Вы находитесь на странице: 1из 158

.

Structural Analysis with FEM

F
2

Gerald Kress
31. May 2014

Composite Materials and Adaptive Structures

Contents
1 Goals and contents
1.1 Significance of the Finite-Element Method . . . . . . . . . . . . . . . .
1.2 Educational objective . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.3 Frame . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.4 Learning matter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.4.1 What is a finite element and which properties does it possess?
1.4.2 FEM standard notation for linear elastic problems . . . . . . .
1.4.3 The search for most efficient finite element formulations . . . .
1.4.4 Numerical preprocessing . . . . . . . . . . . . . . . . . . . . . .
1.4.5 Direct and iterative methods for solving a system of equations
1.4.6 Substructuring . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.4.7 Post-processing with evaluation of the secondary solution . . .
1.4.8 Reduced modeling techniques . . . . . . . . . . . . . . . . . . .
1.4.9 Nonlinear damage progression analysis . . . . . . . . . . . . . .
1.4.10 Beam finite elements and locking effect . . . . . . . . . . . . .
1.4.11 Eigenvalue problems . . . . . . . . . . . . . . . . . . . . . . . .
1.5 Acknowledgment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

2 Direct element equation derivation


2.1 The truss and its element equations . . .
2.1.1 Direct derivation of the truss finite
2.2 The beam and its element equations . . .
2.3 Summary direct methods . . . . . . . . .

1
1
1
1
2
2
3
3
3
4
4
4
4
4
4
5
5

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

7
. 7
. 9
. 11
. 13

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

15
15
15
16
19
21
21
22
22
24

4 Truss element revisited


4.1 Truss problem weak variational form . . . . . . . . . . . . . . . . . . . . . .
4.2 Truss weak variational form discretization . . . . . . . . . . . . . . . . . . .
4.3 Transformation between truss and element coordinates . . . . . . . . . . . .

27
27
28
28

. . . . . . . . . . .
element equations
. . . . . . . . . . .
. . . . . . . . . . .

3 FEM theory in standard notation


3.1 Concept of deriving a numerical system of equations . .
3.2 Basic problem of linear elasticity . . . . . . . . . . . . .
3.3 Derivation of the weak variational form . . . . . . . . .
3.4 Subdividing the problem domain into finite elements . .
3.4.1 Element and system coordinates transformation .
3.4.2 Domain and boundary differentials . . . . . . . .
3.4.3 Iso-parametric elements and Jacobian matrix . .
3.5 Boundary conditions and equivalent nodal forces . . . .
3.6 Polynomial element shape functions . . . . . . . . . . .

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

c ETH Z

urich IMES-ST, March 26, 2014

4.4
4.5
4.6
4.7

Truss finite element numerical equations . . . . .


Assembly of system equations . . . . . . . . . . .
Implementation of essential boundary conditions
Directional transformation of the truss element .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

5 Planar finite elements


5.1 Basic elasticity equations for plane states . . . . . . . . . . . .
5.2 Planar element domain and boundary differentials . . . . . . .
5.3 Higher-order shape functions for quadrilateral elements . . . .
5.4 The linear triangular element . . . . . . . . . . . . . . . . . . .
5.4.1 B-matrix structure for the linear triangular element . .
5.4.2 Linear triangular element domain integral . . . . . . . .
5.4.3 Linear triangular element boundary integrals . . . . . .
5.4.4 Linear triangular element kinematically equivalent nodal
5.5 The bilinear quadrilateral finite element . . . . . . . . . . . . .
5.6 Three- and four-noded elements bending performance . . . . .
5.6.1 Meshing with triangular and quadrilateral elements . . .
5.7 Curvilinear elements . . . . . . . . . . . . . . . . . . . . . . . .
6 Element technology
6.1 Sample problem truss element . . . . . . . . . . . . . . . . . .
6.2 Sample problem quadrilateral element . . . . . . . . . . . . .
6.2.1 Bilinear quadrilateral element behavior . . . . . . . .
6.2.2 Additional incompatible element displacement modes

.
.
.
.

7 Numerical preprocessing
7.1 Element matrices size . . . . . . . . . . . . . . . . . . . . . . .
7.2 Gauss quadrature rule . . . . . . . . . . . . . . . . . . . . . . .
7.2.1 Polynomial order and quadrature rule . . . . . . . . . .
7.2.2 Table and two-dimensional point grating . . . . . . . . .
7.3 Selection of the quadrature rule . . . . . . . . . . . . . . . . . .
7.3.1 Integration of the stiffness matrix . . . . . . . . . . . . .
7.3.2 Integration of the right-hand side . . . . . . . . . . . . .
7.4 Element shape distortion and integration error . . . . . . . . .
7.5 Pre-processing program parts . . . . . . . . . . . . . . . . . . .
7.5.1 Assembly of the global stiffness matrix . . . . . . . . . .
7.5.2 Calculation of element stiffness matrices . . . . . . . . .
7.5.3 Calculation of matrix B and Jacobi-matrix determinant
7.6 Pre-processing numerical effort . . . . . . . . . . . . . . . . . .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
forces
. . . .
. . . .
. . . .
. . . .

.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.

29
29
30
31

.
.
.
.
.
.
.
.
.
.
.
.

33
34
34
35
36
37
38
39
40
41
42
42
43

.
.
.
.

45
45
47
47
51

.
.
.
.
.
.
.
.
.
.
.
.
.

57
57
58
59
60
61
61
62
62
64
64
65
66
67

8 Degree-of-freedom coupling
69
8.1 Sample problem slanted support . . . . . . . . . . . . . . . . . . . . . . . . 69
9 Solution of the system equations
9.1 Band width minimization . . . .
9.2 Direct solution methods . . . . .
9.2.1 Gauss elimination . . . .
9.2.2 Cholesky decomposition .
9.3 Iterative solution methods . . . .
9.3.1 Jacobi iteration . . . . . .

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

71
71
73
74
75
77
77

.
.
.
.
.

79
79
82
82
83

10 Substructure technique
10.1 Partitioning of the system of equations . . . . . . . . . . . . . . . . . . . . .
10.2 Method of continuous elastic foundation . . . . . . . . . . . . . . . . . . . .
10.3 Substructure technique in structural optimization . . . . . . . . . . . . . . .

85
86
88
89

9.4

9.3.2
9.3.3
Vector
9.4.1
9.4.2

Gauss-Seidel iteration . . . . . .
Method of conjugate gradients .
iteration for eigenvalue problems
Vector iteration . . . . . . . . . .
Matrix deflation . . . . . . . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

11 Post-processing
91
11.1 Calculation of strains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
12 Reduced modelling techniques
12.1 Periodic systems . . . . . . . . . . . . . . . . . . . . .
12.1.1 Sandwich structure with honeycomb core . . .
12.2 Generalized plane-strain state . . . . . . . . . . . . . .
12.2.1 Multidirectional FRP laminate test simulation
12.3 Corrugated sheets . . . . . . . . . . . . . . . . . . . .
13 Sequential processes
13.1 Laminated structures damage progression simulation .
13.1.1 Damage modeling . . . . . . . . . . . . . . . .
13.1.2 Damage states and displacement fields . . . . .
13.1.3 Preconditioning of the system of equations . .
13.1.4 Notched tensile test CFRP specimen model . .
13.1.5 Simulated damage states . . . . . . . . . . . . .
13.1.6 Conjugated gradient method solution efficiency

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

14 Finite beam elements


14.1 Equilibrium at a small beam element . . . . . . . . . . . . . . . . .
14.2 Kinematics of bending and shear . . . . . . . . . . . . . . . . . . .
14.3 Integrated constitutive laws . . . . . . . . . . . . . . . . . . . . . .
14.4 Shear-stress distribution and shear correction factor . . . . . . . .
14.5 Differential equations . . . . . . . . . . . . . . . . . . . . . . . . . .
14.6 Shear-rigid beam finite element . . . . . . . . . . . . . . . . . . . .
14.6.1 Weak variational form of the shear-rigid beam problem . .
14.6.2 Boundary terms interpretation . . . . . . . . . . . . . . . .
14.6.3 Conclusions for the discretization step . . . . . . . . . . . .
14.6.4 Polynomial complexity and Hermite shape functions . . . .
14.7 Shear-compliant beam finite element . . . . . . . . . . . . . . . . .
14.7.1 Penalty method and shear-compliant beam weak variational
14.7.2 Shear-compliant beam weak variational form discretization
14.7.3 Shear-compliant beam finite element locking effect . . . . .
14.7.4 Locking effect mitigation by reduced integration . . . . . .

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
form
. . .
. . .
. . .

95
96
97
99
101
102

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

107
. 108
. 108
. 109
. 110
. 110
. 111
. 112

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

115
. 115
. 115
. 116
. 116
. 119
. 119
. 119
. 120
. 120
. 120
. 123
. 123
. 124
. 125
. 125

15 Eigenvalue problems
15.1 Harmonic Vibrations . . . . . . . . . . . . . . . . . . . . . . . . . .
15.1.1 Equations of motion . . . . . . . . . . . . . . . . . . . . . .
15.1.2 Weak variational form and system of equations . . . . . . .
15.1.3 Relation between amplitudes and accelerations . . . . . . .
15.1.4 Further aspects of solving an harmonic vibrations problem .
15.2 Instability and FEM . . . . . . . . . . . . . . . . . . . . . . . . . .
15.3 Critical buckling load and natural frequency . . . . . . . . . . . . .

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

129
129
129
129
129
130
132
134

A Equations of linear elasticity and plane problems


A.1 Basic problem of linear elasticity . . . . . . . . . . . . . .
A.1.1 Equilibrium conditions . . . . . . . . . . . . . . . .
A.1.2 Kinematic or strain displacement relations . . . . .
A.1.3 Generalized Hookes law . . . . . . . . . . . . . . .
A.1.4 Temperature and moisture strains . . . . . . . . .
A.1.5 Displacement formulation of a solid-body problem
A.1.6 Matrix-operator notation . . . . . . . . . . . . . .
A.2 Two-dimensional Problems . . . . . . . . . . . . . . . . .
A.2.1 Plane elasticity problems . . . . . . . . . . . . . .
A.2.2 Axisymmetry . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

143
143
143
144
146
146
147
147
149
149
151

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

Chapter 1

Goals and contents


1.1

Significance of the Finite-Element Method

The Finite-Element Method (FEM) is the basis for powerful analysis tools for solving
continuum mechanical problems. It finds approximations to true Solutions which are
not accessible with analytical methods. With the support of ever increasing computer
resources FEM is routinely used in contemporary engineering practice. The available
commercial programs allow performing complex analyzes in small amounts of time. Their
user interfaces become increasingly translucent, convenient, and, at least for the user with
the necessary theoretical background, self-explanatory.

