 3 (r ) = ( 1 + r )( 1 r ) .
The graphs of the shape functions (4.26) are shown in Fig. 4.5. They have a
unit value at the node with the same index and zero at the other nodes. This is a
general property of the shape functions.
54 FINITE ELEMENT ANALYSIS

The expressions for these shape functions can be written down by


inspection. For example, since N1 = 0 at r = 0 and r = 1 , we know that N1 has to
contain the product r ( 1 r ) , i. e. the left hand part of the equations of the vertical
lines passing through nodes 3 and 2. That is, N1 is of the form N1 = C r ( 1 r ) .
The constant C is obtained from the condition N1 = 1 at r = 1 , which yields
C = 1 2 , resulting in the expression given in (4.26).

Fig. 4.5

The displacement field within the element is written in matrix form as


3
u= N q = N { q } ,
i =1
i i
e
(4.27)

where
N = N1 N 2 N3 and { q }= { q
e
1 q2 q3 }T . (4.28)

At any point within an element the axial displacement can be found by


multiplying the matrix of shape functions by the vector of nodal displacements, as
in (4.27). It is easy to check that u = q 1 at node 1, because N1 = 1 and
N 2 = N3 = 0 . Similarly, u = q 2 at node 2, and u = q 3 at node 3. Thus, u is a
quadratic polynomial passing through q1 , q2 , and q3 (Fig. 4.6). It is obtained by
interpolation, using quadratic shape functions.
4. BARS AND SHAFTS 55

The element strain-displacement row vector B in (4.14) is given by

d d dr 2 1 2r 1+ 2r
B = d x N = d r N d x = l 2 2
2 r . (4.29)
e

Fig. 4.6

The element stiffness matrix (4.19) is


+1

[k ]e Al
= e e
2 B T
Ee B d r
1
or, substituting (4.29),
7 1 8
[k ]e
=
Ee Ae
3l e 1 7 8 . (4.30)
8 8 16

4.1.6 Vector of element nodal forces

Consider an axial load p (x ) , having the units of force per unit length,
distributed along the bar element. The mechanical work of such a force is

W= u p dx = u
T
p dx . (4.31)
le le

Substituting (4.11), equation (4.31) becomes

W = qe { } N T T
p dx . (4.32)
le
It has the form
56 FINITE ELEMENT ANALYSIS

W = qe { } {f }
T e
(4.33)
where the element equivalent load vector is
+1
{ f }= N
e T
p dx =
le
2 1 N
T
p dr . (4.34)
le

For the linear two-node element, if the axial force is uniformly distributed,
p = const . , then
+1

{f }e l
= e p
2 N d r
T
(4.35)
1
or, substituting (4.7),

{ f }= p2l
e e 1
. (4.36)
1
p le
A force , equal to half the total force on the element, is applied at each node.
2
For the quadratic three-node element, if p = const. , substituting the shape
functions (4.26) into (4.35), gives

{ f }= p l
e
e 1 6 1 6 2 3 .
T
(4.37)

4.1.7 Assembly of the global stiffness matrix and load vector

Assembly of the system stiffness matrix for one-dimensional structures


modelled as bars is carried out as shown in sections 3.4 to 3.7 for trusses.
Consider the five-node finite element model in Fig. 4.7, a. Each node has
only one degree of freedom in the x-direction. The nodal displacements are Q1 ,
Q2 ,, Q5 (Fig. 4.7, b). The global vector of nodal displacements is denoted by

{Q } = Q1 Q2 Q3 Q4 Q5 T .

The nodal forces are F1 , F2 ,, F5 . The global load vector is denoted by

{ F } = F1 F2 F3 F4 F5 T .

The unreduced global stiffness matrix [ K ] plays the role of a


proportionality factor between the two global vectors
4. BARS AND SHAFTS 57

{ F } = [ K ] {Q }.

Fig. 4.7

The assembly of [ K ] from the element stiffness matrices can be explained


by an energy approach. Consider the strain energy in, say, element 3. We have

U3 =
1
2
{ q } [k ] { q }
3 T 3 3

or
1 EA3 1 1 Q3
U3 =
2
Q3 Q4
l 3 1
.
1 Q4

We can write U 3 as

0 0 0 0 0
0 0 Q1
0 0 0
EA3 EA

Q2
1 0 0 3 0
U3 = Q1 Q2 Q3 Q4 Q5 l3 l3 Q3
2
EA EA3 Q4
0 0 3 0
l3 l3 Q5
0
0 0 0 0
or

U3 =
1
{ Q }T [ K~ 3 ] { Q },
2
58 FINITE ELEMENT ANALYSIS

~
[ ]
where K 3 is the expanded stiffness matrix of element 3. It has the size of the
[ ]
global stiffness matrix, but the elements of the matrix k 3 occupy the third and
fourth rows and columns of the [ K ] matrix.
The strain energy of the entire structure is equal to the sum of the element
strain energies

e U e = e 2 { Q } [ K e ]{ Q }= 2 { Q } e [ K e ] { Q }
1 T ~ 1 T ~
U=

so that, as in (3.27), the global stiffness matrix is equal to the sum of the expanded
element stiffness matrices

[
[ K ] = K~ e . ]
e
This is a convenient algebraic explanation of the assembly process, which
is never done in practice. The entries of the element matrices k e are placed in [ ]
the appropriate locations of the global [ K ] matrix by the so-called direct method,
based on element connectivity. Overlapping elements are simply added as already
shown in section 3.4 for truss elements. This is based on the simple addition of
element strain energies.
The element matrices can be written
1 2 2 3

[ k ] = ElA 11
1 1 1 1
1 2
, [ k ] = ElA
2 2 1 1 2
1 1 3 ,
1 2
3 4 4 5

[ k ] = ElA
3 3 1 1 3
1 1 4
, [ k ] = ElA
4 4 1 1 4
1 1 5
.
3 4
At the top and on the right of the element stiffness matrices, the numbering
of coordinates in the global stiffness matrix is shown, according to the connectivity
Table 4.1.
Table 4.1
Node
Element
i j
1 1 2
2 2 3
3 3 4
4 4 5
4. BARS AND SHAFTS 59

The result is the following unreduced global stiffness matrix

A1 A1
l 0 0 0
l1
1
A1 A1 A2
+
A2
0 0
l1 l1 l 2 l2

[ K ]= E 0 A
2
A2 A3
+
A3
0 .
l2 l2 l3 l3
A A3 A4 A4
0 0 3 +
l3 l3 l4 l4
0 0 0
A
4
A4
l4 l 4

The global load vector is assembled as

{F }= R1 F2 0 0 F5 T .

The final finite element equations are obtained, as shown in section 3.5,
using the boundary condition, Q 1 = 0 . The reduced (non-singular) stiffness matrix
is obtained deleting the first row and column of the unreduced matrix, and the
reduced load vector - by deleting the first element. In fact, this information is
stored for the subsequent calculation of the reaction R 1 . Stresses are then
calculated from the axial forces obtained using equation (3.41).

4.1.8 Initial strain effects

Let an initial strain 0 be induced in a bar element. It may arise from


thermal action or by forcing members into place that are either too short or too
long, due to fabrication errors.
The stress-strain law in the presence of 0 is of the form

= E ( 0 ) . (4.38)
The mechanical work of external nodal forces applied to suppress the
initial prestressing due to 0 is

W=

Ve
0 dV =

Ve
T
0 dV = 0 Ee Ae

le
T
dx . (4.39)

Substituting (4.13), the expression (4.39) becomes


60 FINITE ELEMENT ANALYSIS

W = qe { } T
0 Ee Ae B
T
dx . (4.40)
le
It has the form (4.33), where the element load vector due to initial
straining is
+1
{ f }=
e
0 Ee Ae B
T l
dx = 0 Ee Ae e
2 B
T
dr (4.41)
le 1
or
+1 +1

{ f }=
T
dN 1

le
d r = 0 Ee Ae
e
0 Ee Ae dr ,
dr 2 1
1 1
hence

{ f }=
e
0 Ee
1
Ae . (4.42)
1
After solving for nodal displacements, stresses are computed as
q2 q1
e = Ee + ( Ee 0 ) . (4.43)
le
In the case of thermal loading,
0 = T , (4.44)
where is the coefficient of thermal expansion and T is the average change in
temperature within the element.

4.2 Plane shaft elements


From Mechanics of Materials it is known that a uniform shaft of diameter
d and length l , from a material with shear modulus of elasticity G, acted upon by a
Mt l d4
torque M t will twist an angle = , where I p = is the polar second
GIp 32
moment of area of the shaft cross section. The shaft torsional stiffness is then
M GIp
K= t = .
l
A two-node shaft finite element, of length l and torsional rigidity G I p , is
shown in Fig. 4.8. The nodal torques M 1 and M 2 can be related to the nodal
rotation angles 1 and 2 using the equilibrium and the torque/rotation equations
4. BARS AND SHAFTS 61

M1 = M 2 = K 1 when 2 = 0,
(4.45)
M 1 = M 2 = K 2 when 1 = 0.

Equations (4.45) may be written in matrix form as


M1 K K 1
= (4.46)
M 2 K K 2

or in shorthand form

{ M }= [ k ] { },
e e e

where

[ k ] = GlI
e p 1 1
1 1 (4.47)

is the stiffness matrix of the shaft element.

Fig. 4.8

The derivation of the shaft stiffness matrix is essentially identical to the


derivation of the stiffness matrix for an axially loaded bar element. Similarity
between these two derivations occurs because the differential equations for both
problems have the same mathematical form. The differential equation for torsional
displacement is
d Mt
= , (4.48)
dx GI p
while for the axial displacement is (4.3)
du N
= . (4.49)
dx E A
The rotation angle of an arbitrary section within the shaft element can be
expressed in terms of the nodal rotations 1 and 2 as

= N1 (r ) 1 + N 2 (r ) 2 . (4.50)
62 FINITE ELEMENT ANALYSIS

where
1
N1 (r ) = ( 1 r ) and N 2 (r ) = 1 ( 1 + r ) (4.51)
2 2
are the shape functions of the shaft element, the same as for the axially loaded bar
element.
Analogous to equation (4.20), the stiffness matrix for the shaft element can
be calculated from

[ k ] = G I ddNx
T
e dN
e pe dx dx (4.52)

le
or
+1 T
2 Ge I p e dN dN
[ ke ] =
le
1
dr dr dr

(4.53)

which yields

[ k ] = G lI
e e pe 1 1
1 1 (4.54)
e
Equation (4.54) is also used to account for torsional effects in grid finite
elements, as equation (4.22) is used to account for axial effects in inclined beam
finite elements. For non-axially-symmetric cross sections, the polar second
moment of area I p is replaced by the torsional constant I t .

Example 4.1
Consider the bar in Fig. E4.1 with d = 5 mm , l = 0.2 m , E = 200 GPa ,
loaded by an axial load F = 2 kN . Determine: a) the displacement of point 2, b) the
stresses in bar, and c) the support reactions.

Fig. E4.1

Solution. a) The bar is divided into three bar finite elements.


4. BARS AND SHAFTS 63

The areas of the cross sections are


d2 ( 1,4 d )2
A1 = = 19.62 mm 2 , A2 = A3 = = 38.47 mm 2 .
4 4
The element stiffness matrices are

[ k ] = 2 1040019.62 11 1 1 1 N ,
5
1
= 9.81 103
1 1
1 mm

[ k ] = [ k ] = 2 10200 38.47 11 1 1 1 N
5
2 3
= 38.47 103 .
1 1
1 mm
The global unreduced stiffness matrix is
9.81 9.81 0 0
9.81 9.81 + 38.47 38.47 0
[ K ] = 10 0
3
38.47 38.47 + 38.47 38.47
.

0 0 38.47 38.47

Including the boundary conditions (degrees of freedom 1 and 4 are fixed),


the finite element equations can be written
9.81 9.81 0 0 0 R1
9.81 48.28 38.47 0 Q2 0

103 = 3 .
0 38.47 76.94 38.47 Q3 2 10

0 0 38.47 38.47 0 R4

Retaining only the second and third equation yields

48.28 38.47 Q2 0
103 = 3
38.47 76.94 Q3 2 10
with solutions
Q2 = 34.42 mm , Q3 = 43.21 mm .

b) The reactions are obtained from the first and fourth equation

R1 = 9.81 Q2 = 9.81 103 34.42 = 337.8 N ,

R4 = 38.47 Q3 = 38.47 103 43.21 = 1662 .2 N .

c) Stresses are
64 FINITE ELEMENT ANALYSIS

Q2 Q1 Q 34.42 N
12 = E 12 = E = E 2 = 2 105 = 17.2 ,
l1 2l 400 mm 2
Q3 Q2 Q Q2 8.79 N
23 = E 23 = E =E 3 = 2 105 = 8.79 ,
l2 l 200 mm 2
Q4 Q3 Q 43.21 N
34 = E 34 = E = E 3 = 2 105 = 43.2 .
l3 l 200 mm 2

Example 4.2
A support, consisting of a steel bar 1 fitted inside a cast iron tube 2, is
loaded by a 60 kN axial force, as in Fig. E4.2. Consider A1 = 200 mm 2 ,
A2 = 800 mm 2 , E1 = 210 GPa , E2 = 120 GPa , l = 0.2 m . Determine: a) the
elongation of the assembly, b) the stress in each material, and c) the internal forces
in bar and tube.

Fig. E4.2

Solution. a) Using a two-node two-element finite element model, the


element stiffness matrices are

[ k ] = 2.110200 200 11 1 1 1 N ,
5
1
= 2.1 105
1 1
1 mm

[ k ] = 1.2 10200 800 11 1 1 1 N


5
2
= 4.8 105 .
1 1
1 mm
The global unreduced stiffness matrix is
1 1 N
[ K ] = 105 2.2.11 + 44..88 2.1 4.8

2.1 + 4.8
= 6.9 105 1 1 mm
.

4. BARS AND SHAFTS 65

If point 1 is fixed, the finite element equations are


1 1 0 R1
6.9 105 = 4 ,
1 1 Q2 6 10
Q2 = 0.087 mm ,
which is equal to the beam shortening.
b) Stresses are
Q2 Q1 Q 0.087 N
1 = E1 1 = E1 = E1 2 = 2.1 105 = 91.3 ,
l l 200 mm 2
Q2 Q1 Q 0.087 N
2 = E2 2 = E2 = E2 2 = 1.2 105 = 52.14 .
l l 200 mm 2
c) Internal forces are

N1 = 1 A1 = 91.3 200 = 18.26 kN ,

N 2 = 2 A2 = 52.14 800 = 41.74 kN .

Example 4.3
The temperature of the bar in Fig. E4.3 is raised 800 C . If E = 130 GPa
and = 17 106 , determine the thermal stresses.

Fig. E4.3

Solution. We should first determine whether contact occurs between the


bar and the wall. If the wall does not exist, the displacement of point 2 is

l = l T = 1800 17 10 6 80 = 2.448 mm > 2 mm


so that the contact does occur.
Denoting by A the cross section area, the temperature forces (4.42) are
66 FINITE ELEMENT ANALYSIS

1 1 1
{ fT } = T E A 6 5
= 17 10 80 1.3 10 A = 176.8 A N .
1 1 1
The stiffness matrix is

A 1 1 N .
5
[ k ] = 1.3 10
1800 1 1 mm

As point 1 is fixed, the finite element equations are


1.3 105 A 1 1 0 R 1 176.8 A
1800 1 1 2 = R + 176.8 A
2
wherefrom we get the reactions
R 1 = 176.8 A 144.4 A = 32.4 A , R 2 = 32.4 A ,

so that the normal stress is


R2 N
= = 32.4 .
A mm 2

Example 4.4
A prismatic bar is made of two different materials, as shown in Fig. E4.4.
The temperature of the central part is raised T 0C . Determine the axial stress in the
bar.

Fig. E4.4

Solution. The bar is divided into three linear finite elements. The element
stiffness matrices are

[ k ] = [ k ] = El A 11
1 3 2 1 ,
1
[ k ] = E2lA 11
2 1 1 .
1

For element 2, the vector of nodal thermal forces is

{ f }= T E A 1
2
1 1T .
4. BARS AND SHAFTS 67

Using the boundary conditions Q1 = Q4 = 0 , the finite element equations


can be written
E2 E2 0 0
E E 0 R1
E2 + 1 1 0 Q T E A
A E2 2 2 2 1
= .
T E1 A
l 0 E1 E1
+ E2 E2 Q3
2 2 0
E2
R4
0 0 E2

The second and third equation yield

E1 E1
A E2 + 2
2 Q2 = T E A 1
l
E E 1
1

1 E2 + 1 Q3
2 2
with solutions
E1 E1
Q2 = T l , Q3 = T l .
E1 + E2 E1 + E2
The strains in elements are
Q2 Q E1 Q3 Q2 E1
1 = 3 = = 4 = T , 2 = =T .
l l E1 + E2 2l E1 + E2
Stresses are
1 1
1 = 3 = E2 1 = T , 2 = E1 2 E1 T = T = 1.
1 1 1 1
+ +
E1 E2 E1 E2

Example 4.5
A steel bolt of active length l = 100 mm and diameter = 10 mm is single
threaded with a 1.6 mm pitch. It is mounted inside a copper tube with diameters
d = 12 mm and D = 18 mm (Fig. E4.5, a). After the nut has been fitted smugly, it
is tightened one-quarter of a full turn. Determine stresses in bolt and tube, if for
steel E1 = 208 GPa and for copper E2 = 100 GPa .

Solution. The assembly is modeled by two bar finite elements as in Fig.


E4.5, b. Both elements have fixed ends at points 1 and 4 so that Q1 = Q4 = 0 .

The problem has a multipoint constraint


68 FINITE ELEMENT ANALYSIS

Q3 Q2 = 0.4 mm .
Such conditions are programmed using a so-called penalty approach.
Herein a simpler straightforward solution is given.

Fig. E4.5

The cross section areas are

A1 =
2
= 78.54 mm 2 , A2 =
(
D2 d 2 ) = 141.37 mm 2
.
4 4
The element stiffness matrices are

[ k ] = 2.08 10100 78.54 11 1 1 1 N ,


5
1
= 1.63 105


1 1 1 mm

[ k ] = 10 141.37 1 1 1 1 N
5
2
= 1.41 105 .
100
1 1 1 1 mm
The finite element equations are
1.63 1.63 0 0 0 R1
1.63 1.63 0 0 Q2 F

105 = .
0 0 1.41 1.41 Q3 F

0 0 1.41 1.41 0 R4

Retaining only the second and third equation yields


1.63 0 Q2 F
105 =
0 1.41 Q3 F
with solutions
4. BARS AND SHAFTS 69

F F
Q2 = 5
, Q3 = .
1.63 10 1.41 105
Substituting in the multipoint constraint condition gives
1 1
F 5
+ 5
= 0.4 ,
1.63 10 1.41 10
or
F = 30314 N .
Stresses are given by
F 30316 N
1 = = = 386 ,
A1 78.54 mm 2
F 30316 N
2 = = = 214.5 .
A2 141.37 mm 2

Example 4.6
For the bar of Fig. E4.6, with linearly variable cross-section
x
A ( x ) = A0 1 + , find the displacement at the free end under the action of force
2l
F, using two tapered bar finite elements.

Fig. E4.6

Solution. The stiffness matrix for a bar element with variable cross section
is

[ k ] = lE
e
2
1 1
1 1 A dx .
e
le
For the bar of Fig. E4.6, divided into two equal length tapered elements,
the element stiffness matrices are
70 FINITE ELEMENT ANALYSIS

[k ]
1 E 1 1
= 2
l 1 1
A0
x
1 + dx =
2l
5 E A0 1 1
4l 1 1
,
0

2l

[k ]
2 E 1 1
= 2
l 1 1
A0

x
1 + dx =
2l
7 E A0
4l
1 1
1 1
.

l

With Q1 = 0 , the finite element equations can be written

5 5 0 0 R1
EA0
5 12 7 Q2 = 0 .
4l
0 7 7 Q3 F

From the second and third equation

EA0 12 7 Q2 0
4l 7 7 Q = F ,
3
wherefrom
4 Fl 48 F l
Q2 = , Q3 = .
5 E A0 35 E A0
The approximate assumed linear variation of the displacement field yields
constant strain elements, hence a stepwise variation of stresses along the bar.

Example 4.7
Write the stiffness matrix of the bar shown in Fig. E4.7, condensing Q3
from condition F3 = 0 . Model the bar by: a) two two-node linear elements; b) a
three-node quadratic element. Comment the results.

Fig. E4.7

Solution. a) Consider the bar divided into two linear elements. The element
stiffness matrices are
4. BARS AND SHAFTS 71

[ k ] = [ k ] = 2EA
1 2 1
l 1
1 .
1

The unreduced global stiffness matrix is
1 1 0
[ K ] = l 1 2 1 .
2 EA

0 1 1

The finite element equations can be written


1 0 1 Q1 F1
2 EA
0 1 1 Q2 = F2 .
l
1 1 2 Q3 F3
For F3 = 0 , the last equation yields

Q1 + Q2 Q
Q3 = = 1 2 1 2 1
2 Q2
which assumes a linear displacement field within the bar.
The first two equations give
2 EA 1 0 Q1 2 EA 1 F1
l 0 1 Q + l 1 Q3 = F
2 2
and upon substitution of Q3

2 EA 1 0 Q1 2 EA 1 Q1 F1
l 0 1 Q + l 1 1 2 1 2 Q = F
2 2 2
which can be written
EA 1 1 Q1 F1
l 1 1 Q = F
2 2
where the left hand side contains the (2 2) stiffness matrix of the two-node linear
element.
b) Consider the bar modeled by a 3-node quadratic bar element. The finite
element equations can be written
72 FINITE ELEMENT ANALYSIS

7 1 8 Q1 F1
EA
1 7 8 Q2 = F2 .
3l
8 8 16 Q3 F3
Condensing Q3 from condition F3 = 0 yields

Q1 + Q2 Q
Q3 = = 1 2 1 2 1 .
2 Q2
The first two equations give
EA 7 1 Q1 EA 8 F1
+ Q3 =
3l 1 7 Q2 3l 8 F2
and upon substitution of Q3

EA 7 1 Q1 EA 8 Q1 F1
+ 1 2 1 2 =
3l 1 7 Q2 3l 8 Q2 F2
which can be written again
EA 1 1 Q1 F1
= ,
l 1 1 Q2 F2

where the matrix in the left hand can be recognized as the conventional bar matrix
developed from linear polynomials.
c) Comments. Substituting
Q1 1 0
Q
Q2 = 0 1 1
Q 1 2 1 2 Q2
3
into the expression of the strain energy (4.18) gives
7 1 8 Q1
1 EA
U e = Q1 Q2 Q3 1 7 8 Q2 ,
2 3l
8 8 16 Q3

7 1 8 1 0
1
U e = Q1 Q2
1 0 1 2 EA Q1
2
0 1 1
2 3l
1 7 8 0 1 Q ,
8 8 16 1 2 1 2 2
4. BARS AND SHAFTS 73

1 EA 1 1 Q1
Ue =
2
Q1 Q2 l 1 1 Q .
2
For p = const . , the work of nodal forces is

F1 1 6

W = Q1 Q2
e
Q3 F2 = Q1 Q2 Q3 1 6 pl e ,
F 2 3
3
1 6
1 0 1 2 1 pl
W = Q1 Q2
e
0 1 1 2 1 6 pl e = Q1 Q2 1 2 e .
2 3

We may conclude that the introduction of the quadratic term in equation
(4.23) does not bring about a change in the conventional stiffness matrix and load
vector. Whenever the assumed functions, used to describe the displacement field,
form the complete homogeneous solution of the differential equation of
equilibrium (4.4), the developed stiffness matrix and the equivalent load vector will
be exact. This is because, as it is shown in a next chapter, only the homogeneous
part of the solution contains the free parameters with respect to which the total
potential energy is minimized. The parameters in the particular part of the solution
are prescribed and do not take part in the process of minimization.
The exactness of the stiffness matrix and load vector also implies that the
computed nodal displacements will also be exact. However, displacements within
elements depend upon the general (homogeneous plus the particular) solution. The
conventional formulation based on a linear polynomial will yield exact
displacements within the elements only when p = 0 . For the case p = const. , exact
displacements within the elements may be obtained from equation (4.25).
However, before using equation (4.25), the variable Q3 must be computed from
the exact nodal displacements (computed for the conventional linear element), via a
constraint equation between Q3 and the remaining nodal variables, obtained by a
minimization of the total potential energy with respect to Q3 at element level.

Example 4.8
Consider a prismatic robot arm, rotating at constant angular velocity
= 30 rad sec (Fig. E4.8, a). Determine the axial stress distribution due to the
centrifugal force in the rod using: a) two quadratic elements, and b) three linear
elements.
Solution. The beam is acted upon by a centrifugal linearly distributed load
74 FINITE ELEMENT ANALYSIS

p ( x ) = p0
x
,
l
where p0 = Al 2 .

Fig. E4.8
4. BARS AND SHAFTS 75

a) A finite element model with two quadratic elements is shown in Fig.


E4.8, b. The model has five degrees of freedom.
The element stiffness matrices are
7 8 1
[k ]= [k ]
1 2
=
2 EA
3l 8 16 8 .
1 8 7

The unreduced global stiffness matrix is


7 8 1 0 0
8 16 8 0 0

[ K ] = 23EA 1 8 7 + 7 8 1 .
l
0 0 8 16 8
0 0 1 8 7

The axial load acting on the two elements can be decomposed as in Fig.
E4.8, c. The nodal forces, equivalent to a linearly distributed load, are given by
equation (4.34)
1 (1 2 )(1 ) 0
{ f }= N
e T
p dx = p0 l e


(1 2 ) d = p0 l e 1 6 ,
4 (1 ) 1 3
le 0
where = x l e and the shape functions (4.26) are defined for convenience over an
interval [0, 1] . For a uniformly distributed load they are given by equation (4.37).

The distributed loads from Fig. E4.8, c are replaced by the kinematically
equivalent nodal forces shown in Fig. E4.8, d. The element nodal load vectors are

{ f }= p12l 0
1 0
1 1 2 T , { f }= p12l 1 2
2 0
3 1T .

Using the boundary condition Q1 = 0 , the finite element equations can be


written
7 8 1 0 0 0 R1
8 16 8 0 Q p0l 12
0 2
2 EA
1 8 14 8 1 Q3 = p0l 24 + p0l 24 .
3l
0 0 8 16 8 Q4 3 p0l 12

0 0 1 8 7 Q5 p0l 12
76 FINITE ELEMENT ANALYSIS

Deleting the first row and column, we obtain the reduced global stiffness
matrix and reduced global load vector which yield
47 p0 l 2 88 p0 l 2 117 p0 l 2 128 p0 l 2
Q2 = , Q3 = , Q4 = , Q5 = .
384 E A 384 E A 384 E A 384 E A
The nodal displacements can be compared with the exact solution, given
by

p0 l 2 x x3
u( x ) = .
l 3l 3
(a)
2E A
We obtain
2 2 2
l 47 p0 l l 11 p0 l 3l 117 p0 l 1 p0 l 2
u = , u = , u = , u (l ) = .
4 384 2 E A 2 48 2 E A 4 384 2 E A 3 2E A
The finite element values of the nodal displacements are exact. This is due
to the nodal equivalence of forces. While the nodal displacements are exact, the
displacements within the elements are approximate because the exact cubic
distribution (a) has been replaced by a quadratic law.
The element strain-displacement row vector B in (4.29) is given by

d 1
B = d x N = l 3 + 4 1 + 4 4 8 .

The strains are calculated as


2
1 = (N 2 Q3 + N 3 Q2 ) = p0l ( 25 6 ) ,
l 48 EA
2
2 = (N1 Q3 + N 2 Q5 + N 3 Q4 ) = p0l ( 19 18 ) ,
l 48 EA
which yields the following nodal values
25 p0l 22 p0 l 19 p0 l 10 p0 l 1 p0 l
1 = , 2 = , 3 = , 4 = , 5 = .
48 EA 48 EA 48 EA 48 EA 48 EA

The corresponding axial stresses are


25 p0l 22 p0 l 19 p0 l 10 p0 l 1 p0 l
1 = , 2 = , 3 = , 4 = , 5 = .
48 A 48 A 48 A 48 A 48 A

The stresses can be compared with the exact solution


4. BARS AND SHAFTS 77

p0 l 2
( x )= 1 x ,

(b)
2A l 2

which yields the following nodal values


24 p0 l l 22.5 p0 l
(0) = , = ,
48 A 4 48 A

l 18 p0 l 3l 10.5 p0 l
= , = , (l ) = 0 .
2 48 A 4 48 A
A graph of the stress distribution is shown in Fig. E4.8, e.
b) A finite element model comprised of three linear elements is shown in
Fig. E4.8, f. The model has four degrees of freedom.
The element stiffness matrices are

[ k ] = [ k ] = [ k ] = 3EA
1 2 3 1
l 1
1 .
1

The axial load acting on the three elements can be decomposed as in Fig.
E4.8, g. The nodal forces, equivalent to a load per unit length, are given by
equation (4.34). If p1 and p2 are the intensities of a linearly distributed load at
nodes 1 and 2, respectively,

p ( ) = p1 + ( p2 p1 ) ,
1

{ f }= N
e T
p dx = l e

1
[ p1 + ( p2 p1 ) ] d =
l e 2 p1 + p2


,
6 p1 + 2 p2
le 0

where = x l e and the shape functions (4.26) are defined for convenience over an
interval [ 0, 1 ] . For p1 = p2 = p , the nodal forces are given by equation (4.36).
The distributed loads from Fig. E4.8, g are replaced by the kinematically
equivalent nodal forces shown in Fig. E4.8, h. The element nodal load vectors are

{ f }= p54l 1
1 0
2 T , { f }= p54l 4
2 0
5 T , { f }= p54l 7
3 0
8 T .

Using the boundary condition Q 1 = 0 , the finite element equations can be


written
78 FINITE ELEMENT ANALYSIS

54 R1
1 1 0 0 0 1+ p l

p0 l
0
3EA 1 2 1 0 Q2 6
= .
l 0 1 2 1 Q3 54 12

0 0 1 1 Q4
8

The last three equations yield


13 p0 l 2 23 p0 l 2 27 p0 l 2 1 p0 l 2
Q2 = , Q3 = , Q4 = = .
81 E A 81 E A 81 E A 3 E A

The element strain-displacement row vector B in (4.29) is given by

d 1
B = d x N = l 1 1 .
e

The strains are calculated as


3 13 p0 l 3 10 p0 l 3 4 p0 l
1 = Q2 = , 2 = ( Q3 Q2 ) = , 3 = ( Q4 Q3 ) = ,
l 27 EA l 27 EA l 27 EA
and are constant within each element.
The corresponding axial stresses are
13 p0l 10 p0 l 4 p0 l
1 = , 2 = , 3 = .
27 A 27 A 27 A
In Fig. E4.8, i they are compared to those given by equation (b).

Example 4.9
Show that the shape functions of the four-node cubic bar element, in local
natural coordinates, are the following

N1 (r ) =
9
( 1 r ) 1 + r 1 r ,
16 3 3

N 2 (r ) =
27
( 1 + r ) ( 1 r ) 1 r ,
16 3

N 3 (r ) =
27
( 1 + r ) ( 1 r ) 1 + r ,
16 3
9 1 1
N 4 (r ) = + r r (1 + r ) .
16 3 3
5.
BEAMS, FRAMES AND GRIDS

Frames are structures with rigidly connected members called beams.


Beams are slender members used to support transverse loading. They are connected
by rigid joints that have determinate rotations and, apart from forces, transmit
bending moments from member to member.
One-dimensional mathematical models of structural beams are constructed
on two beam theories: the Bernoulli-Euler beam theory, that neglects transverse
shear deformations, and the Timoshenko beam theory, that incorporates a first order
correction for transverse shear effects. Beam behaviour is described by fourth order
differential equations and require C1 continuity. This requires that both transverse
displacements and slopes must be continuous over the entire member and, in
particular, between adjacent beam elements. The Timoshenko beam model pertains
to the class of C 0 elements. It is based on the assumption that plane sections
remain plane but not necessarily normal to the deformed neutral surface. This leads
to the introduction of a mean shear distortion, which is constant over the element.
In this section, we first present the finite element formulation for plane
Bernoulli-Euler beams, then extend it to plane frames, and for grids. An inclined
beam element will be referred to as a frame element. Grids are planar frames
subjected to loads applied normally to their plane.

5.1 Finite element discretization

A plane frame is divided into elements, as shown in Fig. 5.1. Each node
has three degrees of freedom, two linear displacements and a rotation. Typically,
the degrees of freedom of node i are Q3 i 2 , Q3 i 1 and Q3 i , defined as the
displacement along the X axis, the displacement along the Y axis and the rotation
about the Z axis, respectively.
80 FINITE ELEMENT ANALYSIS

Nodes are located by their coordinates in the global reference frame XOY
and element connectivity is defined by the indices of the end nodes. Elements are
modelled as uniform beams without shear deformations and not loaded between
ends. Their properties are the bending rigidity E I and the length l .

Fig. 5.1

In the following, the shape functions are established for the plane
Bernoulli-Euler beam element, then the element stiffness matrix is calculated first
in the local coordinate system, then in the global coordinate system. The latter are
expanded to the structure size, then simply added to get the global uncondensed
stiffness matrix. Imposing the boundary conditions, the reduced stiffness matrix
and load vector are calculated and used in the static analysis.

Fig. 5.2
5. BEAMS, FRAMES AND GRIDS 81

Consider an inclined beam element, as illustrated in Fig. 5.2, a, where the


nodal displacements are also shown. In a local physical coordinate system, the x
axis, oriented along the beam, is inclined an angle with respect to the global X
axis. Alternatively, an intrinsic (natural) coordinate system can be used.
The vector of nodal displacements in the local coordinate system is
{ q }= q
e
1 q2 q3 q4 q5 q6 T (5.1)
and the corresponding vector of element nodal forces can be written
{ f }= f
e
1 f2 f3 f4 f5 f6 T . (5.2)

Forces f 2, f 3, f 5, f 6 and the corresponding displacements


q 2 , q 3 , q 5 , q 6 describe the element bending (Fig. 5.2, b) while axial forces
f 1 , f 4 , and axial displacements q 1 , q 4 , describe the element stretching (Fig.
5.2, c). Their action is decoupled so that the respective stiffness matrices can be
calculated separately.

5.2 Static analysis of a uniform beam

Beams with cross sections that are symmetric with respect to the plane of
loading are considered herein (Fig. 5.3). Transverse shear deformations are
neglected, as in the Bernoulli-Euler classical beam theory. Only transverse loads
act upon the beam, axial forces are ignored.

Fig. 5.3

The axial displacement of any point on the section, at a distance y from the
neutral axis, is approximated by
dv
u = y = y, (5.3)
dx
82 FINITE ELEMENT ANALYSIS

where v is the deflection of the centroidal axis at x and = v is the cross section
rotation (or slope) at x . Axial strains are

du d2 v
x = = 2 y = y , (5.4)
dx dx
where v denotes the deformed beam axis curvature.
Normal stresses on the cross section are given by Hookes law
d2 v
x = E x = E y, (5.5)
dx2
where E is Youngs modulus of the material.
The bending moment is the resultant of the stress distribution on the cross
section

d2 v
M (x ) = x y dA = E I z
= EIz (5.6)
A
dx2
where I z is the second moment of area of the section with respect to the neutral
axis z. The negative sign above is introduced because M is considered positive if it
compresses the upper portion of the beam cross section. The product E I z is called
the bending rigidity of the beam.
The shear force is given by
dM d3 v
T (x ) = = E I z = E I z v III . (5.7)
dx dx 3
The transverse load per unit length is
dT d4 v
p (x ) = = EIz = E I z v IV . (5.8)
dx dx4
The differential equation of equilibrium is
d4 v
EIz = p (x ) . (5.9)
dx4
This is a fourth order differential equation and consequently four boundary
conditions are required, two at each end. They can be geometric or kinematic
boundary conditions, involving the transverse displacement and slope, and physical
boundary conditions, involving the shear and bending moment.
5. BEAMS, FRAMES AND GRIDS 83

5.3 Uniform beam not loaded between ends


For a uniform beam not loaded between ends, p = 0 and equation (5.9)
yields d 4 v d x 4 = 0 . Integrating four times, we obtain the deflection v described
by a third order polynomial

v (x ) = a 1 x 3 + a 2 x 2 + a 3 x + a 4 . (5.10)
For a free-free beam element, the four integration constants a 1 , a 2 , a 3 , a 4
in (5.10) can be determined from the geometric boundary conditions at each end
(Fig. 5.4):
at x = x1 , v = v1 = q2 , and d v d x = 1 = q3 ; (5.11, a)

at x = x 2 , v = v2 = q5 , and d v d x = 2 = q6 . (5.11, b)

Fig. 5.4
This gives

v1 x13 x12 x1 1 a1
2
1 3 x1 2 x1 1 0 a2
= 3 .
v2 x2 x 22 x2 1 a3

2 3 x 2 2 x2 0 a 4
2 1
On inversion, the integration constants a 1 , a 2 , a 3 , a 4 can be expressed in
terms of the nodal displacements v1 , 1 , v2 , 2 , so that the beam deflection can
be expressed in terms of the nodal displacements.

5.3.1 Shape functions

The transverse displacement v can be expressed in terms of the nodal


displacements as
84 FINITE ELEMENT ANALYSIS

{ }
v (x ) = N q e , (5.12)

where N is a row vector containing the shape functions, called Hermitian cubic
polynomials, and

{ q }= v
e
1 1 v2 2 T . (5.13)

Using natural coordinates, with r = 1 at node 1 and r = +1 at node 2,


2 x + x2
r= x 1 (5.14)
x 2 x1 2

the beam transverse displacement can be written

dv dv
v (r ) = N1 (r ) v 1 + N 2 (r ) + N 3 (r ) v 2 + N 4 (r ) . (5.15)
d r 1 d r 2

Because the coordinates transform by the relationship (5.14)


x1 + x 2 x 2 x1
x= + r
2 2
and since l e = x 2 x 1 is the length of the element,

le
dx = dr . (5.16)
2
Using the differentiation chain rule

d v le d v
= , (5.17)
dr 2 dx
equation (5.15) becomes

le dv l dv
v (r ) = N1 (r ) v 1 + N 2 (r ) + N 3 (r ) v 2 + N 4 (r ) e (5.18)
2 d x 1 2 d x 2
or
le l
v (r ) = N1 q2 + N 2 q3 + N 3 q5 + e N 4 q6 . (5.19)
2 2
In (5.12) the row vector of shape functions is

le le
N = N1 N2 N3 N4 . (5.20)
2 2
5. BEAMS, FRAMES AND GRIDS 85

The Hermitian shape functions are cubic polynomials which should satisfy
the boundary conditions given in Table 5.1, where primes indicate differentiation
with respect to r.
Table 5.1
N1 N 1 N2 N 2 N3 N 3 N4 N 4
r = 1 1 0 0 1 0 0 0 0
r = +1 0 0 0 0 1 0 0 1

Imposing the above conditions to cubic polynomials with four arbitrary


constants, we obtain the expressions of the beam element shape functions in terms
of r, graphically shown in Fig. 5.5

N 1(r ) =
1
4
( )
( 1 r ) 2 ( 2 + r ) = 1 2 3r + r 3 ,
4

N 2 (r ) =
1
4
(
( 1 r ) 2 ( 1 + r ) = 1 1 r r 2 + r 3 ,
4
) (5.21)

N 3(r ) =
1
4
( )
( 1 + r ) 2 ( 2 r ) = 1 2 + 3r r 3 ,
4

N 4 (r ) =
1
4
(
( 1 + r ) 2 ( 1 r ) = 1 1 + r r 2 r 3 .
4
)

Fig. 5.5
86 FINITE ELEMENT ANALYSIS

d v le
It is easy to check that at node 1, v = q 2 and = q 3 , while at node 2,
dr 2
d v le
v = q 5 and = q6 .
dr 2

5.3.2 Stiffness matrix

The strain energy U e of a beam element is


2
d 2v

E Ie dx .
Ue = (5.22)
2 d x2

e
Equation (5.17) yields
dv 2 d v d 2v 4 d 2 v
= and = .
d x le d r d x 2 l 2e d r 2
Substituting (5.12) we obtain

d 2v
dx 2
=
4 d 2 N

l 2e d r 2
{ q }. e
(5.23)

The square of the above quantity is calculated as

2 T T
d 2v

d x2
2
=d v
d x2
d 2v
= qe
d x2
{ } T 16
l 4e
d 2 N
2 { }
d 2 N e
2 q ,
d r d r

which can also be written


2
d 2v
= qe
d x2
{ } T 16
le
N T N q e .
4 r r
{ } (5.24)

On substituting (5.16) and (5.24) into (5.22) we get the element strain
energy
+1
Ue =
1
2
{q } e T 8E I e
l 3e N r N r d r { q }
T e
(5.25)
1

which has the form


5. BEAMS, FRAMES AND GRIDS 87

Ue =
1
2
{ q } [ k ] { q }.
e T e
B
e
(5.26)

Comparing (5.25) with (5.26) we obtain the element stiffness matrix due to
bending
+1
[ k ] = 8lEI
e
B 3
e T
N r N r d r (5.27)
e 1
or
(N1)2 N1N 2 N1N 3 N1N 4
+1
[ k ] = 8lEI N 2 N1 (N 2 ) N 2 N 3 N 2 N 4
2
e e
dr . (5.28)
B 3 N N N N (N 3 )2 N 3N 4
e
1 3 1 3 2

4 1
N N N 4 N 2 N 4 N 3 (N 4 )2
Substituting the shape functions (5.21) and performing the integration
yields the stiffness matrix due to bending in local coordinates

12 6l 12 6l
6l 4l 2 6l 2l 2
[k ]
e
B
EI
= 3e
l e 12 6l 12 6l
. (5.29)

6l 2l
2
6l 4l 2 e

An alternative way of deriving the matrix (5.29) is based on the general


formula (4.19)

[ k ] = B
e T
Ee B dV . (5.30)
Ve

The curvature (5.23) can be expressed in terms of the nodal


displacements as

=
d 2v
dx 2
=
4 d 2 N

l 2e d r 2
{ q }= B { q },
e e
(5.31)

where the curvature-displacement row vector B is

1 r r
B = l 6 l 3r 1 6 l 3r + 1 . (5.32)
e e e
Substituting (5.32) and d V = Ae d x into (5.30) gives the matrix (5.29).
88 FINITE ELEMENT ANALYSIS

5.3.3 Physical significance of the stiffness matrix

The element stiffness matrix (5.29) relates the vector of nodal forces to the
vector of nodal displacements

f2 12 6l 12 6l q2
f 3 6l 4l 2 6l 2l 2 q
EI e 3
= 3 . (5.33)
f5 le 12 6l 12 6l q5

f 6 e 6l 2 l
2
6l 4l 2 e q6
e

Consider a beam element with end 2 fixed ( q5 = q6 = 0 ) and end 1 having


a unit displacement along the global X-axis ( q2 = 1, q3 = 0) and zero rotation, as
in Fig. 5.6.

Fig. 5.6

Equation (5.33) can be written


f 2 k11 . . . q2 = 1
f k . . . q3 = 0
3 21
= (5.34)
f 5 k31 . . . q5 = 0
f 6 k 41 . .

. q6 = 0
which gives
f 2 = k11 , f 3 = k 21 , f 5 = k31 , f 6 = k 41 . (5.35)
This shows that the first column of the stiffness matrix represents the
forces and moments that must be applied to the beam element to preserve static
equilibrium when q2 = 1 and all other displacements are zero.
For equilibrium
k11 + k31 = 0 , k 21 + k 41 + k31 l = 0 . (5.36)
5. BEAMS, FRAMES AND GRIDS 89

5.4 Uniform beam loaded between ends

For a uniform beam loaded between ends, p 0 , d 4 v d x 4 0 in


equation (5.9) and the beam deflected shape is no more a cubic polynomial.
However, it is the homogeneous solution of the differential equation. Using cubic
polynomials as admissible functions, the computed nodal displacements are exact.
Within the elements, the displacements, moments and shear forces are in error.
When the transverse load is uniformly distributed, p = const . , the general
solution of equation (5.9) is a quartic polynomial. The corresponding five constants
have to be determined from five boundary conditions. An internal node added at
the centre of element will solve the problem, introducing its nodal displacement as
the fifth nodal coordinate.
For a linearly distributed transverse load, v (x ) will be a quintic with six
arbitrary constants. They can be determined, adding the transverse displacement
and slope at the element midpoint to the element nodal coordinates. As already
shown in Chapter 4, rising the power of the function describing the displacement
within a beam element is tantamount to introducing additional internal nodes.
However, the current practice is to use lower order approximate assumed
shape functions that ensure the minimal convergence requirements, as shown in the
following. The cubic shape functions do the job. But they can be used only if the
element has uniform rigidity E I z and is not loaded between nodes. For beams
with transverse forces, the solution is to replace the actual distributed load by
equivalent nodal forces.

5.4.1 Consistent vector of nodal forces

Consider a transverse load p (x ) , having the units of force per unit length,
distributed along the beam element. The mechanical work of such a force is

W=
v p dx = v
le le
T
p dx . (5.37)

Substituting (5.12), equation (5.37) becomes

{ } N
W = qe
T T
p dx . (5.38)
le
It has the form

{ } {f }
W = qe
T e
(5.39)
where the element load vector is
90 FINITE ELEMENT ANALYSIS

+1

{ f } = N
e T l
p ( x ) dx = e
2 N
T
p (r ) d r . (5.40)
le 1

For the Hermitian two-node element, if the transverse force is uniformly


distributed, p = const . , the vector of element consistent nodal forces is
+1
{f } e l
= e p
2 N
T
dr (5.41)
1
or, substituting (5.18),
T
{f } e pl
= e
p l 2e p le

p l 2e
12
. (5.42)
2 12 2
In Fig. 5.7, a it is seen that f 2e is a shear force and f 3e is a moment. They
are called kinematically equivalent nodal forces since they replace a distributed
load p (r ) weighted with the shape functions N i (r ) so that the correct work is
simulated.

a b

c d
Fig. 5.7

The kinematically equivalent loads are those which, if applied in the


opposite direction as constraints, would keep all nodal displacements zero in the
presence of the true loading. To replace p = const . by statically equivalent forces
(Fig. 5.6, b) would be incorrect since the beam element has the ends rigidly jointed,
5. BEAMS, FRAMES AND GRIDS 91

i.e. it is rigidly built in the adjacent beam elements. In order to ensure the C1
continuity across elements, nodal forces must include moments, not only shear
forces. Equivalent nodal forces for linearly distributed loads are given in Figs. 5.6,
c and d.
Kinematically equivalent loads yield displacements which do not coincide
with those produced by actual loading, as shown in Example 5.1. Assuming
approximate deflected shapes instead of the true ones may be imagined as the
result of application of a fake loading, forcing the beam to maintain the
approximate deflection. This is equivalent to applying additional constraints to the
beam, i.e. stiffening it. The deflections of this over-stiff finite element model are
smaller in the mean than the true deflections of the actual structure.
The source of error comes from the arbitrary selection of the shape
functions. Even if these functions are built up to satisfy the geometric boundary
conditions at the ends, the equilibrium within the elements is broken, due to the
difference between the applied load p (x ) and the resistance E I z v IV which gives
rise to a sort of unbalanced residual force.
The smaller the element, the smaller the error, so we would expect to
increase the accuracy by increasing the number of elements modeling the same
structure, or refining the mesh. A correct solution will approach the true value with
monotonically increasing values of displacements. The finite element solution is
therefore referred to as a lower bound. This applies only to the strain energy and
not to the displacement or stress at a point. Local stresses may be higher than the
true ones.
Assuming cubic displacement functions implies linearly varying bending
moments (and hence stresses) in uniform beams, even if it is known that, for
uniform loading, for instance, they have a quadratic distribution.

Example 5.1
Calculate the transverse displacement at the centre of the simply supported
beam shown in Fig. E5.1.

Fig. E5.1
92 FINITE ELEMENT ANALYSIS

Solution. Using a single beam element, the distributed load is replaced by


two concentrated moments at the ends.
Using the boundary conditions and the equivalent nodal loads, the
equations of equilibrium can be written
pl
2
12 6l 12 6l q2 = 0
pl 2
6l 4l 2 6l 2l 2 q
EIz 3 = 12 .
l 3 12 6l 12 6l q5 = 0 pl

6l 2l
2
6l 4l 2 q6 2 2
pl

12
The set of equations in the unknown displacements is
2
EI
( 4q3 + 2q6 ) = pl ,
l 12
2
EI
( 2q3 + 4q6 ) = pl ,
l 12
with solutions
pl3
q6 = q3 = .
24 E I
The displacement at the middle is

l l l pl 4
v = N 2 q3 + N 4 q6 = .
2 2 2 96 E I

The true solution is


5 pl 4
v true =
384 E I
so that the approximate finite element solution is, as expected, smaller
l 4
v = vtrue .
2 5

5.4.2 Higher-degree interpolation functions

Let explore the possibility of approximating the displacements of a fourth


order system by a quadratic function of the form
5. BEAMS, FRAMES AND GRIDS 93

v (x ) = a 1 x 4 + a 2 x 3 + a 3 x 2 + a 4 x + a5 . (5.43)

The five integration constants a 1 , a 2 , a 3 , a 4 , a5 in (5.43) can be


determined from the end displacements and slopes, and the displacement of the
internal node at the centre of the element (Fig. 5.8):

at x = x1 , v = v1 , and d v d x = 1 ,

at x = x 2 , v = v2 , and d v d x = 2 , (5.44)
at x = x3 , v = v3 .

This gives the nodal coordinates in terms of the polynomial coefficients

1 0 0 0 0
v1 a5
0 1 0 0 0
1 a4
1 l l2 l3 l4
v2 = a3 .
0 1 2l 3l 2 4l 3
a2
2 l l2 l3 l4
v3 1 a1
2 4 8 16
On inversion

1 0 0 0 0
0 v1
a5 0 1 0 0
a 11 4 5 1 16 1
4 2
a3 = l
2 l l l l 2 v 2 .
a 18 5 14 3
2 3 2
32
2 l3 l2 l3 l l
a1 8 2 8 2 16 v3
3 4
l 4 l l l3 l 4
The beam deflection is given by an expression of the form (5.12)

v1

1
{ }
v (x ) = N q e = N1 N2 N3 N4 N 5 v2 , (5.45)

2
v3

where the quartic shape functions are


94 FINITE ELEMENT ANALYSIS

N 1( r ) = ( 1 r ) 2 r ( 3 + 2r ) ,
1
4
N 2 ( r ) = ( 1 r ) 2 r ( 1 + r )l ,
1
8
N 3 ( r ) = ( 3 2r ) r ( 1 + r ) 2 ,
1
(5.46)
4
N 4 (r ) = (1 r ) r (1 + r ) 2 l ,
1
8

N 5 (r ) = (1 r ) 2 (1 + r ) 2 .

In equations (5.46), N 5 is called a bubble function, having zero


displacements and slopes at the ends.

Fig. 5.8

Substituting the shape functions (5.46) into equation (5.27) and performing
the integration, we obtain the element stiffness matrix

316 94l e 196 34l e 512


36l 2e 34l e 6l 2e 128l e
[k ]
e
B
EI e
= 3
5l e
316 94l e

512 . (5.47)
36l 2e 128l e
SYM 1024

For p = const . , the element consistent load vector (5.41) is


5. BEAMS, FRAMES AND GRIDS 95

T
{f }
e 7 ple
=
p l 2e 7 p le

p l 2e 8 p le
. (5.48)
30 60 30 60 15

5.4.3 Bending moment and shear force

Using the bending moment expression (5.6) and equation (5.12) we get

M = EIz
d2 v
dx 2
= E I
4 d2 v
z 2
l e dr 2
4
= E I z 2 N q e ,
le
{ }
M=
EIz
6 r (3r 1) l e 6r (3r + 1) l e { q e }. (5.49)
l 2e
The shear force is given by equation (5.7)

T = E I z
d3 v
dx 3
= E I z
8 d3 v
3
l e dr 3
8
= E I z 3 N q e ,
le
{ }
T=
6E I z
l 3e
2 l e { }
2 l e qe . (5.50)

For elements with uniformly distributed load, the end equilibrium loads are
given by
pl e
2
R2 12 6l 12 6l q2
pl2
R 6l 4l 2 6l 2l 2 q e
3 EI e 3 + 12 . (5.51)
= 3
R
5 le 12 6l 12 6 l q5 p l e

R6 6l 2l
2
6l 4l e q6 e 22
2
e pl e

12
The first term on the right is k Be [ ]{ q }. The second term consists of
e

elements that are called fixed-end reactions.


The shear forces at the two ends of the beam element are T1 = R2 and
T2 = R5 . The end bending moments are M 1 = R3 and M 2 = R 6 . For p = 0 , they
are obtained substituting r = 1 and r = +1 in (5.49) and (5.50).
When p = const . , the exact moment and shear force within the element are

p l 2e p l e l e 2
Mp =M + ( r + 1) + p l e ( r + 1) 2 ,
12 2 2 2 4
96 FINITE ELEMENT ANALYSIS

pl e
Tp = T r,
2
where M and T are the expressions (5.49) and (5.50), respectively, valid for p = 0 .

5.5 Basic convergence requirements

As element sizes are reduced, the sequence of solutions to a problem is


expected to converge to the correct result if the assumed element displacement
fields satisfy the following criteria:
1. An element should describe rigid body modes exactly.
Although equilibrium is not satisfied exactly at every interior point or
across interfaces, an element as a whole should be in equilibrium, because the
structure as a whole should be in equilibrium. When nodal displacements are given
values corresponding to a state of rigid body motion, the element must exhibit zero
strain and therefore zero nodal forces, and the interior points should correspond to
the assumed rigid body displacement.

a. Vertical translation. If the nodes are given unit vertical displacements


(Fig. 5.9, a), equation (5.12) gives
1
0

v (x ) = N
1
1
4
( ) (
1
)
= N1 + N 3 = 2 3 r + r + 2 + 3 r r = 1 = const .
3
4
3

which shows that the element has indeed a vertical displacement as a rigid body.

b. Rotation. If the nodes are given unit rotations, with node 1 fixed and
node 2 having a vertical displacement l e (Fig. 5.9, b), equation (5.12) gives

0
1 l e
l l
v (x ) = N = N 2 + l e N 3 + e N 4 = e ( 1 + r ) = linear ,
l 2 2 2
1

which shows that the element has indeed an anticlockwise rotation as a rigid body.
5. BEAMS, FRAMES AND GRIDS 97

a b c
Fig. 5.9

2. An element should simulate constant strain states.


In the case of beams, when element sizes shrink to zero, they must have at
least constant curvature. Assuming zero nodal vertical displacements and unit
rotations in opposite directions (Fig. 5.9, c)
0
1

v (x ) = N
le
=
l l
N2 e N4 = e 1 r 2 , ( )
0 2 2 4
1
so that the second derivative (curvature) is constant
v = - l 2 = const .

5.6 Frame element

As shown in Fig. 5.2, an inclined beam element should include longitudinal


displacements since, apart from moments and shear forces, it is acted upon by axial
forces. Because there is no coupling between the bending and stretching
displacements, the two stiffness matrices can be added taking into account the
proper location of their elements.

5.6.1 Axial effects

The axial nodal forces are related to the nodal displacements by equation
f1
[ ]
e q1
= kS (5.52)
f4 q4
where the stiffness matrix for stretching (4.22) is

[ k ]= ElA 11
e
S
e 1
1
. (5.53)
e
98 FINITE ELEMENT ANALYSIS

5.6.2 Stiffness matrix and load vector in local coordinates

For the frame element, combining equations (5.53) and (5.29), and
arranging the elements in proper locations we get the element stiffness matrix
given by
EA EA
l 0 0 0 0
l
12 E I 6E I 12 E I 6E I
0 0
l3 l2 l3 l2
6E I 4E I 6E I 2E I
2
[k ] 0 0
e
= l2 l l l . (5.54)

E A 0 0
EA
0 0
l l
12 E I 6E I 12 E I 6E I
0 0 2
l3 l2 l3 l
0 6E I 2E I 6E I 4 E I
0 2
l2 l l l e

The ratio of the bending terms to the stretching terms in (5.54) is of order
(i l ) 2
, where i is the relevant radius of gyration. For slender beams, this ratio
may be as small as 1 20 or 1 50 , so the stiffness matrix may possibly be
numerically ill-conditioned.
If there is a uniformly distributed load on a member, the vector of
consistent nodal forces is
T
{ f }= 0
e p le p l 2e
0
p le

p l 2e
. (5.55)
2 12 2 12

5.6.3 Coordinate transformation

A frame element is shown in Fig. 5.10 both in the initial and deformed
state. For node 1, the local linear displacements q 1 and q 2 are related to the
global linear displacements Q 1 and Q 2 by the equations

q 1 = Q 1 cos + Q 2 sin ,
. (5.56)
q 2 = Q 1 sin + Q 2 cos .
Equations (5.56) can be written in matrix form as
q1 c s Q1
q =
c Q 2
(5.57)
2 s
5. BEAMS, FRAMES AND GRIDS 99

where c = cos and s = sin .


The angular displacements (rotations) are the same in both coordinate
systems
q 3 = Q3 . (5.58)

Fig. 5.10

Adding the similar relationships for node 2

q4 c s Q4
q = , q6 = Q6 ,
c Q 5
(5.59)
5 s
we obtain

{ q } = [ T ] { Q }.
e e e
(5.60)

In (5.60), { q } is the beam element displacement vector in the local


e

coordinate system, { Q } is the beam element displacement vector in the global


e

coordinate system and


c s 0 0 0 0
s c 0 0 0 0

[T ]e
0
=
0 1 0 0 0
c s 0
(5.61)
0 0 0
0 0 0 s c 0

0 0 0 0 0 1
is the local-to-global coordinate transformation matrix for plane frames.
100 FINITE ELEMENT ANALYSIS

5.6.4 Stiffness matrix and load vector in global coordinates

Using the same procedure as in section 3.7, the stiffness matrix of the
frame element in global coordinates is obtained as

[ K ]= [T ] [ k ][T ] .
e e T e e
(5.62)

The nodal loads due to a uniformly distributed load p are given by

{F }= [T ] { f }
e e T e
(5.63)

{ } are added to the global load vector.


The values of F e

5.7 Assembly of the global stiffness matrix

The global stiffness matrix, [K ],


is assembled from element matrices
[ ]
K e using element connectivity matrices
~
[ ]
T e , that relate the nodal
displacements at element level with the nodal displacements at the entire structure
level, by equations of the form

{ } [ ]
Q e = T e { Q }.
~
(5.64)
The global uncondensed stiffness matrix is equal to the sum of the
expanded element stiffness matrices

[ ]
[ K ] = K~ e , (5.65)
e
where
[ K~ ] = [T~ ] [ K ][T~ ] .
e e T e e
(5.66)
The equations of equilibrium can be regarded as having been derived as
soon as the reduced stiffness matrix and load vector have been calculated using the
boundary conditions.
Having solved [ K ] { Q } = { F } it is routine to backtrack and recover the
eth element strains, using equations (5.4), (5.12) and (5.60)

{ }
e = y v = y N q e = y N T e [ ] { Q }.
e
(5.67)
5. BEAMS, FRAMES AND GRIDS 101

Strains, and hence stresses, are not accurate. Strains are derivatives of an
approximate displacement (in beams - second derivatives) and differentiation
inevitably decreases accuracy.

Example 5.2
Calculate the transverse displacement at the free end of the cantilever
stepped beam shown in Fig. E5.2.

Fig. E5.2

Solution. Consider the beam divided into two cubic Hermitian elements.
The element stiffness matrices are

12 6l 12 6l 12 6l 12 6l
6l 4l 2 6l 2l 2 6l 4l 2 6l 2l 2
[k ]
1 2E I
= 3
l 12 6l 12 6 l
,

[k ] 2 EI
= 3
l 12 6 l 12 6 l
.

2 2
6l 2l
2
6l 4l 6l 2l
2
6l 4l

The unreduced global stiffness matrix is


24 12l 24 12l 0 0
12l 8l 2 12l 4l 2 0 0

E I 24 12l 36 6l 12 6l
[ K ]= 3 2
6l 12l 2 6l 2l 2
.
l 12l 4l
0 0 12 6l 12 6l

0 0 6l 2l 2 6l 4l 2

Using the boundary conditions at the fixed end Q1 = Q2 = 0 , and omitting


the first two rows and columns, we obtain the finite element equations
102 FINITE ELEMENT ANALYSIS

36 6l 12 6l Q3 0

EI 6l 12l
2
6l 2l 2 Q4 0



= .
l 3 12 6l 12 6l Q5 F
0
6l 2l 2 6l 4l 2 Q6

The three equations with zero right hand side can be written
36 6l 6l Q3 12
6l 12l 2 Q
2l 2 4 = 6l Q5 .
6l 2l 2 4l 2 Q6 6l

or
Q3 11l 2 9l 21l 12 60l
1 1
Q4 = 2 9l 27 27 6l Q5 = 108 Q5 .
Q 216l 21l 27 99 6l 216 l
6 180

By substitution into the equation with non-zero right hand side


EI
( 12 Q3 6l Q4 + 12 Q5 6l Q6 ) = F
l3
the transverse displacement in 3 is obtained as
F l3
v3 = Q5 = 1.5 .
EI

Example 5.3
Calculate the transverse displacement at 2 and the support reactions for the
beam shown in Fig. E5.3.

Solution. Consider the beam modeled by two Bernoulli-Euler beam


elements. Assembling the element stiffness matrices (5.29), the unreduced global
stiffness matrix is obtained as
12 6l 12 6l 0 0
6l 4l 2 6l 2l 2
0 0
12 6l
[ K ] = E3I 612 6l 24
2
0
2
6l 2l 2
.
l l 2l 0 8l
0 0 12 6l 12 6l

0 0 6l 2l 2 6l 4l 2
5. BEAMS, FRAMES AND GRIDS 103

At the fixed end Q1 = Q2 = 0 ; at the simply supported end Q5 = 0 .


Omitting the corresponding rows and columns, we obtain the finite element
equations
24 0 6l Q3 F
EI
3
0 8l 2l 2 Q4 = 0 .
2
l
6l 2l 2 4l 2 Q6 0

The last two equations give


8l 2 2l 2 Q4 0
2 = Q3
2l 4l 2 Q6 6l
or
3 12
Q4 = Q3 , Q6 = Q3 .
7l 7l

Fig. E5.3

Substitution of Q4 and Q6 into the first equation gives

3 7l F l3
24 Q3 + 0 6l Q3 =
12 7l EI
or
7 F l3
v2 = Q3 = .
96 E I
The rotations are
3 F l2 12 F l 2
2 = Q4 = , 3 = Q6 = .
96 E I 96 E I
The support reactions are given by
104 FINITE ELEMENT ANALYSIS

12 6l 0 Q3 V1
EI
3
6l 2l 2
0 Q4 = M1
l
12 6l 6l Q6 V
3
which yield
11 3 5
V1 = F, M1 = F l , V3 = F.
16 8 16

Example 5.4
Find the transverse displacement at 2 and the support reactions for the
beam with fixed ends shown in Fig. E5.4.

Fig. E5.4

Solution. Consider the beam modeled by two Bernoulli-Euler beam


elements. Using the boundary conditions Q1 = Q2 = Q5 = Q6 = 0 , the finite element
equations can be written
12 6l 12 6l 0 0 0 V1
6l 4l 2
6l 2l 2
0 0 0 M 1
E I 12 6l 13.5 4.5l 1.5 1.5l Q3 F
= .
l 3 6l 2l
2
4.5l 6l 2 1.5l l 2 Q4 0
0 0 1.5 1.5l 1.5 1.5l 0 V3

0 0 1.5l l 2
1.5l 2l 2 0 M 3

From the third and fourth rows we obtain


E I 13.5 4.5l Q3 F
2 =
l 3 4.5l 6l Q4 0
with solutions
8 F l3 2 F l2
v2 = Q3 = , 2 = Q4 = .
81 E I 27 E I
5. BEAMS, FRAMES AND GRIDS 105

Substituting the displacements into the other four equations yields


12 6l V1
6l 2l 2 l
8

81 F l = M 1 .
2
EI
l 3 1.5 1.5l 2 E I V3

1.5l l 2 27 M 3

The support reactions are


20 4 7 2
V1 = F , M 1 = F l , V3 = F , M3 = Fl .
27 9 27 9

Example 5.5
Calculate the rotations at supports and the reaction forces for the two-span
beam shown in Fig. E5.5 where the right span carries a uniformly distributed load.

Fig. E5.5

Solution. The continuous beam is modeled by two Bernoulli-Euler beam


elements. Using the boundary conditions Q1 = Q3 = Q5 = 0 , the finite element
equations can be written
12 6l 12 6l 0 0 0 V1
6l 4l 2 6l 2l 2 Q
0 0 2 0
E I 12 6l 24 0 12 6l 0 V2 p l 2
= .
l 3 6l 2l
2
0 8l 2 6l 2l 2 Q4 p l 2 12
0 0 12 6l 12 6l 0 V3 p l 2

0 0 6l 2l 2 6l 4l 2 Q6 p l 2 12

Omitting the first, third and fifth rows and columns, we obtain
4l 2 2l 2 0 Q2 0
EI 2
8l 2 2l 2 Q4 = p l 12 .
2
2l
l3 0 2l 2 4l 2 Q6 p l 2 12


106 FINITE ELEMENT ANALYSIS

The first equation above yields


Q
Q2 = 1 2 0 4 (a)
Q6
which, substituted into the second and third equation

EI 2l 2 8l 2 2l 2 Q4 p l 2 1
Q2 + 2 = ,
l3 0 2l 4l 2 Q6 12 1

gives
pl3 pl3
2 = Q4 = , 3 = Q6 = . (b)
48 E I 32 E I
Substituting the rotations (b) in (a) we obtain
pl3
1 = Q2 = .
96 E I
The reaction forces are calculated from
6l 6l 0 Q2 V1
EI
3
6l 0 6l Q4 = V2 p l 2
l
0 6l 6l Q6 V pl 2
3
obtaining
pl 5 7
V1 = , V2 = pl , V3 = pl .
16 8 16

Example 5.6
Find the transverse displacement and the angle of rotation at 2 for the beam
shown in Fig. E5.6.

Fig. E5.6
5. BEAMS, FRAMES AND GRIDS 107

Solution. Two Bernoulli-Euler beam elements are used to model the


system. Using the boundary conditions Q1 = Q2 = Q5 = 0 , the finite element
equations can be written

1.5 1.5l 1.5 1.5l 0 0 0 V1 0.7 p0 l



1.5l
2l 2
1.5l l 2
0 0 0 M 1 0.2 p0 l 2
E I 1.5 1.5l 13.5 4.5l 12 6l Q3 0.3 p0 l
= .
l 3 1.5l l2 4.5l 6l 2 6l 2l 2 Q4 0.1333 p0 l 2
0 0 12 6l 12 6l 0 V3

0 0 6l 2l 2 6l 4l 2 Q6 0

Deleting the first, second and fifth rows and columns we obtain

13.5 4.5l 6l Q3 0 .3 p 0 l
EI
3
4.5l 6l 2 2l 2 Q4 = 0.1333 p0l
2
.
l
6l 2l 2 4l 2 Q6
0

The last equation yields


3 1
Q6 = Q3 Q4
2l 2
which, substituted into the first two equations

EI 13.5 4.5l Q3 6l 0.3 p0l


4.5l 6l 2 Q + 2l 2 Q6 = 0.1333 p l 2 ,
l3
4 0

gives
p0 l 4 p0 l 3
v2 = Q3 = 0.084 , 2 = Q4 = 0.0518 .
EI EI

Example 5.7
Find the shape functions for the 3-node beam element shown in Fig. E5.7.

Answer.

N 1( r ) =
1 2
4
(
r 4 5r 2r 2 + 3r 3 , ) N 2 (r ) =
1 2
4
( )
r 1 r r 2 + r 3 ,
108 FINITE ELEMENT ANALYSIS

(
N 3(r ) = 1 r 2 ) 2
, (
N 4 (r ) = r 1 r 2 ),
2

N 5 (r ) =
1 2
4
(
r 4 + 5r 2r 2 3r 3 , ) N 6 (r ) =
1 2
4
( )
r 1 r +r 2 + r3 .

Fig. E5.7, a

Fig. E5.7, b
5. BEAMS, FRAMES AND GRIDS 109

Example 5.8
The planar frame shown in Fig. E5.8, a has pinned ends and is loaded in 2
by a force F = 1000 N . It has E = 2 1011 N m 2 , A = 1600 mm 2 and
I = 2 10 mm . Determine the nodal displacements, plot the deformed shape and
5 4

the diagrams of axial force, shear and bending moment.