1.2

Educational objective

Users of FEM software should have a thorough understanding of the theory, the mediation
of which is the objective of this course. Application of the theory to practical problems
by use of commercial software completes the theoretical understanding and is addressed
by some of the exercises. Knowledge of the FEM include the awareness of its limitations
and the risk of obtaining incorrect results through modeling errors. The lecture class
imparts all building blocks of understanding which are necessary to develop dedicated
finite elements for solving particular problems and write a FEM program.

1.3

Frame

The lecture class Structural Analysis with FEM considers static and harmonic problems
within linear elasticity. The simulation of progressing damage in a notched test specimen
made of laminated composite materials is the one nonlinear problem considered here.

c ETH Z

urich IMES-ST, March 26, 2014

Goals and contents

1.4

Learning matter

The learning matter comprises all significant aspects of a structural analysis with FEM:

Basic principle of FEM


what is a finite element and what properties does it possess
mapping of mechanical basic equations onto a finite element
limits of element mapping capabilities and element technology
assembly of the system equations from the element contributions
boundary conditions and degree-of-freedom coupling
direct and iterative methods for solving the system of equations
substructuring
postprocessing with evaluation of the secondary solution
reduced modeling techniques
nonlinear damage progression analysis
beam elements and locking effect
eigenvalue problems and vector iteration solution method
modeling strategies

The contents up to and including post processing are mirrored in the in Fig. 1.1 sketched
V o r la u fr e c h n u n g

D a te n e in le s e n
G e o m e tr ie e r s te lle n
G e o m e tr ie v e r n e tz e n
E le m e n tg le ic h u n g e n
S y s te m g le ic h u n g e n
R a n d b e d in g u n g e n

H a u p tre c h n u n g
N a c h la u fr e c h n u n g

G le ic h u n g s s y s te m

l s e n

E le m e n tv e r z e r r u n g e n
S p a n n u n g e n
F e s tig k e its k r ite r ie n
V is u a lis ie r u n g

Figure 1.1: Steps of an FEM analysis


sequence of a FEM program run. The combination of the learning matters with some
knowledge of any programming environment allows the students to write their own FEM
programs. Appendix A contains basic equations of the theory of linear elasticity and in
Appendix B one finds the simplified equations for structural elements such as truss, beam,
or plate. The following sections describe the contents in more detail.

1.4.1

What is a finite element and which properties does it possess?

A finite element provides a linear relation between nodal forces and nodal displacements,
where a finite-element mesh can be understood to be analogous to a spring, only that the
FEM mesh involves many degrees-of-freedom instead of just one. As an introductory piece
of learning material, the derivation of the element equations of the structural elements truss
and beam are shown with the direct method explained in Chapter 2.
c ETH Z

urich IMES-ST, March 26, 2014

1.4 Learning matter

1.4.2

FEM standard notation for linear elastic problems

Chapter 3 recalls the basic problem of linear elasticity, explains the principle of the minimum of the total potential energy, derives the weak variational form, shows the equivalency of the principle with solving the basic problem by obtaining the Euler equations,
explains the discretization step with element-wise defined approximation of the true solution, and explains the necessary coordinate transformation for referring the normalized
master-element coordinates to the Cartesian system. Isoparametric elements are defined
and the proper use of boundary conditions is also explained. The FEM schematic in standard notation, derived for three-dimensional problems, is applied to simplified mechanical
situations such as the truss in Chapter 4 and simple plane problems in Chapter 5.

1.4.3

The search for most efficient finite element formulations

Simple elements with nodes only on element corners are afflicted with limitations of their
mapping abilities. The more complex elements with higher-order shape functions are more
powerful in terms of being able to geometrically approximate curved boundaries and map
more complex stress distributions on the penalty of requiring significantly more numerical
effort in the pre- as well as the postprocessing steps. Section 6 Element Technology analyzes
L in ie n e le m e n t
lin e a r o d e r k u b is c h

F l c h e n e le m e n t
3 K n o te n , lin e a r

F l c h e n e le m e n t
4 K n o te n , lin e a r

F l c h e n e le m e n t
6 K n o te n , q u a d r a tis c h

F l c h e n e le m e n t
8 K n o te n , q u a d r a tis c h

V o lu m e n e le m e n t
4 K n o te n , lin e a r

V o lu m e n e le m e n t
8 K n o te n , lin e a r

V o lu m e n e le m e n t
1 0 K n o te n , q u a d r a tis c h

V o lu m e n e le m e n t
2 0 K n o te n , q u a d r a tis c h

Figure 1.2: Elements of various spatial and shape-function complexities


the limitations, of simple elements regarding certain mechanical situations such as planestrain deformation of nearly incompressible materials or bending of thin structures, and
explains mitigation of the problems so that the simple elements can perform nearly as well
as higher-order elements with more degrees-of-freedom.

1.4.4

Numerical preprocessing

The automated preprocessing steps explained in Chapter 7 include numerical integration


by Gauss quadrature, where the source of possible integration errors is also illuminated,
and the assembly of the system equations. The implementation of degree-of-freedom coupling and essential boundary conditions into the assembled system of equations is explained
in Chapter 8.
c ETH Z

urich IMES-ST, March 26, 2014

1.4.5

Goals and contents

Direct and iterative methods for solving a system of equations

Although not at the heart of the FEM theory, the solution methods are interesting in
context with structural analysis with FEM. A few of the important concepts are presented
in Chapter 9 and it is outlined how their numerical effort depends on size and population
structure of the stiffness matrix. The same Chapter also contains the vector iteration
method for solving eigenvalue problems.

1.4.6

Substructuring

Substructuring can improve computational efficiency if only a small part of the total
problem domain is subject to an iterative design optimization process, but the elastic
coupling with the other part of the domain has a significant influence on the stressing of the
first part. The degrees-of-freedom of the other part are eliminated to give a modified system
of equations with less degrees-of-freedom. Chapter 10 explains two different schemes for
substructuring. The problem, that the reduced system may have a higher bandwidth than
the original one, is also addressed.

1.4.7

Post-processing with evaluation of the secondary solution

The student learns how the strains are calculated on element level from the nodal-point
displacements and that they agree best with a proposed true solution at certain points
within the element. A scheme for finding the optimum evaluation points is explained in
Chapter 11.

1.4.8

Reduced modeling techniques

A number of modeling simplifications are well-known. These include plane states or use
of structural elements such as truss, beam, plate or shell. Chapter 12 introduces unit-cell
modeling of periodic systems and generalized plane-strain situations.

1.4.9

Nonlinear damage progression analysis

The topic, explored in [2] and presented in Chapter 13, takes up several other learning
matters of the lecture class and places these in context with each other. In addition, the
backdrop of damage progression analysis builds a bridge to other lecture classes of the
major Advanced Structures and Aerospace Engineering, namely Mechanics of Composite
Materials and Manufacturing of Polymer Composites. The Chapter provides an understanding of how a suitable conditioning of the system equations and well-chosen starting
vectors can make iterative solution methods superior to direct ones.

1.4.10

Beam finite elements and locking effect

Even though depending on only one spacial variable, there is much material in the topic of
beam finite elements and problems need to be discussed. Chapter 14 explains systematic
derivation of shear-rigid and shear-deformable beam finite elements by using the displacement formulation. The latter offers the opportunity to understand the shear-locking effect
and its mitigation by using a reduced technique for integration of the shear stiffness matrix.
c ETH Z

urich IMES-ST, March 26, 2014

1.5 Acknowledgment

1.4.11

Eigenvalue problems

Harmonic vibrations as well as instability problems appear quite often in practice and can
often be treated as eigenvalue problems. Chapter 15 explains the respective FEM formulations and how the eigenvalue problems arise. By using the sample of truss instability
the influence of longitudinal load on eigen frequency is illuminated. The solution method
by use of vector iteration is explained in Chapter 9.4.

1.5

Acknowledgment

The larger parts of the present script have been translated into English from the previous
German version. I would like to thank Prof. P. Ermanni for giving me the time-wise
opportunity for the work. Dr. M. Jabareen, now a staff member of the Faculty of Civil
and Environmental Engineering at the Technion in Haifa, held in 2008 a lecture on element
technology which motivated me to include, with his agreement, the topic in this lecture
class. The topic of reduced modeling techniques in Chapter 12 is new and appears first in
the present English version of the script.

Zurich, March 2014

c ETH Z

urich IMES-ST, March 26, 2014

c ETH Z

urich IMES-ST, March 26, 2014

Goals and contents

Chapter 2

Direct element equation derivation


The finite-element method converts a given continuum-mechanics problem with domain
and boundary conditions into a numerical system of equations. Each finite element contributes a number of equations to the system. The element equations are quite systematically and elegantly constructed by using the variational methods explained in textbooks,
for instance [3]. For beginners not yet familiar with variational methods, it may be a helpful preliminary learning step to see how the finite element equations for simple structural
members, such as truss or beam, can alternatively be derived by using strength-of-materials
approaches, which are sometimes called direct method.
Before one starts with developing element equations for a truss or a beam one should be
able to answer the questions:
1. What is a truss?
2. What is a beam?
3. What is a finite element?
The definitions of the structural members are recalled in the following Sections 2.1 and
2.2, respectively.
For the third question we note that a finite element features an element domain, boundary, and nodes. The displacement formulation of linear solid mechanics (see Appendix
A.1) seeks the solution first in terms of the displacements. Discrete nodal-point displacements then take the role of primary solution, or degrees-of-freedom. The nodal points also
accept external forces or transmit inner forces. A finite element creates a linear system
containing the
of equations of the form K
u = r with stiffness matrix K, the vector u
displacement degrees-of-freedom, and a right-hand-side vector r containing the forces responding to body forces, boundary tractions, or concentrated loads. A finite element is
the more complex analog to the spring model.
Fig. illustrates that a truss finite element has a line domain and its boundary consists of
its two end points.