Answer. The frame is modeled with 6 elements and 7 nodes.

Nodal data
Node Restr Restr Restr Coord Coord Displ Displ Rotation
nr X Y Z X Y X Y Z
1 1 1 0 0 0 0 0 -1.232e-2
2 0 0 0 0 1 0.010 8.500e-7 -5.531e-3
3 0 0 0 0 2 1.0699e-2 1.700e-6 2.353e-3
4 0 0 0 0.5 2 1.0698e-2 1.309e-3 2.592e-3
5 0 0 0 1.4 2 1.0697e-2 2.321e-3 -1.260e-3
6 0 0 0 2 2 1.0696e-2 -8.500e-7 -6.889e-3
7 1 1 0 2 1 0 0 -1.260e-2

The deformed shape is shown in Fig. E5.8, b.


The diagrams of the axial force N , shear T and bending moment M are
shown in Figs. E5.8, c, d, e.

a b

c d e

Fig. E5.8
110 FINITE ELEMENT ANALYSIS

Example 5.9

The planar frame shown in Fig. E5.9, a has fixed ends in 1, 5, 10, 13. It is
loaded by a couple M 3 = 2 106 N mm and two point loads F6 = 2 10 4 N and
F12 = 10 4 N . For E = 2 105 N mm 2 , A = 400 mm 2 , I = 2 104 mm 4 ,
l1 2 = 100 mm and l 9 10 = l 9 11 = 200 mm , find the displacements in 6 and plot
the deformed shape.

Answer. The frame is modeled with 12 elements and 13 nodes.

Fig. E5.9, a

The displacements in 6 are

h 6 = 0.2476 mm , v 6 = 0.9673 mm , 6 = 0.00623 rad .

The deformed shape is presented in Fig. E5.9, b.

Fig. E5.9, b
5. BEAMS, FRAMES AND GRIDS 111

5.8 Grids

Grids or grillages are planar structural systems subjected to loads applied


normally to their plane. They are special cases of tree-dimensional frames in which
each joint has only three nodal displacements, a translation and two rotations,
describing bending and torsional effects.

Fig. 5.11

5.8.1 Finite element discretization

The grid is divided into finite elements, as shown in Fig. 5.11. Each node
has three degrees of freedom, two rotations and a linear displacement. Typically,
the degrees of freedom of node i are Q3 i 2 , Q3 i 1 and Q3 i , defined as the rotation
about the X axis, the rotation about the Y axis and the displacement along the Z
axis, respectively.
Nodes are located by their coordinates in the global reference frame XOY
and element connectivity is defined by the indices of the end nodes. Elements are
modelled as uniform rods with bending and torsional flexibility, without shear
deformations and not loaded between ends. Their properties are the flexural rigidity
E I , the torsional rigidity G I t and the length l . Only cross sections whose shear
centre coincides with the centroid are considered.

5.8.2 Element stiffness matrix in local coordinates

Consider an inclined grid element, as illustrated in Fig. 5.12, a, where the


nodal displacements are also shown.
112 FINITE ELEMENT ANALYSIS

In a local physical coordinate system, the x axis, oriented along the beam,
is inclined an angle with respect to the global X axis. The z axis for the local
coordinate system is collinear with the Z axis for the global system. Alternatively,
an intrinsic (natural) coordinate system can be used.

Fig. 5.12

The vector of element nodal displacements is

{ q }= q
e
1 q2 q3 q4 q5 q6 T (5.68)

and the corresponding vector of element nodal forces can be written

{ f }= f
e
1 f2 f3 f4 f5 f6 T . (5.69)

In (5.69) f 3 and f 6 are transverse forces, while f 2 and f 5 are


couples producing bending (Fig. 5.12, b). The corresponding displacements
q 3 , q 6 are translations, while q 2 , q 5 are rotations. Their column vectors are
related by the flexural stiffness matrix.

Rearranging the matrix (5.29) we obtain

f2 4l 2 6l 2l 2 6l q 2

f3 EI e 6l 12 6l 12 q 3
= 3 2 . (5.70)
f5 l e 2l 6l 4l 2 6l q 5

f 6 6l 12 6l 12 e q 6

The axial nodal forces f 1 , f 4 are torques and the nodal displacements
q 1 , q 4 are twist angles. They describe torsional effects so that their action is
decoupled from bending. The respective stiffness matrix can be calculated
5. BEAMS, FRAMES AND GRIDS 113

separately. The derivation of this matrix is essentially identical to the derivation of


the stiffness matrix for axial effects in a frame element or in a truss element.
The twist angle can be expressed in terms of the shape functions (4.51) as

(r ) = N1 (r ) q 1 + N 2 (r ) q 4 , (5.71)

which substituted into the strain energy


2
G It e
Ue =
2
e
d x
x
(5.72)

yields, after the change of coordinates, the element stiffness matrix


+1

[ ]=
kte
2 G It e
le N
r
T
N r d r . (5.73)
1
As a consequence of this analogy, the nodal forces are related to the nodal
displacements by equation
f1 e q1
= kt [ ] (5.74)
f4 q4
where the stiffness matrix for torsional effects is

[ k ] = GlI
t
e te 1 1
1 1 . (5.75)
e
In (5.75), G is the shear modulus of elasticity and I t e is the torsional
constant of the cross section. For axially symmetrical cross sections the latter is the
polar second moment of area.
For the grid element, combining the stiffness matrices from equations
(5.70) and (5.75), we get the stiffness matrix in local coordinates relating the nodal
forces (5.69) and the nodal displacements (5.68)

a 0 0 a 0 0
0 4l 2 6l
0 2l 2
6l

[k ]
e EI 0
= 3e
l e a
6l
0
12
0
0
a
6l 12
0 0
(5.76)

0 2l 2 6l 0 4l 2 6l

0 6l 12 0 6l 12 e

where a = G I t e l 2e E I e .
114 FINITE ELEMENT ANALYSIS

5.8.3 Coordinate transformation

It is necessary to transform the matrix (5.76) from the local to the global
system of coordinates before its assemblage in the stiffness matrix for the complete
grid. As has been indicated, the z direction for local axes coincides with the Z
direction for the global axes, so that only the rotational components of
displacements should be converted. The transformation of coordinates is defined
by equation

{ q } = [ T ] { Q },
e e e
(5.77)

{ } is the element displacement vector (5.68) in the local coordinate


where q e
system,

{ Q } = Q
e
1 Q2 Q3 Q4 Q5 Q6 T

is the element displacement vector in the global coordinate system (Fig. 5.12) and

c s 0 0 0
0
s
c 0 0 0 0

[T ]
e
0
=
0 1 0 0 0
c s 0
, (5.78)
0 0 0
0 0 0 s c 0

0 0 0 0 0 1
where c = cos and s = sin , is the local-to-global coordinate transformation
matrix. The same transformation matrix (5.78) serves to transform the nodal forces
from local to global coordinates.

5.8.4 Element stiffness matrix in global coordinates

Using the same procedure as for frame elements, we obtain the stiffness
matrix of the grid element in global coordinates as

[ K ]= [T ] [ k ][T ] .
e e T e e
(5.79)

It is used to assemble the unreduced global stiffness matrix [ K ] using


~
[ ]
element connectivity matrices T e that relate the nodal displacements at element
level with the nodal displacements at the complete structure level, by equations of
the form (5.64).
5. BEAMS, FRAMES AND GRIDS 115

For grounded systems the unreduced matrix [ K ] is then condensed using


the boundary conditions. The effect of lumped springs can be accounted for by
adding their values along the main diagonal at the appropriate locations in the
global stiffness matrix.

Example 5.10
The grid shown in Fig. E5.10 is fixed at points 1 and 2, has E = 210 GPa ,
G = 81 GPa , l = 1 m and diameter d = 20 mm . Find the vertical displacement of
point 7 when the grid is loaded by forces F7 = F8 = 500 N and draw the spatial
deflected shape.

a b
Fig. E5.10

Answer. The grid is modeled with 14 elements and 8 nodes, having 18


dofs. The deflected shape is presented in Fig. E5.10, b. The deflection is
w7 = 0.436 m .

Example 5.11

The grid shown in Fig. E5.11, a is fixed at points 1 and 2, and has l = 1 m ,
I = 0.785 10 8 m 4 , I t = 1.57 10 8 m 4 , E = 210 GPa and G = 81 GPa . Find the
vertical displacement of point 5 when the grid is loaded by a force F5 = 103 N .
Draw the deflected shape and the diagrams of the bending moment and torque.
116 FINITE ELEMENT ANALYSIS

a b

c d
Fig. E5.11

Answer. The grid is modeled with 5 elements and 5 nodes, having 9 dofs.
The largest displacement is w5 = 0.4 m . The deflected shape is presented in Fig.
E5.11, b. The bending moment and torque diagrams are shown in Figs. E5.11, c, d.

5.9 Deep beam bending element

Shear deformation becomes important when analyzing deep beams, for


which Bernoullis hypothesis is no more valid. The nonlinear distribution of shear
stresses produces the warping of the cross section.
A simplifying hypothesis (Poncelet, 1825) considers an average shear
strain, constant over the cross section. This way, planar cross sections remain
undistorted and plane (warping neglected) but no more perpendicular to the
centroidal axis. The assumption is adopted in the formulation of the Timoshenko
beam element used in vibration studies.
5. BEAMS, FRAMES AND GRIDS 117

5.9.1 Static analysis of a uniform beam

Beams with cross sections that are symmetric with respect to the plane of
loading are considered herein (Fig. 5.13, a). Only transverse loads act upon the
beam, axial forces are ignored.

Fig. 5.13

The axial displacement of any point on the section, at a distance y from the
neutral axis, is approximated by
u (x, y ) = y (x ) , (5.80)
where is the cross section rotation at position x .

The strain components x and xy are given by

du d
x = = y = y , (5.81)
dx dx
u v dv
xy = + = + = + v , (5.82)
y x dx

where v is the slope of the deformed beam axis.


Note that the slope v is no more equal to the rotation , as in the
Bernoulli-Euler theory.
Normal stresses on the cross section are given by Hookes law
d
x = E x = Ey, (5.83)
dx
where E is Youngs modulus of the material.
The bending moment is the resultant of the normal stress distribution on
the cross section
118 FINITE ELEMENT ANALYSIS

d

M (x ) = x y dA = E I z
A
dx
= E I z (5.84)

where I z is the second moment of area of the cross section.


The sign convention used here (Fig. 5.13, b) is that positive internal forces
and moments act in positive (negative) coordinate directions on beam cross
sections with a positive (negative) outward normal.
The shear force is given by

T (x ) =
dM
. (5.85)
dx
The transverse load per unit length is

p (x ) =
dT
. (5.86)
dx
The average shear strain is
T T
xy = = (5.87)
G As AG

where G is the shear modulus of elasticity, is a shear factor and As is the


effective shear area calculated as

As =
[ dA ] 2

. (5.88)

2
dA

Equations (5.82) and (5.87) give

T = G As ( v ) . (5.89)

For a uniform beam not loaded between ends ( p = 0 ) , elimination of M


and T gives the differential equations of equilibrium

G As v G As = 0 , (5.90)
G As v + E I z G As = 0 . (5.91)

5.9.2 Shape functions

Consider a prismatic beam element (Fig. 5.14) of length l (the index e is


omitted) not loaded between ends ( p = 0 ) . Using natural coordinates, r = 2 x l .
5. BEAMS, FRAMES AND GRIDS 119

Fig. 5.14

Eliminating and v in turn in (5.90) and (5.91) gives

d4v d 3
=0, and = 0. (5.92)
d x4 d x3
Changing to the r coordinate yields
d4v d 3
=0, and = 0. (5.93)
d r4 d r3
The general solutions of these equations are

v (r ) = a 4 r 3 + a 3 r 2 + a 2 r + a 1 , (5.94)

(r ) = b 3 r 2 + b 2 r + b 1 . (5.95)

The seven constants of integration are not independent since the above
solutions must also satisfy equation (5.91) which, in the new variable r, becomes
2 dv d 2
+ = 0, (5.96)
l dr d r2
where
4E I z
= . (5.97)
G As l 2
Substituting (5.94) and (5.95) into (5.96) gives
6 4 2 12
b3 = a4 , b2 = a3 , b1 = a2 + a 4 . (5.98)
l l l l
This leaves only four independent constants which can be determined by
evaluating (5.94) and (5.95) at r = 1 . The resulting displacement functions are
120 FINITE ELEMENT ANALYSIS


v = N1

l
2
N2 N3
l
N4
2
{ q } = N { q },
e e
(5.99)

2 ~
= N1 N 2
l
~ 2 ~
l
N3
~
N4

{ q } = N~ { q },
e e
(5.100)

where

{ q }= v
e
1 1 v2 2 T . (5.101)

The shape functions are

N 1( r ) =
1
4 ( 1 + 3 )
[ 2 3r + r 3
+ 6 ( 1 r ) , ]
N 2 (r ) =
1
4 ( 1 + 3 )
[1 r r 2
(
+ r 3 + 3 1 r 2 , )]
N 3(r ) =
1
4 ( 1 + 3 )
[ 2 + 3r r 3
+ 6 ( 1 + r ) , ]
N 4 (r ) =
1
4 ( 1 + 3 )
[ 1 r + r 2
(
+ r 3 + 3 1 + r 2 , )]
~
N1 ( r ) =
1
4 ( 1 + 3 )
( 3 + 3r ) , 2
(5.102)

~
N 2 (r ) =
1
4 ( 1 + 3 )
[ 1 2r + 3r 2
+ 6 ( 1 r ) , ]
~
N3 ( r ) =
1
4 ( 1 + 3 )
( 3 3r ),2

~
N 4 (r ) =
1
4 ( 1 + 3 )
[ 1 +2r + 3r 2
+ 6 ( 1 + r ) . ]
For = 0 , the first four functions (5.102) become the third degree
Hermitic polynomials (5.21) and the last four functions satisfy the relationship

~ 2 Ni
N i (r ) = ( i = 1,..., 4 ) . (5.103)
l r
It is useful to introduce a third set of shape functions which are used in the
derivation of the element stiffness matrix, and defined by
5. BEAMS, FRAMES AND GRIDS 121

2 dN
N = N~ l . (5.104)
dr
They are
3 1
N1 (r ) = N 2 (r ) = N 3 (r ) = N 4 (r ) = . (5.105)
1 + 3 l

5.9.3 Stiffness matrix

The strain energy for a beam element with shear effects included is
l2 l 2
2 2
d dv

E Iz G As
Ue = d x + dx . (5.106)
2 dx 2 d x
l 2 l 2
As
d 2 d l
= and dx = dr , (5.107)
dx l dr 2
the contribution due to bending is
+1 2
d

E Iz 2
U eB = d r , (5.108)
2 l dr
1

and the contribution due to shear is


+1 2
2 dv

G As l
U eS = dr . (5.109)
2 2 l d r
1
On substituting (5.100) into (5.108) we get
+1

U eB =
1
2
{q } e T
EIz
2
l N N dr { q }
~ T ~ e
(5.110)
1

wherefrom we obtain the element stiffness matrix due to bending


+1
[ k ] = E I 2l N~ N~ dr .
e
B z
T
(5.111)
1

On substituting (5.100) and (5.99) into (5.109) we get


122 FINITE ELEMENT ANALYSIS

+1

U eS =
1
2
{q } e T
G As
l
2 N
T
N d r { q e } (5.112)
1

wherefrom we obtain the element stiffness matrix due to shear


+1
[ k ] = G A 2l N N dr .
e
S s
T
(5.113)
1

Substituting the shape functions (5.102) and (5.105) and performing the
integration yields the stiffness matrix
12 6l 12 6l
(4 + 3 ) l (2 3 ) l 2
[k ] 6l
2
e 1 EI z 6l .
= (5.114)
1 + 3 l 3 12 6l 12 6l
2
6l (2 3 ) l 2 6l (4 + 3 ) l
The vector of consistent nodal forces is identical to the corresponding
vector (5.42) derived for a slender beam.
6.
LINEAR ELASTICITY

In this chapter the fundamental concepts from the linear theory of elasticity
are recalled, with emphasis on two-dimensional problems. The four main groups of
equations are written in the matrix notation used in FEA: (a) equations of
equilibrium, (b) equations of compatibility or strain/displacement relations, (c)
stress/strain relations or Hookes law, and (d) boundary conditions.

6.1 Matrix notation for loads, stresses and strains

An arbitrarily shaped tree-dimensional body of volume V, in equilibrium


under the action of external loads and the reactions in supports, is shown in Fig.
6.1. The total surface S of the body has two distinct parts: S u , the portion of the
boundary on which displacements are prescribed, and S , the portion on which
surface forces are prescribed.
Points in the body are located by x, y, z coordinates. Any point on the
surface has a local outward-pointing normal n whose orientation is usually
described by its three direction cosines n x , n y , n z .
In general there may be three sets of applied forces: (a) internal body
forces, (b) surface forces, and (c) concentrated forces.

Internal body forces


Internal body forces inside the volume V can be inertial forces, like
centrifugal or gravity forces. Their the magnitude per unit volume is denoted by
components pv x , pv y , pv z . It is convenient to write these components as a single
body force vector
{ pv } = pv x pv y pv z T . (6.1)
124 FINITE ELEMENT ANALYSIS

Surface tractions
Likewise there could be surface forces (not necessarily normal pressures)
on the surface S , defined by the magnitude per unit surface area, also having
three components
{ ps } = ps x ps y ps z T , (6.2)

Concentrated loads
Concentrated loads are defined by their three components
{ Fi } = Fi x Fi y Fi z T . (6.3)

Any system of loads has to fall into categories (6.1) to (6.3).

Fig. 6.1

Displacements
It is natural to form the single displacement vector
{ u} = u v w T , (6.4)
where u , v , w are the displacement components inside the body or on the surface
S with unprescribed displacements.

Stresses and strains


The stresses inside V will have two types of component, the direct stress
components x , y , z and the shear stresses xy , yz , zx . It is convenient to
represent both stress and strain components as single column matrices. Thus
{ } = x y z xy yz zx T . (6.5)
6. LINEAR ELASTICITY 125

In two-dimensional problems
{ } = x y xy T . (6.5, a)

Strains are represented in vector form as


{ } = x y z xy yz zx T . (6.6)

In two-dimensional problems
{ } = x y xy T . (6.6, a)

6.2 Equations of equilibrium inside V


Figure 6.2 shows stresses acting on an infinitesimal element in the plane
xOy. The small linear increments in stresses are equilibrated by the applied body
forces yielding the following equilibrium equations, where xy = yx

x yx
+ + pv x = 0,
x y
(6.7)
xy y
+ + p v y = 0.
x y

Fig. 6.2

In matrix form, equations (6.7) can be written


126 FINITE ELEMENT ANALYSIS


x 0 x
y p vx
y = (6.8)
0 pvy
y x xy

or, denoting the matrix of differential operators



0
x

[ ] = 0 , (6.9)
y


y x
and using (6.1)
[ ]T { } + { pv } = { 0 } . (6.10)

6.2 Equations of equilibrium on the surface S

Let now consider an element near the loaded boundary S . The


equilibrium along the two axes directions yields
x l + yx m = ps x ,
(6.11)
xy l + y m = ps y .

Fig. 6.3

In (6.11) the direction cosines for the outward normal n are


6. LINEAR ELASTICITY 127

n
l = cos (n , x ) = = nx ,
x
(6.12)
n
m = cos (n , y ) = = ny .
y
In matrix form, equations (6.11) can be written

nx 0 n y x ps x
0 n n y = p , (6.13)
x sy
xy
y

or in condensed form

[ n ] T { } = { p s } , (6.14)
sometimes written

[ n ] { } = { p
T
s }, (6.15)

where T n implies that the operators in [ ] T are acting upon n.

6.3 Strain-displacement relations

Figure 6.4 gives the deformation of the dx dy face for small


deformations.
The equations of compatibility have the familiar form
u v v u
x = ; y = , xy = + . (6.16)
x y x y
In matrix form

0
x x
u
y = 0 , (6.17)
y v
xy

y x
and can be summarized as

{ } = [ ] { u } . (6.18)
128 FINITE ELEMENT ANALYSIS

Fig. 6.4

6.4 Stress-strain relations

For linear isotropic elastic materials, the stress-strain relations come from
the generalized Hookes law
{ } = [ C ] { } , (6.19)
where the material elastic compliance matrix is
1 0 0 0
1 0 0 0

1 0 0 0
[C ]= 1 . (6.20)
E 2 (1 + ) 0 0
SYM 2 (1 + ) 0

2 ( 1 + )
The inverse relation is

{ } = [ C ] 1{ } = [ D ] { } . (6.21)

The inverse of [ C ] is the material stiffness matrix


6. LINEAR ELASTICITY 129

1 0 0 0
1 0 0 0

1 0 0 0
1
[D] = E 0 0 .(6.22)
(1 + )(1 2 ) 2
SYM
1
0
2
1

2

Plane stress
A thin planar body subjected to in-plane loading on its edge surface is said
to be in plane stress. If stresses z , xz , and yz are set as zero, discarding rows
and columns 3, 5 and 6 in (6.20), the Hookes law can be written
x 1 0 x
1
y = 1 0 y . (6.23)
E 0 2 (1 + ) xy
xy 0
The inverse relations are given by

x
E 1 0 x
y = 1 0 y (6.24)
1 1
2

xy 0 0 xy
2
which is used as { } = [ D ] { }.

Plane strain
If a long body of uniform cross section is subjected to transverse loading
along its length, a small thickness in the loaded area can be treated as subjected to
plane strain. If strains z , xz , and yz are set as zero, from (6.21) and discarding
rows and columns 3, 5 and 6 in (6.22), we obtain

x
E 1 0 x
y = 1 0 y (6.25)
(1 + )(1 2 ) 1
xy 0 0 xy
2
130 FINITE ELEMENT ANALYSIS

which is also used as { } = [ D ] { } but here [ D ] has a different expression as in


(6.24).

6.5 Temperature effects

The temperature strains are represented as initial strains


{ 0 } = T T T 0 0 0 T , (6.26)
where T is the temperature rise and is the coefficient of linear expansion of the
material.
The stress-strain relations become
{ } = [ D ] ( { } { 0 } ) . (6.27)
In plane stress
{ 0 } = T T 0 T . (6.28)
In plane strain
{ 0 } = (1 + ) T T 0 T . (6.29)

6.6 Strain energy

For linear elastic materials, the strain energy per unit volume in the body is

U0 =
1
{ }T { } = 1 { }T { }. (6.30)
2 2
For the elastic body shown in Fig. 6.1, the total strain energy is given by

{ } { } dV .
1 T
U= (6.31)
2
V

Substituting (6.21) we obtain the strain energy in terms of strains

{ } [ D ] { } dV .
1 T
U= (6.32)
2
V
7.
ENERGY METHODS

The finite element method can be considered a Rayleigh-Ritz method.


The classical Rayleigh-Ritz technique represents a variational approach
whereby a distributed system is approximated by a discrete one by assuming a
solution of the differential eigenvalue problem as a finite series of admissible
functions. Unfortunately, systems with complex geometry or complex boundary
conditions cannot be accomodated easily by global admissible functions.
In the finite element method, the approximate solution is constructed using
local admissible functions, defined over small subdomains of the structure. Good
approximations can be realized with low-degree polynomials. Displacements are
calculated by methods based on the principle of virtual work and/or the principle of
minimum total potential energy.
Instead of solving differential equations with complicated boundary
conditions, the finite element method evaluates integrals of relatively simple
polynomial functions. Variational methods put less strict conditions on the
functions approximating the displacement field than the analytical methods based
on differential equations.

7.1 Principle of virtual work (PVW)


PVW is basically a statement of the static equilibrium of a mechanical
system. In the following, the form known as the principle of virtual displacements
(PVD) will be used, as applied to elastic bodies.

7.1.1 Virtual displacements

By definition, virtual displacements are:


a) arbitrary (fictitious, virtual);
b) infinitesimal (follow the rules of differential calculus);
132 FINITE ELEMENT ANALYSIS

c) not related to either the actual displacements or to the forces


producing them;
d) continuous in the interior and on the surface of the body;

A continuity C 0 is generally required for bars and elasticity problems,


while a continuity C1 is imposed for beams, plates and shells. Exceptions do exist.
Remember that a function of several variables is said to be of class C m in a
domain V if all its partial derivatives, up to the mth order inclusive, exist and are
continuous in the domain V.
e) kinematically admissible, i.e. consistent with the system
kinematic boundary conditions (geometric constraints).
If the differential equation of the problem is of order m = 2n , the
admissible functions must have continuity C n 1 , i.e. the geometrical boundary
conditions must be satisfied to the (n 1)th derivative. For bars, m = 2 and the
assumed functions must have continuity C 0 .
For beams m = 4 and the approximating functions must have continuity
C . Because the continuity required is reduced from C 2 in the governing
1

differential equation to C1 in the variational equation, the functional is said to have


a weak form.
A virtual displacement will be denoted by in front of a letter, e.g. u .
The symbol was introduced by Lagrange to emphasize the virtual character of
the variations, as opposed to the symbol d which designates actual differentials of
position coordinates.
Denoting by
{ u} = u v w T ,

the displacement vector (6.4) inside the body or on the surface S with
unprescribed displacements (Fig. 6.1), the vector of virtual displacements is
{u } = u v w T . (7.1)
The vector of the corresponding virtual strains will be

{ } = [ B ] { u } . (7.2)

where [ B ] is the strain-displacement matrix.


7. ENERGY METHODS 133

7.1.2 Virtual work of external loads

For a bar in tension (Fig. 7.1, a), the virtual work of the external force F is
WE = F u . (7.3)
It has the same value whether the bar material is linear elastic (Fig. 7.1, b)
or nonlinear elastic (Fig. 7.1, c). Note that it is simply (force displacement),
because the force is constant along the virtual displacement, the latter being
arbitrary, hence independent of the force.
In the general case of loading by conservative body forces (6.1), surface
tractions (6.2) and point forces (6.3), the virtual work of external loads is

{ u } { pv }dV + { u } { ps }d A + i { ui } { Fi } .
T T T
WE = (7.4)
V S

Note that the scalar product under the first integral is

{ u }T { pv } = u pv x + v pv y + w pv z .

a b c
Fig. 7.1
Note also the absence of the factor 1 2 which occurs in the expression of
the work of elastic forces, because the external loads remain constant during the
action along the virtual displacements.

7.1.3 Virtual work of internal forces

For a three-dimensional continuum, the virtual work of internal stresses is

{ } { }dV .
T
WI = (7.5)
V

as shown in Fig. 7.2 for the uniaxial case.


134 FINITE ELEMENT ANALYSIS

Fig. 7.2

Again, stresses remain constant during the action on virtual strains.

7.1.4 Principle of virtual displacements

For elastic bodies, the principle of virtual displacements states that:


If a system is in equilibrium, then during an arbitrary small displacement
from the equilibrium position, the virtual work of applied loads equals the virtual
work of internal forces
WE = WI . (7.6)
Also: A body is in equilibrium if the internal virtual work equals the
external virtual work for every kinematically admissible displacement field.

{ } { }dV { u } { pv }dV { u } { ps }d A i { ui } { Fi } = 0 .
T T T T

V V S
(7.6, a)
In (7.6) WE is the work of external loads on the virtual displacements
{ u } which are independent of loading and kinematically admissible.
If stresses are expressed in terms of a set of parameters defining
completely the displacement pattern the nodal displacements, then equilibrium
relations can be obtained and the displacement parameters determined. The nodal
displacements do not permit the fully equilibrating position to be reached, so that
the PVD will ensure approximate equilibrium.
Note that the virtual work of reaction forces at supports is zero. Since the
principle of virtual displacements is an equilibrium requirement, it is independent
of material behaviour, i.e. whether the material is elastic or inelastic. It applies only
7. ENERGY METHODS 135

for loading by conservative forces, which do not change direction during the action
on the virtual displacements. The external work is independent of the path taken.

Example 7.1
For the three-bar pin-jointed framework shown in Fig. 7.3, loaded by a
force F, find the internal bar forces and the displacement of point 4.

Fig. 7.3

Solution. Consider three states of the analyzed system:


1. The initial state, in which bars are not loaded by external forces and are
not prestressed (Fig. 7.4, a).
2. The final state of static equilibrium, in which the external force F, of
components F1 = F sin and F2 = F cos , produces a displacement of the joint 4,
of components u1 and u2 (Fig. 7.4, b).

The joint 4 is acted upon by the external forces F1 , F2 and by internal


forces T1 , T2 , T3 (Fig. 7.4, c). The joint reacts with forces equal in magnitude but
of opposite sign, producing the elongations 1 , 2 , 3 (Fig. 7.4, d).

3. An imaginary state, in which the joint 4 is given a virtual displacement


of components u1 and u2 (Fig. 7.4, e), which produce virtual elongations in bars
1 , 2 , 3 (Fig. 7.4, f), the applied forces remaining constant.

The virtual displacements u1 and u2 and the virtual bar extensions 1 ,


2 , 3 satisfy the compatibility equations (2.19)
136 FINITE ELEMENT ANALYSIS

1 = u1 sin + u 2 cos ,
2 = u 2 , (7.7)
3 = u1 sin + u 2 cos .
For the three bars, the force-elongation equations (2.21) can be written
T1 l T2 l cos T3 l
1= , 2= , 3= . (7.8)
EA EA EA

Fig. 7.4

Equating internal work to external work (7.6) using the products of real
forces and virtual displacements we obtain

T1 1+ T2 2 + T3 3 = F1 u1 + F2 u2 . (7.9)

Substituting (7.7) into (7.9) and collecting coefficients of u1 and u2


gives
7. ENERGY METHODS 137

u1 (T1 sin T3 sin F1 ) + u2 (T1 cos + T2 + T3 cos F2 ) = 0 . (7.10)

Since u1 and u2 are unrelated to each other, we could put either to zero.
Equation (7.10) must be zero whatever the values of u1 and u2 . This can only
be true if their coefficients vanish

T1 sin T3 sin = F1 ,
(7.11)
T1 cos + T2 + T3 cos = F2 .
These are indeed the equations of equilibrium. It is confirmed that the
principle of virtual work is an equivalent statement of statical equilibrium.
Substituting (7.8) in the finite form of (7.7), then in (7.11), the components
of the displacement of point 4 can be determined from the following equations

2 EA 2
sin u 1 = F1 ,
l
(7.12)
EA 2 1
2 cos + u 2 = F2 .
l cos

7.1.5 Proof that PDV is equivalent to equilibrium equations

Consider a form of equation (7.6, a) without point forces

{ } { }dV = { u } { pv }dV + { u } { ps }d A .
T T T
(7.6, b)
V V S

Convert { } to { u } using the integration by parts

u
x x dV = x x d x dy dz = x (u ) d x l d A =
x
V V S

x
= x d(u ) l d A = x u l d A u
x
dx dy dz ,
dV
S S V

y
y y dV = y v m d A v y
dV ,
V S V

xy xy
xy xy dV = xy ( v l + u m)d A
x
v +
y
u dV ,

V S V
138 FINITE ELEMENT ANALYSIS

where l and m are direction cosines of the outward normal at the surface.

Adding together

( x x + y y + xy xy )dV = [ u ( x l + xy m)+ v ( xy l + y m) ] d A
V S

x xy y
u
x
+
y
+ v xy +
x y
dV .

V

In matrix form (Gauss theorem)

{ } { }dV = { u } [ n ] { }d A { u } [ ] { }dV ,
T T T T T

V S V

which is true provided [ ] T { } and { } are finite in V, that is the stress and
displacement fields are continuous. This applies only within a single element and
up to its surface.

The componets of stresses at element interfaces may not balance at a point,


but only in the mean. Most finite elements in use today do not achieve continuous
stresses across interfaces.

On substituting in (7.6, b)

{ u } ( [ ] { } + { pv }) dV { u } ( [ n ] { } { ps }) d A = 0 .
T T T T

V S

Because { u } are arbitrary, its coefficients must vanish. The equations of


equilibrium emerge from the brackets

[ ]T { } + { pv } = { 0 } in V, [ n ]T { } { ps } = { 0 } on S .

The PVD supplies equilibrium conditions both within and on the surface of
the body. So, when using approximate functions for { u }, it is not necessary to
worry about the equilibrium boundary conditions. Part of the surface, i.e. S , is
supported in some way and there the tractions { ps } will be unknown reactions and
not specified loads. It is conventional to remove the unknown reactions by
choosing the virtual displacements { u } to be zero over S .
7. ENERGY METHODS 139

7.2 Principle of minimum total potential energy

The total potential energy of an elastic body is defined as the sum of


the strain energy U and the work potential of external loads WP

= U + WP . (7.13)

7.2.1 Strain energy

Consider the strain energy (6.32)


1
U= { }
T
[ D ] { } dV .
2
V
For a virtual strain { } , the virtual increase of strain energy is

U =
1
{ }
T
[ D ] { } dV + 1 { }T [ D ] { } dV .
2 2
V V

Because

({ } T
[ D ] { } )
T
= { } T [ D ] { } ,

U = V { }
T
[ D ] { } dV = { }T { } dV = WI .
V

U = WI . (7.14)
For an elastic body, the virtual variation of the strain energy is equal to
the virtual work of the internal stresses on virtual strains.
du
For a bar in tension, substituting = E and = , yields
dx
d u du
U =
dV = E A d x =
V l l
dx
EA
dx
dx . (7.15)

d2v
For a beam in bending, substituting = y (5.4), gives
d x2

2v 2 v 2 d 2v d2v
U =

l
x2
E
x 2
A
y dA

dx =


l
d x2
E I
d x2
dx . (7.16)

The above expressions can be obtained directly from the strain energies
140 FINITE ELEMENT ANALYSIS

2
E A du
for a bar U=

l
dx ,
2 d x
(7.17)

2
EI d2v
for a beam U=

l
2
x.
d x2 d

(7.18)

7.2.2 External potential energy

The work potential of external loads WP is equal to the negative product of


external forces by the corresponding displacement
WP = WE . (7.19)
The negative sign appears because the external loads lose some of their
capacity for doing work when displaced in the direction they act.
For example, a gravitational force F = m g acts in the opposite direction to
a vertical displacement h and the potential becomes m g h .