2.1

The truss and its element equations

The idealized truss transmits forces only which are aligned with its longitudinal directions. However, even very complex structures such as the Crymlin viaduct at Gwent,
South Wales, shown in in Fig. 2.1 can be assembled form real structural elements as the
detail in Fig. 2.2 testifies. It can be seen from the figure that, after overcoming a certain
c ETH Z

urich IMES-ST, March 26, 2014

Direct element equation derivation

Figure 2.1: The Crymlin viaduct at Gwent as a truss frame

Figure 2.2: Trusses and connections


friction resistance, rotation of the members against each other is possible. In idealization
the ends of the members can freely rotate in one plane wherefore the part of the structure
can be treated mathematically as a two-dimensional truss frame.
Fig. 2.3 summarizes the mechanical equations describing a truss. They are simplificap
L

c o n v e n tio n s

N , u

s ,s+ f = 0

e q u ilib r iu m
H o o k e 's la w

s = E e

k in e m a tic s

e = u ,s

d iffe r e n tia l e q u a tio n

E A u , ss + p = 0 , p = A f

Figure 2.3: Structural member truss


tions of the general basic equations of linear elasticity for the one-dimensional case. The
displacement differential equation u is already integrated over the cross-sectional area by
simply multiplying Youngs modulus E and the body force f with the cross-sectional area
A, which presupposes that A remains constant along the truss coordinate s.
c ETH Z

urich IMES-ST, March 26, 2014

2.1 The truss and its element equations

2.1.1

Direct derivation of the truss finite element equations

At the top of Fig. 2.4 the truss domain with length L is depicted. If variable loads or
L

L
1

1
1

2
2

e 1

L
2

,
L

3
3

u
e

e 1

e 2

e 2

Figure 2.4: Truss and FEM model: Domain (top), subdivision into finite elements (center),
Element (bottom)
properties along the domain must be considered, the latter can be subdivided into several
finite elements of respective lengths Le . The sample appearing in the figure is a FE model
with three elements, four nodes, and an equal number of degrees-of-freedom.
The element equations for a truss element are now derived by mapping the definition of
the mechanical truss model onto that of the truss finite element. That is achieved with

e 1

= N = E A /L
1

~
e u = N ,
K

u
N

e 1

e = - 1 /L

= 1

N = -E A /L

u
e

N
e

K
e

= 0
= -N = -E A /L

e ,

E A 1 - 1
,
L e - 1 1
e = 1 /L

e ,

N
N

N =
u

N
e

= 0

= N = -E A /L
2

u~
u~ = ~ 1 ,
u 2

N = E A /L

e 2

e 2

1
2

= 1

= N = E A /L
e

Figure 2.5: Mapping of the truss onto a finite element


performing the theoretical stiffness measurements illustrated in Fig. 2.5. According to the
sketch at the top the displacement at node 2 is at first specified to zero, ue2 = 0. The
other node undergoes a forced unit displacement ue1 = 1 auf. At constant cross-sectional
stiffness EA this obtains a constant and compressive strain  = 1/Le . The integrated
constitutive law then gives the constant compressive internal force N = EA/Le , which
corresponds with a positive nodal force N1 = EA/Le , because the latter must act in the
positive direction (to the right) to effect the positive displacement ue1 = 1:
N1 =

EA
u1 .
L

(2.1)
c ETH Z

urich IMES-ST, March 26, 2014

10

Direct element equation derivation

At the constrained node one finds the reaction force N2 = N1 :


N2 = N1 =

EA
u1 .
L

(2.2)

Next, we constrain node 1 and impose at node 2 the unit displacement oriented in positive
direction, conduct the analogue thought experiment and receive from that the equations
EA
u2
L

N2 =

(2.3)

and
N1 = N2 =

EA
u2 .
L

(2.4)

The equations of the truss finite element are given in matrix form on the left at the center
region of Fig. 2.5. Further to the right the stiffness matrix K, the unknown solution vector
, and the right-hand-side with nodal forces are identified. The tilde symbol distinguishes
u
the discrete nodal point displacements from the continuous displacement field. It remains
to implement the line force p, illustrated in Fig. 2.6, into the element equations. If p is
F = p L
e

p 1= p L e/2

p L
e

=
p

p 2= p L e/2

+
p

p L
2
=

1

1

Figure 2.6: Line loads and Nodal forces of the truss element
assumed constant over the element length L, it integrates to the resulting force F = pLe
which is obviously symmetrically, which means in equal parts, distributed on the the two
element nodes.
This completes the element equations of a truss finite element which are presented in Fig.
2.7. It is important to understand why the stiffness increases with increasing Youngss

E A 1

L e - 1

- 1 u~ 1
=

1 u ~ 2

1
2

p L
2
e

1

1

lin e lo a d
b o u n d a r y c o n d itio n s
n o d e d is p la c e m e n ts
s tiffn e s s m a tr ix

( s y m m e tr ic , n o t in v e r tib le )

Figure 2.7: Truss finite element equations


c ETH Z

urich IMES-ST, March 26, 2014

2.2 The beam and its element equations

11

modulus E and cross-sectional area A and decreasing length L. Another point of understanding is why the stiffness matrix is singular and thus not invertible so that the equations
in the present form can not be resolved. The the right-hand-side is composed of the nodal
forces N, which can stem from externally applied concentrated forces, and the contribution
of the line force p. The equations shown in 2.7 should be present as a preparation for the
oral exam. Finally it is interesting to note that the here applied direct method for deriving
element equations does without knowledge of the displacement differential equation given
in Fig. 2.3.

2.2

The beam and its element equations

The straight beam is a simple structural element. By definition, it carries force only
transverse to its longitudinal direction and the force evokes bending moment. Thus, the
beam finite element must transfer through its nodal poins transverse force Q as well as
bending moment M . They correspond with the deflection w and rotation . One speaks
of conjugate pairs Q, w and M, . Together they participate in the work done by external
forces and the relation between generalized forces and displacements must be defined via
a stiffness matrix in the same way as for the truss and as indicated in Fig. 2.8. The
q

Q
M

Q
M

M , b
L
y

Q , w

E Iw
z

iv

- q = 0

?
?

?
?

2
2

?
?

w 1
b 1
w
2

b
2

Figure 2.8: Beam element conventions and generalized displacements and forces relations
structure of the relation anticipates that the beam finite element, just as the truss finite
element, possesses only two nodes at its ends. It is further assumed that the beam is very
thin compared if to its length, which justifies neglect of shear deformation so that the
rotation is the negative of the spatial derivative of the deflection line w0 or = w0 .
For finding the displacement-force relations one proceeds analogous to the truss, as a unit
displacement is applied to one of the four degrees-of-freedom at a time while the remaining
three are set equal to zero. Then the forces and moments are found on all nodes. For this,
one begins with the polynomial form and its spatial derivatives for the deflection line w
for a shear-rigid beam element,
w(x)
w0
w00
w000
w0000

=
a0
=
a1
= 2a2
= 6a3
= 24a4

+
a1 x +
a2 x2 + a3 x3 + a4 x4
+ 2a2 x + 3a3 x2 + 4a4 x3
+ 6a3 x + 12a4 x2
,
+ 24a4 x

(2.5)

were the free coefficients must be fit to satisfy the respective boundary conditions illustrated in Fig. 2.9. The coefficient a4 follows from satisfying the differential equation
imparted with Fig. 2.8 and remains the same for all four situations:
EI24a4 q = 0

a4 =

q
.
24EI

(2.6)

c ETH Z

urich IMES-ST, March 26, 2014

12

Direct element equation derivation

k
k

k
k

1 1

2 1

k
k

1 1

2 1

k
k

4 2

1 2

k
k

k
k
k

4 3

1 3

k
k

3 4
4 4

1 4
2 4

3 3
4 3

1 4
2 4

3 3

2 3

3 2
4 2

1 3
2 3

3 2

2 2

3 1
4 1

2 2

3 1
4 1

1 2

3 4
4 4

k
k

=
k
k

1
0

0

0

0
0

1

0

1 1

2 1

k
k

=
k
k

3 1
4 1

1 3

2 3

3 3
4 3

k
k

k
k

k
2

1 1

4 2

1 1

4 3

k
k

2 2

3 1

4 2

3 4

4 4

1 3

2 3

3 2

2 4

3 3

1 2

2 1

4 1

1 4

2 3

3 2

1 3

2 2

3 1
4 1

1 2

2 1

2 4

3 3
4 3

1 4

3 4
4 4

0
1

0

0

0
0

0

1

k
k

=
k
k

1 2

2 2

k
k

=
k
k

3 2
4 2

1 4

2 4

3 4
4 4

Figure 2.9: Beam boundary conditions


However, for determining the stiffness matrix form the support forces stemming from the
geometric boundary conditions, of which always one is inhomogeneous, the line load q is
set equal to zero. Thus a4 vanishes as well. The general relations between deformations
and inner forces of the beam are:
M = EI = EIw00 ,

Q = EIw000 .

(2.7)

For finding the stiffness matrix entries the sign conventions for internal forces and for nodal
forces must be considered:
Q1 = Q(0) = EIw000 (0)
M1 = M (0) = EIw00 (0)

Q2 = Q(L) = EIw000 (L)


.
M2 = M (L) = EIw00 (L)

(2.8)

The situation 1, depicted at the upper left in Fig. 2.9, determines the first column of the
stiffness matrix:
w(0)
w0 (0)
w(L)
w0 (L)

=
=
=
=

1
0
0
0

a0

a1

a2 L2 + a3 L3

2a2 L + 3a3 L2

=
1
=
0
.
= 1
=
0

(2.9)

Resolution of the latter two equations yields a2 = 3/L2 und a3 = 2/L3 . With this the
loading situation 1 creates the deflection line and its spatial derivatives:
x2
x3
x
x2
6
x
12
0
+2
,
w
(x)
=
6
+6
, w00 (x) = 2 +12 3 , w000 (x) = 3 . (2.10)
L2
L3
L2
L3
L
L
L
With reference to the definitions (2.8) the deformations at the nodal points,
w(x) = 13

6
12
6
, w000 (0) = 3 , w00 (L) = 2 ,
2
L
L
L
deliver the first column of the stiffness matrix K:

2EI 3L
Ki1 = 3
.
6
L

3L
w00 (0) =

w000 (0) =

12
,
L3

(2.11)

(2.12)

Carrying out the analogous steps for the other three situations depicted in Fig. 2.9, one
receives the remaining columns of the stiffness matrix

6 3L 6 3L
2EI 3L 2L2 3L
L2
.
K= 3
(2.13)
3L
6
3L
L 6
3L
L2
3L 2L2
c ETH Z

urich IMES-ST, March 26, 2014

2.3 Summary direct methods

13

The effect of a homogeneous line load q = const, acting on the beam element, must be
captured with so-called kinematically equivalent nodal forces Q and -moments M . For the
analysis one considers the beam element clamped at both ends and first finds the deflection
line coefficients due to q,
w(0)
w0 (0)
w(L)
w0 (L)

=
=
=
=

0
0
0
0

a0

a1

a2 L2 + a3 L3

2a2 L + 3a3 L2

=
=
=
=

0
0
.
qL4
24EI
qL3
6EI

(2.14)

From (2.14) one obtains a0 = a1 = 0, a2 = qL2 /24EI, and a3 = qL/12EI. With these
results one finds from (2.5) the deformations at the ends x = 0 and x = L which must
then be inserted into (2.9) to find the support reactions caused by the line load q. The
support reactions are the desired nodal forces and moments equivalent to q. The complete
element equations are given in Fig. 2.10.

6 - 3 L
2 E I - 3 L 2 L

3 L
L 3e - 6
L
- 3 L
2

- 6

- 3 L
L
3 L
6
3 L
3 L 2 L
2

w
b

w
b

1
2
2

Q
M
= Q

q L e - L

1 2 6
L

+
1

2
2

lin e lo a d
b o u n d a r y c o n d itio n s
n o d e d is p la c e m e n ts
s tiffn e s s m a tr ix

( s y m m e tr ic , n o t in v e r tib le )

Figure 2.10: Element equations of the shear-rigid beam

2.3

Summary direct methods

The chapter corresponds with the first lecture class and the most important learning goal
is the understanding of the term finite element by studying the examples of the truss and
the shear-rigid beam.
The element equations could be derived by using the so-called direct method. It can
be easily comprehended without having to first understand new concepts, but it is less
suited for finding the equations of more complex finite elements. The much more powerful
method for deriving finite elements, based on the variational method, is introduced in the
following Chapter 3 and the applied to the truss and plane situation in Chapters 4 and 5,
respectively.

c ETH Z

urich IMES-ST, March 26, 2014

14

c ETH Z

urich IMES-ST, March 26, 2014

Direct element equation derivation

Chapter 3

FEM theory in standard notation


3.1

Concept of deriving a numerical system of equations

This chapter presents FEM theory for linear-elastic static problems in general terms, i.e.
independent of a specific element, and standard notation. It starts with Section 3.2 from
the basic problem of linear elasticity, derives with Section 3.3 the weak variational form
based on the principle of the minimum of the total potential energy, introduces in Section
3.4 with the finite element mesh the discretization step, and establishes the linear system
of equations with a symmetric stiffness matrix according to the Bubnov-Galerkin method
[4]. Programming of the method requires transformations between the model reference
and the local element coordinates by use of Jacobi matrices; the transformations concern
displacement derivatives as well as domain and boundary differentials. The realization of
natural boundary conditions is outlined in Section 3.5 and a scheme for finding element
shape functions is explained in general terms and on a tri-linear volume element in Section
3.6.

3.2

Basic problem of linear elasticity

In order to render the following equations and derivation steps concise, the summary of
elasticity basic equations is taken as an opportunity to introduce with Fig. 3.1 the Voight
te n s o r n o ta tio n

e q u ilib r iu m

k in e m a tic s

d iffe r e n tia l e q u a tio n

m a te r ia l la w
1
2

ijk l

(u

, j+ fi = 0

ij

ij

ij

m a tr ix - o p e r a to r n o ta tio n

(u
= C

, j+ u j,i
ijk l

k l

, lj + u l , k j ) + f i = 0

s + f= 0
e = L u

s = C e
L T C L u + f = 0

Figure 3.1: Basic equations of elasticity in tensor and in matrix-operator notations


notation and the differential operator L. If the stress tensor is symmetric, which is the
case in the absence of body moments, stresses as well as strains  can be written as
one-dimensional arrays and the constitutive law as a two-dimensional array. On the other
c ETH Z

urich IMES-ST, March 26, 2014

16

FEM theory in standard notation

u x

u = u y ,
u
z

e
e

e
g
g

e =

s
s

c
c

zx

s
t
t

zx

x y

x y

y z

s =

C =

y z

1 1

2 1

c
c

5 2

4 4

5 3

1 5

2 5

3 4

6 3

1 4
2 4

4 3

6 2

3 3

4 2

1 3
2 3

3 2

5 1
6 1

2 2

3 1
4 1

1 2

5 4

5 5

6 4

6 5

L =

5 6

6 6

4 6

c
c

3 6

4 5

2 6

3 5

1 6

0
0

Figure 3.2: Symbols of matrix-operator notation


hand the definition of the linear differential operator L in Fig. 3.2 allows dropping of
explicit index writing.
An elasticity problem requires solution of the basic equations within the problem domain
and displacements u
specified
, sketched in Fig. 3.3, and under the external stresses
on its boundary . There may also be non-zero body forces f . An exact solution of such a
z

d o m a in

b o u n d a ry
n o r m a l u n it v e c to r

d W
n

b o d y fo rc e

e x te rn a l s tre s s

s
d G

Figure 3.3: General solid-body problem


solid-body problem can be achieved only where domain geometry and boundary conditions
are simple. Engineering practice must generally rely on approximate solutions. The FEM
delivers such solutions which can, depending on the numerical effort, come very close to
true solutions. The power of the method comes from subdividing even complex domains
into many smaller low-complexity sub-domains, the finite elements.

3.3

Derivation of the weak variational form

Energy methods facilitate finding of solution approximations by replacing the requirement


to satisfy the basic equations exactly at every point of the domain by the weaker
requirement of satisfying the equations in an average sense. Thereby the mean deviation
from a true solution over the domain, or residual, is minimum and characteristic for the
approximation quality.
The displacement formulation of linear elasticity problems allows finding of approximations
by minimizing, with respect to the displacement solution parameters, the total potential
energy = min:
Z
Z
Z
Z
1
T
T
T
ud
d ,
=U W =
 d
f ud

T u
(3.1)
2

u
where U is the deformation energy and W the work done by body forces f and the exter and displacements u
together with the corresponding unknown conjugate
nal stresses
c ETH Z

urich IMES-ST, March 26, 2014

3.3 Derivation of the weak variational form

17

symbolizes a stress vector


quantities on the boundary . Thereby it must be noted that
with components
x ,
y , and
z along the spatial directions and that these components
must not be confused with those of the double indexed stress tensor, which appear with
single indices in the Voight notation.
Principle of the minimum of the total potential energy: If the admissible displacement fields u satisfy all basic equations, is a minimum. If u does not include the true
solution, is a minimum for the best approximation. At the minimum point the variation
of with respect to u and Lu, i.e. its derivative with respect to the solution parameters
of u, vanishes.
For beginners a preliminary understanding of the principle is facilitated by considering the
discrete numerical system:
1 T
K
T r
=U W = u
uu
2

(3.2)

The total potential energy is a scalar (number) which depends on the solution parameters
. is minimum where its derivative with respect to the solution parameters u
vanishes:
u

= = K
ur=0

(3.3)

is called variation . Fig. 3.4


More generally, the derivative of with respect to u
illustrates for a one-dimensional problem that the minimum of the total potential energy
U

u~

u~
0

Figure 3.4: Total potential energy


is negative. Trivially, both the deformation energy and the external work vanish with the
displacement. With increasing absolute displacement values the external work and the
deformation energy increase linearly and quadratically, respectively. The best solution
coincides with the minimum of the total potential energy (3.1).
In order to develop from the basic equations, Fig. 3.3, a finite element by the displacement
formulation, one writes the functional (3.1) in terms of displacements and displacement
derivatives by eliminating the strains and stresses:
Z
Z
Z
Z

1
T T
T
T

T d. (3.4)
u d
=U W =
u L C (Lu) d
u f d
u
2

u
Now a variation of (3.4) analog to (3.3) must be taken but presently there are no solution
parameters of the continuous displacement fields u defined. The problem is mitigated by
c ETH Z

urich IMES-ST, March 26, 2014

18

FEM theory in standard notation

assuming that any candidate solution u may deviate from the unknown best solution u0
by a variation u:
u = u0 + u = u0 + v.
(3.5)
The variation u is the product of an arbitrary test function v and a small parameter :
u = v.

(3.6)

The test function is an arbitrary displacement field which must be non-zero in the whole
are specified. Now the
domain except for the boundary u on which displacements u
variation of the functional is represented by a simple derivative with respect to the one
parameter :
Z
Z
Z

d

=
=
(3.7)
vT LT CL (u0 + v) d
vT f d
vT d.
d

The work done by the specified displacements vanishes because these can not be varied.
The multiplication with allows writing the result by using the variation symbol u.
Because of
(u, Lu) (u0 , Lu0 )
(3.8)
the value of must be small as to obtain, in the limit 0, the solution u0 by letting
the variation vanish:

Z
Z
Z

d
T T
T

=
=
u L CLu0 d
u f d
uT d
=0
(3.9)
d

The operation demonstrated with (3.7) and (3.9) is called Gateaux derivative [5] of ,


d
(u0 , Lu0 ; v, Lv) =
,
(3.10)
(u0 + v, Lu0 + Lv)
d
=0
and the result (3.9) is the so-called weak variational form which is so central for the
finite-element method. Another often used notation for the weak variational form is




(u, u) = u
(3.11)
= Lu
=0
u
Lu
u0

u0

A transformation of the weak variational form verifies the principle of the minimum of the
total potential energy. Partial integration is applied to the total deformation energy so
that the derivative of the displacement variation (or virtual displacement) vanishes:
Z
Z
Z



T T
T
T
u L C (Lu) d =
u N CLu d
uT LT CLu d.
(3.12)

The matrix N arises from the divergence theorem [1] and contains the components of the
on the boundary with the same pattern as the the differentiations
unit normal vector n
in the linear operator L,

nx 0 0 0 nz ny
NT = 0 ny 0 nz 0 nx .
(3.13)
0 0 nz ny nx 0
It establishes the relation, depicted in Fig. 3.5, between the components of the stress
c ETH Z

urich IMES-ST, March 26, 2014

3.4 Subdividing the problem domain into finite elements

19

z
C

d z
t
s

s
i

y x

y y

y z

d x

B
d y

x
Figure 3.5: Infinitesimal tetrahedral element for visualization of the force equilibrium at
the boundary
tensor = CLu and those of the stress vector
. Thus one identifies
Z
Z

T
T

u N CLu d =
uT d

(3.14)

and substituting (3.12) for the deformation energy term in (3.8) obtains
Z

uT LT CLu + f d = 0

(3.15)

The expression in parentheses, called Euler equation, is identical with the displacement
differential equation which combines all equations of the the basic elasticity problem, and
thus verifies the principle of the minimum of the total potential energy. The equation also
states the principal of virtual displacements, as the equilibrium of forces at an infinitesimal
volume element is multiplied with virtual displacements and the result is integrated over
the domain and required to vanish. However, this form of variational form is not well suited
for deriving finite elements since the unknown displacement solution must be differentiated
twice and the boundary terms are lacking. Here, the meaning of the name weak variational
form becomes apparent: The requirement on the continuous differentiability of the solution
function is weakened. The principle of virtual displacements (3.15) recovers the weak
variational form (3.9) by applying the identity (3.12). The scheme shown in Fig. 3.6
illustrates how the energy methods are interrelated. It also contains the principle of
virtual work :
Z
Z
Z
T T
T

=
 d
u f d
uT d
=0
(3.16)

The principal of virtual work leads to the weak variational form in terms of displacements
by expressing the stresses through strains via Hookes law and be expressing the strains
through the displacements via the kinematic relations.