An external point load F j has potential energy F j u j ( ) instead of


1
F j u j , because this potential arises from the magnitude of force and its
2
capacity to do work when it moves, being independent of the linear properties of
the body on which it acts.
For a three-dimensional continuum

{u } { pv }dV {u } { ps }d A i {ui } { Fi } .
T T T
WP = (7.20)
V S

For virtual displacements {u }

WP = WE . (7.21)

7.2.3 Total potential energy

The total potential energy (7.13) can be written

= { }T { } dV { u }T { pv }dV {u } { ps }d A i { ui } { Fi } .
T T

V V S
(7.13, a)
Its variation is
7. ENERGY METHODS 141

= U + WP = WI WE . (7.22)

Based on equation (7.6) it follows that

= 0 , (7.22, a)
hence, at equilibrium, the total potential energy has a stationary value. If
2 > 0 , the stationary value is a minimum, the equilibrium is stable.

The principle of minimum total potential energy states that:

If a deformable body is in equilibrium under the action of external loads


and reaction forces, then the total potential energy has a minimum value.

Reciprocally, if under the action of external loads and reaction forces the
total potential energy of a deformable body is a minimum, then it is in a stable
equilibrium state.
Thus, it can be considered that (7.22, a) is a condition that establishes or
defines the equilibrium, rather than a result of the equilibrium.

An equivalent statement is: For conservative systems, of all possible


kinematically admissible displacement fields, the one satisfying equilibrium
corresponds to a minimum value of the total potential energy.

Reciprocally, any kinematically admissible displacement field which


minimizes the total potential energy represents a stable equilibrium configuration.

Example 7.2

For the truss shown in Fig. 7.3, the strain energy for a bar is

1 1 E Ai 2
U i = Ti i = i ,
2 2 li
and the external potential energy is

WP = F u .
i
i i

Expressing the elongations in terms of displacements, according to the


compatibility relations, the total potential energy can be written
142 FINITE ELEMENT ANALYSIS

EA EA
= ( u1 sin + u2 cos )2 + u 22 +
2l 2 l cos
EA
+ ( u1 sin + u2 cos )2 F1 u1 F2 u2 .
2l

Cancelling the derivatives of with respect to each independent variable


= 0, =0,
u1 u2

we obtain the equilibrium equations (7.11).

Example 7.3

Apply the principle of minimum total potential energy to a beam in


bending, subjected to a distributed load and to the end bending moments and shear
forces as shown in Fig. 7.5. Show that PMTPE is equivalent to the equilibrium
conditions inside and at the ends of the beam.

Fig. 7.5

Solution. For a beam segment loaded as shown, the total potential energy is

1
E I (v) d x p v dx + M
2
= 0 v0 M l vl T0 v0 + Tl vl .
2
l l

At equilibrium, is stationary, = 0 or

E I v (v)d x p v dx + M
l l
0 v0 M l vl T0 v0 + Tl vl = 0 . (a)

Integrating by parts the first term gives


7. ENERGY METHODS 143

d dv
l
E I v (v) d x =

l
E I v dx =
d x d x E I v d (v) =
l
d
d x (E I v) dx v = E I v v (E I v) v dx.
l l
= E I v v 0 0
l l

Integrating by parts the last term gives


d
(E I v) v dx = (E I v) d x (v)dx = (E I v) d (v)=
l l l
d
= (E I v) v (E I v) dx v = (E I v) v
(E I v) v dx.
l l
0
0

dx
l l

Equation (a ) becomes

E I v v 0 (E I v) v
(E I v) v dx p v dx M v
l l l l
0
+ 0
+ T v 0
=0
l l
or

((E I v) p ) v dx + ( E I v - M ) v (( E I v ) T )v
l
l
=0.
0
0
l

The coefficient of v in the integrand gives the equation of equilibrium

(E I v) = p (x ) .
As v is arbitrary, the other terms deliver the equilibrium conditions at the
beam ends

( E I v) 0 = M 0 , or v0 = 0 , ( E I v)0 = T0 , or v0 = 0 ,

( E I v) l = M l , or vl = 0 , and ( E I v)l = Tl , or vl = 0 ,

which are the boundary conditions.

7.3 The Rayleigh-Ritz method

The Rayleigh-Ritz method involves the construction of an assumed


displacement field. For a beam, the transverse displacement v (x ) is approximated
by a finite series
144 FINITE ELEMENT ANALYSIS

n
v (x ) a (x )
j =1
j j (7.23)

where a j are undetermined constants called generalized coordinates, and j ( x )


are prescribed functions of x , called admissible functions, that satisfy the
kinematic (geometric) boundary conditions and are continuous within the
definition interval.
Substituting the displacements (7.23) into the expression of the total
potential energy , the latter becomes a function of the parameters a j , whose
values are determined from the stationarity conditions

= aj j
a j = 0 .

Because a j are arbitrary,


=0, ( j = 1,..., n ) , (7.24)
aj
which is a linear algebraic set of equations in the constants a j .
The solutions are back-substituted into (7.23) which represents an
approximate deflected shape, which is more accurate the more terms are selected in
the respective series.
The necessary requirements for the convergence of the Rayleigh-Ritz
method are the following:
a) The approximating functions must be continuous to one order less the
highest derivative in the integrand.
b) The functions must individually satisfy the geometric boundary
conditions, i.e. to be admissible functions.
c) The sequence of functions must be complete.
If the functions are not selected from the domain space of the operator of
the equation being solved (completeness property) the resulting solution could be
either zero or wrong.

Example 7.4
For the beam shown in Fig. 7.6 find the vertical displacement of point 2.
Consider: E = 210 MPa , I = 1600 mm4 , l = 3 m , F = 100 N and q = 200 N m .
Solution. The total potential energy is
7. ENERGY METHODS 145

l 2 l
d2v
1 d x q v (x ) d x F v l .
= EI
2
0

d x2


2l 3
2
(7.25)

The geometric boundary conditions are


2l
v (0) = 0 , v (0) = 0 , v =0, v (l ) = 0 . (7.26)
3
In the following, for simplicity, the transverse displacements are
approximated by a series consisting of only two terms
v ( x ) = a1 1 (x ) + a2 2 (x ) , (7.27)
where the functions

x 2 (3 x 2 l )( x l ) x 3 (3 x 2 l )( x l )
1 (x ) = , 2 (x ) = , (7.28)
l4 l5
satisfy all geometrical boundary conditions (7.26).

Fig. 7.6

Substituting (7.27) into (7.25) we obtain the functional


l l

( a1 1 + a2 2 ) dx q ( a1 1 + a2 2 ) dx .
1 2
= EI
2
0 2l 3
F [ a1 1(l 2) + a2 2 (l 2) ].
Requiring to be stationary with respect to a1 and a 2 , leads to two
equations relating the generalized coordinates
l l

( a1 1 + a2 2 ) 1 d x q 1 d x F 1(l 2) = 0 ,

= EI
a1
0 2l 3
146 FINITE ELEMENT ANALYSIS

l l

( a + a ) d x q

= EI 1 1 2 2 2 2 d x F 2 (l 2 ) = 0 .
a2
0 2l 3

Because is of second order in and , the equations are linear in


aj .

Substituting the functions (7.28), for F = q l 6 , we obtain

56 E I EI 11
3
a 1 + 10 3 a2 = ql,
5 l l 4320
EI 72 E I 269
10 a +
3 1 3
a2 = ql.
l 7 l 46656

For example, the coefficient of a1 in the first equation is


l l

l ( 36 x )
1 2 56 E I
( )
2 2
EI 1 dx = EI 8
30 l x + 4 l 2 dx = .
5 l3
0 0

The solutions are


q l4 q l4
a1 = 0.00207 , a2 = 0.00257 . (7.29)
EI E I
In 2, the displacement is

v (l 2 ) = a1 1 (l 2) + a2 2 (l 2) =
1 1 q l4
= a1 + a2 = 4.9 10 5 = 2.48 mm.
16 32 E Iy

Example 7.5
Consider a linear spring of stiffness k (Fig. 7.7, a) subjected to a load F.
Comment on the approximations of the Rayleigh-Ritz method.
Answer. The total potential energy (Fig. 7.7, b) is
1 2
= ku Fu . (7.30)
2
For a small virtual displacement u , the variation of the total potential
energy is
= k u u F u = ( k u F ) u .
7. ENERGY METHODS 147

The equilibrium equation is obtained by requiring to be zero ( be


stationary) for arbitrary u .

Fig. 7.7

For =0 we obtain

k ueq F = 0 . (7.31)

As seen in Fig. 7.7, b, the exact solution corresponds to an absolute


minimum value of .
The total potential energy of the equilibrium configuration is
1 1 1 F2
eq = F ueq F ueq = F ueq = . (7.32)
2 2 2 k
The stiffness is
1 F2 1 F2
k = = . (7.33)
2 eq 2 eq

The strain energy is


1 2 1 F2
U eq = k ueq = = eq . (7.34)
2 2 k
If F is prescribed and the resulting u has been approximated by a Rayleigh-
Ritz solution uapp ueq , equations (7.32) (7.34) and Fig. 7.7, b indicate that

a) Because eq is a minimum, the potential energy for an approximate


displacement which satisfies the kinematic boundary conditions is greater than the
true value
app > eq .
In magnitude
app < eq ,
148 FINITE ELEMENT ANALYSIS

the approximate total potential energy is underestimated.


b) The approximate stiffness is overestimated
1 F2
k = . (7.35)
2 eq

c) The approximate displacement is underestimated


F
u = . (7.36)
k
An approximate compatible displacement field corresponds to a structure
which is stiffer than the actual structure and therefore will give a lower bound on
displcement.

7.4 F.E.M. - a localized version of the Rayleigh-Ritz method

Instead of finding an admissible function satisfying the boundary


conditions for the entire domain, which is often difficult, in the FEM the
admissible functions are defined over small size subdomains.

7.4.1 F.E.M. in Structural Mechanics

a) Problem. Given a geometrically complex structure (including the


boundary conditions) and the external loads { pv } , { ps } , { Fi } , find the
displacement field { u } within V and on the surface S (Fig. 6.1). Then determine
stresses, internal forces, reaction forces, etc.
b) Solution approach. Use PVD or PMTPE as an approximate method for
solving the boundary-value problem. Admissible functions are defined over small
size finite elements, with simple geometry and well identified structural behaviour.
With these individually defined functions matching each other at certain points
(nodes) at the element interfaces, the unknown function is approximated piecewise
over the entire domain (continuity at global level).
c) Procedure. The geometric shape and the internal displacement field are
described by a series of discrete quantities (like nodal coordinates and nodal
displacements) distributed through the structure. For this a matrix notation is used.
d) Tools. Computers are used to store long lists of separate numbers and to
manipulate them, to present output data in an engineering format, taking advantage
of graphical and animation facilities.
7. ENERGY METHODS 149

7.4.2 Discretization

The structure is divided into finite elements (Fig. 7.8) that define the mesh.
Elements are defined by their nodal coordinates and some physical parameters.

Fig. 7.8

7.4.3 Principle of virtual displacements

For the entire structure, equation (7.6, a) can be written (considering only
surface tractions)

{ } { }dV { u } { ps }d A = 0 .
T T
= (7.37)
V S
As a summation of virtual works on all elements, PVD yields


= e
= { } { }dV { u } { ps }d A = 0 .
Ve
T

T


(7.38)
e e S e

In the following, only e will be considered. The aim is the calculation


of the element stiffness matrix and load vector.

7.4.4 Approximating functions for the element

In the Rayleigh-Ritz method, the trial function is expressed as a finite


expansion
150 FINITE ELEMENT ANALYSIS

a1
n a

u (x )
j =1

L

a j j (x ) = 1 2 L n 2 = { a }

(7.39)

an
where the undetermined constants a j have no direct evident signification.

The basic idea of FEM is to choose the constants - the displacement


{ }
unknowns at the nodes { a } = Q e and to prescribe admissible functions denoted
= N so that

u N u 0 0 { }
Qu
e

v = 0 N v { }

0 Qve
w 0
0 { }
N w Qw
e

or
{ u } = [ N ] {Q e }, (7.40)

where [ N ] is the matrix of shape functions (interpolation functions).


The reason is that elements are small enough so that the shape of the
displacement field can be approximated without too much error and only the
{ }
magnitude, defined by Q e remains to be found.
The proper selection of shape functions ensure the continuity of the
displacement field at global level. A finite element described by admissible shape
functions (integrable in the interior and with equal values of generalized
coordinates at element interfaces) is referred to as co-deformable or conforming.

7.4.5 Compatibility between strains and nodal displacements

From the compatibility relationship (6.18)

{ } = [ ] { u } = [ ][ N ] {Q e }= [ B ] {Q e }, (7.41)

where [ B ] is the matrix of differentiated shape functions.


The strain virtual variation is

{ } = [ B ] {Q e }.
7. ENERGY METHODS 151

7.4.6 Element stiffness matrix and load vector

Using the constitutive equation { } = [ D ] { } , the virtual work for an


element is

{ Q } [ B ] [ D ][ B ]{Q }dV { Q } [ N ] { ps }d A = 0
e T T e T T
e = e

Ve Ae
or

= { Q } [ B ] {Q } [ N ]

e e T T
[ D ][ B ] dV e T
{ ps }d A = 0 .

Ve Ae

{ }
As Q e are arbitrary and non-zero, cancelation of the bracket yields the
element equilibrium equation

[ K ] {Q }= {F },
e e e
(7.42)
where the element stiffness matrix is

[ K ]= [ B ]
e T
[ D ][ B ] dV (7.43)
Ve

and the vector of consistent nodal forces is

{F }= [ N ]
e T
{ ps }d A . (7.44)
Ae

7.4.7 Assembly of the global stiffness matrix and load vector

In the next step, all individual elements are assembled together so that the
displacements are continuous across element interfaces and the boundary
conditions are satisfied.
The kinematic connectivity is expressed by the relationship between
element and global displacements
~
{ } [ ]
Q e = T e { Q }, (7.45)

{ }
where Q e is the vector of nodal element displacements, {Q } is the vector of the
~
[ ]
global displacements of the structure and T e is a connectivity matrix, containing
ones at the nodal displacements of element nodes and zeros elsewhere.
The variation of element displacements is
152 FINITE ELEMENT ANALYSIS

{ Q }= [ T~ ] { Q }.
e e
(7.46)
The PVD equation for the entire structure is

{Q } [ K ] {Q }= {Q } {F },
e T e e e T e

e e
or using (7.45) and (7.46)

{Q }T [ T~ ] [ K ][ T~ ] {Q } = {Q } [ T~ ] {Q }.
e T e e T e T e

e e
As { Q } are arbitrary and non-zero, the unreduced global equilibrium
equations are
[ K ] {Q } = { F } , (7.47)
where the global stiffness matrix is

[ K ]= [ T~ ] [ K ][ T~ ]
e T e e
(7.48)
e
and the global load vector is

{ F } = [ T~ e ] T {Q e }. (7.49)
e
Applying the boundary conditions, the condensed equilibrium equations
are
[ K ] {Q } = { F } . (7.50)
The above procedure is never used in practice. It has been used only to
show algebraically how to assemble a global stiffness matrix. The assembly is done
by directly placing the nonzero entries of element stiffness matrices in the right
locations of the global stiffness matrix based on element connectivity.

7.4.8 Solution and back-substitution

Nodal displacements are determined by solving the linear equations (7.50).


In the back-substitution phase, element stresses are evaluated as

{ }
{ } = [ D ] { } = [ D ][ B ] Q e = [ D ][ B ] T~ e { Q } [ ]
where [ D ] is given by (6.24) or (6.25).
8.
TWO-DIMENSIONAL ELEMENTS

Many engineering structures can be modeled as two-dimensional flat


plates, designed to be primarily loaded in their plane and to resist loads by
membrane action rather than bending. In-plane displacement, strain and stress
components are uniform through the plate thickness, which is considered constant.
Only transversely homogeneous plates will be analyzed herein, composite and
sandwich plates being studied in other courses. This chapter presents the element
stiffness matrices and consistent force vectors for triangular and rectangular
elements, that allow closed form derivations, without the need for numerical
integration. The very first approximate finite element developed in 1956 to model
delta wing skin panels, the three-noded triangle with constant strain field, is treated
separately.

8.1 The plane constant-strain triangle (CST)

Before the advent of arbitrarily shaped isoparametric elements, discussed


in the next chapter, the CST was one of the most widely used elements and is still
available in systems today. It is a much more adaptable shape than the rectangle
and it allows the user to tailor the element mesh to suit any structural geometry. A
large number of small elements can be densely packed into a region of expected
high stress gradients, and uniformly stressed regions can be left with a small
number of larger triangles.

8.1.1 Discretization of structure

The plate is divided into a number of straight-sided triangles (Fig. 8.1, a),
joined together at their corners (nodes), so that the corners of adjacent elements
have common displacements. The elements fill the entire region except of a small
region at the boundary. This unfilled region exists for curved boundaries and it can
be reduced by choosing smaller elements.
154 FINITE ELEMENT ANALYSIS

The three nodes of the isolated element from Fig. 8.1, b are numbered
locally as 1, 2 and 3. The corresponding nodal coordinates are designated as
( x1 , y1 ) , ( x2 , y2 ) and ( x3 , y3 ) . The numbering is in anticlockwise direction to
avoid calculating a negative area.

a b
Fig. 8.1

Each node is permitted to displace in the two directions x and y. Thus, each
node has two degrees of freedom. The displacement components of a local node j
are denoted as u j in the x direction and v j in the y direction. The vector of
element nodal displacements is defined as

{ q }= u
e
1 v1 u2 v2 u3 v3 T
. (8.1)

8.1.2 Polynomial approximation of the displacement field

The displacements u and v of a point within the triangle are expressed in


terms of the nodal displacements. Because for two displacements there are six
boundary conditions, the assumed displacement field is linear
u (x , y ) = a1 + a2 x + a3 y ,
(8.2)
v ( x , y ) = a4 + a5 x + a6 y
with six arbitrary parameters.
8. TWO-DIMENSIONAL MEMBRANES 155

The strains (6.16) are


u v
x = = a2 = const . , y = = a6 = const . ,
x y

u v
xy = + = a3 + a5 = const . ,
y x
hence the name constant strain triangle.

8.1.3 Nodal approximation of the displacement field

In the polynomial approximation (8.2) the constants ai have no physical


meaning. A nodal approximation is preferred, in which the constants are the nodal
displacements and the displacement field is obtained by interpolation based on
values of corner displacements.
Since the functions for u and v are of the same form, only one need be
considered in detail. We can write
a1

u = 1 x y a2 = 1 x y { a } . (8.3)
a
3
Evaluating the expression for u at the three nodes gives the nodal
displacements in terms of the polynomial coefficients

u1 1 x1 y1 a1
a
u2 = 1 x2 y2 2 , (8.4)
u 1 x y3 a3
3 3
or

{ u }= [ A ] { a } ,
e
(8.4, a)
where
1 x1 y1
[A ] = 1 x2 y2 . (8.5)
1 x3 y3

By inversion

{ a } = [ A ] 1 { u e }. (8.6)
where
156 FINITE ELEMENT ANALYSIS

1 2 3
1
[A ] 1
= 1 2 3 . (8.7)
2A
1 2 3

in which
i = x j yk xk y j , i = y j yk , i = xk x j (8.8)

and the subscripts i , j , k permute in a natural order.


The area of the triangle A is equal to one half the magnitude of the
determinant of [ A ]

1 x1 y1
2 A = 1 x2 y2 = (x2 y3 x3 y2 ) + ( x3 y1 x1 y3 ) + (x1 y2 x2 y1 ) . (8.9)
1 x3 y3

The determinant is positive if nodes 1, 2, 3 are labeled anticlockwise


around the element.
Substituting (8.6) into (8.3) gives

u = N {u }e
(8.10)
where the row vector of shape functions

N = N1 N 2 N 3 = 1 x y [ A ] .
1
(8.11)

By transposition

N = [A ] 1 x y ,
T T T

N1 1 1 1 1
1
N2 = 2 2 2 x (8.12)
N 2 A 3 3 y
3 3
or
1
Ni = ( i + i x + i y ) . ( i = 1, 2, 3 ) (8.12, a)
2A

Similarly

v = N v e .{ } (8.13)
8. TWO-DIMENSIONAL MEMBRANES 157

The shape functions (8.12, a) can also be written

1
N1 (x , y ) = [ x2 y3 x3 y2 + ( y2 y3 ) x + (x3 x2 ) y ] ,
2A
1
N 2 (x , y ) = [ x3 y1 x1 y3 + ( y3 y1 ) x + (x1 x3 ) y ] , (8.12, b)
2A

1
N 3 (x , y ) = [ x1 y2 x2 y1 + ( y1 y2 ) x + (x2 x1 ) y ] .
2A

Fig. 8.2

The shape functions N i vary linearly, have a unit value at node i and zero
values at the other two nodes, as illustrated in Fig. 8.2:

N i ( xi , yi ) = i j , ( i , j = 1, 2, 3 ) (8.14)

and
3
Ni = 1 . (8.15)
i =1

Combining (8.10) and (8.14) yields


158 FINITE ELEMENT ANALYSIS

u1
v
1
u N1 0 N2 0 N3 0 u2
= (8.16)
v 0 N1 0 N2 0 N 3 v2
u3

v3
or

u
=[N ] { q e }. (8.16, a)
v

Fig. 8.3

The displacement u (x,y ) =u1 N1 (x,y ) + u2 N 2 (x,y ) + u3 N 3 (x,y ) determines a


plane surface passing through u1 , u2 and u3 , as shown in Fig. 8.3.

8.1.4 The matrix [B ]

Strains are expressed in terms of displacements as



0 0
x x
{ } e
= [ ] { u }= 0

u
= 0
y v

y
[N ] { q e }= [ B ] { q e }.


y x y x
8. TWO-DIMENSIONAL MEMBRANES 159

The matrix of the derivatives of shape functions is


N1 N2 N3
0 0 0
x x x
N1 N2 N3
[ B ] = 0 0 0 ,
y y y
N N1 N2 N2 N3 N3
1
y x y x y x

1 0 2 0 3 0
1
[ B ]= 0 1 0 2 0 3 ,
2A
1 1 2 2 3 3

y2 y3 0 y3 y1 0 y1 y2 0
1
[ B ]= 0 x3 x2 0 x1 x3 0 x2 x1 . (8.17)
2A
x3 x2 y 2 y3 x1 x3 y3 y1 x2 x1 y1 y2

The matrix [B] is constant for a given CST element. Sometimes it is


written simply
y23 0 y31 0 y12 0
1
[ B ]= 0 x32 0 x13 0 x21 (8.17, a)
2A
x32 y23 x13 y31 x21 y12

where the notation is obvious.

8.1.5 Element stiffness matrix and load vector

The element stiffness matrix (7.43) is

[k ]= [ B ]
e T
[ D ][ B ] dV = [ B ]T [ D ][ B ] t e A, (8.18)
Ve

where t e is the element thickness, A is the element area and [ D ] is the material
stiffness matrix given by (6.24) for plane stress and by (6.25) for plane strain
conditions.
The consistent nodal forces due to traction loads acting on a portion of the
boundary are calculated as for a linear two-node element. Along an edge 1-2, the
shape function N 3 is zero while N1 and N 2 are similar to the shape functions in
one dimension, satisfying N1 + N 2 = 1 . When the surface load distribution (per unit
160 FINITE ELEMENT ANALYSIS

area) is linear, varying from p1 at node 1 to p2 at node 2 (Fig. 8.4), the nodal
forces are
lte lte
f1e = ( 2 p1 + p2 ) , f 2e = ( p1 + 2 p2 ) (8.19)
6 6

where t e is the thickness of the element. Because of linearity they coincide with
the static resultants.

Fig. 8.4
The nodal forces associated with the weight of an element are equally
distributed at the nodes.

8.1.6 Remarks

A mesh like in Fig. 8.5, a is clearly a directionally sensitive assembly, and


this could be corrected by using the union jack pattern of Fig. 8.5, b which
however produces a larger bandwidth. Benchmark tests using triangular elements
have shown that CST elements, even in a fine mesh, are much inferior to higher
order elements in a coarse mesh.

Fig. 8.5
8. TWO-DIMENSIONAL MEMBRANES 161

A drawback of the displacement form of the finite element method is that


equilibrium is only satisfied in the mean or over the element. This means that along
an edge which is common to two elements the stresses are different across the
edge, where they should be continuous. Most programs contain facilities for
averaging the stresses. The simplest form of averaging consists of simply
connecting the centroids of two adjacent triangles and to assign the mean stress
value to the crossing point of this line with the common edge.
A simple example of stress averaging is shown in Fig. 8.6 for a square
plate with a circular hole. Taking advantage of symmetry, only one quarter of the
plate is considered. The diagram compares the theoretical stress distribution along
the marked line with the averaged values calculated using CSTs.

Fig. 8.6
162 FINITE ELEMENT ANALYSIS

The procedure will be used in Example E8.3

Example 8.1
A square plate of thickness t and Youngs modulus E is pin-jointed at
three corners and subjected to a force F at the free corner (Fig. E8.1, a). Let
= 0 . Divide the plate into four CST elements and find: a) the nodal
displacements; b) stresses in elements; and c) the support reactions.

a b
Fig. E8.1

Solution. The material stiffness matrix (6.24) is


1 0 0
[ D ] = E 0 1 0 . (a)
0 0 1 2

With the origin of coordinate axes at the plate centre, the nodal coordinates
of element 1 are
x1 = y1 = 0 , x2 = l , y2 = 0 , x3 = 0 , y3 = l .

The area is A = l 2 2 and the matrix (8.17) is

1 0 1 0 0 0
[B ]
1 1
= 0 1 0 0 0 1 .
l
(b)
1 1 0 1 1 0

The element stiffness matrix (8.18) is


8. TWO-DIMENSIONAL MEMBRANES 163

[ K ]= [k ]= [ B ]
1 1 1 T
[ D ] [ B1 ] t A , (c)

3 2 1 2 1 1 2 1 2 0
32 0 1 2 1 2 1

[K ]
1
=
Et
2

1 0
12 12
0 0
0
. (d)

SYM 12 0

1

Element 2 can be obtained by rotating 900 anticlockwise element 1. In


general, the transformation matrix for a rotation with an angle is

c s 0 0 0 0
s c 0 0 0 0

[T ]
e
0
=
0
0 s c
c s 0 0
(e)
0 0 0
0 0 0 0 c s

0 0 0 0 s c
where c = cos and s = sin .

The stiffness matrix of the rotated element is

[ K ] = [T ] [ k ] [T ]. .
e e T e e
(f)

For = 900 we obtain

0 1 0 0 0 0
1 0 0 0 0 0

[T ]
2
0
=
0
0 1 0
0 1 0 0
(g)
0 0 0
0 0 0 0 0 1

0 0 0 0 1 0

The stiffness matrix of element 2 is

[ K ] = [T ] [ k ] [T ]. ,
2 2 T 1 2
164 FINITE ELEMENT ANALYSIS

3 2 1 2 1 2 0 1 1 2
32 1 2 1 0 1 2

[K ]2
=
Et
2

12 0 0 1 2
1 0 0
. (h)

SYM 1 0

1 2
The same matrix is obtained by substituting the nodal coordinates
x1 = y1 = 0 , x2 = 0 , y2 = 1 , x3 = 1 , y3 = 0
in (8.17) and performing the product (8.18).
Due to symmetry, only the upper half of the plate can be considered (Fig.
E8.1, b). Using the appropriate boundary conditions, the finite element equations
can be written
3 0 1 1 2 1 0 1 1 2 u1 0
0
3 0 1 2 0 2 0 1 2 0 V1
1 0 1 0 0 0 0 0 u2 F 2

Et 1 2 1 2 0 1 2 12 0 0 0 0 V2
= .
2 1 0 0 12 1 0 0 1 2 0 H3

0 2 0 0 0 2 0 0 0 V3
1
0 0 0 0 0 1 0 0 H 4

1 2 1 2 0 0 1 2 0 0 1 2 0 V4

Solving
E t 3 1 u1 0
=
2 1 1 u2 F 2

we obtain the nodal displacements


F 3F
u1 = , u2 = .
2E t 2E t
The reaction forces for half the plate are obtained as

H3 = F 4 , H 4 = F 4 ,
so that for the entire plate

H4 = F 2 , H5 = F 4 .
8. TWO-DIMENSIONAL MEMBRANES 165

Strains and stresses in elements are


u1
0
1 0 1 0 0 0
{ }= [ B ]{ q }
1 1 1 1
= 0 1 0 0 0 1 2
l
u
0

,
1 1 0 1 1 0
0

0
1 1 2
{ }
1 1 u1
= 0 0 =
l
F
0 .
u 2 2l E t
1 0
1

x 1 0 0 2 2

y

{ }
=[D ] = E
1 0 1 0 F 0 = F 0 ,
2l E t 2 l t
0 0 1 2
xy 1 1 1 2

u1
0
1 0 0 0 1 0 1
{ }= [ B ] { q }
2 2 2 1
= 0 1 0 1 0 0 =
l
0 F
0 ,
0 2l E t
1 1 0 0 1 1
0 1

0

x 1 0 0 1 1

{ }

y =[D ] = E 0 1 0
2 F F
0 = 0 .
2l E t 2 l t

0 0 1 2
xy 2 1 1 2

Example 8.2

A thin triangular plate is fixed along the edge 5-4 and loaded by forces
F1 = 1 and F2 = 2 along the upper edge in its plane (Fig. E8.2). Assume the
cantilever to have unit thickness t = 1 and be in plane stress, with Youngs modulus
E = 1 and Poissons ratio = 0.3 . Divide the plate into three CST elements and
find: a) the nodal displacements; b) stresses in elements; and c) the support
reactions. The units are coherent.
166 FINITE ELEMENT ANALYSIS

Fig. E8.2

Solution. The input data are given below

For each element, the stiffness matrix is calculated longhand from equation
(8.18), where the matrix [ B ] is obtained from (8.17), based on the nodal
coordinates, and the matrix [ D ] is given by (6.24).
8. TWO-DIMENSIONAL MEMBRANES 167

The element stiffness matrices are the following:

The condensed global stiffness matrix is

The condensed finite element equations can be written

0.317 0 0.317 0.165 0 0.165 u1 0



0
0.111 0.192 0.111 0.192 0 v1 1

0.317 0.192 1.301 0 0.666 0 u2 0
= .
0.165 0.111 0 2.125 0 1.903 v2 2
0 0.192 0.666 0 1.301 0 u3 0

0.165 0 0 1.903 0 2.125 v3 0

The solution to these equations gives


u1 = 7.712 , u2 = 6.542 , u3 = 2.686 ,
v1 = 40.823 , v2 = 15.835 , v3 = 13.582 .
168 FINITE ELEMENT ANALYSIS

The output data are presented below:

The reaction forces at nodes 4 and 5 are obtained from the unused
unreduced equations.

Note that the adopted discretization is very crude, to allow longhand


calculation, and this leads to misleading results. Along the edge 4-5, a distribution
of bending stresses from compressive at 4 to tensile at 5 is expected. However,
element 3, being a constant strain element, gives only one value of x , which is
wrong. Normally, the plate should be modeled by many more elements. Ascribing
stress values to the element centroids and averaging them as shown in section
8.1.6, a more realistic stress distribution along the edge 4-5 is obtained.

Example 8.3

A thin rectangular plate, containing a circular hole of radius a = 10 mm , is


subjected to loads that produce uniform tensile stresses 0 = 5 MPa at its ends
(Fig. E8.3, a). The plate has length l = 60 mm , width b = 40 mm , thickness
t = 5 mm , E = 210 GPa and = 0.3 . Determine: a) the deformed shape of the
hole; b) the location and magnitude of the maximum von Mises stress in the plate;
c) the distribution of x stresses in the midsection. Compare the stress values at
the periphery of the hole obtained by FEM and from the theory of elasticity.
Solution. Taking advantage of the symmetry of geometry and symmetry of
loading, we can analyze only one-quarter of the plate (upper right).
8. TWO-DIMENSIONAL MEMBRANES 169

Fig. E8.3, a

A 55-node, 81-element mesh is created as shown in Fig. E8.3, b. Let x and


y represent the axes of symmetry. The points along the x axis are constrained in the
y direction, and points along the y axis are constrained along the x direction.

Fig. E8.3, b

The applied nodal forces are shown, but the element numbering is omitted
for clarity. The centroids of the elements near the midsection are marked, and the
crossing points of the lines connecting the centroids with the common sides (where
stresses are averaged), are denoted a to f.
The deformed shape is shown in Fig. E8.3, c. The hole is elongated in the
direction of the loading axis.
170 FINITE ELEMENT ANALYSIS

Fig. E8.3, c

The calculation of stresses is summarized in Fig. E8.3, d. Elements are


hatched according to the value of von Mises stresses, using five intervals with
limits shown in the legend.

Fig. E8.3, d

The maximum von Mises stress is 19.6 MPa and occurs in element 1.
Stress values in elements near the midsection are given in Table E8.3.
8. TWO-DIMENSIONAL MEMBRANES 171

Table E8.3

Element x y xy eq

1 20.0178 0.8845 -0.5706 19.6155


2 11.4577 1.9228 -0.6221 10.6821
3 11.3275 0.815 0.0595 10.9433
4 8.068 1.5768 0.927 7.5786
5 7.6445 0.1651 0.174 7.5693
6 5.7162 0.2632 0.6667 5.7073
7 4.0752 0.763 0.3595 3.8037

Stresses x , averaged at points a to f, have the following values:

Point a b c d e f

x , MPa 15.78 11.39 9.69 7.85 6.68 4.89

The theoretical distribution of stresses x in the midsection of a plate


containing a small circular hole and subjected to uniaxial tension is approximated
by
a2 3 a4
x = 0 1 + 2 + 4 ,
2r 2r

where r is the distance from the x axis (Fig. E8.3, a).


The maximum stress occurs at the periphery of the hole (r = a ) and is

xmax = 3 0 = 15 MPa .

The averaged value of elements 1 and 2, in point a, is xa = 15.78 MPa ,


which is surprisingly accurate for the rough mesh used in the exercise.

Example 8.4

An axially loaded thin plate with a circular fillet (Fig. E8.4, a) has length
150 mm , width of the reduced portion 40 mm , width of the extended portion
80 mm , thickness 5 mm , fillet radius 20 mm , E = 2 105 MPa and = 0.25 . The
normal stress at a point far to the right of the fillet has a uniformly distributed value
172 FINITE ELEMENT ANALYSIS

of 440 MPa . Find the location and magnitude of the maximum von Mises stress in
the plate. Determine the stress concentration factor for the circular fillet.

Fig. E8.4, a

Answer. Taking advantage of the symmetry, only half of the plate is


considered. The points along the symmetry axis are constrained in the vertical
direction. Points along the left edge are constrained in the horizontal direction. A
model comprising 95 nodes and 144 CST elements is constructed as shown in Fig.
E8.4, b.