3.4

Subdividing the problem domain into finite elements

The finite element method subdivides the problem domain of arbitrary geometric complexity into a mesh of smaller sub-domains of low complexity, called finite elements; hence the
name of the method. The mesh generally contains finite elements of different shapes, orientations, and sizes which are all referred to a so-called master element. Shape functions
c ETH Z

urich IMES-ST, March 26, 2014

20

FEM theory in standard notation

p r in c ip le o f th e m in im u m o f
th e to ta l p o te n tia l e n e r g y

p r in c ip le o f
v ir tu a l d is p la c e m e n ts

v a r ia tio n o f th e fu n c tio n a l

in te g r a tio n b y p a r ts

F E M d is c r e tis a tio n :
n u m e r ic a l e q u a tio n s

w e a k v a r ia tio n a l fo r m

p r in c ip le o f v ir tu a l w o r k

Figure 3.6: Summary energy methods


within the master element domain depend on local element coordinates , , . The
shape functions are normalized to assume a unit value on an associated nodal point and
vanish on all other nodal points. Thus, element-wise displacement field approximations
u (, , ) are obtained by weighting the element shape functions (, , ) with respective
:
nodal-point discrete displacement values u

u = T u

(3.17)

Parameterizing the virtual displacements u with the same shape functions as the unknown
displacement solution u, namely
,
u = T u

(3.18)

obtains the Bubnov-Galerkin method [4]. Since adjacent finite elements are connected
with each other through common nodes, the element displacement fields are continuous at
all nodes but their spatial derivatives are generally not continuous. This kind of problemdomain displacement approximation is called C 0 continuous. The finite-element strains
are presented with
= B
 = Lu = LT u
u,

B = B (, , )

(3.19)

Substituting (3.17), (3.18), and (3.19) into the weak variational form (3.9) gives the discretized weak variational form,
"
T
= u

Ne Z
X
k=1

BT CBd u

Z
f d

#

d
=0,

(3.20)

where the integral over the problem domain must be broken down into the sum of the
integrals over the element domains since the displacements are approximated elementwise. Since the virtual displacement degrees-of-freedom are arbitrary, the contents of the
brackets must vanish to satisfy the minimality condition. After evaluation of the element
integrals one obtains the numerical system of linear equations:

K
u = r;

K=

Ne Z
X
k=1

c ETH Z

urich IMES-ST, March 26, 2014

B CBd;

r=

Ne Z
X
k=1

Z
f d +

(3.21)

3.4 Subdividing the problem domain into finite elements

3.4.1

21

Element and system coordinates transformation

The linear operator L, prescribing derivatives with respect to the reference coordinates
xyz, is applied to shape functions parameterized with local element coordinates .
The derivatives with respect to the reference coordinates must therefore be expressed with

,x ,x ,x ,
,x
,y
,
= ,y ,y ,y

,z
,z ,z ,z
,

(3.22)

where the matrix contains the entries of the inverse of the Jacobian J,


x, y, z,
(x, y, z)
= x, y, z, .
J=
(, , )
x, y, z,


3.4.2

(3.23)

Domain and boundary differentials

In a Cartesian reference basis ex ey ez the differential d of a three-dimensional domain


is expressed in terms of length differentials with d = dxdydz. For expressing the same
volume differential in terms of the local element length differentials ddd, it is useful to
recollect the geometric interpretation of the scalar triple product of vectors:
d = dV = dxex (dyey dzez )

(3.24)

d
dx
d
dy
=J

d
dz

(3.25)

Because of

the volume differential in terms of the element coordinates is written as:


T

J12 d J13 d
J11 d
J22 d
J21 d
J23 d
d = dV =

J32 d
J33 d
J31 d

(3.26)

Evaluating this form obtains that


Z

|J|ddd

dxdydz =

(3.27)

where it is considered that the master element is a cube with side length 2 and its local
normalized element coordinates originate at its center.
On the domain boundary an area differential d is calculated as the cross-product of two
line differentials ds and dt, namely d = ds dt. The orientation of the line differentials in
the Cartesian reference space follows from the element node positions. Depending on which
of its surfaces lies within the boundary, one of its local coordinates remains a constant.
The corresponding local differential vanishes, which gives one of the following possible
c ETH Z

urich IMES-ST, March 26, 2014

22

FEM theory in standard notation

relations:

J12

d (, ) = J22
J32

J11

d (, ) = J21
J31

J11

d (, ) = J21
J31

3.4.3

J13
J23


J33

J13
J23


J33

J12
J22


J32

1 J12 J13

dd = det 1 J22 J23

1 J32 J33

J11 1 J13

dd = det J21 1 J23

J31 1 J33

J11 J12 1

dd = det J21 J22 1

J31 J32 1

dd

dd

(3.28)

dd

Iso-parametric elements and Jacobian matrix

The geometry of iso-parametric elements is parameterized with the same shape functions
as the unknown displacement solution:
x
ux = T u
T
y
uy = u
z
uz = T u

x = T x
T

y = y

z = T z

(3.29)

y
z
are the element node coordinates. Such elements allow straight-forward comwhere x
putation of the Jacobian matrix entries with:

(, )T y
(, )T z

(, )T x
(, )T y
(, )T z

J = (, )T x
T
T
T
(, ) y
(, ) z

(, ) x

3.5

(3.30)

Boundary conditions and equivalent nodal forces

Fig. 3.7 illustrates that both types of boundary conditions, describing either displacements
or forces, appear in the literature under several synonymic names and can contribute to
external work. The natural boundary conditions are to be implemented with the weak
variational form boundary term whereas the essential boundary conditions must be implemented into the assembled system equations. Note that on the same parts of the
boundary only one of the two boundary condition types can be prescribed; the boundary
falls therefore into the distinct parts u and . Also, no part of the boundary exists
without prescribed boundary conditions: A stress-free surface is a boundary underlying
= 0!
the natural boundary condition
The boundary term of the discretized weak variational form translates surface tractions
into kinematically equivalent nodal forces. Prescribed pressure p always acts along the
unit vector n perpendicular to the boundary wherefore the Cartesian surface-traction
= np. Again, depending on which of the element
components are then calculated with
c ETH Z

urich IMES-ST, March 26, 2014

3.5 Boundary conditions and equivalent nodal forces

23

c o n ju g a te d p a ir s
u

w '

e s s e n tia l b o u n d a r y c o n d itio n s

n a tu r a l b o u n d a r y c o n d itio n s

g e o m
k in e m
V e rs c
D ir ic h

s ta
k in
fo r
N e

e tr is c h e
a tis c h e
h ie b u n g s le t

tic
e tic
c e o r s tre s s
u m a n n

h o m o g e n e o u s a n d in h o m o g e n e o u s

Figure 3.7: Boundary conditions


coordinates remains a constant within the boundary, the unit normal vector is given with:

J11
J21
n (,) = 2 1 2 2
J11 +J21 +J31

J31

J12
J22
n (,) = 2 1 2 2
(3.31)
J12 +J22 +J32

J32

J13
J23
n (,) = 2 1 2 2
J13 +J23 +J33

J33

c ETH Z

urich IMES-ST, March 26, 2014

24

FEM theory in standard notation

3.6

Polynomial element shape functions

The subdivision of the domain into a number of finite elements with chosen shape functions
replaces the original solutions space, having an infinite number of degrees-of-freedom, with
the discrete solutions space with a finite number of degrees-of-freedom. Thus, the true
solution of a problem may not be contained in the numerical model solutions space. This is
not a problem for all practical purposes as long as the numerical model solution converges
with increasing mesh density against the true solution. To guarantee such convergence,
the shape functions must satisfy some conditions:
1. if the node displacements arise from a rigid-body motion, the element shape functions
must not create any strains
2. if the node displacements arise from a homogeneous strain distribution, the element
shape functions must reproduce exactly that strain state
3. strains along boundaries must remain finite
Typical finite element shape functions correspond with a general polynomial having as
many powers as the element has nodes. The shape functions derived from general polynomials must include the constant and linear terms in order to satisfy the general conditions 1
and 2. Moreover, the polynomials must be complete, i.e., all terms up to the highest power
used must be contained in it. The set of shape functions i are generally described with
the sets of coefficients ai weighting the polynomial powers p of the element coordinates;
considering a linear volume finite element with eight nodes it holds that
i () = aTi p() = a1i + a2i + a3i + a4i + a5i + a6i + a7i + a8i ; (3.32)
the corresponding element is defined with Fig. 3.8. and its node coordinates are given in

Figure 3.8: Volume element with 8 nodes


Table 3.1.
node

1
-1
-1
-1

2
1
-1
-1

3
1
1
-1

4
-1
1
-1

5
-1
-1
1

6
1
-1
1

7
1
1
1

8
-1
1
1

Table 3.1: Nodal point coordinates of the element Fig. 3.8

c ETH Z

urich IMES-ST, March 26, 2014

3.6 Polynomial element shape functions

25

Zienkewicz and Taylor [6] explain a scheme for deriving standard element shape functions . Recall with (3.17) that the displacement solution is approximated element-wise
with the shape functions and nodal-point displacements:
= AT p
u = T u
u

(3.33)

Element shape functions i must have a unit value on the respective associated nodes
j = i and vanish on all other respective nodes j 6= i. In order to derive the shape functions
from the local node coordinates and the general polynomial form (3.32) one writes:
I = AP

(3.34)

where the rows of A contain the sets of polynomial coefficients, and the columns of P
contain the polynomial powers at the respective nodes. Inversion yields formally
A = P1 ,

(3.35)

wherefore (3.33) can be written as

u = pT P1 u

= PT p .

(3.36)

The result of the operations for the tri-linear volume element is here given in terms of a
concise writing of the shape functions i :
1 =
3 =
5 =
7 =

1
8
1
8
1
8
1
8

(1 ) (1 ) (1 )

2 =

(1 + ) (1 + ) (1 )

4 =

(1 ) (1 ) (1 + )

6 =

(1 + ) (1 + ) (1 + )

8 =

1
8
1
8
1
8
1
8

(1 + ) (1 ) (1 )
(1 ) (1 + ) (1 )
(1 + ) (1 ) (1 + )

(3.37)

(1 ) (1 + ) (1 + )

Zienkewicz and Taylor [6] also point out that in some cases the inverse of P may not exist
and then the scheme cannot be used.

c ETH Z

urich IMES-ST, March 26, 2014

26

c ETH Z

urich IMES-ST, March 26, 2014

FEM theory in standard notation

Chapter 4

Truss element revisited


The truss element is considered here to have a simple case of using standard notation to
derive a concrete finite element. In addition to the material covered in Chapter 3 the
assembly of the system equations from the element equations and the implementation
of boundary conditions are explained. The truss element is generally used in complex
truss frame structures and therefore the decomposition of the truss equations into components along the directions of the Cartesian reference system must be achieved. If the
one-dimensional truss element is embedded in a three-dimensional space, it can not be considered an isoparametric element, and the transformations are achieved by using another
schematic than by the Jacobian matrix.

4.1

Truss problem weak variational form

An objective of this chapter is to show that application of the standard representation


FEM theory, together with the simplifications done for the truss, quickly leads to the
finite truss element.
The equations of the basic problem of linear elasticity shown in Fig. 3.2 are significantly
simplified:
( )
u = us ,  = , = , C = EA, L = ( )0 =
.
(4.1)
s
The subscript s reminds of the coordinate s along the truss direction. The total potential
energy (3.1) becomes
1
(u, , ) =
2

Z
Ads

L
L

u
upds uN
N
0

(4.2)

Eliminating the strains and stresses obtains the form in terms of displacements only which
is analog to (3.4):
1
(u, u ) =
2
0

u EAu ds
0

L
L

u
N .
upds uN
0

(4.3)

The weak variational form (3.12) becomes:


0

u EAu ds

(u, u ) =
0

L
.
upds uN
0

(4.4)

c ETH Z

urich IMES-ST, March 26, 2014

28

Truss element revisited

Partial integration transforms the term of the deformation energy so that the unknown
displacement solution u appears in its second derivative,
L

Z
 L
u0 EAu0 ds = uEAu0

uEAu00 ds .

(4.5)

Substituting the result into the variational form (4.4) and setting it to zero gives:
L

Z
0

L

 L
.
u EAu00 + p ds = uEAu0 uN
0

(4.6)

The two terms on the right-hand side are identical with opposite signs and cancel out. Thus
one obtains the result corresponding with the principle of virtual displacements (3.15)
L


u EAu00 + p ds = 0

(4.7)

where the Euler equation of the truss element appears in parentheses.

4.2

Truss weak variational form discretization

The shape functions of a linear truss element are illustrated with Fig. 4.1. The Bubnov-

u~
1

u~

1
F

F
1

2
1

-1
x

{
=

F
1

~
~u = u 1
~
u 2
u = F

u~

u ,x = F

, x u~

F
=

u~ 1 + F

1
1

, x u~ 1 + F
2

u~
, x u~

2
2

Figure 4.1: Linear truss element shape functions


Galerkin discretized weak variational form corresponding with (3.20) of the truss element
is
!
!
L
Ne Z L
Ne Z L
X
X

0
0T

= u

N :
u
EA ds u
pds + u
(4.8)
k=1

4.3

k=1

Transformation between truss and element coordinates

For the derivatives with respect to the truss coordinate s it holds that
0 =

c ETH Z

urich IMES-ST, March 26, 2014

d d
d
=
.
ds
d ds

(4.9)

4.4 Truss finite element numerical equations

29

The Jacobian matrix reduces to the scalar value d/ds which follows from the length Le
of the respective finite element within the mesh:
Le
d
2
(4.10)
d

=
2
ds
Le
When modeling spatial truss frames the problem arises that the truss coordinate s must
be embedded in the reference system x, y, z. This is not achieved with Jacobian type
transformations as the truss is a one-dimensional element. This type of transformations
is applied to the element stiffness matrix and explained later in Section 4.7.
ds =

4.4

Truss finite element numerical equations

The discretized weak variational form (3.21) written in local element coordinates is
!
1
Ne Z 1
Ne Z 1
X
X

2
Le
T
=
, EA,
d u
p d + N
(4.11)
Le
2
1
1
1
k=1

k=1

With the help of the chosen shape functions and their derivatives,




1
1 1
1
0
,
=
=
1
2 1+
Le

(4.12)

it can be converted into a numerical system of equations. First we write the integrand of
the deformation energy term in detail as a matrix,


EA
1 1
T 2
, EA,
,
(4.13)
=
1
Le
2Le 1
and then carry out the integration:


Z 1
EA
1 1
T 2
, EA,
d =
.
1
1
Le
Le
1

4.5

(4.14)

Assembly of system equations

The system equations of a finite-element mesh arise from superposition of the individual
element equations. Thereby the connectivity of the local element nodes with the globally
numbered mesh nodes must be observed. The connectivity information is stored in a
connectivity matrix either for nodes or for degrees-of-freedom, as one node typically carries
several degrees-of-freedom. Fig. 4.2 illustrates a system consisting of three collinear truss
elements. The mesh nodes are numbered from 1 through 4 whereas the local element nodes
are identified with a and b. The illustration obviates the connectivity and the connectivity
matrix is shown to the right in the same figure. The lower part of it illustrates how the 22
stiffness matrices of the respective elements are added to the system matrix (addressed
addition). Entries consisting of the sum of two element contributions are indicated with
the mixed colors orange and green.
If the system length is L, all elements are of equal length, and the whole truss is made of
the same material, the system equations are written as:

N1
A1
A1
0
0
u
1
p1

u
3E
A
A
+
A
A
0

p
+
p
0
e
2
1
2
1
1
2
2

.
=
+
u
3
p2 + p3
0
0
A2 A2 + A3 A3

Le
6

p3
N4
0
0
A3
A3
u
4
(4.15)
c ETH Z

urich IMES-ST, March 26, 2014

30

Truss element revisited

a
b

2
1

3
2
3
4

=
+

Figure 4.2: Simplistic truss frame work: Connectivity and system matrix
They simulate a truss with tapered cross-sections and different values of the line force p
in each element.
The system is not solvable and the next step is to understand how implement the effect
of a, at least statically determinate, support into the system equations.

4.6

Implementation of essential boundary conditions

We consider the same system as in the previous Section 4.5 and introduce with Fig. 4.3
a statically determinate support by specifying the node displacement u
1 with u
1 = u
.
The special case of a homogeneous essential boundary condition, u
= 0, simulates a fixed

u~ 1 = u

u~

u~
2

2
1

u~
4

L
Figure 4.3: Statically determinate truss frame
support. Before implementation of the boundary condition the system equations read:

k11 k12
0
0
u
1
r1

k12 k22 k23


0
u
2
r2

=
.
(4.16)
0 k23 k33 k34 u

r

3

3

0
0 k34 k44
u
4
r4
An algorithm for implementing essential boundary conditions,

kij = kji = 0, i 6= j
u
j = u

rj = u
, ri = ri kij u
,
,
kjj = 1, i = j

(4.17)

transforms the original unsolvable system of equations (4.16) into the solvable system:

u
1
u
1
1
0
0
0

0 k22 k23
u
0

k
u

2
2
12 1

=
.
(4.18)
0 k23 k33 k34 u

r3

0
0 k34 k44
u
4
r4
c ETH Z

urich IMES-ST, March 26, 2014

4.7 Directional transformation of the truss element

31

The sample problem illustrates that the first equation is replaced with u
1 = u
so that
solution of the system guarantees observation of the imposed boundary condition. Since
u
1 is no longer an unknown, the couplings with other degrees-of-freedom in the respective
equations are respected by transferring the products of u
and the respective stiffness matrix
entries to the right-hand side, so that the stiffness matrix maintains symmetry.
As a rule statically over-determinate support conditions must be simulated. If a significant
sub-set of all degrees-of-freedom is specified with essential boundary conditions it may be
worthwhile to reduce the system of equations by removing the respective equations. Within
the stiffness matrix then all rows and columns carrying a unit value on the main diagonal
would disappear.
At this point we have understood the essentials of all steps until arriving at a solvable
system of equations. With regard to the currently used backdrop of the truss element,
it is now useful to extend its applicability to arbitrary spatial orientation within a threedimensional Cartesian system.

4.7

Directional transformation of the truss element

The one-dimensional truss element can be used to simulate complex three-dimensional


truss frame works. To this end the global reference system xyz must be transformed onto
the coordinate s which is fixed to the truss longitudinal direction. The decomposition,
illustrated with Fig. 4.4, of the displacement usi into its components u
, v, and w
along
1

v~ i =
y

( ~y
2

- ~y

u~

1 )

v~

si

si

u~
2

u~ i =
y

u~

u~
1

( ~x

u~

si

L
1

u~

- ~x
2

v~

si

( ~x
2

- ~x
1

) 2 + ( ~y
2

- ~y
1

)2

x
1

Figure 4.4: Directional transformation at the truss element


xyz at the two nodes i = 1, 2 is:
u
i =

1
(
x2 x
1 )
usi ,
Le

vi =

1
(
y2 y1 )
usi ,
Le

w
i =

1
(
z2 z1 )
usi .
Le

(4.19)

Inverting these relations gives:


u
s1 =

1
Le

[(
x2 x
1 )
u1 + (
y2 y1 )
v1 + (
z2 z1 )w
1 ]

u
s2 =

1
Le

[(
x2 x
1 )
u2 + (
y2 y1 )
v2 + (
z2 z1 )w
2 ]

(4.20)

For the benefit of a more compact representation we choose the abbreviations


cx =

1
(
x2 x
1 ),
Le

cy =

1
(
y2 y1 ),
Le

cz =

1
(
z2 z1 )
Le

(4.21)

c ETH Z

urich IMES-ST, March 26, 2014

32

Truss element revisited

With it, the transformation (4.20) is conveniently written in matrix form:

u
1

v1


 


u
s1
cx cy cz 0 0 0
w
1
=

.
u
s2
0 0 0 cx cy cz
u
2

w
2

(4.22)

Substitution of the the node degrees-of-freedom and forces caused by line loads defined in
the truss coordinate s by the components in reference coordinates by use of the transformation (4.21) obtains the element equations in reference coordinates (cij = ci cj ):

EA

Le

cxx
cxy
cxz cxx cxy cxz
cyx
cyy
cyz cyx cyy cyz
czx
czy
czz czx czy czz
cxx cxy cxz
cxx
cxy
cxz
cyx cyy cyz
cyx
cyy
cyz
czx czy czz
czx
czy
czz

c ETH Z

urich IMES-ST, March 26, 2014

u
1
v1
w
1
u
2
v2
w
2

c
F
x
1x

y
1y

Le p
cz
F1z
=
+
cx
F2x

c
F2y
y

cz
F2z
(4.23)