Fig. E8.4, b

The nodal loads are calculated from equation (8.19). The right edge has
four elements, each with a surface of 25 mm 2 , subjected to a normal stress of
440 N mm 2 . The nodal loads for each element are 440 25 / 2 = 5500 N . The
resulting five nodal forces, for the quarter plate, have magnitudes of 5500 N at the
upper and lower node, and 11000 N at the three middle nodes.
The distribution of von Mises stresses is shown in Fig. E8.4, c for five
stress intervals given in the legend. The largest equivalent stress is 604.7 MPa .
8. TWO-DIMENSIONAL MEMBRANES 173

Fig. E8.4, c

Values of the principal stress 1 around the fillet are presented in


the upper part. The largest principal stress 1 is 622.7 MPa . The stress
concentration factor relative to the value 440 MPa is 1.415. This value is
reasonably close to the theoretical true value of 1.42 given by Singer (1962).

Example 8.5

The nodal coordinates and the nodal displacements of element 96 in Fig.


E8.4, b are given below:
x60 = 92.35 , y60 = 21.52 , x64 = 95 , y64 = 18 , x65 = 95.5 , y65 = 20.5 ;
u60 = 0.11224 , v60 = 3.97e 3 , u64 = 0.12367 , v64 = 3.63e 3 ,
u65 = 0.12247 , v65 = 4.92e 3 .

E = 2 105 MPa and = 0.25 . Calculate the element stresses.

Solution. The triangle area is A96 = 4.1925 mm 2 .

The matrix [ B ] is (8.17)

0.2982 0 0.1216 0 0.4198 0



[ B ]= 0 0.0596 0 0.3757 0 0.3160 .
0.0596 0.2982 0.3757 0.1216 0.3160 0.4198
174 FINITE ELEMENT ANALYSIS

The matrix [ D ] is (6.24)

2.1333 0.5333 0
[ D ] = 10 0.5333 2.1333 0 .
5

0 0 0.8

The vector of nodal displacements (8.1) is


{ q } = 0.11224 3.97e 3 0.12367 3.63e 3 0.12247 4.92e 3 T .

Element stresses are given by


{ } = [ D ][ B ] { q },
{ } = x y xy T = 596.72 63.58 120.08 T .

Example 8.6

Find the deformed shape and the stress distribution in the loaded gear tooth
shown in Fig. E8.6, a. The load is not applied at the tooth tip, but is distributed
over a larger area than for the mating contact of two teeth, to concentrate only on
the influence of fillet geometry.

Fig. E8.6, a
8. TWO-DIMENSIONAL MEMBRANES 175

Answer. The adopted finite element mesh has 126 nodes and 212
CST elements. Note the restraints which model the tooth built-in end. The
deformed shape is presented in Fig. E8.6, b.

Fig. E8.6, b

The stress distribution is shown in Fig. E8.6, c. Note the stress


concentration at both fillet areas, where the equivalent von Mises stresses are
indicated.

Fig. E8.6, c
176 FINITE ELEMENT ANALYSIS

8.2 Rectangular elements

Rectangular elements are the easiest to discuss and are used to model
stiffened thinwalled panel structures, built up beams and boxes.

8.2.1 The four-node rectangle (linear)

Consider the 4-node element in Fig. 8.7.

Fig. 8.7

We have to find two-dimensional shape functions which have unit values


at one selected node, but zero at all other nodes on the element
( )
N i x j , y j = ij . (8.20)

They can be generated using the equations of the sides. For example, N1 is
identically zero on lines x = a and y = b , and has a unit value at the node with the
same index N1 ( x1 , y1 ) = 1 , hence it must be of the form

N1 ( x , y ) = c1 ( a x ) ( b y ) .
1
From condition N1 ( x1 , y1 ) = N1 ( 0,0 ) = 1 we get c1 = , hence
ab
N1 ( x , y ) =
1
( a x ) ( b y ) = 1 x 1 y .
ab a b
Likewise, the remaining three shape functions are
x y x y y x
N2 ( x, y ) = 1 , N3 ( x, y ) = , N4 ( x, y ) = 1 .
a b a b b a
It is convenient to transform to dimensionless central coordinates
8. TWO-DIMENSIONAL MEMBRANES 177

2x 2y
r= 1, s= 1. (8.21)
a b
The equations of element sides become
1 r = 0 ,. 1 s = 0 . (8.22)
The required shape functions are
1
N1 = (1 r ) (1 s ) , N2 = 1 (1 + r ) (1 s ) ,
4 4 (8.23)
1 1
N3 = ( 1 + r ) ( 1 + s ) , N 4 = ( 1 r ) ( 1 + s ) .
4 4
The displacements inside the element can be expressed in terms of the
nodal displacements as
u N1
=
0 N2 0 N3 0 N4 0 e
N 4
{ }
q , (8.24)
v 0 N1 0 N2 0 N3 0

where
{q }e T
= u1 v1 u2 v2 u3 v3 u4 v4 . (8.25)

The use of the shape functions (8.23) has one drawback they are not
complete. That is, all like powers in x and y are not present. In this case x y is
present but x 2 and y 2 are not. Thus the variation of strain does not have the same
order in all directions.
The displacements u and v of a point within the rectangle can be
expressed in terms of the nodal displacements by a polynomial approximation.
Because, for two displacements, there are eight boundary conditions, the assumed
displacement field is
u (x , y ) = a 1 + a2 x + a3 y + a4 x y ,
(8.26)
v ( x , y ) = b1 + b 2 x + b 3 y + b4 x y .

The direct strain x = a2 + a4 y is constant in the x direction whereas the


shear strain xy = a3 + a4 x + b2 + b4 y varies linearly with x and y .

Some improvement can be gained by diminishing the shear variation


through the use of reduced integration on the shear contribution in the element
stiffness matrix. Many users prefer to use higher order elements rather than use a
larger number of small 4-noded rectangles. The 4-noded rectangular element is
exploited in non-linear problems like elasto-plastic behaviour, where the stiffness
has to be reevaluated as the load increments are applied and plasticity spreads.
178 FINITE ELEMENT ANALYSIS

The element stiffness matrix (7.39) is

[k ]= [ B ]
e T
[ D ][ B ] dV ,
Ve
where
[ B ] = [ ][ N ] .
The poor performance of the 4-noded element is shown below, in an
example where it is expected to behave like a slender beam in which bending
stresses are dominant.

Fig. 8.8

Figure 8.8, a shows the deformed shape of a single element in pure


bending. In comparison, Fig. 8.8, b shows the expected circular deformed shape
free of shear deformations. The 4-node element has flat sides. At the element ends
the shear strain is xy = d b and only the centre has no shear deformation. The
direct strains are x = d a on the upper and lower surface. The ratio
spurious shears a
= = aspect ratio .
bending strains b
This explains the poor results obtained with high aspect ratio elements
( 3) which have also badly conditioned stiffness matrices. The stress
discontinuities at element interfaces can be comparable with the mean values and
the free edge stresses are not zero. Generally, because the strain energy is
underestimated, stresses are lower than the true values.

8.2.2 The eight-node rectangle (quadratic)

The higher order 8-node rectangular element is shown in Fig. 8.9. Adding
four nodes means adding four supplementary boundary conditions (or nodal
displacements) so that in the polynomial approximation we can add four more
(higher order) terms.
8. TWO-DIMENSIONAL MEMBRANES 179

The assumed displacement field is cubic

u (x , y ) = a 1 + a2 x + a3 y + a4 x 2 + a5 x y + a6 y 2 + a7 x 2 y + a8 x y 2 ,
(8.27)
v ( x , y ) = b1 + b2 x + b3 y + b4 x 2 + b5 x y + b6 y 2 + b7 x 2 y + b8 x y 2 ,

so that the strains are quadratic. Note that choosing the last terms in x 3 and y 3
would make the element more anisotropic. The 16 constants can be expressed
(interpolated) in terms of the 16 nodal displacements and the nodal coordinates.
Using the nodal approximation, the displacement field is expressed as

u = N1 u1 + N 2 u2 + N 3 u3 + ... + N8 u8 ,
(8.28)
v = N1 v1 + N 2 v2 + N 3 v3 + ... + N8 v8 ,

where the shape functions N i can be easily built up based on the equations of the
lines passing through the nodes (serendipity approach).

Fig. 8.9

In Fig. 8.9 only the lines through the new nodes are shown for clarity, the
old ones are presented in Fig. 8.7. In the natural system of coordinates r and s, the
shape functions have the following expressions

N1 =
1
( 1 r ) ( 1 s ) (1 + r + s ) , N 2 = 1 ( 1 + r ) ( 1 s ) (1 r + s ) ,
4 4
1 1
N 3 = ( 1 + r ) ( 1 + s ) ( 1 r s ), N 4 = ( 1 r ) ( 1 + s ) (1 + r s ) ,
4 4
(8.29)
1 1
N5 = ( 1 r ) ( 1 + r ) ( 1 s ) , N6 = ( 1 + r ) ( 1 s ) ( 1 + s ) ,
2 2
1 1
N 7 = ( 1 r ) ( 1 + r ) (1 + s ) , N 8 = ( 1 r ) ( 1 s ) ( 1 + s ).
2 4
180 FINITE ELEMENT ANALYSIS

The interpolation function N1 should be zero at nodes 2, 3,..., 8 and have


a value of unity at node 1. Equivalently, N1 should vanish on the sides defined by
the equations of lines passing through nodes 2 to 8: 1 s = 0 , 1 r = 0 ,
1 + r + s = 0 . Therefore N1 is of the form
N1 = c1 ( 1 r ) ( 1 s ) (1 + r + s ) ,

where c1 is a constant which is determined so as to yield N1 ( 1,1) = 1 , i.e.


c1 = 1 4 , which gives the expression of N1 in (8.29).

The displacements inside the element can be expressed in terms of the


nodal displacements as

u N1
=
0 N2 0 L L N8 0 e
N 8
{ }
q , (8.30)
v 0 N1 0 N2 L L 0

where

{ q }= u
e
1 v1 u2 v 2 L L u8 v8 T . (8.31)

Having generated the shape functions [ N ] , the functions [ B ] = [ ][ N ]


[ ]
follow and the element stiffness matrix k e can be evaluated using [ D ] .

8.3 Triangular elements


In order to obtain accurate results for stresses with a constant strain
discretization, one has to use a large number of elements. Considerable effort has
been devoted to develop refined elements, i.e. elements having linear, quadratic
or higher-order strain expansions.

8.3.1 Area coordinates

For a triangle it is possible to define a completely symmetrical coordinate


system known as area coordinates.
The position of any point M in the triangle is identified by the
perpendicular distances h1 , h2 , h3 from the three sides (Fig. 8.10, a), and non-
dimensionalized, by division to the triangle heights, as
h1 h2 h3
1 = , 2 = , 3 = ,
H1 H2 H3
8. TWO-DIMENSIONAL MEMBRANES 181

or
A1 A2 A3
1 = , 2 = , 3 = . (8.32)
A A A
As A1 + A2 + A3 = A (Fig. 8.10, b), one obtains

1 + 2 + 3 = 1 , (8.33)
so the three coordinates are not independent and they behave like the shape
functions.

Fig. 8.10

The physical coordinates of point M can be expressed in terms of the nodal


coordinates as
x = 1 x1 + 2 x 2 + 3 x3 ,
(8.34)
y = 1 y1 + 2 y2 + 3 y3 ,
or in matrix form
1 1 1 1 1

x = x1 x2 x3 2 .
y y y2 y3 3
1
By inversion, we obtain equation (8.12)
1 1 1 1 1
1
2 = 2 2 2 x
2 A 3 3 y
3 3
or
182 FINITE ELEMENT ANALYSIS

1
i = ( i + i x + i y ) . ( i = 1, 2, 3 ) (8.35)
2A
where A is given by (8.9) and the expressions (8.8) are
i = x j yk xk y j , i = y j yk , i = xk x j .

8.3.2 Linear strain triangle (LST)

The linear strain triangle is a six-noded element, obtained adding three


mid-side nodes to the CST. In Fig. 8.11, the equations of the three sides and the
lines through the mid-side nodes are shown using area coordinates.

Fig. 8.11

The shape functions are readily seen to be


1 1 1
N1 = 2 1 1 , N 2 = 2 2 2 , N 3 = 2 3 3 ,
2 2 2 (8.36)
N 4 = 4 1 2 , N 5 = 4 2 3 , N 6 = 4 3 1 .

The displacements in terms of the nodal displacements are

u N1
=
0 N2 0 L L N6 0 e
N 6
{ }
q , (8.37)
v 0 N1 0 N2 L L 0
where
{ q }= u
e
1 v1 u2 v 2 L L u6 v6 T . (8.38)
8. TWO-DIMENSIONAL MEMBRANES 183

To form [ B ] = [ ][ N ] we need to find the derivatives of shape functions


with respect to the physical coordinates, so we must transform from x , y to
1 , 2 , 3 . Using (8.33) we can eliminate 1 in (8.34) obtaining

x = (x 2 x1 ) 2 + (x3 x1 ) 3 + x1 ,
(8.39)
y = ( y2 y1 ) 2 + ( y3 y1 ) 3 + y1 .
Using the chain rule for partial derivatives
x y
2 x

2
= x
2
y
=[J ] x , (8.40)

3 3 3 y y

where [ J ] is the Jacobian of the transformation.


The inverse relationship is

x 1 2 3 2
=[J ] 1 2 = , (8.41)
2A 2 3
y 3 3

where A is the triangle area (8.9) and i , i are defined by (8.8).


The element stiffness matrix is

[k ]= [ B ]
e T
[ D ][ B ] t e d A , (8.42)
A
where
d A = l 3 d 2 l 2 d 3 sin = 2 A d 2 d 3 .
The integration is frequently performed numerically. However, it can be
solved explicitly in terms of a, b and [ D ] using the integration formula for
monomials
a! b!

a b
i j dA = 2 A . (8.43)
A
(2 + a + b ) !
The results for first- and second-degree terms are
A A A

2
i dA = , i j dA = , i dA = . (8.44)
3 12 6
A A A
184 FINITE ELEMENT ANALYSIS

The LST element is based on expanding the displacements in a complete


second-degree polynomial in 2 and 3

u = a1 + a2 2 + a3 3 + a4 22 + a5 2 3 + a6 32 ,
(8.45)
v = b1 + b2 2 + b3 3 + b4 22 + b5 2 3 + b6 32 .
The displacement field is complete, containing all possible products, so the
element is truly isotropic, without recourse to computationally inefficient internal
nodes. There are twelve nodal displacements, six for each component. In order to
satisfy inter-element displacement compatibility, the displacement expansion on a
boundary must involve only the nodal quantities for that boundary. There are three
constants for this case (the function is quadratic) and an additional interior node is
required for each boundary. It is convenient to locate these interior nodes at the
mid-points of the sides.
Solving for the constants in terms of the nodal displacements, the nodal
approximation (8.37) is obtained.
Denoting
2 = r, 3 = s, 1 = 1 2 3 = 1 r s (8.46)
we can introduce oblique triangular coordinates (Fig. 8.12).
The physical coordinates of a point within the triangle are
x = ( 1 r s ) x1 + r x 2 + s x3 ,
(8.47)
y = ( 1 r s ) y1 + r y2 + s y3 .

Fig. 8.12

The coefficients of the nodal coordinates are genuine geometric


interpolation functions
N1 = 1 r s , N2 = r , N3 = s , (8.48)
8. TWO-DIMENSIONAL MEMBRANES 185

The nodal displacements for the 3-node triangle can also be written
u = ( 1 r s ) u1 + r u2 + s u3 ,
(8.49)
v = ( 1 r s ) v1 + r v2 + s v3 ,
so that the same functions can be used as interpolation functions. This is the basic
idea behind the isoparametric formulation, treated in the next chapter.

8.3.3 Quadratic strain triangle

A triangular element with a quadratic displacement field can be built up in


two ways. One possibility is to work with corner point nodes and two interior
nodes per side, taking as nodal quantities the displacement components. This
element is shown in Fig. 8.13, a.

Fig. 8.13

The displacements are expressed as complete cubic polynomials. In order


to satisfy inter-element displacement compatibility, the displacement function for a
side must depend only on the nodal displacement quantities for the side. Since the
function is cubic, four nodal quantities are required to define the distribution on a
side. The side nodes are located at the third points. An additional interior node is
needed to maintain completeness of the polynomial since, if the polynomial is not
complete, the stiffness will have preferred direction which is not desirable. It is
convenient to take the interior node at the centroid.
The displacement nodal expansion has the form
9 9
u = N i ui + N c u c , v = N i vi + N c v c , (8.50)
i =1 i =1
where
186 FINITE ELEMENT ANALYSIS

1 1
N1 = 1 ( 3 1 1 ) ( 3 1 2 ) , N 2 = 2 ( 3 2 1 ) ( 3 2 2 ) ,
2 2
1
N3 = 3 ( 3 3 1 ) ( 3 3 2 ) ,
2
9 9
N4 = 1 2 ( 3 1 1 ) , N 5 = 1 2 ( 3 2 1 ),
2 2 (8.51)
9 9
N6 = 2 3 ( 3 2 1 ), N 7 = 2 3 ( 3 3 1 ),
2 2
9 9
N8 = 3 1 ( 3 3 1 ) , N 9 = 3 1 ( 3 1 1 ) ,
2 2
Nc = 27 1 2 3 .

Another possibility is to work only with corner nodes, in order to reduce


the bandwidth of the stiffness matrix. The solution is to include displacement
u u v v
derivatives , , , as nodal quantities. At each corner, there are six
x y x y
nodal variables, two displacements and four first derivatives, a total of 18
parameters (Fig. 8.13, b). Two additional displacement quantities not associated
with the boundaries are required for completeness. It is convenient to take the
displacement components of the centroid (uc , vc ) as the remaining parameters.
The nodal expansion for u has the form
u = N1 u1 + N 2 ux1 + N 3 uy1 + N 4 u2 + N 5 ux 2 + ... + + N 9 uy 3 + N cuc . (8.52)

In (8.52) the interpolation polynomials, expressed in terms of triangular


coordinates, are
N1 = 12 ( 1 + 3 2 + 3 3 ) 7 1 2 3 ,
N 2 = 12 ( 3 2 2 3 ) + ( 2 3 ) 1 2 3 ,
N 3 = 12 ( 2 3 3 2 ) + ( 3 2 ) 1 2 3 ,
N 4 = 22 ( 2 + 3 3 + 3 1 ) 7 1 2 3 ,
N 5 = 22 ( 1 3 3 1 ) + ( 3 1 ) 1 2 3 ,
(8.53)
N 6 = 22 ( 3 1 1 3 ) + ( 1 3 ) 1 2 3 ,
N 7 = 32 ( 3 + 3 1 + 3 2 ) 7 1 2 3 ,
N8 = 32 ( 2 1 1 2 ) + ( 1 2 ) 1 2 3 ,
N 9 = 32 ( 1 2 2 1 ) + ( 2 1 ) 1 2 3 ,
N c = 27 1 2 3 ,
8. TWO-DIMENSIONAL MEMBRANES 187

where i , i are defined by (8.8) and the subscript c refers to the centroid.
The same interpolation functions are valid for v .

8.4 Equilibrium, convergence and compatibility

It is useful to make some general comments on the fulfilment of the


equilibrium and compatibility conditions in a finite element solution based on
assumed displacement fields [33].

8.4.1 Equilibrium vs. compatibility

In a FEM solution based on assumed displacement fields, one can say that:

a) Within elements, compatibility is satisfied if the assumed element


displacement field is continuous, but equilibrium is usually not satisfied.
Let write the equilibrium equations in terms of displacements for the case
pv x = pv y = 0 . Substituting the stress-strain relations (6.24) into the equilibrium
equations (6.7) we obtain

x
(
x + y + )
1
2 y
xy = 0,
(8.54)
1
2 x
xy +

y
(
y + x = 0. )
Substituting the strain-displacement relations (6.16) in (8.54) gives

2u 2u 1 + 2u 2 v
+ = ,
x2 y 2 2 y 2 x y
(8.55)
2 v 2 v 1 + 2 v 2 u
+ = .
x2 y 2 2 x 2 x y

Equations (8.55) are not satisfied by the assumed displacement field (8.26)
u ( x , y ) = a1 + a2 x + a3 y + a4 x y ,
(8.56)
v ( x , y ) = b1 + b2 x + b3 y + b4 x y .
of the rectangular element.

For example,
188 FINITE ELEMENT ANALYSIS

u u 2u 2u 2v
= a2 + a4 y , = a3 + a4 x , = = 0 , and = b4
x y x2 y2 x y
which is not zero, so the first equation (8.55) is not satisfied.
The rectangle would satisfy equilibrium if a4 = b4 = 0 , but this is the case
only in a field of constant strain. But as already shown, the direct strain x is linear
in the y direction whereas the shear strain xy varies linearly with x and y .

However, equilibrium is satisfied within the CST because of its extreme


simplicity. One cannot conclude from this that, in general, the rectangle is inferior
to the triangle. In some cases it can give better results.

b) Between elements, compatibility may or may not be satisfied, and


equilibrium is usually not satisfied.
For example, for both the 3-node triangle and the 4-node rectangle, u and
v are linear in x (or y ) along element edges. So, for any nodal displacement, the
edges remain straight, and adjacent elements do not overlap or separate, the
elements fit together. Inter-element equilibrium is obviously violated in the CST,
where stresses are constant within the element but differ from one element to
another. However, inter-element stress continuity may exist, as in the case of
uniform beams loaded only at nodes.
When inter-element compatibility is satisfied, the finite element solution
gives an upper bound on the total potential energy. As the discretization is refined,
the solution will converge monotonically to the true solution, provided that the new
mesh contains all the nodes of the previous meshes.

c) At nodes, compatibility is enforced by joining elements at these


locations, and equilibrium of nodal forces and moments is satisfied.
The finite element equations are a set of equilibrium equations and the
solution is such that resultant forces and moments acting on each node are zero.
Happily, in a proper finite element solution, any violations of equilibrium
and compatibility tend to vanish as more and more elements are used in the mesh.
Moreover, the convergence of displacements with the mesh refinement must be
monotonic from below.

8.4.2 Convergence and compatibility

As the mesh of elements is refined, the sequence of solutions to a problem


is expected to converge to the correct result if the assumed element displacement
fields satisfy the following criteria:
8. TWO-DIMENSIONAL MEMBRANES 189

a) The displacement field within an element must be continuous. This is


normally ensured by the selection of shape functions.

b) When the nodal degrees of freedom are given values corresponding to a


state of constant strain, the displacement field must produce the constant strain
state throughout the element.
The check is done using the so called patch test. The model consists of
an assembly of several elements arranged so that at least one node is completely
surrounded by elements. The boundary nodes are then given either displacements
or forces consistent with a constant strain state. Internal nodes are to be neither
loaded nor restrained. The computed displacements, strains, and stresses within
elements should be consistent with the constant strain state. If not, the element type
is invalid or at least suspect (it may happen that an element is valid in certain
configurations only).

c) Rigid body modes must be represented. When nodal degrees of freedom


are given values corresponding to a state of rigid body motion, the element must
exhibit zero strain and therefore zero nodal forces.
If this requirement is violated, extraneous nodal forces appear, and the
equations of nodal equilibrium are altered. To satisfy the requirements on both
rigid body modes and constant strain rates, the expansion must be at least a
complete polynomial of order equal to the highest derivative occurring in the
strain-displacement relations.

d) Compatibility must exist between elements. Elements must not overlap or


separate. In the case of beam, plate and shell elements, the slope must be
continuous across interelement boundaries.
This requirement is violated by many successful non-conforming elements.
Such elements do satisfy inter-element compatibility in the limit of mesh
refinement, as each element approaches a state of constant strain. However, non-
compatible elements do not provide a bound on the potential energy, i.e. we do not
know whether the potential energy corresponding to a particular non-conforming
element is higher or lower than the true value. Also, it is not possible to construct a
minimizing sequence with non-compatible elements. This means that by suitably
specializing the nodal displacements for the nth discretization, we will not be able
to reproduce the displacement patterns corresponding to the n 1 previous
discretizations.

e) The element should have no preferred directions. Elements should be


invariant with respect to the orientation of the load system. Invariance exists if
complete polynomials are used for element displacement fields. It is achieved even
when based on incomplete polynomials, if there is a balanced representation of
terms in the polynomial expansion.
190 FINITE ELEMENT ANALYSIS

Example 8.7

For the assembly of four triangular elements shown in Fig. E8.6, a simple
patch test is carried out as follows.

Fig. E8.7

Assume a displacement field


u=v= x+ y
for which the strains are

{ } = 1 1 2 T .
The resulting nodal displacements are
u1 = 0 , u2 = 1 , u3 = 4 , u4 = 1 ,
v1 = 0 , v2 = 1 , v3 = 4 , v4 = 1 .
If these are the boundary conditions imposed to the model, the internal
node must behave accordingly.
The condensed set of finite element equations has the form

[K ]
u5
= {F } ,
v5
where [K ] depends on the material properties and {F } depends on both the
material properties and the prescribed nodal displacements.
Its solution must be
u5 = v5 = 1.25

and the strains at any point must be { } = 1 1 2 T .


9.
ISOPARAMETRIC ELEMENTS

Simple triangular and rectangular elements allow closed form derivations


of stiffness matrices and load vectors. The construction of shape functions and
evaluation of stiffness matrices for quadrilateral and higher-order elements with
curved sides faces difficulties which are overcome by the use of isoparametric
elements and numerical integration. Elements with curved sides provide a better fit
to curved edges of an actual structure.
For isoparametric elements, the same interpolation functions are used to
define the element shape as are used to define the displacement field within the
element. The constant strain triangle is an isoparametric element though it was not
treated like that. In fact, the shapes of the interpolation functions and not the
parameters are the same. It is possible to construct subparametric elements whose
geometry is determined by a lower order model than the displacements. When the
eight node rectangle is transformed into a quadrilateral with straight sides, we
obtain a subparametric element. When it is mapped into a quadrilateral with
parabolically curved sides, the result is an isoparametric element. If we develop
higher order triangular elements while keeping straight sides, they are
subparametric elements, because only the displacement expansion is refined
whereas the geometry definition remains the same.

9.1 Linear quadrilateral element

Consider the general quadrilateral element shown in Fig. 9.1, a. The local
nodes 1, 2, 3 and 4 are labelled counterclockwise. The coordinates of node i are
( xi , yi ) . Each node has two degrees of freedom. The displacement components of
a local node i are denoted as ui in the x direction and vi in the y direction. The
vector of element nodal displacements is defined as

{ q }= u
e
1 v1 u2 v2 u3 v3 u4 v4 T . (9.1)
192 FINITE ELEMENT ANALYSIS

9.1.1 Natural coordinates

A natural coordinate system can be attached to a quadrilateral element as


illustrated in Fig. 9.1, b. The coordinates r and s vary from 1 on one side to + 1
at the other, taking the value zero over the quadrilateral medians. They are called
quadrilateral coordinates.

a b
Fig. 9.1

In the development of isoparametric elements it is useful to visualize the


quadrilateral coordinates plotted as cartesian coordinates in the { r , s } plane. This
is called the reference plane. In the reference plane, quadrilateral elements become
a square of side 2 (Fig. 9.2, a), called the reference element (or master element),
which extends over r [ 1, 1 ] , s [ 1, 1 ] . The transformation between the
natural coordinates { r , s } in the reference plane and the cartesian coordinates
{ x , y } is called the isoparametric mapping.

a b
Fig. 9.2
9. ISOPARAMETRIC ELEMENTS 193

Each quadrilateral child in the { x , y } is generated by the parent or


reference element from the { r , s } plane (Fig. 9.2, b). The advantage is that the
interpolation functions defined for the reference element are the same for all actual
elements and have simple expressions. The drawback is that the mapping is a
change of coordinates which implies complications in the evaluation of the integral
in the element stiffness matrix.

9.1.2 Shape functions

The interpolation functions (8.23) developed for the rectangle are directly
applicable
1 1
N1 = ( 1 r ) ( 1 s ) , N 2 = ( 1 + r ) ( 1 s ) ,
4 4 (9.2)
1 1
N3 = ( 1 + r ) ( 1 + s ) , N 4 = ( 1 r ) ( 1 + s ) .
4 4
The following compact representation is useful for the implementation in a
computer program
1
Ni ( r , s ) = ( 1 + r ri )( 1 + s si ) , (9.3)
4
where ( ri , si ) are the coordinates of node i .
The coordinates of a point within the element are expressed in terms of the
nodal coordinates as
4
x = N1 x1 + N 2 x 2 + N 3 x3 + N 4 x 4 = N i x i ,
i =1
(9.4)
4
y = N1 y1 + N 2 y2 + N 3 y3 + N 4 y4 = N i y i .
i =1
In matrix form
x1
x2
{ }
x = N x e = N1 N2 N3

N4
x3
, (9.4, a)

x4

y1

{ }
y = N y e = N1 N2 N3

N4
y2
y3

. (9.4, b)

y4
194 FINITE ELEMENT ANALYSIS

9.1.3 The displacement field

In the isoparametric formulation, the same shape functions are used to


express the displacements within the element in terms of the nodal values as are
used to define the element shape.
If u and v are the displacement components of a point located at ( r , s ) ,
then
4
u = N1 u 1 + N 2 u 2 + N 3 u 3 + N 4 u 4 = N i u i ,
i =1
(9.5)
4
v = N1 v1 + N 2 v2 + N 3 v3 + N 4 v4 = N i v i ,
i =1

which can be written in matrix form

{ u } = [ N ] { q e }, (9.6)
where

[ N ] =
N1 0 N2 0 N3 0 N4 0
. (9.7)
0 N1 0 N2 0 N3 0 N 4

As the shape functions in natural coordinates (9.2) fulfil continuity of


geometry and displacements both within the element and between adjacent
elements, the compatibility requirement is satisfied in cartesian coordinates too. As
mentioned in Chapter 8, polynomial models are inherently continuous within the
element. Then it is easy to show that the displacements along a side of the element
depend only upon the displacements of the nodes occurring on that side.
If the u displacement is approximated as
u ( r , s ) = a1 + a2 r + a3 s + a4 r s , (9.8)
along each side the approximation is linear, because r (or s) is constant.
For example, along the side 2-3 (Fig. 9.2, a), r = 1 and
u r =1
= a1 + a2 + ( a3 + a4 ) s ,

1 s 1+ s
u r =1
= u2 + u3 .
2 2
The four-node quadrilateral is termed a linear (sometimes bi-linear)
element because its sides remain straight during deformation. This ensures inter-
element compatibility, i.e. there are no openings, overlaps or discontinuities
9. ISOPARAMETRIC ELEMENTS 195

between the elements. The 4-node isoparametric quadrilateral is a conforming


element.
Also, because the interpolation displacement model provides rigid body
displacements in the natural coordinate system, the conditions of both rigid body
displacements and constant strain states are satisfied in cartesian coordinates too.
On the contrary, the 4-node quadrilateral is a non-conforming element.
If the u displacement is approximated as
u ( x , y ) = b1 + b2 x + b3 y + b4 x y , (9.9)
along each side the approximation is quadratic.
For example, the equation of the side 2-3 (Fig. 9.1, a) has the form
y = m x + c , where m is the slope of 2-3, so that

u ( x) 2 3
= b1 + cb3 + ( b2 + m b3 + c b4 ) x + m b4 x 2 .

Along 2-3, u ( x ) varies quadratically and cannot be uniquely defined as a


function of u 2 and u 3 . The approximation (9.9) is non-conforming.

9.1.4 Mapping from natural to cartesian coordinates

Subsequently, we need to express the derivatives of a function in { x , y }


coordinates in terms of its derivatives in { r , s } coordinates. This is done
considering the function f = f ( x , y ) to be an implicit function of r and s as
f = f [ x (r , s ) , y (r , s ) ] .

Transformation of differential operators


Using the chain rule of differentiation, we have
x y
r r x
= x
r
y
=[J ] x , (9.10)

s s s y y

where

[ J ] =
J11 J12
(9.11)
J 21 J 22

is the Jacobian matrix. Using (9.10) and (9.4), it can also be expressed as
196 FINITE ELEMENT ANALYSIS


r
[ J ] = r x y =
N { x e } N { y
e
} =

s s
(9.12)

r N
=
{ x e } { y e } .

s

N

For the four-node quadrilateral element
x1 y1
(1 s ) (1 s ) (1 + s ) (1 + s ) x 2
[ J ] = 1
y2 , (9.13)
4 (1 r ) (1 + r ) (1 + r ) (1 r ) x3 y3

x4 y4
x1 + x 2 + x3 x 4 + y1 + y2 + y3 y4 +

1 + s (x1 x 2 + x3 x 4 ) + s ( y1 y2 + y3 y4 )
[ J ]= . (9.13, a)
4 x1 x 2 + x3 + x 4 + y1 y2 + y3 + y4 +

+ r (x1 x 2 + x3 x 4 ) + r ( y1 y2 + y3 y4 )

Analogously, the inverse relationship is


r s
x x x r r
= r s
= [ j ] , (9.14)

y y y s s

where [ j ] = [ J ] 1 is the inverse of the Jacobian matrix


1 J 22 J12
[ j ] =
j11 j12
= . (9.15)
j21 j22 det [J ] J 21 J11

For the four-node quadrilateral element

det [ J ] = A0 + A1 r + A2 s , (9.16)
where
1
A0 = [ ( y4 y2 )( x3 x1 ) ( y3 y1 )( x4 x2 ) ] ,
8
9. ISOPARAMETRIC ELEMENTS 197

1
A1 = [ ( y3 y4 )( x2 x1 ) ( y2 y1 )( x3 x4 ) ],
8
1
A2 = [ ( y4 y1 )( x3 x2 ) ( y3 y2 )( x4 x1 ) ] .
8
f f
Usually [ j] is used because we want to express , in terms of
x y
f f
, , determined on the reference element. Because r = r ( x , y ) , s = s ( x , y )
r s
are not explicitly known, we need [ j ].
The above expressions will be used in the derivation of the element
stiffness matrix.