Chapter 5

Planar finite elements


Fig. 5.1 illustrates that the domain of a two-dimensional element is an area and its
boundary consists of the lines connecting the nodes. Let each node carry two degrees-

v~
G

y , v

3
4

u~

h
W

1
2

x , u

~x , ~y
2
2

Figure 5.1: Planar four-node element


of-freedom, namely the displacement components along the reference system directions.
Even though the element can have a general shape as indicated in the figure, its master
element appears as a square in local element coordinates and . It is convenient to place
the local coordinates origin at the element center.
Planar elements can simulate problems under the states of plane stress, plane stress, or
rotational symmetry. Chapter 12 is dedicated to more complex mechanical situations
which can also be analyzed with planar elements.
Fig. 5.2 illustrates a typical plane-stress problem, namely a flat test coupon with a central
bore. Because of the double symmetry inherent to the problem the simulation domain is
only one quarter of the geometric domain and the sketched essential boundary conditions
guarantee that the quarter model behaves as it were a part of a full model and the correct
information is obtained from it. The loading situation provided by the testing machine
enforces at the moving set of clamps a displacement along the horizontal direction which
does not depend on the vertical direction. A realistic mapping of this loading situation
is easily achieved by implementing inhomogeneous boundary conditions. The sketched
subdivision of the domain into finite elements is much too rough for finding a good stress
approximation.
c ETH Z

urich IMES-ST, March 26, 2014

34

Planar finite elements

1 0

31

1
32

3
7
1 1
5

2
3

1 4

1 5

4
6

8
5

1 2
8
9

1 6

9
1 3

1 7

Figure 5.2: Test situation with notched specimen

5.1

Basic elasticity equations for plane states

The general equations of linear elasticity, given in Figs. 3.1 and 3.2, are here written
for two-dimensional problems including plane-stess (z = yz = xz = 0) and special
plane-strain (z = yz = xz = 0) states.
(
)

,
+

,
+
f
x
x
xy
y
x
LT + f =
= 0.
(5.1)
xy ,x + y ,y + fy


0


x
x


0
y
(5.2)
 = Lu,
=
,
L=
y
,

xy
y
x

x
C11 C12 0

y
C12 C22 0
= C,
=
,
C=
(5.3)

xy
0
0 C66
Plane-stress and plane state equations differ only in the definition of the material-law
entries. Fig. 5.3 compares symbolic and expanded notations.

5.2

Planar element domain and boundary differentials

The domain differential d becomes an area differential dxdy which is expressed with the
cross product

 

J11 d
J12 d
d = dx
ex dy
ey =

= |J| dd .
(5.4)
J21 d
J22 d
The boundary differential d becomes a line differential ds which is the length of the
differential vector ds:


dxex p
= dx2 + dy 2
(5.5)
ds = |ds| =
dyey
Expressed in local element coordinates:
p
2 + J 2 d
J11
12
p
2
2 d
ds() =
J21 + J22
ds() =

c ETH Z

urich IMES-ST, March 26, 2014

(5.6)

5.3 Higher-order shape functions for quadrilateral elements

35

L T C L u + f = 0

[C

1 1

u ,x + C

[C (u
6 6

1 2

,y + v ,

v ,
y

)]

], + [C (u
x

,y + v ,

6 6

[C
+

1 2

u ,x + C

2 2

)]
x

v ,

+ fx = 0
y

],
y

+ fy = 0

d iffe r e n tia l e q u a tio n

d u

+ d v

{[C

1 1

u ,x+ C

(L
T

d u

v ,

1 2

C L u + f

], + [C (u
x

6 6

(d

d W = 0

,y + v ,

)]
x

,y + f

}
x

{d
W

u ,

+ d v ,

{ [ C (u
6 6

,y + v ,
x

)]

,x+

[C

1 2

u ,x+ C

2 2

v ,
y

],
y

+ f

}
y

d x d y = 0
=
W

)
T

(d

(
C

[C

1 1

u ,x + C

[C (u
6 6

L u

v ,

1 2

,y + v ,

u fx + d v f
y

d W =

)]+

]+

d u ,
d v ,

d x d y +

[C (u

[C
y

d u T fd W +
W

(d

6 6

,y + v ,

u ,x+ C

1 2

2 2

v ,

u s x + d v s

d u Ts d G
G

)]

]}

d x d y

d s

w e a k v a r ia tio n a l fo r m

p r in z ip le o f v ir tu a l d is p la c e m e n ts

Figure 5.3: Symbolic notations and expansions

5.3

Higher-order shape functions for quadrilateral elements

The simplest elements carry nodes on their vertices only. Higher-order elements require
addional nodes. A general polynomial for planar elements including higher powers of up
to the third order is:
i = a1i
+ a5i

+ a2i
2

+ a6i

+ a9i 2 2 + a10i 3
+ a13i 3

+ a3i
+ a7i

+ a11i 3

+ a4i
+ a8i 2
+ a12i 3

(5.7)

+ a14i 3 2 + a15i 2 3 + a16i 3 3

The complete polynomial corresponds with the element with 16 nodes sketched at the
lower right in Fig. 5.4. It belongs to the Lagrange element family where nodes exist also
in the interior of the element domain. The Lagrange elements are distinguished from the
Serendipity element family (The Three Princes of Serendip by Horace Walpole) where
nodes appear on the element boundary only. Shape functions of Lagrange elements are
composed of the polynomials the powers of which are chosen from Pasquals triangle,
Even though the internal nodes can improve the approximation to the true solution, they
do not play a role in the continuity of the primary solution at the interface of adjacent
elements. In practice Serendipity-type elements tend to be preferred because they are
generally more efficient in terms of numerical effort and solution quality. On the other
hand, efficient elements maybe derived from Lagrange elements by eliminating the internal
degrees-of-freedom of these.
c ETH Z

urich IMES-ST, March 26, 2014

36

Planar finite elements

c o n s ta n t

c o n s ta n t
1

lin e a r
x

in c o m p le te q u a d r a tic
in c o m p le te k u b ic
x

x y
2

x 2y
3

lin e a r
y

x 3y

x y
x 2y

x 3y

x y
x 2y

x 3y

y
2

c o m p le te q u a d r a tic
2

c o m p le te k u b ic

S e r e n d ip ity e le m e n ts

L a g r a n g e e le m e n ts

Figure 5.4: Pasquals triangle and quadrilateral elements

5.4

The linear triangular element

The linear triangular element with nodes at its vertices only is very translucent; particularly, the integral of the strain-energy density over the element domain can be written by
hand. Its shape functions
i = a1i + a2i + a3i
(5.8)
are contained in the general polynomial and defined for the reference element shown at
the upper left in Fig. 5.5. The element coordinates originate at the lower left vertex and
are normalized 0 , 1. Points within the element domain satisfy the restriction
+ 1.

(5.9)

The shape functions are illustrated in Fig. 5.5 at the upper right and at the bottom. In
matrix notation they are arranged in a vector

1 1
2

=
=
(5.10)

The sketches indicating the shape functions indicate that a meshed triangular element will
generally have a different shape and size than the master element.

c ETH Z

urich IMES-ST, March 26, 2014

5.4 The linear triangular element

( 0 ,1 )
3

G 2: F
2

37

= 0
A

G 1: F
A

A
1

h
3

( 0 ,0 )

( 0 ,1 )

= 0

G 3: F

= 0
3

( 1 ,0 )

( 0 ,0 )

= h
3

( 0 ,1 )

( 0 ,1 )
3

h
1

( 0 ,0 )

( 1 ,0 )

h
x

F
1

= 1 - x - h

( 1 ,0 )

( 0 ,0 )

F
2

= x

( 1 ,0 )

Figure 5.5: Shape functions of the linear triangular element

5.4.1

B-matrix structure for the linear triangular element

Let the two displacement degrees-of-freedom on each node according to Fig. 5.6 be labeled
with ui and vi . An obvious choice for the arrangement of the node-displacements vector
u~

u~
~x , ~ y
1

~x , ~ y
3
1

v~
1

v~
3

u~
~x , ~ y
2
2

v~
2

Figure 5.6: Linear triangular element node coordinates and displacements


and the matrix , containing the individual shape functions, is
u


u=

u
v

= u
=

1 0 2 0 3
0 1 0 2 0

u
1

v1

u
2
0
3
v2

u
3

v3

(5.11)

Then follows, from the structure of the differential operator L (see Fig. 3.2) and from the
definition of the matrix B in (3.19), the structure of the latter for the linear triangular
element:

1 ,x
0
2 ,x
0
3 ,x
0
1 ,y
0
2 ,y
0
3 ,y .
B= 0
(5.12)
1 ,y 1 ,x 2 ,y 2 ,x 3 ,y 3 ,x
c ETH Z

urich IMES-ST, March 26, 2014

38

5.4.2

Planar finite elements

Linear triangular element domain integral

Due to its simplicity the integration of the deformation energy density terms over the
element domain can be carried out with closed-from expressions. Following Section 3.4.1
first one calculates the derivatives of with respect to the element coordinates:

1
1
1
0
, =
,
, =
.
(5.13)

0
1
The derivatives give the concrete result for the Jacobian matrix J,
 T
 

x
, y
T ,
x
2 x
1 y2 y1
J=
=
,
x
T , y
T ,
x
3 x
1 y3 y1

(5.14)

and its determinant


J = (
x2 x
1 ) (
y3 y1 ) (
y2 y1 ) (
x3 x
1 ) .

(5.15)

The inverse J1 of the Jacobian matrix is






1 y3 y1 y1 y2
,x ,x
1
.
J =
=
1 x
3 x
2 x
1
,y ,y
J x
The matrix B in terms of the nodal point coordinates is obtained by
together with (5.16):

1 y y
y2 y3
1 y y
1
1
2
3
1
0
1
y3 y1
+
=
,x =

J
J
1
0
y1 y2

(5.16)
applying (3.22)

(5.17)

1
x
3 x
2

1 x

x
2 x
1
1 x
3
1
0
1
x
1 x
3
+
=
,y =
.

J
J
J
1
0
x
2 x
1

(5.18)

Since the shape functions are strictly linear the B-matrix entries are constants. Therefore
the integral of the deformation-energy density can be directly implemented into a computer
program.

c ETH Z

urich IMES-ST, March 26, 2014

5.4 The linear triangular element

5.4.3

39

Linear triangular element boundary integrals

The boundary consists of the three parts 1 , 2 , and 3 and the vector line element
ds = dx + dy shown in Fig. 5.7 must be expressed in local coordinates. Using (5.6) and
( 0 ,1 )
3

G 2: F
2

= 0
A

y
3

G 3: F
3

= 0

d x
d y

A
( 0 ,0 )

( 0 ,1 )

1 = 0

h
1

G 1: F

d s

( 1 ,0 )

( 0 ,0 )

x , x

( 1 ,0 )

Figure 5.7: Vector line element in local coordinates


observing the constraint (5.9), which in the limit describes the edge 1 , obtains:
1 d = d
2 d =

0 .