Transformation of an infinitesimal area

An additional result that will be needed is the relation


dx dy = det [ J ] dr ds . (9.17)
It is needed because, for the evaluation of the element stiffness matrix, the
integration on the real element is replaced by the simpler integration over the
reference element.
In cartesian coordinates { x , y }, the elementary area dA is given by the
modulus of the cross product d x d y , where d x = d x i , d y = d y j , and i , j
are base vectors.
In a curvilinear system { r , s } , the elementary area dA is given by the
modulus of the cross product d r d s . It is equal to the area of the elemental
parallelogram enclosed by the two vectors d r and d s directed tangentially to the
r = const . and s = const . contours respectively. The components of these vectors
in a cartesian coordinate system are
x y
dr = d r i + d r j = ( J11 i + J12 j ) d r , d s = ( J 21 i + J 22 j ) d s .
r r
By equating the moduli of the cross products
d x d y = d x d y k ,
198 FINITE ELEMENT ANALYSIS

i j k
d r d s = J11 J12 0 d r d s = det [J ] d r d s k
J 21 J 22 0

we obtain equation (9.17).

9.1.5 Element stiffness matrix

The element stiffness matrix (7.43) is

[k ]= t [ B ]
e e T
[ D ][ B ] d A , (9.18)
A

where [ D ] is the material stiffness matrix and t e is the element thickness.

The matrix of differentiated shape functions [ B ] is (7.41)

[ B ] = [ ][ N ] , (9.19)

where [ ] is the matrix of differential operators (6.9) and [ N ] is the matrix of


shape functions (9.7).
Substituting (6.9) into (7.41) gives the strain-displacement matrix

0
x

[ B ] = [ ] [ N ] = 0 [ N ]. (9.20)
y


y x

Using the transformation (9.14) we obtain

[ B ] = [ B1 ][ B2 ] , (9.21)
where
j11 j12 0 0
[ B1 ] = 0 0 j21 j22 ,
(9.22)
j21 j22 j11 j12
9. ISOPARAMETRIC ELEMENTS 199


r 0

0
and [ B2 ] = s

[ N ], (9.23)
0
r

0 s

or

(1 s ) 0 (1 s ) 0 (1 + s ) 0 (1 + s ) 0
(1 r )
0 (1 + r ) 0 (1 + r ) 0 (1 r ) 0
[ B2 ] = 1 .
4 0 (1 s ) 0 (1 s ) 0 (1 + s ) 0 (1 + s )

0 (1 r ) 0 (1 + r ) 0 (1 + r ) 0 (1 r )

Using equation (9.17), the element stiffness matrix can be written

+1 +1
[k ]= t [ B ]
e e T
[ D ][ B ] det [ J ] dr ds . (9.24)
1 1

Because [ B ] and det [ J ] are involved functions of r and s, the above


integration has to be performed numerically, as shown in the following.

9.1.6 Element load vectors

Because the displacements along a side of the quadrilateral isoparametric


element are linear, the consistent nodal forces applied along that side are calculated
as for a linear two-node element.
If there is a distributed load having components ( px , p y ) per unit length
along side 2-3 in Fig. 9.1, then the equivalent nodal force vector is

{ f }= 12 l 0
e
23 0 px py px py 0 0 T , (9.25)

if p x and p y are constants and l 2 3 is the length of the side 2-3.


200 FINITE ELEMENT ANALYSIS

9.2 Numerical integration

The numerical evaluation of stiffness integrals is usually done by Gaussian


quadrature. It is more efficient than other methods, as for example the Newton-
Cotes integration, because it involves only a half of the sample values of the
integrand required by the latter. For higher order elements, the stiffness integrals
become progressively more complicated. As the order of the displacement field
increases, the differentiated shape functions in the matrix [ B ] grow algebraically
more cumbersome. The isoparametric mapping introduces det [ J ] in the integrand
so that closed form evaluation of stiffness integrals becomes impossible.

9.2.1 One dimensional Gauss quadrature

The one-dimensional version of the Gauss method relies on the concept


that any function f ( r ) can be represented approximately over the interval
r [ 1, 1 ] by a polynomial which can be integrated exactly. A polynomial of
order (2n 1) can be fitted to f ( r ) by imposing n weights wi and n sampling
points in such a way that the summation

+1 n

f (r ) d r = wi f ( r i ) (9.26)
1 i =1

is exact up to the chosen order.


The particular positions of these sampling points are known as Gauss
Points. They turn out to be the roots of Legendre polynomials and so the method is
referred to as a Gauss-Legendre integration.
The actual area represented by the integral is replaced by a series of
rectangles of unequal widths, whose heights are equal to the function values at the
sampling points. In other words, the integral of a polynomial function is replaced
by a linear combination of its values at the integration points ri :
+1

f ( r )d r = w1 f (r1 ) + w2 f (r 2 ) + ... + wi f (ri ) + ... + wn f (r n ) .


1

The 2n coefficients are determined from the condition that the above
equation is satisfied for a polynomial of order 2n 1 of the form

f ( r ) = a1 + a2 r + a3 r 2 + ... + a2 n r 2 n 1 . (9.27)
9. ISOPARAMETRIC ELEMENTS 201

On substitution above we get

+1 +1 +1
a1 d r + a2 r d r + ... + a2 n r
2 n 1
d r = a1 ( w1 + w2 + ... + wn ) +
1 1 1 (9.28)
( ) (
+ a2 w1 r1 + w2 r 2 + ... + wn r n + ... + a2 n w1 r12 n 1 + w2 r 2 2 n 1 + ... + wn r n 2 n 1 . )
Identifying the terms

+1 n +1 n +1 n
d r = wi , r d r = wi r i , r d r = wi ri .
1 i =1 1 i =1 1 i =1

Generally
+1 0 if i is odd


1
r dr =


2
+1
if i is even
(9.29)

so that
2 = w1 + w2 + ... + wn ,
0 = w1 r1 + w2 r 2 + ... + wn r n ,
2
= w1 r12 + w2 r 22 + ... + wn r n2 , (9.30)
3
L L L L L L
0 = w1 r12 n 1 + w2 r 22 n 1 + ... + wn r n2 n 1 .

This is a set of 2 n equations, linear in wi and nonlinear in r i . The 2 n


parameters are determined from conditions
wi > 0
i = 1, 2 ,..., n .
1 < r i < +1

One-point formula

For n = 1 we have the midpoint rule (Fig. 9.3, a)


+1

f ( r )d r w1 f (r1 )
1
202 FINITE ELEMENT ANALYSIS

which is exact only when f (r ) is a polynomial of order 1. However, it is employed


in some selective integration schemes, e.g. to separate shear from extension effects,
avoiding the so-called shear locking by using reduced integration for shearing.

Two-point formula

For a two-point integration


+1

f ( r )d r = w1 f (r1 ) + w2 f (r 2 ) . (9.31)
1

For a polynomial of order 2n 1 = 2 2 1 = 3 , equations (9.30) give


2 = w1 + w2 ,
0 = w1 r1 + w2 r 2 ,
2
= w1 r12 + w2 r 22 ,
3
0 = w1 r13 + w2 r 23 .

The solution to this set of equations is


1
w1 = w2 = 1 , r1 = r 2 = = 0.5773502691
3
so that (Fig. 9.3, b)
+1

f ( r )d r = 1 f ( 0.57735 ) + 1 f ( 0.57735 ) .
1

Fig. 9.3
9. ISOPARAMETRIC ELEMENTS 203

Three-point formula

For a three-point integration (Fig. 9.3, c)

wi f (ri ) = 0.555 f ( 0.774) + 0.888 f ( 0 ) + 0.555 f ( 0.774 ) .


3

i =1

Gauss points and weights are given in Table 9.1 for the first four orders.

Table 9.1
Number of Order of polynomial Gauss Points, Weights,
Gauss points integrated exactly ri wi
1 1 0 2.0
2 3 0.57735 1.0
0 0.8888
3 5
0.774596 0.5555
0.339981 0.652145
4 7 0.347854
0.861136

9.2.2 Two dimensional Gauss quadrature

The one dimensional version of Gaussian quadrature is readily extended to


two dimensions if we use the rectangles or isoparametric quadrilaterals of section
9.1:
+1 +1 +1 n n n
I= f ( r , s )d r d s j
w f r ,(s j ds =) i w j f ri , s j ,
w ( )
1 1 1 j =1 i =1 j =1

+1 +1 n n
I= (
f ( r , s ) d r d s = wi w j f ri , s j . ) (9.32)
1 1 i =1 j =1

For n = 2
w1 = w2 = 1 , r1 = s 1 = 0.57735 , r 2 = s 2 = 0.57735 ,

so that
I = f ( r1 , s1 ) + f ( r1 , s2 ) + f ( r2 , s1 ) + f ( r2 , s2 ) .

Gauss quadrature formulae for quadrilateral elements are shown in Table. 9.2.
204 FINITE ELEMENT ANALYSIS

Table 9.2
Number of
Gauss Points, Weights,
n Gauss points
nn ri , s j wi , w j

1
1 (11) 4 (= 2 2) at centre

4
2 (2 2) 1 (= 1 1) at points 1, 2, 3, 4

25 5 5
= at points 1, 2, 3, 4
81 9 9
9 40 5 8
3 (3 3) = at points 1, 2 , 3, 4
81 9 9
64 8 8
= at points 1, 2, 3, 4
81 9 9

9.2.3 Stiffness integration


The element stiffness matrix (9.24) is

[k ]= t [ B ]
1 1
e e T
[ D ][ B ] det [ J ] dr ds
1 1

which means that the integral above actually consists of the integral of each
element (below or above the main diagonal) in an ( 8 8 ) matrix.

In general, det [ J ] is a linear function of r and s . For a rectangle or


parallelogram it is a constant. The elements of [ B ] are obtained by dividing a bi-
linear function of r and s by a linear function. Therefore, the elements of
9. ISOPARAMETRIC ELEMENTS 205

[ B ]T [ D ][ B ] det [ J ] are bi-quadratic functions divided by a linear function. This


[ ]
means that k e cannot be evaluated exactly using numerical integration.
From practical considerations, it is best to use as few integration points as
is possible without causing numerical difficulties. An alternative is to use reduced
integration at fewer points than necessary, with the aim of decreasing the stiffness
and so compensating for the overstiff finite element model. This is cheaper as well.
A lower limit on the number of integration points can be obtained by
observing that as the mesh is refined, the state of constant strain is reached within
an element. In this case, the stiffness matrix equation (9.24) becomes
+1 +1

[k ] [ B ]
e T
[ D ][ B ] t det [ J ] dr ds .
e
(9.33)
1 1

The integral in (9.33) represents the volume of the element. Therefore, the
minimum number of integration points, is the number required to evaluate exactly
the volume of the element. Taking the thickness, t e , to be constant and noting that
det [ J ] is linear, indicates that the volume can be evaluated exactly using one
integration point.

Fig. 9.4

However, in the present case, one integration point is unacceptable since it


gives rise to zero-energy deformation modes (Fig. 9.4). These are modes of
deformation which give rise to zero strain energy. This will be the case if one of
these modes gives zero strain at the integration point. The existence of these modes
is indicated by the stiffness matrix having more zero eigenvalues than rigid body
modes.
Experience has shown that the best order of integration for the 4-node
quadrilateral element is a ( 2 2 ) array of points.
Using the 2 2 rule (9.33), we can write
206 FINITE ELEMENT ANALYSIS

[ k ] t [ B ( r , s ) ] [ [ D ] det [J ( r , s )] ] [ B ( r , s ) ]+
e e
1 1
T
1 1 1 1

+t e
[ B ( r1 , s2 ) ] [ [ D ] det [J ( r1 , s2 )] ] [ B ( r1 , s2 ) ]+
T

+ t e [ B ( r2 , s 1 ) ]T [ [ D ] det [ J ( r2 , s 1 )] ] [ B ( r2 , s 1 ) ]+ (9.34)
+ t e [ B ( r2 , s2 ) ] T [ [ D ] det [ J ( r2 , s2 )] ] [ B ( r2 , s2 ) ] .

The calculation consists of four basic steps:


a) Calculation of [ J ] at the Gauss points (9.11)

x1 y1

[ J ( ri ,s j ) ] = 14 ( 1 s j ) ( 1 s j ) ( 1 + s j ) ( 1 + s j ) x2

y 2
( 1 ri ) ( 1 + ri ) ( 1 + ri ) ( 1 ri ) y3
x 3
x4 y4

and its determinant.

b) Calculation of the inverse [ j ] = [ J ] 1 at the Gauss points.


c) Calculation of the [ B ] matrix (9.21)

j11 j12 0 0
(
1 s j ) 0 1 s j 0 1+ s j 0 (
1+ s j ) 0

(1 r ) 0 (1 + ri ) 0 1 + ri 0 1 ri 0
[ B ] = 1 0 i .
4
j 21
0
j 22
j 21
j11
j 22
j12


0 1 s j ( ) 0 1 s j 0 1+ s j 0 (
1+ s j )

0 (1 ri ) 0 (1 + ri ) 0 1 + ri 0 1 ri

d) Calculation of [k ] e using the 2 2 rule (9.34).

Example 9.1
Consider the quadrilateral element from Fig. 9.1 with the following nodal
coordinates
x1 y1 10 10
x y2 20 15
2 = .
x3 y3 25 30

x4 y4 8 25

For convenience, consider approximately the Gauss Points at


r i , s j = 0.5 instead of r i , s j = 0.57735 . For the first term in (9.34) we need
det [ J ] and [ B ] calculated at GP1.
9. ISOPARAMETRIC ELEMENTS 207

At GP1, the Jacobian matrix (9.11) is

10 10
0.5 0.5 1.5 1.5 20 15 1
[ J ] = 1 =
61 20
13 60 .
4 0.5 1.5 1.5 0.5 25 30 8

8 25

Its determinant and inverse are

1 60 20
det [ J ]= 53.125 , [ j ] = [ J ] 1 = .
425 13 61

At GP1, the matrix [ B ] is given by (9.21)


0.5 0 0.5 0 1. 5 0 1. 5 0
60 20 0 0 0.5 0 1.5 0 1.5 0 0.5 0
[ B ] = 1 0 0 13 61
0
,
1700 0.5 0 0.5 0 1. 5 0 1. 5
13 61 60 20
0 0. 5 0 1.5 0 1.5 0 0.5

10 0 30 0 30 0 50 0
1
[ B ] = 0 12 0 49 0 36 0 25 .
850
12 10 49 30 36 30 25 50

9.2.4 Stress calculations

In the quadrilateral element, stresses

{ } = [ D ] [ B ] {q e } (9.35)

vary within the element.


In practice, stresses are evaluated at the Gauss points, where they are found
to be accurate (Barlow, 1976). Some programs which use Gauss point stresses
extrapolate to the element nodes and then output the mean value if several elements
meet at a node.
Maximum stresses usually occur at edges of plates or at other
discontinuities at the element boundaries. In order not to miss these peak stresses, a
refined mesh should be used in these regions. However, Gauss point values are
ideal for constructing internal stress contours.
208 FINITE ELEMENT ANALYSIS

9.3 Eight-node quadrilateral

This element is the isoparametric version of the eight-node rectangle


presented in section 8.2.2. The difference is that in cartesian coordinates it has
curved sides (Fig. 9.5, a) and the master element is a square (Fig. 9.5, b).

Fig. 9.5

The displacement functions in simple polynomial form (8.27) are

u = a1 + a2 r + a3 s + a4 r 2 + a5 r s + a6 s 2 + a7 r 2 s + a8 r s 2 ,
(9.36)
v = b1 + b2 r + b3 s + b4 r 2 + b5 r s + b6 s 2 + b7 r 2 s + b8 r s 2 .

The element has three nodes along one edge, so that the displacement
variation should be parabolic (three constants) to satisfy compatibility.
The eight-node quadrilateral employs displacement fields quadratic in r
and s , as are the stresses. The product [ B ] T [ D ][ B ] is therefore fourth order and
det [ J ] is cubic. It would seem necessary to use a 4 4 Gauss quadrature for
exact stiffness integration.
For a single 8-node rectangular element, even if it is supported
conventionally, using a 2 2 integration scheme would result in a singular
stiffness matrix. In practice, large groups of elements will not suffer from such
mechanisms, and 2 2 integration is normally used for this popular element. The
optimal sampling points for this element are the Gauss points for the 2 2 scheme
9. ISOPARAMETRIC ELEMENTS 209

and not those for the higher order 3 3 scheme, so the advantages of reduced
integration are compounded.
For an 8-node quadrilateral element, better displacements are obtained
using a 3 3 Gauss point mesh, due to reduced integration.
It was shown that, in order to evaluate the minimum order of the numerical
integration, it is sufficient to examine the Jacobian determinant. For the 8-node
quadratic element, det [ J ] is of third order hence the minimum number of
integration points is 4 (2 2) .
The eight-node curved quad is ideal for nonlinear problems and can be
easily adapted to model cracks by moving the midside nodes. Care is needed
because the accuracy declines with excessive shape distortion.

9.3.1 Shape functions

The shape functions (8.29) have the following expressions

N1 =
1
4
( 1 r ) ( 1 s ) (1 + r + s ) , N5 =
1
2
( )
1 r 2 (1 s ) ,

1
N 2 = ( 1 + r ) ( 1 s ) (1 r + s ) ,
4
1
(
N6 = (1 + r ) 1 s 2 ,
2
)
1
N3 = ( 1 + r ) ( 1 + s ) ( 1 r s ),
4
1
( )
N 7 = 1 r 2 ( 1 + s ) , (9.37)
2
1
N 4 = ( 1 r ) ( 1 + s ) (1 + r s ) ,
4
1
(
N8 = ( 1 r ) 1 s 2 .
4
)
As they are constructed based on the equations of the lines passing through
the nodal points, they belong to the serendipity *) family of elements.
For ease in programming, one can also use the following equivalent
formulae in which ( ri , si ) are the natural coordinates of node i:
for corner nodes

1
Ni = ( 1 + r ri ) ( 1 + s si ) (r ri + s si 1) ,
4
(9.38)
N i ri
= ( 1 + s si ) (2r ri + s si ) , N i = si ( 1 + r ri ) (2s si + r ri ) ;
r 4 s 4
___________
*) After the extraordinary discoveries of the princes of Serendip from the horror novels by
Horace Walpole (1717-1797).
210 FINITE ELEMENT ANALYSIS

for mid-side nodes, when ri = 0


1
Ni =
2
( )
1 r 2 ( 1 + s si ) ,
(9.39)
Ni
r
= r ( 1 + s si ) ,
Ni 1
s
= si 1 r 2 ;
2
( )
for mid-side nodes, when si = 0
1
2
(
1 s 2 ( 1 + r ri ) ,
Ni = )
(9.40)
Ni 1
r
= ri 1 s 2 ,
2
( Ni
s
)
= s ( 1 + r ri ).

The displacements inside the element can be expressed in terms of the


nodal displacements as
u
=[N ] q ,
v
e
{ } (9.41)

where

[ N ] = 1
N 0 N 2 0 L L N8 0
, (9.42)
0 N1 0 N 2 L L 0 N 8

{ q }= u
e
1 v1 u2 v 2 L L u8 v8 T . (9.43)

Having generated the shape functions [ N ] , the matrix [ B ] = [ ][ N ] is

N1 N2 N8
0 0 L 0
x x x
N1 N2 N8
[ B ] = 0 0 L 0 .

(9.44)
y y y
N N1 N2 N2 N8 N8
1 L
y x y x y x

9.3.2 Shape function derivatives

Use of the isoparametric formulation


8 8
x = N i (r , s ) x i , y = N i (r , s ) y i (9.45)
i =1 i =1

leads to
9. ISOPARAMETRIC ELEMENTS 211

Ni Ni y y Ni
x r r
1 s r
N = [ j ] N = , (9.46)
i i det [ J ] x x Ni

y s s r s

where
8 8
N i (r , s )
x
r
=
i =1
r
xi ,
x
s
= N(sr , s ) x
i =1
i
i ,
(9.47)
8 8
N i (r , s ) N i (r , s )
y
r
=
i =1
r
yi ,
y
s
= i =1
s
yi .

9.3.3 Determinant of the Jacobian matrix

The Jacobian matrix is


x y

[ J ] = xr r
y
, (9.48)

s s

and its determinant is


x y x y
det [ J ] = . (9.49)
r s s r

9.3.4 Element stiffness matrix

The element stiffness matrix is


+1 +1

[k ]= t [ B ]
e e T
[ D ][ B ] det [ J ] d r d s (9.50)
1 1

where t is the thickness of the plate and the material stiffness matrix [ D ] has the
e

form
D1 D 2 0
[ D ] = D 2 D 4 0 . (9.51)
0 0 D 5

212 FINITE ELEMENT ANALYSIS

Substituting (9.44), the stiffness matrix (9.50) can be written


T
N1 N1
0
x
y
0 N1 N1
y x
N1 N2 N8
N2 N2 0 0 L 0
+1 +1 0 x
x y x x


N1 N2 N8
ke = te
N2 N2 [D ] 0 0 L 0 det [ J ] d r d s .
0 y y y
y x
1 1
L
N1 N1 N2 N2
L
N8 N8
L L y x y x y x
N N8
8 0
x y
N8 N8
0
y x
(9.52)
The (16 16) element stiffness matrix contains submatrices [ k ] i j which,
evaluated by numerical integration, have the form
N j
0
Ni Ni x
n n
x 0
y N j
[k ] i j = t e
w p wq [D] 0 det [ J ] , (9.53)
0 Ni Ni y
p =1 q =1
y x N j N j
y x

k k 2
or [ k ] i j = k 1 k 4
. (9.53, a)
3 ij

Denoting

w = t e w p wq det [ J ] , (9.54)

the elements of the matrix (9.53, a) can be determined as


Ni N j Ni N j
k 1 i j = w D 1 + D5 ,
x x y y

Ni N j Ni N j
k2 = w D 2 + D5 , (9.55)
ij
x y y x

Ni N j Ni N j
k3 = w D 2 + D5 ,
ij
y x x y
9. ISOPARAMETRIC ELEMENTS 213

Ni N j Ni N j
k 4 i j = w D 4 + D5 .
y y x x

When i = j , k 2 = k 3 .

A close inspection of equations (9.55) shows that all the terms are of the
form
(cons tan t ) (material property) ( product of shape function derivatives ) .
Therefore, providing that the material properties are constant throughout
the element, it is useful to store separately in an array [ S ] i j the sum of the
products of the shape function derivatives multiplied by w , such that
Ni N j Ni N j
w w
x x x y
[ S ]i j =
N j N j
(9.56)
Ni N
w y x
w yi y

and then to work out equations (9.55) by incorporating the relevant material
constants.
This method (K. A. Gupta & B. Mohraz, 1972) reduces the computing time
by a factor of nine over the full matrix multiplication method.

9.3.5 Stress calculation

The vector of element stresses is

{ } = [ D ][ B ]{ q e }. (9.57)
The explicit form of equation (9.57) for the quadratic 8-node element is
N1 N2 N8 u 1
0 0 L 0
x D1 D2 0 x x x v1
N1 N2 N8 u
y = D 2 D4 0 0 0 0
2
L
y y y
0 0 D 5 N L
xy N1 N2 N2 N8 N8
1 L v
y x y x y x 8
wherefrom
8 8
Ni Ni
x = D1 ui + D2 vi ,
i =1 x i =1 y
214 FINITE ELEMENT ANALYSIS

8 8
Ni Ni
y = D2 ui + D4 vi , (9.58)
i =1 x i =1 y

8 Ni Ni
xy = D 5 ui + v i ,
i =1 y x
In some programs the stresses are determined at nodes, since the nodal
positions are readily located and it is convenient to output the displacements and
stresses at the same points. It has been found that nodal stresses from an 8-node
quadratic element are usually incorrect. However, if the stresses of all elements
meeting at a node are averaged, closer results to the true values are obtained.
A better alternative is to calculate stresses at the Gauss points, in which
superior accuracy is obtained and averaging is not necessary. This is due to the fact
that the element stiffness is calculated by sampling at the Gauss points, and it is
therefore reasonable to expect the most accurate stresses and strains occurring at
the same points.

9.3.6 Consistent nodal forces

Unlike the triangular element, in which all loads can be reduced to nodes
intuitively or by statics, for a quadratic isoparametric element the nodal forces due
to distributed loads must be computed in accordance with equation (7.44)

{ f }= [ N ]
e T
{ p }d V . (9.59)
Ve
The equivalent nodal forces are added, element by element, into the global
load vector which represents the right hand side of the linear set of equations to be
solved for displacements.

Edge pressure

When coding the load vectors in a computer program, the actual pressure
distribution along an element edge is replaced by a parabolic distribution defined
by the pressure values at each of the three nodes along that edge. All intermediate
values can be calculated using the shape functions. It is usual to use all the nodes of
the element in the computation, so that there is no need to sort out the appropriate
shape functions for the three nodes with given pressure values.
Consider a distributed load p , specified in force per unit length, acting
along the s = +1 edge of an element. The components of the force p d r acting
upon an elemental length d r are
9. ISOPARAMETRIC ELEMENTS 215

y
px r
= p x dr . (9.60)
py
r

The consistent load for p is given by a modified form of equation (9.59) in


which the volume integral has been reduced to a line integral

+1
y
r
{ f }=
e
p [ N ]T
x dr (9.61)
1

r
where [ N ] is given by (9.42).
The above vector is usually integrated numerically and is written as
y
n
r
{ }
fe = wi [ N ] T (9.62)
pi
x
i =1
i r
where pi is the pressure at the Gauss point i along the s = +1 edge and is
computed by equation
8
p= N (r ,s ) p
k =1
k k . (9.63)

If the same pressure acts on the edge s = 1 , the sign of the force vector
{ f } is reversed, since a positive pressure is assumed to act towards the centre of
e

the element. A more general equation applicable to the two edges s = 1 is


y
r
{ }
n 8
f e = wi ws [N ]T N k p k (9.64)
x
i =1 k =1
r i
where w s takes up the value of the s coordinate for the loaded edge.
A similar expression for pressure loads on the r = 1 edges is
y
s
{ f }=
n 8
e
wi wr [N ] N k p k
T
. (9.65)
x
i =1 k =1
s i
216 FINITE ELEMENT ANALYSIS

For a rectangular isoparametric element with edges parallel to the x , y


axes, it is possible to integrate the equations for the equivalent nodal forces
explicitly.

Example 9.2
Consider the element of Fig. E9.2, a with a uniform pressure load along the
top edge. Find the equivalent nodal forces.

Fig. E9.2

Solution.
For the edge s = +1 we have N1 = N 5 = N 2 = N 6 = N8 = 0 .

y x
For the rectangular element = 0 and = a . Equation (9.61) gives
r r

f y3 +1 N3
{f }e
= f y7


= p

N7

( a ) d r . (9.66)
f 1 N
y4 4

Denoting the total pressure load P = 2 p a , we can write


+1
P 1
f y3 =
2 4 ( 1 + r ri ) ( 1 + s si ) ( r ri + s si 1) d r ,
1
+1

4 ( 1 + s si ) ( 1 r ) d r ,
P 1 2
f y7 =
2
1
9. ISOPARAMETRIC ELEMENTS 217

+1
P 1
f y4 = (1 + r ri ) (1 + s si ) ( r ri + s si 1) d r .
2 4
1
In the above expressions, substituting r i = 1 and s i = 1 for node 3, r i = 0
and s i = 1 for node 7, and r i = 1 and s i = 1 for node 4, yields
+1
P 1
f y3 = (1 + r )( 2)( r + 1 1) d r = P ,
2 4 6
1
+1
f y7 =
P
2
1
4
( )
( 2 ) 1 r 2 dr = 2P ,
3
1
+1
P 1
f y4 = (1 r )(2)( r + 1 1) d r = P .
2 4 6
1
Hence a uniform load acting along an edge is not distributed in the
intuitive ratio of 1 4 : 1 2 : 1 4 , but in the ratio 1 6 : 2 3 : 1 6 (Fig. E9.2, b).

The same result could have been obtained noticing that the integrands in
the above equations are the shape functions of the three-node isoparametric one-
dimensional element (4.26).

Gravity loading

For gravity loading (downwards negative) the vector of equivalent nodal


forces (9.59) is
+1 +1
{ f }= m [ N ]
e T 0
d x d y = m [N ]
T 0
det [ J ] d r d s (9.67)
g 1 1
g
where m is the mass per unit area and g the acceleration in the y direction.
The integration is carried out numerically and equation (9.67) takes the
form

{ f }= w w
n n
0
e
i j m [ N ] T det [ J ] . (9.68)
i =1 j =1 g

Example 9.3
Consider the rectangular element of Fig. E9.3, a with gravity loading. Find
the equivalent nodal forces.
Solution. Because all the equivalent nodal forces act in the y direction,
218 FINITE ELEMENT ANALYSIS

f y1 N1
f y 2
+1 +1 N

= m g 2 det [ J ] d r d s . (9.69)
L L
1 1
f y8 N 8

For a rectangular element det [ J ] = ab .
For a corner node i the integral becomes
+1 +1
1
f yi = m g 4 (1 + r ri ) (1 + s si ) ( r ri + s si 1) det [ J ] d r d s . (9.70)
1 1

Denoting the total weight of the element W = m g 2a 2b , the force (9.70)


can be written
W
f yi = ( I1 + I 2 + I 3 ) , (9.71)
16
where
+1 +1
4
I1 = (1 + r ri ) (1 + s si ) r ri d r d s = 3 (9.72)
1 1
since ri = 1 for all corner nodes,
+1 +1
4
I2 = (1 + r ri ) (1 + s si ) s si d r d s = 3 (9.73)
1 1
since si = 1 for all corner nodes, and
+1 +1
I3 = (1 + r ri ) (1 + s si ) d r d s = 4 . (9.74)
1 1
Therefore, the equivalent nodal forces for gravity loading at a corner node i
are given by
f yi =
W
( I1 + I 2 + I 3 ) = W 4 + 4 4 = W . (9.75)
16 16 3 3 12
For a midside node i with ri = 0 , the integral is
+1 +1

m g 2 (1 r ) (1 + s si ) det [ J ] d r d s = 3 .
1 2 W
f yi = (9.76)
1 1
For a midside node i with si = 0 , the integral is
1 1

m g 2 (1 s ) (1 + r ri ) det [ J ] d r d s = 3 .
1 2 W
f yi = (9.77)
1 1
9. ISOPARAMETRIC ELEMENTS 219

Fig. E9.3

Hence for gravity loading the equivalent nodal forces are as shown in Fig.
E9.3, b. Again, they are different from the values of W 12 and W 6 , which
would have been assigned intuitively to the corner and midside nodes respectively.

9.4 Nine-node quadrilateral


This element belongs to the Lagrange family of elements. Apart from the
eight nodes located on the boundary, it contains an internal node. The local node
numbers for this element are shown in Fig. 9.6, a. The master element is presented
in Fig. 9.6, b.

Fig. 9.6
220 FINITE ELEMENT ANALYSIS

The associated displacement functions in polynomial form are

u = a1 + a2 r + a3 s + a4 r 2 + a5 r s + a6 s 2 + a7 r 2 s + a8 r s 2 + a9 r 2 s 2 ,
(9.78)
v = b1 + b2 r + b3 s + b4 r 2 + b5 r s + b6 s 2 + b7 r 2 s + b8 r s 2 + b9 r 2 s 2 .
Along the sides of the element, the polynomial is quadratic (with three
terms - as can be seen setting s = 0 in u ), and is determined by its values at the
three nodes on that side.
Higher-order rectangular elements can be systematically developed with
the help of the so-called Pascals triangle, which contains the terms of polynomials
of various degrees in the two variables r and s , as shown in Fig. 9.7.

Fig. 9.7

Since a linear quadrilateral element has four nodes, the polynomial should
have the first four terms 1, r, s, and rs . In general, a pth-order Lagrange
rectangular element has ( p + 1) 2 nodes ( p = 0,1, ...) . The quadratic quadrilateral
element has 9 nodes. The polynomial is incomplete. It contains the complete
polynomial of the second degree (6 terms) plus other three terms which have to be
located symmetrically: the third degree terms r 2 s and r s 2 and also an r 2 s 2 term.
The shape functions are defined as follows. Considering the r axis alone,
as shown at the bottom of Fig. 9.6, b, we can define generic shape functions L1 ,
L2 and L3 having unit value at the node with the same index and zero at the other
two nodes

r (1 r ) r (1 + r )
L1 (r ) = , L 2 (r ) = ( 1 + r )( 1 r ) , L 3 (r ) = . (9.79)
2 2

They turn out to be Lagrange polynomials, which, for three points, have
the expressions
9. ISOPARAMETRIC ELEMENTS 221

L 1 (r ) =
( r r 2 ) ( r r3 ) , L 2 (r ) =
( r r1 ) ( r r 3 ) , L 3 (r ) =
( r r1 ) ( r r 2 ) .
( r1 r 2 ) ( r1 r3 ) ( r 2 r1 ) ( r 2 r 3 ) ( r 3 r1 ) ( r 3 r 2 )
Similarly, generic shape functions can be defined along the s axis, as
shown at the right of Fig. 9.6, b

s (1 s ) s (1 + s )
L1 (s ) = , L 2 (s ) = ( 1 + s ) ( 1 s ) , L 3 (s ) = . (9.80)
2 2

The shape functions N i can be constructed as products of the above one-


dimensional functions

N1 N8 N 4 L 1 (r )
N
5 N9 N 7 = L 2 (r ) L 1 (s ) L 2 (s ) L 3 (s ) . (9.81)
N 2 N6 N 3 L 3 (r )

It should be cautioned that the subscripts of N i refer to the node


numbering used in Fig. 9.6, a.
Since the internal nodes of the higher-order elements of the Lagrange
family do not contribute to the inter-element connectivity, they can be condensed
out at the element level to reduce the size of the element matrices. Alternatively,
one can use the serendipity elements, but their shape functions cannot be obtained
using products of one-dimensional interpolation functions.

9.5 Six-node triangle

Higher-order triangular elements can be developed using Pascals triangle


(Fig. 9.8). One can view the position of the terms as the nodes of the triangle, with
the constant term and the first and last terms of a given row being the vertices of
the triangle. The six node triangle (Fig. 9.9, a) is an element of order 2 (i.e., the
degree of the polynomial is 2), as can be seen from the top three rows of Pascals
triangle. The polynomial involves six constants, which can be expressed in terms of
the nodal values of the variable being interpolated.
By referring to the master element in Fig. 9.9, b, the shape functions can be
expressed in terms of area coordinates as for the linear strain triangle (8.36)
N1 = 1 ( 2 1 1 ) , N 2 = 2 ( 2 2 1 ) , N 3 = 3 ( 2 3 1 ) ,
(9.82)
N 4 = 4 1 2 , N5 = 4 2 3 , N 6 = 4 3 1 ,
222 FINITE ELEMENT ANALYSIS

where 1 = 1 2 3 . Because of terms 22 , 32 in the shape functions, this


element is also called a quadratic triangle.

Fig. 9.8

The isoparametric representation is (8.37)


u N1
=
0 N2 0 L L N6 0 e
N 6
{ }
q , (9.83)
v 0 N1 0 N2 L L 0

where { q }= u
e
1 v1 u2 v 2 L L u6 v6 T , and
6 6
x = Ni x i , y = Ni y i . (9.84)
i =1 i =1

Fig. 9.9
9. ISOPARAMETRIC ELEMENTS 223

The element stiffness matrix, which has to be integrated numerically, can


be evaluated using (9.24) if the (r , s ) coordinates are defined by (8.46). This gives


= , = . (9.85)
r N 2 N1 s N 3 N1

Details on numerical integration schemes for triangles are given in [18,


39]. The Gauss quadrature formulas for a triangle differ from those considered
earlier for the rectangle.

9.6 Jacobian positiveness

In the higher-order isoparametric elements discussed above, we note the


presence of midside nodes. The midside node should be as near as possible to the
centre of the side. It must be placed inside the middle third of a side. This condition
ensures that det [ J ] does not attain a value of zero in the element.

The sign of det [ J ] should be checked. If the Jacobian becomes negative


at any location, a warning message will be signaled indicating the nonuniqueness
of mapping. Note that while the Jacobian is always computed at Gauss points, it is
more likely to be negative at corners, where stresses may be computed.
A necessary requirement for applying equation (9.13) to shape functions is
that [ J ] can be inverted. This inverse exists if there is no excessive distortion of
the element such that lines of constant r or s intersect inside or on the element
boundaries or there are re-entrant angles. When the element is degenerated into a
triangle by increasing an internal angle to 1800 then [ J ] is singular at that corner.
A similar situation occurs when two adjacent corner nodes are made coincident to
produce a triangular element. Therefore to ensure that [ J ] can be inverted, any
internal angle of each corner node of the element should be less than 1800 , and, as
an internal angle approaches 1800 there is a loss of accuracy in the element stress,
particularly at that corner.
If the determinant det [ J ] 0 , then [ j ] and the operators in [ ] will
increase without limit and consequently produce infinite strains.
The above statements are illustrated for the one-dimensional quadratic
isoparametric element shown in Fig. 4.4. The row vector of shape functions is
1 1
N = N1 N 2 N 3 = 2 r ( r 1) ( 1 + r )( 1 r ) 2 r ( r + 1) .

224 FINITE ELEMENT ANALYSIS

The Jacobian matrix is


x1 0
[ J ] = J11 = N1 N 2 N 3
x2
2r 1
= 2r
2r + 1
x2

.
r r r 2 2
x3 l
The determinant of the Jacobian matrix is
det [ J ] = J11 = ( r + 0.5 ) l 2 r x2 .
For any x2 this determinant vanishes at the point

0.5 l
r0 = .
2x2 l

Fig. 9.10

This point lies within the element (Fig. 9.10) if 1 r 0 1 , i.e. if

0 .5 l
1 1,
2x2 l
where 0 x 2 l .

3l
< x 2 < , then det [ J ] does not vanish on the element. This explains
l
If
4 4
why it is recommended to place precautiously the midside node inside the middle
third of a side.
10.
PLATE BENDING

Flat plate structures, such as the floors of buildings and aircraft, enclosures
surrounding machinery and bridge decks are subject to loads normal to their plane.
Such structures can be analyzed by dividing the plate into an assemblage of two-
dimensional finite elements called plate bending elements. These elements may be
either triangular, rectangular or quadrilateral in shape. In this chapter, finite
element displacement models for the flat plate bending problem are discussed.
Thin plates with transverse shear neglected are analyzed based on
Kirchhoffs classical theory. Plates with a constant shear deformation through the
plate thickness are treated in the Reissner-Mindlin plate theory. For thick or
composite plates some higher-order shear deformation plate theories are available,
in which on the faces of plate the shear strains are equal to zero.

10.1 Thin plate theory (Kirchhoff)

A plate is described as a structure in which the thickness is very small


compared with the other dimensions, that is the thickness-to-span is h l 0.1 . For
this case, it can be assumed that the plate deformation may be expressed by the
deformation state at the middle surface, which is the plane midway between the
faces of the plate.
For thin plates, Kirchhoffs hypotheses are adopted: a) there is no
deformation (stretching) in the middle plane of the plate; b) normals to the middle
plane of the undeformed plate remain straight and normal to the middle surface of
the plate during deformation; and c) the direct stress in the transverse direction can
be disregarded.
The x-y plane is taken to coincide with the middle surface of the plate (Fig.
10.1) and the positive z-direction is upwards. The plate has constant thickness h
and is subject to distributed surface loads.
226 FINITE ELEMENT ANALYSIS

The displacements parallel to the undeformed middle surface are given by


w w
u ( x, y, z ) = z , v ( x, y, z ) = z , (10.1)
x y
where w ( x, y ) denotes the displacement of the middle surface in the z-direction
(Fig. 10.2, a).

Fig. 10.1

The components of the strain are given as follows:


u 2w v 2w u v 2w
x = = z 2 , y = = z 2 , x y = + = 2 z ,
x x y y y x x y

v w w u
yz = + =0, zx = + = 0. (10.2)
z y x z

a b
Fig. 10.2

The strain vector can be written in the form

{ } = x y x y T = z { }, (10.3)
where the curvature vector
10. PLATE BENDING 227

T
2 2w 2w
{ } = w2 2 . (10.4)
x y2 x y

Since z = 0 , the stress-strain relations take the form

{ } = [ D ] { } = z [ D ] { } , (10.5)
where

{ } = x y x y T , (10.6)
and
d11 d12 d16
[ D ] = d12 d 22 d 26 .
(10.7)
d16 d 26 d 66

The strain energy in the plate can be written as


h3

1
U= { }T { }dV = 1 { }T [ D ] { }d A . (10.8)
2 2 12
V A

Fig. 10.3

The positive sense of internal bending and twisting moments (per unit
length) is shown in Fig. 10.3. Their definig equations are
h2

M x My Mxy =
z
h 2
x y xy dz . (10.9)

Denoting
h2

Di j =
d
h 2
ij z2 d z (10.10)
228 FINITE ELEMENT ANALYSIS

the moment - curvature equations can be expressed as follows


2w

x2
Mx D11 D12 D16 2
w
My = D12 D22 D26 (10.11)
y2
M
xy
D
16 D26 D66 2
w

x y
or
{ M } = [ D ]{ } . (10.12)

For isotropic materials



1 0
[ D ] = D 1 0 . (10.13)
1
0 0
2
where
E h3
D= , (10.14)
(
12 1 2
)
E is Youngs modulus and is the Poissons ratio.

The equilibrium equations are


M x M y x
+ Qx = 0 ,
x y
M x y M y
+ Qy = 0 , (10.15)
x y
Qx Q y
+ + pz = 0 ,
x y

where p z is the lateral distributed load (per unit surface) and Qx , Q y are the shear
forces (per unit length).
From (10.11) and (10.15) it can be shown that

2 w 2 w
, 2 w 2 w
.
Qx = D + Qy = D + (10.16)
x x 2 y 2
y x 2 y 2

10. PLATE BENDING 229

By substituting (10.16) into the last equation (10.15), the differential


equilibrium equation for isotropic materials can be obtained as
4w 4w 4 w pz
+ 2 + = . (10.17)
x4 x2 y2 y4 D

The bending and torsion stresses are given by (10.9)


12 z
x y xy =
h3
M x My Mxy (10.18)

and the shear stresses by


3 2
z x yz = 1 4z Qx Qy , (10.19)
2h h2

where a parabolic distribution is assumed.
The strain energy (10.8) can be expressed as
2 2 2 2
w + w + 2 ( 1
2
w 2w 2w
)
D
U= 2 dx dy .
2 x2 y 2 x y x y 2
A
(10.20)
The potential of the external load is


W P = p z w dx dy .
A
(10.21)

10.2 Thick plate theory (Reissner-Mindlin)

For moderately thick plates, the thickness-to-span ratio is not small enough
to neglect transverse shear deformations and Kirchhoffs assumption is no longer
valid. To overcome this problem, the classical hypothesis of zero transverse shear
strain is relaxed. First, Reissner proposed to introduce the rotations of the normal to
the plate midsurface in the xOz and yOz plane as independent variables. Then,
Mindlin simplified Reissners assumption considering that normals to the plate
midsurface before deformation remain straight but not necessarily normal to the
plate after deformation. The displacements of the middle surface are independent
of the rotations of the normal (Fig. 10.2, b). The transverse normal stress is
disregarded as in the Kirchhoffs theory.
According to Reissner-Mindlin assumptions, the displacements parallel to
the middle surface can be expressed as
230 FINITE ELEMENT ANALYSIS

u (x , y , z ) = z y (x , y ) , v (x , y , z ) = z x (x , y ) , (10.22)

where x , y are the rotations about the Ox and Oy axes of lines originally normal
to the middle plane before deformation.
The in-plane strains are now given by

{ } = z { }, (10.23)
where the curvature vector is
T
x x y
{ }= y . (10.24)
x y x y

The transverse shear strains are


v w w u
yz = + , zx = + . (10.25)
z y x z

Substituting for the displacements from (10.22) gives


w
+
y z x y
{ } = = w . (10.26)
zx y +
x
w
Note that when y z = 0 , z x = 0 , equation (10.26) gives x = ,
y
w
y = , and equation (10.24) reduces to (10.4).
x
The average shear stresses are given by

{ } =
yz
= D
S
[ ]{ } . (10.27)
zx
where is a shear correction factor and

[ D ] = G0
S 0 E 1 0
= 2 (1 + ) 0 1 . (10.28)
G
The strain energy in the plate can be written as the sum of the energies due
to bending and shear deformation

{ } { }dV + 2
1 1
U= T
{ }T { }dV . (10.29)
2
V V
10. PLATE BENDING 231

or, integrating through the thickness

h { } [ D ]{ }d A .
h3

1
U= { }T [ D ] { }d A + 1 T S
(10.30)
2 12 2
A A

For isotropic materials, equation (10.11) becomes


y

Mx x

1 0 x


My
=D 1 0 (10.31)
M 1 y
xy 0 0 y
2
x y
x


or (10.12), where { } is given by (10.24).
The transverse shear constitutive equation is
w
x +
Qy
y
=G w
. (10.32)
Qx y +
x

The strain energy (10.30) can be expressed as

y

1
U= M x + M x y x y + M y x dx dy +
2 x x y y
A
w w

1
+ Qx y + + Q y x + dx dy .
2 x y
A

w w
If x = and y = , the above expression reduces to (10.20).
y x
The main drawback of this theory is the arbitrary averaging of the shear
strains. If a constant shear strain is considered through the plate thickness, on the
faces of plate the shear strains are not zero, which is not true. This assumption is
equivalent to the introduction of parasitic shear stresses that force the normal to
remain a straight line. As the thickness of the plate becomes extremely thin, the
shear strain energy predicted by the finite element analysis can be magnified
unreasonably and a shear locking occurs for large span-to-thickness ratios.
232 FINITE ELEMENT ANALYSIS

10.3 Rectangular plate bending elements

For a thin plate bending element, the strain energy is given by (10.20) and
the potential of the external load by (10.21). The highest derivative appearing in
these expressions is the second. Hence, for convergence, it will be necessary to
w w
ensure that w and its first derivatives and are continuous between
x y
elements. These three quantities are, therefore, taken as degrees of freedom at each
node. Also, complete polynomials of at least degree two should be used, according
to the convergence criteria of the Rayleigh-Ritz method. The assumed form of the
displacement function, irrespective of the element shape, is

w = a1 + a2 x + a3 y + a4 x 2 + a5 x y + a6 y 2 + higher degree terms. (10.33)


Figure 10.4 shows a thin rectangular element of thickness h, having four node
points, one at each corner.

Fig. 10.4

The dimensionless coordinates = x a and = y b will be used in the


following with the origin at the plate centre.

10.3.1 ACM element (non-conforming)

Figure 10.5 shows the ACM element (Adini, Clough, Melosh - 1961,
1963). There are three degrees of freedom at each node, namely, the transverse
w w
displacement w , and the two rotations x = and y = . In terms of the
y x
dimensionless coordinates and , these become
10. PLATE BENDING 233

1 w 1 w
x = , y = . (10.34)
b a
Since the element has 12 degrees of freedom, the displacement function
can be represented by a polynomial having twelve terms, that is

w = 1 + 2 + 3 + 4 2 + 5 + 6 2 +
(10.35)
+ 7 3 + 8 2 + 9 2 + 10 3 + 11 3 + 12 3 .

Note that this function is a complete cubic to which two quartic terms 3
and 3 have been added, which are symetrically located in Pascals triangle. This
ensures that the element is geometrically invariant.

Fig. 10.5

The expression (10.35) can be written in the following matrix form


w = 1 2 2 3 2 2 3 3 3 { }, (10.36)

w = P ( , ) { } , (10.36, a)
where

{ }T = 1 2 3 L 12 . (10.37)

Differentiating (10.36) gives


w


= 0 1 0 2 0 3 2 2 2 0 3 2 3 { } , (10.38)


= 0 0 1 0 2 0 2 2 3 2 3 3 2 { }. (10.39)
234 FINITE ELEMENT ANALYSIS

Evaluating (10.36), (10.38) and (10.39) at = 1 , = 1 gives

{ q }= [ A ]{ },
e e
(10.40)
where

{ q } = w b a L w b a { }, (10.41)
e T
1 x1 y1 4 x4 y4

and the matrix [ A ] is given in [87] but not reproduced here, for conciseness.
e

Solving (10.40) for { } gives

{ } = [ Ae ] {q } .
1 e
(10.42)

Substituting (10.42) into (10.36, a) yields

{ }
w = N ( , ) q e = N1 N 2 N 3 N 4 {q },
e
(10.43)

where the vector of nodal displacements is

{ q } = w
e T
1 x1 y1 L w4 x 4 y 4 , (10.44)

and the shape functions are defined as


1
( )(
8 ( 1 + i ) 1 + j 2 + i + i
2 2
)

N i ( , )
T
=
b
8
( 1 + i )( i + ) 2 1 ( ) , (10.45)



a
8
(
( i + ) 2 1 ( 1 + i ) )

where ( i ,i ) are the coordinates of node i.


The element is non-conforming. Indeed, evaluating (10.45) on the side 2-3
(for = +1 ) gives
0

(
(1 4 )( 1 ) 2 2 )
N 1
T
= 0 , N 2
T
= (b 4 )( 1 + ) 2 1 ( ) ,
0 0

(
(1 4 )( 1 + ) 2 + 2 ) 0
N 3
T
= (b 4 )( 1 + ) 2 1 ( ) , N 4
T
= 0 .
0 0

10. PLATE BENDING 235

This indicates that the displacement w , and hence the rotation x , are
uniquely determined by their values at nodes 2 and 3. If the element is attached to
another rectangular element at nodes 2 and 3, then w and x will be continuous
along the common side. Unfortunately, this is not the case with the rotation y .

The rotation y is given by (10.34)

y =
1 w
a
1 N
= 1 L
a
N 4 e

q . { } (10.46)

Substituting (10.45) into (10.46) gives



i ( i )
N i ( , ) T b

= (
i ( i + ) 2 1 . ) (10.47)
8
a ( 2 + 2 ) ( 1 + )
8 i i

Evaluating (10.47) along = +1 gives

( 1 ) ( 1 )
N1 T

(
= (b 8) ( 1 + ) 2 1 ) N 2 T
,

(
= (b 8) ( 1 + ) 2 1 )
,
0 (a 2 ) ( 1 )

( 1 ) ( 1 )
N 3 T

(
= (b 8) ( 1 + ) 2 1 )
,
N 4 T

(
= (b 8) ( 1 + ) 2 1 )
.
(a 2 ) ( 1 + ) 0

The above expressions indicate that y is determined by the values of w
and x at nodes 1, 2, 3, and 4 as well as by y at nodes 2 and 3. The rotation y is
not continuous across the side 2-3 and the element is non-conforming.
Substituting (10.43) into (10.4) and (10.8) gives

Ue =
1
2
{ q } [k ] { q },
e T e e
(10.48)

where the element stiffness matrix is

[ k ] = 12h [ B ]
3
e T
[ D ] [ B ]dA (10.49)
A
and the strain-displacement matrix is
236 FINITE ELEMENT ANALYSIS

2 1 2
2 2 2
x a
2 1 2
[ B ] = 2 2
N = 2 N ( , ) . (10.50)
y b
2 2 2
2
x y ab

Substituting the shape functions from (10.45) and integrating gives the
element stiffness matrix

[ k11 ] [ k12 ] [ k13 ] [ k14 ]


[ k22 ] [ k23 ] [ k24 ]
[k ]
3
Eh
e
= (10.51)
(
48 1 a b
2

) [ k33 ] [ k34 ]
SYM [ k44 ]
where the submatrices have the following expressions

( 2
) 2 1
4 + + 5 (7 2 ) 2 + 5 (1 + 4 ) b
2 2

2 1
2 (1 + 4 ) a
5

4 2 4 2 ,
[ k11 ] = + (1 ) b a b
3 15

4 4 2
+ ( 1 ) a
2
SYM
3 15


( 2
)
2 2 + 5 (7 2 )
2 2 2 1
(1 + 4 ) b
5
1
2 2 + (1 ) a
5

2 1 2 2 4 2 ,
[ k12 ] = (1 + 4 ) b (1 ) b 0
5 3 15

1 2 2 1 2
2 + (1 ) a (1 ) a
2
0
5 3 15


(2 2 2
)
2 + + 5 (7 2 )
2 1
( 1 ) b
5
2 1
+ ( 1 ) a
5

2 1 1 2 1 2 ,
[ k13 ] = + ( 1 ) b + ( 1 ) b 0
5 3 15

1 1 2 1 2
( 1 ) a + ( 1 ) a
2
0
5 3 15
10. PLATE BENDING 237


( 2
) 2 1
2 2 5 (7 2 ) 2 + 5 (1 ) b
2 2

2 1
+ (1 + 4 ) a
5

2 1 2 2 1 2
[ k 41 ] = 2 (1 ) b (1 ) b 0
5 3 15

1
2 + (1 + 4 ) a 2 2 4 2
0 (1 ) a
5 3 15
and
a b
= , = .
b a
The remaining submatrices of (10.51) are

[ k22 ] = [ I 3 ]T [ k11 ][ I 3 ] , [ k23 ] = [ I 3 ]T [ k14 ][ I 3 ] , [ k24 ] = [ I 3 ]T [ k23 ][ I 3 ] ,


[ k33 ] = [ I1 ]T [ k11 ][ I1 ] , [ k34 ] = [ I1 ]T [ k12 ][ I1 ] ,
[ k44 ] = [ I 2 ]T [ k11 ][ I 2 ] ,
(10.52)
where
1 0 0 1 0 0 1 0 0
[ I1 ] = 0 1 0 , [ I 2 ] = 0 1 0 , [ I 3 ] = 0 1 0 . (10.53)
0 0 1 0 0 1 0 0 1

The above relationships are presented in reference [87].


In deriving this result, it is simpler to use the expression (10.36) for w and
substitute for { } after performing the integration. A typical integration is then of
the form
+1 +1 0 m or n odd


1 1
m n d d = 4
(m + 1)(n + 1)

m and n even

For p z = constant , substituting the shape functions into

{ f } = N
e T
pz d A (10.54)
A
and integrating, the vector of equivalent nodal forces is obtained as

{ f }= p
e ab
3
3 b a 3 b a 3 b a 3 b a .
z
T

Stresses at any point in the plate are given by (10.5). In terms of the nodal
displacements they can be expressed as
238 FINITE ELEMENT ANALYSIS

{ } = z [ D ][ B ]{q e }, (10.55)

where [ B ] is defined in (10.50). The most accurate values are at the Gauss points
of a (2 2) numerical integration scheme.

10.3.2 BFS element (conforming)

A conforming rectangular element, of the form shown in Fig. 10.4,


commonly referred to as the BFS element (Bogner, Fox and Schmit - 1966), can be
obtained using products of separate one-dimensional Hermitian shape functions
(5.21) as for uniform slender beams. It is a four-node thin plate bending element.
The nodal expansion of the displacement function has the form (10.43)
with
f i ( ) f i ( )

N i ( , ) = b f i ( ) g i ( ) ,
T
(10.56)
a g ( ) f ( )
i i
where

f i ( ) =
1
4
( ) 1
(
2 + 3 i i 3 , g i ( ) = i + i 2 + 3 ,
4
)
(10.57)
1
( ) 1
(
f i ( ) = 2 + 3i i 3 , g i ( ) = i +i 2 + 3 ,
4 4
)
in which ( i ,i ) are the coordinates of node i.
Unfortunately, when we examine the derivatives of these products we find
that the twist 2 w is zero at all four corners and that there is no constant
component to this second derivative. As this controls the shear strain in equation
(10.2), this violates the fundamental requirement that the element can represent all
constant strain states. In the limit, as an increasing number of elements is used, the
plate will tend towards a zero twist condition.
A compromise is to use the cubic Hermitian polynomials for the deflection
shape functions but reduce the order to linear for the rotation shape functions. This
element does have constant strain but is non-conforming, and discontinuities in
slope occur at interfaces. However, as the mesh is refined and the element size is
decreased, the results are convergent.

The solution used for the BFS element is the introduction of 2 w as


an additional degree of freedom at each node. In this case, the displacement
function is of the form (10.43) but with 16 terms
10. PLATE BENDING 239

{ q } = w
e T
1 x1 y1 wx y1 L w4 x 4 y 4 wx y 4 , (10.58)

where wx y 2 w x y and
f i ( ) f i ( )
b f ( ) g ( )

N i ( , ) =
T i i
. (10.59)
a g i ( ) f i ( )
ab g i ( ) g i ( )

It can be shown that this element can perform rigid body movements
without deformation and can describe pure bending behaviour in the x- and y-
directions. This is ensured by the presence in the functions (10.59) of the first six
terms in (10.33).
The element stiffness matrix and consistent vector of nodal forces are
given by (10.49) and (10.53) where the matrix N is defined by (10.43) and
(10.59).
Although the BFS element is more accurate than the ACM element, it is
difficult to use in conjunction with other types of elements in built-up structures
due to the presence of the degree of freedom 2 w x y . This is overcome in the
WB element (Wilson, Brebbia 1971) by introducing the approximations

wx y1 =
1
2b
(
y1 y 4 , ) wx y 2 =
1
2a
( )
x 2 x1 ,

wx y 3 =
1
2b
(
y 2 y3 , ) wx y 4 =
1
2a
( )
x3 x 4 .

Applying the above constraints to the BFS element makes it a non-


conforming one. The transverse displacement and tangential slope are continuous
between elements but the normal slope is not.

10.3.3 HTK thick rectangular element

When the thickness is greater than about a tenth of the plate width, the
shear deformations become important and a Reissner-Mindlin plate model is
adopted. For a thick plate-bending element, the strain energy expression is (10.30)
and the potential of the external load is (10.21), with { } and { } given by
(10.24) and (10.26). The highest derivative of w , x and y appearing in these
expressions is the first. Hence, for convergence, w , x and y are the only
degrees of freedom required at the nodal points (Fig. 10.5).
240 FINITE ELEMENT ANALYSIS

The four-node HTK element (Hughes, Taylor and Kanoknukulcha - 1977)


expands separately w , x and y in terms of their nodal values. It represents the
shear deformation directly, without having to infer it as the derivative of the
bending moment.
The displacement functions are of the form
4 4 4
w = Ni w i , x = Ni x i , y = Ni y i , (10.60)
i =1 i =1 i =1

where the functions N i are defined by (9.3), that is

Ni =
1
4
(
1 + i 1 + i . )( ) (10.61)

These functions ensure that w , x and y are continuous between


elements.
In matrix form, expressions (10.60) can be written
w

x

=[N] q ,
e
{ } (10.62)

y
where

{ q } = w
e T
1 x1 y1 L w4 x 4 y 4 , (10.44)

and
N1 0 0 L N4 0 0
[ N ] = 0 N1 0 L 0 N4 0 . (10.63)
0 0 N1 L 0 0 N 4

Substituting (10.62) into (10.24), (10.26) and (10.30) gives

Ue =
1
2
{ q } [ k ] { q },
e T e e
(10.64)

[ ]
where the element stiffness matrix k e can be written as the sum of the matrices
due to bending and shear

[ k ] = [ k ] + [ k ].
e B S
(10.65)
In (10.65)
10. PLATE BENDING 241

[ k ] = 12h [ B ] [ D ] [BB ] d A
3 T
B B
(10.66)
A
and

[ k ] = h [ B ] [ D ][ B ] d A .
S S T S S
(10.67)
A

The strain-displacement matrix [ B ] is of the form B

[ B ]= [ [ B ] [ B ] [ B ] [ B ] ]
B
1
B B
2
B
3
B
4 (10.68)
where
0 0 Ni x
[ ]
BiB
= 0 Ni y 0 .
(10.69)
0 N i x N i y

[ ]
The strain-displacement matrix B S is of the form

[ B ]= [ [ B ] [ B ] [ B ] [ B ] ]
S S
1
S
2
S
3
S
4 (10.70)
where

[ B ] = NN
i
S i x 0
y Ni
Ni
0
. (10.71)
i

Substituting (10.61) into (10.69) and (10.71) gives

0 0 i 1 + i 4a ( )
[ ]
BiB

(
= 0 1 + i i 4b ) 0

(10.72)
( )
0 i 1 + i 4 a 1 + i i 4b
( )
and

[ B ] = (1 +(1+ ) ) 44ba (1 + )(01 + ) 4 (1 + )(01 + ) 4 .(10.73)


i
S i i i i

i i i i

Substituting (10.72) into (10.68) and the resulting matrices into (10.66)
gives the element stiffness matrix due to bending
242 FINITE ELEMENT ANALYSIS

[ ] [ k ] [ k ] [ k ]
kB
11
B
12
B
13
B
14

[k ]= E h3

[ k ] [ k ] [ k ]
B
22
B
23
B
24

[ k ] [ k ]
B
(10.74)
(
48 1 2 ) ab
B
33
B
34

SYM [ k ] B
44

where the submatrices are

0 0 0


[k ]
B
11
= 0

4
3
2 1
2
2
+ ( 1 ) b

1
(1 + )a b
2
,

4 2 1
1 2
0 (1 + )a b + ( 1 ) a
2 3 2

0 0 0

[k ]
B
12

= 0
2
3
{
2 ( 1 ) b 2 } 1
2

( 3 1) a b ,

0

1
( 3 1 ) a b
2
1
3
{ 4 + ( 1 ) a
2 2

}

0 0 0


[k ]
B
13
= 0

2
3
2 1
2
2
( 1 ) b

1
2
( 1 + )a b ,

2
1 2 1 2
0 (1 + )a b ( 1 ) a
2 3 2

0 0 0

[k ]
B
14
1
= 0
3
{
4 2 + ( 1 ) b 2 } 1
( 3 1) a b ,
2

0

1
2
( 3 1 ) a b
3
{
2 2
( 1 ) a 2

}
and
a b
= , = .
b a
10. PLATE BENDING 243

The remaining submatrices of (10.74) are given by relationships


corresponding to (10.52).
Substituting (10.73) into (10.71) and the resulting matrices into (10.67)
gives the element stiffness matrix due to shear

[ ] [ k ] [ k ] [ k ]
kS
11
S
12
S
13
S
14

[k ] =
E h3

[ k ] [ k ] [ k ] ,
S
22
S
23
S
24

[ k ] [ k ]
S
(10.75)
48 S a b
S S
33 34

SYM [ k ] S
44

where S = E h 3 12 G b 2 is a shear parameter similar to (5.97) and

1 + 2 2b a 1 + 2 2b a
[k ]
S
11

= 2b 2 b 2

0 , [k ]
S
12

= 2b 2 b2

0 ,
a 0 a2 a 0 a2

1 2 2b a 1 2 2b a
[k ]
S
13

= 2b 2 b 2

0 , [k ]
S
14

= 2b 2 b 2

0 .
a 0 a2 a 0 a2

The remaining submatrices of (10.75) are given by relationships


corresponding to (10.52).
The above expressions are presented in reference [87].
The vector of equivalent nodal forces is
pz
{ }
f = N 0 d A
e T

(10.76)

A 0
For pz = constant, substituting the shape functions from (10.63) and
(10.61) into (10.76) and integrating gives

{ f }= p a b 1
e
z 0 0 1 0 0 1 0 0 1 0 0 T (10.77)

which means that one quarter of the total force is concentrated at each node.
The bending and twisting moments per unit length are given by
244 FINITE ELEMENT ANALYSIS

Mx

My

=
h3
12
[ ]{q },
[ I3 ] [ D ] B B e
(10.78)
M
xy
where [ I 3 ] is defined by (10.53). The most accurate values are at the Gauss points
of a (2 2) numerical integration scheme.
The shear forces per unit length are
Qx
= h [ D ][ B ]{q }.
S S e
(10.79)
Qy
The most accurate values are at the centre of the element.
Benchmark problems have shown that the HTK element yields accurate
solutions for simply supported or clamped plates. However, large errors can occur
in the case of cantilever plates.
The above analysis can be easily extended to 8-node or higher-order plate
elements. The bending strains in the 8-node version are recovered accurately if
sampled at the reduced (2 2) Gauss points. In very thin plates, the shear strains
become very small and near-zero values in (10.26) thus imply a dependent
relationship between w , x and y which is not true for thick plates. This is
avoided by selective integration, lowering the integration order for the stiffness
matrix due to shear. However, the standard (2 2) integration is reported to give
good results for width-to-thickness ratios up to 50.
There is no perfect rectangular plate-bending element. There are different
formulations using mixtures of corner and mid-side freedoms in order to achieve
near complete polynomials. Alternative methods, considering rather difficult to
generate a displacement function valid over the entire rectangular element, suggest
to subdivide the element into regions (e.g., into four triangles) and work with
different displacement functions over each region. Obviously, the individual
functions must be continuous (up to the first derivatives) across the interior
boundaries as well as the exterior boundaries.

10.4 Triangular plate bending elements

In this section we outline the development of some plate-bending


triangular elements.
10. PLATE BENDING 245

10.4.1 Thin triangular element (non-conforming)

Figure 10.6 shows the T element (Tocher - 1962). The local x-axis lies
along the side 1-2 and the local y-axis is perpendicular to it. Nodes 1, 2 and 3 have
coordinates (0,0) , (x2 ,0) and (x3 , y3 ) . There are three degrees of freedom at each
w
node, namely, the transverse displacement w , and the two rotations x = and
y
w
y = with respect to the local axes.
x
Since the element has 9 degrees of freedom, the displacement function can
be represented by a polynomial having nine terms, that is

w = 1 + 2 x + 3 y + 4 x 2 + 5 x y + 6 y 2 +
(10.80)
( )
+ 7 x3 + 8 x 2 y + x y 2 + 9 y 3 .
Note that a complete cubic has ten terms so that the polynomial (10.80) is
incomplete. In order to maintain symmetry, the coefficients of x 2 y and x y 2 are
taken to be equal.

Fig. 10.6

The expression (10.80) can be written in the following matrix form


w= 1 x y x2 xy y2 x3 (x 2
y + x y2 ) y3 { } , (10.81)

w = P ( x , y ) { } , (10.81, a)
where

{ }T = 1 2 3 L 9 . (10.82)

Differentiating (10.81) gives


246 FINITE ELEMENT ANALYSIS

w 1 x y x2 x y y2 x3 x2 y + x y2 y3

x =0 0 1 0 x 2y 0 x2 + 2x y 3y2 { }. (10.83)
0 1 0 2 x y 0 3x 2 2x y y2 0
y

Evaluating (10.83) at the nodal points gives

{ q } = [ A ] { } ,
e e
(10.84)
where

{ q } = w
e T
1 x1 y1 L w3 x 3 y 3 , (10.85)

and the matrix


1 0 0 0 0 0 0 0 0
0 0 1 0 0 0 0 0 0

0 1 0 0 0 0 0 0 0

1 x2 0 x22 0 0 x23 0 0
[A ]
e
=0 0 1

0 x2 0 0 x22 0 . (10.86)

0 1 0 2 x2 0 0 3x22 0 0
1 x y3 x32 x3 y3 y32 x33 x32 y3 + x3 y32 y33
3
0 0 1 0 x3 2 y3 0 x32 + 2 x3 y3 3 y32

0 1 0 2 x3 y3 0 3x32 2 x3 y3 y32 0
Solving (10.84) for { } gives

{ } = [ Ae ] {q } .
1 e
(10.87)

Substituting (10.87) into (10.81, a) yields

[ ] { q }.
w = P (x , y ) Ae
1 e
(10.88)

Unfortunately, the matrix [ A ] is singular whenever


e

x2 2 x3 y3 = 0 (10.89)
and, therefore, cannot be inverted. If this occurs, the locations of the nodes should
be altered to avoid this condition.
This element is non-conforming. Evaluating (10.83) along the side 1-2, of
equation y = 0 , gives
10. PLATE BENDING 247

w 1 x 0 x2 0 0 x3 0 0

x =0 0 1 0 x 0 0 x2 0 { } . (10.90)
0 1 0 2 x 0 0 3x 2 0 0
y 1 2

The rotation x is a quadratic function having coefficients 3 , 5 and


8 . These cannot be determined using the values of x at nodes 1 and 2 only.
Therefore the normal slope is not continuous between elements. Moreover, the
assumed function (10.80) is not invariant with respect to the choice of coordinate
axes, due to the x 2 y and x y 2 terms.

The edge 1-2 was taken along the x-axis. In this case x is a tangential
component and y is a normal component of rotation. The transverse displacement
w and the normal component y are continuous between elements, while the
tangential component x is not. To alleviate this, in some plate-bending elements a
linear variation of the tangential component of rotation is adopted.
Substituting (10.88) into (10.4) and (10.8) gives

Ue =
1
2
{ q } [ k ]{ q }
e T e e
(10.91)

where the element stiffness matrix in the local coordinate system is

[ k ] = [ A ] 12h [ B ] [A ]
3
e e T T
[ D ] [ B ]d A e 1
(10.92)
A
and the strain-displacement matrix is

0 0 0 2 0 0 6x 2y 0
[ B ] = 0 0 0 0 0 2 0 2 x 6 y .
(10.93)
0 0 0 0 2 0 0 4 ( x + y ) 0

x
m n
A typical element of the integrand in (10.92) is of the form y dA ,
A
which can be evaluated analytically.
The vector of consistent nodal forces is given by

{ f }= [ A ]
e e T
P p z d A .
T
(10.94)
A
Note that nodal coordinates are usually given in global axes. Calculation of
the element stiffness matrix referred to local axes requires the local coordinates of
248 FINITE ELEMENT ANALYSIS

nodes 2 and 3. These can be obtained from their global coordinates using the
corresponding transformation matrix.
In order to assemble the global stiffness matrix, the element matrices have
to be first transformed from local to global axes through a matrix triple product
which is one of the more time-consuming procedures in finite element analysis.

10.4.2 Thick triangular element (conforming)

Consider a thick triangular element (THT element Henschel, Tocher,


1969) referred to a global coordinate system. Nodes 1, 2 and 3 have area
coordinates 1 , 2 and 3 (8.35). There are three independent degrees of freedom
at each node, namely, the transverse displacement w , and the two rotations X
and Y with respect to the global axes.
The displacement functions are assumed to be
3 3 3
w = i wi , X = i X i , Y = i Y i , (10.95)
i =1 i =1 i =1

where wi , X i and Y i are the degrees of freedom at node i. They ensure that w ,
X and Y are continuous between elements.
In matrix form, expressions (10.95) are written

w

X

=[N ]{ q e }, (10.96)

Y

where the vector of nodal displacements is

{ q } = w
e T
1 X 1 Y 1 L w3 X 3 Y 3 , (10.97)

and
1 0 0 2 0 0 3 0 0
[ N ] = 0 1 0 0 2 0 0 3 0 . (10.98)
0 0 1 0 0 2 0 0 3

Functions expressed in area coordinates can be differentiated with respect


to Cartesian coordinates using
10. PLATE BENDING 249

1 3 1 3
=
x 2 A i =1
i
i
, = i
y 2 A i =1 i
, (10.99)

where i and i are defined in (8.8).

Substituting (10.96) into (10.24) and using (10.99) gives the strain-
displacement matrix due to bending

0 0 1 0 0 2 0 0 3
[B ]
B
=
1
2 A
0 1 0 0 2 0 0 3 0 .

(10.100)
0 1 1 0 2 2 0 3 3

As this matrix is constant, the stiffness matrix due to bending is

[ k ] = 12h A [ B ] [ D ] [ B B ].
3 T
B B
(10.101)

Substituting (10.96) into (10.24) and using (10.99) shows that the strain-
displacement matrix due to shear is

1 2 3
0 1 0 2 0 3
[B ]
S
=

2A
1
2A
2
2A
3
.

(10.102)
1 0 2 0 3 0
2A 2A 2A

Substituting (10.102) into (10.67) and integrating using the formula


m ! n!p !
1 2 3 dA =
m n p
2A (10.103)
(m + n + p + 2)!
A

gives the stiffness matrix due to shear

kS
11
[ ] [ k ] [ k ]
S
12
S
13

[k ] S
= G h

[ k ] [ k ] ,
S
22
S
23
(10.104)
SYM

[ k ] S
33

where
250 FINITE ELEMENT ANALYSIS

12 + 12 1 1 2 1 + 2 1 1 1

4A 6 6 4A 6 6
[k ]
S
11

=
1
6
A
6

0 , [k ]
S
12

=
2
6
A
12

0 ,

1 A 2 A
0 0
6 6 6 12

3 1 + 3 1 1 1 22 + 22 2 2

4A 6 6 4A 6 6
[k ]
S
13

=
3
6
A
12

0 , [k ]
S
22

=
2
6
A
6

0 ,

3 A 2 A
0 0
6 12 6 6

3 2 + 3 2 2 2 23 + 32 3 3

4A 6 6 4A 6 6
[k ]
S
23

=
3
6
A
12

0 , [k ]
S
33

=
3
6
A
6

0 .

3
0
A 3
0
A
6 12 6 6

The complete element stiffness matrix is the sum (10.65).
The above relationships are presented in reference [87].
For pz = constant, substituting the shape functions from (10.98) into
(10.76) and integrating gives the vector of equivalent nodal forces

{ f }= p
e
z
A
3
1 0 0 1 0 0 1 0 0 T (10.105)

which shows that one third of the total force is concentrated at each node.

10.4.3 Discrete Kirchhoff triangles (DKT)

The difficulties to formulate simple (minimum degrees of freedom) and


high-performing elements based on the Kirchhoff theory at the continuum level
(which requires a C1 continuity for w ) led to the development of the so-called
Discrete Kirchhoff (DK) technique (Wempner-1969, Stricklin et al.-1969, Dhatt-
1967).
10. PLATE BENDING 251

In the formulation of various DK plate elements, only the bending strain


energy is considered, in which the curvatures are expressed, as in the Reissner-
Mindlin plate theory, in terms of the first derivatives of the rotations (10.24). In
this case only C 0 approximations of these rotations can be considered. The shear
strain energy is neglected.
The Kirchhoff constraints are imposed in a discrete manner on the element
or/and on the sides. For instance, the transverse shear strains are either taken equal
to zero at the mid-side points (collocation on sides) or their integral along each
edge is taken equal to zero. The aim is to preserve the C 0 continuity of the
tangential components of rotations (normal slopes). Constant curvature patch-tests
are needed to check the validity of the elements.
The T element presented in section 10.4.1 can be called a continuous
Kirchhoff triangle. It was shown that, along a side, the rotations vary quadratically
and their tangential component (normal slope) is not continuous between elements.
To remedy this, in the discrete Kirchhoff triangles the two components of
rotations are assumed independent of one another. The tangential component s is
assumed to vary linearly along each edge, while the normal component n varies
quadratically (Fig. 10.7, a). The latter condition means that the transverse
displacement w can vary cubically.

a b
Fig. 10.7

For a side i j , of length l k , the tangential and normal components of


rotations s k and n k at the mid-side node k, are defined in terms of the
components along the coordinate axes by
s k ck sk x k
= (10.106)
n k sk ck y k
where
252 FINITE ELEMENT ANALYSIS

(
ck = x j x i l k , ) ( )
sk = y j y i l k . (10.107)

Consider a discrete Kirchhoff triangle (DKT) with three nodes and three
degrees of freedom per node (sometimes called DKT9).
Initially, it is assumed that it has six nodes (Fig. 10.7, b). Then, the degrees
of freedom at mid-side nodes are eliminated.
The rotation n varies parabolically (Fig. 10.8, a). Its expression in terms
of the dimensionless coordinate
s = s l k
can be written as
n = ( 1 s ) ni + s n j + 4s ( 1 s ) k .
At the mid-side node k

nk =
1
2
(
ni + n j + k ,)
where the first term in the right hand side is expressed in terms of the rotations at
the corner nodes. The parameter k has to be eliminated.

The rotation s is assumed to vary linearly (Fig. 10.8, b)

s = ( 1 s ) s i + s s j .

Fig. 10.8

In order to eliminate the degrees of freedom at mid-side nodes, the three


parameters 4 , 5 , 6 must be expressed in terms of the degrees of freedom at
the corner nodes. This is done requiring the shear strains s z to vanish along each
side. One kind of discrete Kirchhoff constraint is formulated in integral form as
j j
w
i
s z ds =

i

s
+ n d s = 0 .

(10.108)
10. PLATE BENDING 253

Equation (10.108) can be written


j lk

dw +
i 0
n ds = 0 ,

(w j wi ) + l k [ ( 1 s ) ni + s n j + 4s ( 1 s ) k ] d s = 0 ,
0

wherefrom we obtain

k =
3
2l k
(3
) (
wi w j n i + n j ,
4
)
or, in terms of the degrees of freedom of the corner points

k =
3
2l k
( 3
) (
wi w j s k x i + c k y i s k x j + c k y j .
4
) (10.109)

The equation (10.109) can also be obtained by adopting a Hermitian cubic


polynomial for w (s ) and using the discrete Kirchhoff collocation constraint (zero
w
shear strain) s z = + n = 0 at points i, j, k. Then, the condition s z = 0 will be
s
satisfied at all points along the contour, because w is cubical and n is quadratic.

The rotations x and y can be expressed as

x = N { x } + Px { },
(10.110)
y = N { y }+ Py { },
where
N = N1 N 2 N3 = 1 ,
{ x } = x1 x 2 x3 T , { y }= y1 y 2 y 3 T ,

{ } = 4 5 6 T , (10.111)

Px = P4 s4 P5 s5 P6 s6 , Py = P4c4 P5c5 P6 c6 ,

P4 P5 P6 = 4 ( 1 ) 4 4 ( 1 ) .
Substituting (10.109) into (10.110) we obtain the explicit expressions of
x and y in terms of the corner nodal variables
254 FINITE ELEMENT ANALYSIS

x
=[N ] {q e }, (10.112)
y
where
{ q } = w
e T
1 x1 y1 L w3 x 3 y 3 , (10.113)

[ N ] =
N 2x N3x ,
Nx
1
(10.114)
N1 N 2 N 3
y y y

N1x = N ix1 N ix2


N ix3 , N1y = Niy1 N iy2
N iy3 . (10.115)

In (10.115), the shape functions have the following expressions


3 3 3 3
N ix1 = Pk sk + Pm sm , N iy1 = Pk ck Pm cm ,
2l k 2l m 2l k 2l m

3 3
N ix2 = N i Pk sk2 Pm sm2 , N iy2 = N ix3 , (10.116)
4 4
3 3 3 3
N ix3 = Pk sk ck + Pm sm cm , N iy3 = N i Pk ck2 Pm cm2 .
4 4 4 4
In the above expressions the indices k and m relate to the two edges having
the common corner point i, as shown in Table 10.1.
Taking into account the hypotheses adopted in formulating the DKT, the
nodal rotations x i and y i have the expressions used in Kirchhoffs theory

w w
x i = , y i = , ( i = 1, 2, 3 )
y i x i
allowing the introduction of Kirchhoff-type boundary conditions.

Table 10.1
Corner point, i Side k ( i j ) Side m ( i j )
1 4 (1 2 ) 6 (3 1 )
2 5 (23 ) 4 (1 2 )
3 6 (3 1 ) 5 (23 )
10. PLATE BENDING 255

Because the parameters k are eliminated using an expression which is a


function of the nodal variables of the side k only, the continuity C 0 of n is
maintained.
The stiffness matrix of the DKT element can be expressed in explicit form,
in a local coordinate system [18], using a Hammer integration rule. In the
following, it will be derived in global coordinates as in [68].
Equation (10.112) can be split as

{ }
x = G q e , { }
y = H q e . (10.117)

The shape function row vectors G and H , presented explicitly in


terms of the local oblique coordinates and , can be rewritten as

G = 1 2 2 [ G ] ,
(10.118)
H = 1 2 2 [ H ] .

The curvature vector (10.24) is written for convenience as


T
x x y
{ }= y = 1 2 12 T . (10.119)
x y x y

The elements of the curvature vector can be expressed in terms of the rows
of the ( 6 9 ) matrices [ G ] and [ H ] as follows

1 =
1
1 { X }{ q e },
2A

2 =
1
2A
1 {Y } q ,
e
{ } (10.120)

12 =
1
2A
1 { Z } q ,
e
{ }
where A is the area of the triangle with vertices (x1 , y1 ) , (x2 , y2 ) and (x3 , y3 ) .
In (10.120)

2 G + 3 G 2 H + 3 H
2 3 2 3

{ X } = 2 2 G 4 + 3 G 5 , {Y } = 2 2 H 4 + 3 H 5 ,
G + 2 G H + 2 H
2 5 3 6 2 5 3 6
256 FINITE ELEMENT ANALYSIS

2 G + 3 G + 2 H + 3 H
2 3 2 3

{ Z } = 2 2 G 4 + 3 G 5 + 2 2 H 4 + 3 H 5 , (10.121)
G + 2 G + H + 2 H
2 5 3 6 2 5 3 6

where G i and H i represent the ith row of [ G ] and [ H ] , respectively, and


i = y j yk and i = xk x j as in (8.8), with i, j, k taking values 1, 2, 3
cyclically.
Substituting for the curvature terms into the bending strain energy
expression
Ue =
1
{ }T [ D ] { }d x d y ,
(10.122)
2
A

where the material matrix [ D ] has the form (10.13), U e can be written as

Ue =
1
2
{ q } [ K ]{ q },
e T e e

[ ]
where K e is the element stiffness matrix in global coordinates, given by

[ X ] D11 [ R ] D12 [ R ] D16 [ R ] [ X ]


T

[K ] e
=
1
2A [ Y ]


D22 [ R ] D26 [ R ] [ Y ] , (10.123)
[ Z ] SYM
D66 [ R ] [ Z ]

with
12 4 4
1
[ R ] = 4 2 1 . (10.124)
24
4 1 2

[ ]
The matrix K e is of rank 6, so there are 3 deformation modes with zero
strain energy.
The integrations in (10.122) have the form


m ! n!
m n d d == .
(m + n + 2)!
0 0

The presented procedure avoids the matrix triple products needed when the
element stiffness matrix is given in the local coordinate system.
References
1. * * * A Finite Element Primer, National Engineering Laboratory, Glasgow,
1986.
2. * * * ANSYS Users Manual for Revision 5.0, Swanson Analysis Systems Inc.,
Houston, Pa, 1992.
3. * * * , Author Index of Finite Element Books, Linkping Universiteit,
http://ohio.ikp.liu.se/fe/auth.html.
4. * * * Guide de validation des progiciels de calcul de structures, Socit
Franaise des Mcaniciens, Afnor, Paris, 1990.
5. * * * The Standard NAFEMS Benchmarks, National Agency for Finite Element
Methods and Standards, Glasgow, 1989.
6. Adini, A. and Clough, R. W., Analysis of plate bending by the finite element
method, Report G7337 to Nat. Sci. Found. U.S.A., 1961.
7. Akin, J. E., Finite Element Analysis with Error Estimators, Butterworth
Heinemann, Amsterdam, 2005.
8. Argyris, J. H., Energy theorems and structural analysis, Aircraft Engineering,
26, Oct.-Nov. 1954; 27, Feb.-May 1955.
9. Argyris, J. H. and Kelsey, S., Energy Theorems and Structural Analysis,
Butterworth, London, 1960.
10. Argyris, J. H., Continua and discontinua, Proc. 1st Conf. on Matrix Methods in
Structural Mechanics, Report AFFDL-TR-66-80, Air Force Institute of
Technology, Dayton, Ohio, 10-170, 1965.
11. Argyris, J. and Mlejnek, H.-P., Die Methode der finiten Elemente, 3 vol.,
Vieweg, Braunschweig, 1986, 1987, 1988.
12. Archer. J. S., Consistent mass matrix for distributed mass systems, J. Str. Div.
Proc. ASCE, 89, ST4, 161-178, 1963.
13. Archer. J. S., Consistent mass matrix formulation for structural analysis using
finite element techniques, AIAA J., 3, 1910-1918, 1965.
14. Babuska I. and Strouboulis, T., The Finite Element Method and Its Reliability,
Oxford University Press, Oxford, 2001.
15. Barlow, J., Optimal stress locations in finite element models, Int. J. Num. Meth.
Engrg., 10, 243-251, 1976.
258 FINITE ELEMENT ANALYSIS

16. Bathe, K.-J., Finite Element Procedures in Engineering Analysis, Prentice Hall,
Englewood Cliffs, NJ, 1982.
17. Bathe, K.-J. and Wilson, E. L., Numerical Methods in Finite Element Analysis,
Prentice Hall, Englewood Cliffs, NJ, 1976.
18. Batoz, J. L., Bathe, K.-J. and Ho, L. W., A study of three-node triangular plate
bending elements, Int. J. Num. Meth. Engrg., 15, 1771-1812, 1980.
19. Batoz, J. L. and Dhatt, G., Modelisation des structures par lements finis, 3
vol., Hermes, Paris, 1990, 1992.
20. Belytschko, T. and Hughes T. J. R. (eds.), Computational Methods for
Transient Analysis, North-Holland, Amsterdam, 1983.
21. Blumenfeld, M., Introducere n metoda elementelor finite, Editura tehnic,
Bucureti, 1995.
22. Blumenfeld, M., Ioni, A. i Mare, C., Metoda elementelor finite aplicaii i
programe introductive, U.P.B., 1992.
23. Bogner, F. K., Fox, R. L. and Schmit, L. A., The generation of interelement
compatible stiffness and mass matrices by the use of interpolation formulas,
Proc. 1st Conf. on Matrix Methods in Structural Mechanics, Air Force Institute
of Technology, Report AFFDL TR 66-80, Dayton, Ohio, Nov. 1965.
24. Brebbia, C. A. and Connor, J. J., Fundamentals of Finite Element Techniques,
Butterworths, London, 1973.
25. Brown, D. K., An Introduction to the Finite Element Method Using BASIC
Programs, Surrey University Press, Glasgow, 1984.
26. Buchanan, G. R., Schaums Outline of Theory and Problems of Finite Element
Analysis, Schaum, New York, 1995.
27. Buell, W. R. and Bush, B. A., Mesh generation - A survey, J. Eng. Industry,
Trans. ASME, Series B, 95, 332-338, 1973.
28. Castigliano, A., Nuova teoria intorno allequilibrio dei sistemi elastici, Trans.
Acad. Sci., Torino, 1876.
29. Chandrupatla, T. R. and Belegundu, A. D., Intoduction to Finite Elements in
Engineering, Prentice Hall International, London, 1991.
30. Cheung, Y. K. and Yeo, M. F., A Practical Introduction to Finite Element
Analysis, Pitman, London, 1979.
31. Clough, R. W., The finite element method in plane stress analysis, Proc 2nd
ASCE Conf. on Electronic Computation, Pittsburgh, Pa, 1960.
32. Clough, R. W., The finite element method after twenty-five years: a personal
view, Computers and Structures, 12, 361-370, 1980.
REFERENCES 259

33. Cook, R. D., Malkus, D. S. and Plesha, M. E., Concepts and Applications of
Finite Element Analysis, 3rd ed., Wiley, New York,1989, 2001 (1st ed. 1974).
34. Cook, R. D., Finite Element Modeling for Stress Analysis, Wiley, New York,
1995.
35. Courant, R., Variational methods for the solution of problems in equilibrium
and vibrations, Bull. Amer. Math. Soc., 49, 1-23, 1943.
36. Crisfield, M. A., Non-linear Finite Element Analysis of Solids and Structures,
vol.1, 2, Wiley, Chichester, 1991, 1997.
37. Cuteanu, E. i Marinov, R., Metoda elementelor finite n proiectarea
structurilor, Editura Facla, Timioara, 1980.
38. Desai, C. S. and Abel, J. F., Introduction to the Finite Element Method, Van
Nostrand Reinhold, New York, 1972.
39. Dhatt, G. and Touzot, G., Une prsentation de la mthode des lments finis, 2e
d., Maloine, Paris, 1984.
40. Eisley, J. G., Mechanics of Elastic Structures, Prentice Hall, Englewood Cliffs,
NJ, .1989.
41. Ergatoudis, I., Irons, B. M. and Zienkiewicz, O. C., Curved isoparametric
quadrilateral elements for finite element analysis, Int. J. Solids Struct., 4, 31-
42, 1968.
42. Felippa, C. A., Solution of equations with skyline-stored symmetric coefficient
matrix, Computers & Structures, 5, 13-25, 1975.
43. Felippa, C. A., web-posted Lectures in Introductory Finite Element Methods, at
http://caswww.colorado.edu/courses.d/IFEM.d/Home.html.
44. Fraeijs de Veubeke, B. M., Displacement and equilibrium models, in Stress
Analysis, ed. by O. C. Zienkiewicz and G. Hollister, Wiley, London, 145-197,
1965.
45. Fraeijs de Veubeke, B. M., A conforming finite element for plate bending, Int.
J. Solids Struct., 4, 95-108, 1968.
46. Fried, I., Shear in C 0 and C1 plate bending elements, Int. J. Solids Struct., 9,
449-460, 1973.
47. Galerkin, B. G., Series solutions of some problems of elastic equilibrium of
rods and plates, Vestnik Injenerov, Petrograd, 1915 (in Russian).
48. Gallagher, R. H., Finite Element Analysis-Fundamentals, Prentice Hall,
Englewood Cliffs, NJ, 1975.
49. Grbea, D., Analiza cu elemente finite, Editura tehnic, Bucureti, 1990.
50. Guyan, R. J., Reduction of stiffness and mass matrices, AIAA J., 3, 380, 1965.
260 FINITE ELEMENT ANALYSIS

51. Hahn, H. G., Methode der finiten Elemente in der Festigkeitslehre,


Akademische Verlagsgesellschaft, Frankfurt am Main, 1975.
52. Hinton, E. and Owen, D. R. J., An Introduction to Finite Element Computation,
Pineridge Press, Swansea, 1979.
53. Hitchings, D. (ed.), A Finite Element Dynamics Primer, NAFEMS, Glasgow,
1986.
54. Hrenikoff, A., Solution of problems in elasticity by the framework method, J.
Appl. Mech., 8, 169-175, 1941.
55. Huang, H. C., Defect-Free Shell Elements, PhD Thesis, University College of
Swansea, 1986.
56. Huang, H. C. and Hinton, E., A nine node Lagrangian Mindlin plate element
with enhanced shear interpolation, Eng. Comput., 1, 369-379, 1984.
57. Huebner, K. H. and Thornton, E. A., The Finite Element Method for Engineers,
2nd ed, Wiley-Interscience, New York, 1982.
58. Hughes, T. J. R., The Finite Element Method: Linear Static and Dynamic Finite
Element Analysis, Prentice Hall, Englewood Cliffs, NJ, 1987, 2000.
59. Hughes, T. J. R. and Tezduduyar, T. E., Finite elements based upon Mindlin
plate theory with particular reference to the four-node bilinear isoparametric
element, J. Appl. Mech., 48, 587-596, 1981.
60. Hughes, T. J. R., Taylor, R. L. and Knoknukulchai, W., A simple and efficient
finite element for plate bending, Int. J. Num. Meth. Engrg., 11, 1529-1543,
1977.
61. Imbert, J. F., Analyse des structures par elements finis, 3e d, Cpadus-
ditions, Toulouse, 1991.
62. Irons, B. M. and Ahmad, S., Techniques of Finite Elements, Ellis Horwood,
Chichester, 1978.
63. Irons, B. M., Comments on Matrices for the direct stiffness method, by R. J.
Melosh, AIAA J., 2, 403, 1964.
64. Irons, B. M. Engineering application of numerical integration in stiffness
methods, AIAA J., 4, 2035-2037, 1966.
65. Irons, B. M., A frontal solution program for finite element analysis, Int. J.
Numer. Meth. Engrg., 2, 5-32, 1970.
66. Irons, B. M. and Shrive N., Finite Element Primer, Wiley, New York, 1983.
67. Irons, B. M. and Zienkiewicz, O. C., The isoparametric finite element system
a new concept in finite element analysis, Proc. Conf. Recent Advances in Stress
Analysis, Royal Aero. Soc., London, 1968.
REFERENCES 261

68. Jeyachandrabose, C., Kirkhope, J. and Babu, C. R., An alternative explicit


formulation for the DKT plate-bending element, Int. J. Num. Meth. Engng., 21,
1289-1293, 1985.
69. Kardestuncer, H. and Norrie, D. H. (ed), Finite Element Handbook, McGraw-
Hill, New York, 1987.
70. Kikuchi, N., Finite Element Methods in Mechanics, Cambridge University
Press, Cambridge, 1986.
71. Krishnamoorthy, C. S., Finite Element Analysis. Theory and Programming,
2nd ed, Tata McGraw-Hill, New Delhi, 1994 (1st ed 1987).
72. Levy, S., Computation of influence coefficients for aircraft structures with
discontinuities and sweptback, J. Aero. Sci., 14, 547-560, 1947.
73. Levy, S., Structural analysis and influence coefficients for delta wings, J. Aero.
Sci., 20, 677-684, 1953.
74. Link, M., Finite Elemente in der Statik und Dynamik, 3.Auflage, Teubner,
Stuttgart, 2002.
75. MacNeal, R. H., (ed), The NASTRAN Theoretical Manual, NASA SP-221,
1970.
76. MacNeal, R. H., Derivation of element stiffness matrices by assumed strain
distributions, Nucl. Engrg. Design, 70, 3-12, 1982.
77. MacNeal, R. H. and Mc Cormik, C. W., The NASTRAN Computer Program
for Structural Analysis, Computers and Structures, 1, 389-412, 1971.
78. Martin, H. C. and Carey, G. F., Introduction to Finite Element Analysis,.
Theory and Application, McGraw-Hill, New York, 1972.
79. Maxwell, J. C., On the calculations of the equilibrium and stiffness of frames,
Phil. Mag., 27, 294, 1864.
80. Melosh, R. J., Bases for the derivation of matrices for the direct stiffness
method, AIAA J., 1, 1631-1637, 1963.
81. Moaveni, S., Finite Element Analysis. Theory and Application with ANSYS,
Prentice Hall, Upper Saddle River, NJ, 1999, 2003.
82. Norrie, D. H. and de Vries, G., An Introduction to the Finite Element Method,
2nd ed, Academic Press, New York, 1978.
83. Ostenfeld, A., Die Deformationsmethode, Springer, Berlin, 1926.
84. Owen D. R. J. and Hinton, E., A Simple Guide to Finite Elements, Pineridge
Press, Swansea, 1980.
85. Pascariu I., Elemente finite Concepte, aplicaii, Editura militar, Bucureti,
1985.
262 FINITE ELEMENT ANALYSIS

86. Paz, M., Structural Dynamics. Theory and Computation, Van Nostrand
Reinhold, New York, 1980.
87. Petyt, M., Introduction to Finite Element Vibration Analysis, Cambridge
University Press, 1990.
88. Pian, T. H. H., Derivation of element stiffness matrices, AIAA J., 2, 576-577,
1964.
89. Przemieniecki, J. S., Theory of Matrix Structural Analysis, McGraw-Hill, New
York, 1968.
90. Rao, S. S., The Finite Element Method in Engineering, Pergamon Press,
Oxford, 1982.
91. Reddy, J. N., An Introduction to the Finite Element Method, McGraw-Hill
International Student Edition, New York, 1984.
92. Ritz, W., ber eine neue Methode zur Lsung gewisser Variationsprobleme der
mathematischen Physik, J. reine angew. Math., 35, 1-61, 1908.
93. Ross, C. T. F., Computational Methods in Structural and Continuum
Mechanics, Ellis Horwood, Chichester, 1982.
94. Rossmanith, H. P., Finite Elemente in der Bruchmechanik, Springer, Berlin,
1982.
95. Schwarz, H. R., Methode der finiten Elemente, Teubner, Stuttgart, 1980.
96. Shames, I. H. and Dym, C. L., Energy and Finite Element Methods in
Structural Mechanics, Taylor & Francis, Levittown, Pa, 1985.
97. Smith, I. M. and Griffiths, V., Programming the Finite Element Method,
Wiley, New York, 1982, 1988, 1998, 2004.
98. Sorohan, t., Metoda elementelor finite n ingineria mecanic. Programe i
aplicaii, vol.1, U. P. B., 1996.
99. Sorohan, t. i Constantinescu, I. N., Practica modelrii i analizei cu
elemente finite, Editura Politehnica Press, Bucureti, 2003.
100. Stematiu, D., Calculul structurilor hidrotehnice prin metoda elementelor
finite, Editura tehnic, Bucureti, 1988.
101. Strang, G. and Fix, G., An Analysis of the Finite Element Method, Prentice
Hall, Englewood Cliffs, NJ, 1973.
102. Szabo, B. and Babuska, I., Finite Element Analysis, Wiley, New York, 1991.
103. Taig, I. C. and Kerr, R. I., Some problems in the discrete element
representation of aircraft structures, in Matrix Methods of Structural Analysis,
ed. by B. M. Fraeijs de Veubeke, Pergamon Press, London, 1964.
REFERENCES 263

104. Turner, M. J., The direct stiffness method of structural analysis, Structural
and Materials Panel Paper, AGARD Meeting, Aachen, 1959.
105. Turner, M. J., Clough, R. W., Martin, H. C. and Topp, L. J., Stiffness and
deflection analysis of complex structures, J. Aero. Sci., 23, 805-824, 1956.
106. Turner, M. J., Martin, H. C. and Weikel, R. C., Further development and
applications of the stiffness method, in Matrix Methods of Structural Analysis,
ed. by B. M. Fraeijs de Veubeke, Pergamon Press, London, 203-266, 1964.
107. Uhrig, R., Zur Berechnung der Steifigkeitsmatrizen des Balkens, Der
Stahlbau, 4, 123-125, 1965.
108. Waller, H. and Krings, W., Matrizenmethoden in der Maschinen- und
Bauwerksdynamik, B.I.-Wissenschaftsverlag, Mannheim, 1975.
109. Wilson, E. L., Finite element analysis of two-dimensional structures, PhD
Dissertation, Dept. of Civil Engrg, Univ. of California at Berkeley, 1963.
110. Wilson, E. L., The static condensation algorithm, Int. J. Numer. Meth. Engrg.,
8, 198-203, 1974.
111. Wilson, E. L., Automation of the finite element method - a historical view,
Finite Elements Anal. Des., 13, 91-104, 1993.
112. Zienkiewicz, O., Introductory Lectures on the Finite Element Method, CISM
Udine Course No. 130, Springer, Wien, 1973.
113. Zienkiewicz, O., Taylor, R. L. and Too, J. M., Reduced integration technique
in general analysis of plates and shells, Int. J. Num. Meth. Engrg., 3, 275-290,
1971.
114. Zienkiewicz, O. C. and Cheung, Y. K., The Finite Element Method in
Engineering Science, McGraw-Hill, London, 1967, 1968, 1970.
115. Zienkiewicz, O. C. and Taylor, R. L., The Finite Element Method, 4th ed,
McGraw-Hill, London, vol.1, 1988, vol.2, 1993.
116. Zienkiewicz, O. C., Irons, B. M., Ergatoudis, J., Ahmad, S. and Scott, F. C.,
Isoparametric and associated element family for two- and three-dimensional
analysis, in Finite Element Methods in Stress Analysis, eds. I. Holland and K.
Bell, Tapir, New York, 1969.
117. Zienkiewicz, O. C. and Phillips, D. V., An automatic mesh generation scheme
for plane and curved surfaces by isoparametric coordinates, Int. J. Num. Meth.
Engrg., 3, 519-528, 1971.
118. Zurmhl, R., Ein Matrizenverfahren zur Behandlung von Biegeschwingungen
nach der Deformationsmethode, Ingenieur Archiv, 22, 201-213, 1963.
Index

ACM plate-bending element 232 8-node quad 214


Admissible functions 144 ACM plate-bending element 237
Area coordinates 180 T plate-bending element 247
Aspect ratio 178 THT plate-bending element 249
Assembly of global matrices 100, 151 Expanded stiffness matrix 29
Axial effects 97 External potential energy 140
Band storage 38 Force transformation 20
Bars 47 Four-node rectangle 176
Bar element 47 Frame element 97
Beams 79
Gauss points 200
Beam element 83
Gauss quadrature 200
stiffness matrix 87
vector of nodal forces 89 one-dimensional 200
Bernoulli-Euler beam theory 79
two-dimensional 203
Gravity loading 217
BFS plate-bending element 238
Grids 111
Compatibility 28, 150, 187
Hermitian polynomials 85
Conforming elements 150, 238, 248
HTK plate-bending element 239
Consistent nodal forces 214
Consistent vector of nodal forces 89 Initial strain effects 59
Constant strain triangle 153 Internal forces 35
Continuity 132 Inverse of Jacobian matrix 196
Convergence 97, 187 Isoparametric elements 191
Coordinate transformation 19, 98
Jacobian matrix 195
Deep beam 117 Jacobian positiveness 223
Direct method 26 Joint force equilibrium equations 31
Direct stiffness method 17
Discrete Kirchhoff constraints 252 Kinematic equivalent forces 47
Discrete Kirchhoff triangle 250 Kirchhoffs hypotheses 225
Displacement method 9 Lagrange polynomials 220
Linear elasticity 123
Effective shear area 118
Eight-node quad 208 Linear strain triangle 182
Eight-node rectangle 178 Master element 192
Energy methods 131
Equation of equilibrium 199 Natural coordinates 192
Equivalent nodal forces, vector of Nine-node quad 219
beam 90, 95 Nodal approximation 155
frame 98 Nodal forces 55
constant strain triangle 160 Node numbering 37
linear quad 199 Non-conforming elements 232, 245
Numerical integration 200
266 FINITE ELEMENT ANALYSIS

Object of FEA 1 Skyline storage 39


Stiffness matrix 17
Pascals triangle 220
bars 17, 22, 52
Patch test 190
shaft 61
Plate bending elements 225
beam 87, 94, 122
rectangular 232 constant strain triangle 159
triangular 244 frame 98
Polynomial approximation 154
grid 113
Principle of
linear strain triangle 183
minimum total potential energy 139
linear quadrilateral 199
virtual displacements 134 eight-node quad 211
virtual work 131 T plate-bending element 247
Quadratic strain triangle 185 ACM plate-bending element 236
triangle 222 HTK plate-bending element 242
Quadrilateral membrane element 191 THT plate-bending element 249
eight-node 208 Strain-displacement matrix 158, 236
linear 191 relations 121
nine-node 219 Strain energy 124 ..139
bending 231
Rayleigh-Ritz method 143 shear 232
Reactions 35 Stress averaging 161
Rectangular membrane element 176 Stress-strain relations 122
thin plate element 232 Subparametric elements 191
thick plate elements 239
Reduced global stiffness matrix 33 T plate-bending element 245
integration 205 Temperature effects 124
Reference element 192 Thick plate theory 229
Reissner-Mindlin plate theory 229 Thermal loads 36
Thin plate theory 225
Serendipity elements 209
Total potential energy 140
Shaft elements 60 Transformation of
Shape functions differential operators 195
bars 50, 53 infinitesimal area 197
shaft 62 Triangle area 156
beam 85, 107, 120 quadratic strain triangle 185
constant strain triangle 157 Triangular membrane element 180
four-node rectangle 177 linear strain triangle 182
eight-node rectangle 179 quadratic strain triangle 185
linear strain triangle 182 six-node 221
quadratic strain triangle 186 Triangular thin plate element 245
linear quadrilateral 193 thick plate element 248
eight-node quad 209 Two-dimensional elements 153
nine-node quad 221 Two-point formula 202
six-node triangle 221
ACM plate-bending element 234 Unreduced global stiffness matrix 30
BFS plate-bending element 238 Virtual displacements 131
HTK plate-bending element 240 work of external loads 133
THT plate-bending element 248
work of internal forces 193
Shear correction factor 118, 230
Six-node triangle 221 WB plate-bending element 239

Vous aimerez peut-être aussi