3 d =

(5.19)

The vectorial line element


ds = (x, d + x, d) ex + (y, d + y, d) ey

(5.20)

obtains for the individual parts of the boundary


1 ds = [(
x2 x
3 ) ex + (
y2 y3 ) ey ] d
2 ds = [(
x3 x
1 ) ex + (
y3 y1 ) ey ] d .

(5.21)

3 ds = [(
x2 x
1 ) ex + (
y2 y1 ) ey ] d

The length ds of the line elements ds is calculated with ds = dsT ds:


q
1 ds =
(
x2 x
3 )2 + (
y2 y3 )2 d
q
2 ds =
(
x3 x
1 )2 + (
y3 y1 )2 d .
q
3 ds =
(
x2 x
1 )2 + (
y2 y1 )2 d

(5.22)

c ETH Z

urich IMES-ST, March 26, 2014

40

5.4.4

Planar finite elements

Linear triangular element kinematically equivalent nodal forces

The boundary tractions must be transformed into so-called kinematically equivalent nodal
forces, which is automatically achieved by evaluating the boundary term
Z

nd

(5.23)

of the weak variational form (3.21). The shape functions along the element edges are
shown in Fig. 5.8. Fig. 5.9 explains the transformation of constant and of linear external

( 0 ,1 )
3

( 0 ,1 )
3

1 - x
0
x

1 - x

=
1

F
F

F
F
F

0
=

1 - h
0
h

=
3

( 1 ,0 )

( 0 ,0 )

=
1

= x
2

G 2 :

( 1 ,0 )

( 0 ,0 )

= 1 - x - h
1

G1 :

( 1 ,0 )

( 0 ,0 )

( 0 ,1 )

= h

G 3 :
F

=
0
=
2

1 - x
x

=
0

F
F

Figure 5.8: Shape functions along element edges


stress distributions to kinematically equivalent nodal forces.

G = G

r =

s0 +

1 - x

- 2 x

2 x
2

[s 0 + 2 (x -

1 - x

- x
+ 3 x - 1

0
0
1
1

r =
1 s 0 +
1 s
6
2

1
- 1

F s d x =

)s 1 ]d
2

1 2

s1 d x = 2 x

x - 12 x

s 0 +

3
3

x
2

x
2

- x

s1

x
1

x
1

Figure 5.9: Linear triangular element kinematically equivalent nodal forces

c ETH Z

urich IMES-ST, March 26, 2014

5.5 The bilinear quadrilateral finite element

5.5

41

The bilinear quadrilateral finite element

The simplest quadrilateral finite element possesses nodes at its vertices only. Fig. 5.10
defines the numbering of the nodes and the local coordinates. The general polynomial

1
4

h
-1

1
x

-1

Figure 5.10: Bilinear quadrilateral element conventions


includes the first four powers in (5.7):
p = {1 }

(5.24)

The shape functions

(1 )(1 )
1 (1 + )(1 )
=
(1 + )(1 + )
4

(1 )(1 + )

(5.25)

are depicted in Fig. 5.11.

( 0 ,1 )
4
F

( 1 ,1 )

1
( - 1 ,- 1 )

22
( 1 ,- 1 )

=
4

1
( - 1 ,- 1 )

22
( 1 ,- 1 )
1

=
4

(1 - x ) (1 - h )

=
4

(1 + x ) (1 + h )
( 0 ,1 )
4

x
1

x
3

2
( 1 ,- 1 )

3
( 1 ,1 )

h
F

1
( - 1 ,- 1 )

(1 - x )(1 + h )
( 0 ,1 )
4

( 0 ,1 )
4

( 1 ,1 )

1
( - 1 ,- 1 )

h
x
2

3
( 1 ,1 )

2
( 1 ,- 1 )
2

=
4

(1 + x ) (1 - h )

Figure 5.11: Bilinear quadrilateral element shape functions


Analog to Section 5.4.1 the structure of the matrix B is:

1 ,x
0
2 ,x
0
3 ,x
0
4 ,x
0
1 ,y
0
2 ,y
0
3 ,y
0
4 ,y .
B= 0
1 ,y 1 ,x 2 ,y 2 ,x 3 ,y 3 ,x 4 ,y 4 ,x

(5.26)

c ETH Z

urich IMES-ST, March 26, 2014

42

5.6

Planar finite elements

Three- and four-noded elements bending performance

The two elements obtain different results in loading situations dominated by bending
deformation such as shown in Fig. 5.12, where a cantilever beam is subject to a single
1 :

n u m b e r o f e le m e n ts

2 :

n u m b e r o f d e g re e s -o f-fre e d o m

3 :

m a x im u m

d e fle c tio n

4 :

m a x im u m

b e n d in g s tr e s s
1
2
3

1 5

-0 .3 8 1

1 2 0

4 0

1 0 0

-0 .3 8 0

1 1 4 .3

8 0

1 0 0

-0 .1 6 7

5 3 .2

Figure 5.12: Results of various models for simulating a cantilever beam loading situation
transverse force at its free end. If the influence of shear deformation on the end deflection
may be neglected the thin-beam finite element mesh obtains best results as the analytical
deflection line result is contained in its solutions space. If the beams width is small
enough, a plane-stress problem results which can be modeled with planar elements. The
two meshes with triangular and quadrilateral elements, respectively, have the same number
of degrees-of-freedom. While the mesh with quadrilateral elements agrees well with the
beam finite element results, the mesh with triangular elements gives a much smaller end
deflection. The answer to the question, as to why the triangular element behaves too stiff
in bending, lies in the element shape functions as illustrated with Fig. 5.13. It catches the
-1 , 1
h

1 , 1

F
x

-1 ,-1

1 ,-1

(1

1 (1
=

4 (1
( 1
F

0 ,1

F
0 ,0

1 ,0

= a

- x ) (1
+ x ) (1
+ x ) (1
- x ) (1
1 i

+ a

1 - x - h

x
=
,

- h )
- h )

+ h )
+ h )
2 i

x + a

- 1 + h
1 - h

,
1 + h
- 1 - h

1
F , =

x
4

3 i

h + a

4 i

- 1

1 - 1
F , =

h
4 1
1

+ x
- x

+ x
- x

x h

- 1

F , = 1 ,
x
0

- 1

F , = 0
h
1

Figure 5.13: Ansatzfunktionen Drei- und Viereckelement


eye that the triangular element strains must be constant and can not, per se, map linear
bending strain (or stress) distributions.

5.6.1

Meshing with triangular and quadrilateral elements

Triangular elements are more flexible in terms of domain geometry approximation including local mesh refinement as Fig. 5.14 indicates. That the quadrilaterals remain
c ETH Z

urich IMES-ST, March 26, 2014

5.7 Curvilinear elements

43

Figure 5.14: Quarter-circle approximation with triangular and quadrilateral elements


rectangular is an arbitrarily choice and not typical for the meshes created by numerical
mesh generators.

5.7

Curvilinear elements

Edges of elements with higher-order shape functions may be curved in global reference
coordinates. Fig. 5.15 demonstrates the aptness of curvilinear elements for approximating
R e fe r e n c e E le m e n ts
b
c

c
a

a
q u a d r a tic

c
d

b
a
lin e a r

Figure 5.15: Circular domain approximation with bilinear and curvilinear elements
domains with curved boundaries.

c ETH Z

urich IMES-ST, March 26, 2014

44

Planar finite elements

The shape functions of the eight-node Serendipity element along with their derivatives are
given in Table 5.1 and the shape functions are illustrated in Fig. 5.16.
Knoten

1
4 (1

)(1 )( 1)

1
4 (2

+ )(1 )

1
4 (2

+ )(1 )

1
4 (1

+ )(1 )( 1)

1
4 (2

)(1 )

1
4 (2

)(1 + )

1
4 (1

+ )(1 + )( + 1)

1
4 (2

+ )(1 + )

1
4 (2

+ )(1 + )

1
4 (1

)(1 + )( + 1)

1
4 (2

)(1 + )

1
4 (2

)(1 )

1
2 (1

2 )(1 )

1
2 (1

+ )(1 2 )

1
2 (1

2 )(1 + )

(1 + )

1
2 (1

+ )(1 2 )

12 (1 2 )

(1 )
1
2 (1

2)

12 (1 2 )
(1 + )
1
2 (1

2)

(1 )

Table 5.1: Four-node Serendipity element shape functions and derivatives

h
8

x
1

=
a
+

a 5 ix

a 2 ix
+

a 3 ih

a 6 ix 2h
+

a 7 ih

1 i
2

a 4 ix h
+

a 8 ih

Figure 5.16: Four-node Serendipity element shape functions illustration

c ETH Z

urich IMES-ST, March 26, 2014

Chapter 6

Element technology
The selection of finite elements for simulations underlies a conflict of objectives: those
with low-complexity shape functions can be numerically more efficient, as their number
of degrees-of-freedom and, more significantly, their inherent band width are smaller, while
those with higher-complexity shape functions can better map more complex solutions
distributions such as high stress gradients. Element technology is concerned with the
improvement, or mitigation, of the limitations of simple elements with a small number of
degrees-of-freedom. The sub-structuring, here presented in Chapter 9, is mentioned by
Zienkiewicz [6] in context with complex finite elements. Elimination of interior degrees-offreedom can obtain more efficient element types. He reports that a quadrilateral element,
composed of four triangular elements, gives, after elimination of the central node shown
in Fig. 6.1, an economic advantage over the direct use of simple triangular elements. The

Figure 6.1: Quadrilateral element construction, after [6]


same method can be used to eliminate the inner degrees-of-freedom of a Lagrange element,
resulting in elements which have, like Serendipity elements, explicit degrees-of-freedom on
the element edges only and possess better properties than these. In the following the truss
element and the four-node quadrilateral element are considered.

6.1

Sample problem truss element

The linear truss element can not map the true linear internal force distribution caused by a
constant line load p. Therefore we construct a more complex truss element with quadratic
shape functions. It requires a third node and Fig. 6.2 illustrates the shape functions
2

1
2 +
=
,

2
2 2 2

2 1

1
2 + 1
, =
.

2
4

(6.1)

c ETH Z

urich IMES-ST, March 26, 2014

46

Element technology

f
1

f
f

-1

x
1

Figure 6.2: Quadratic truss-element shape functions


The derivation steps shown in 4.4 obtain the element equations

1
7
1 8 u
1
Le p
AE
u
2
1
7 8
1
=
.

3Le
6
u
3
8 8 16
4

(6.2)

From these the degree-of-freedom u


3 on the central node 3 will now be eliminated. Therefore the equations are written in expanded form:
k11 u
1 + k12 u
2 +k13 u
3 = r1

(a)

k12 u
1 + k22 u
2 +k23 u
3 = r2

(b)

k13 u
1 + k23 u
2 +k33 u
3 = r3

(c)

The third equation is resolved for u


3 ,
1
u
3 =
(r3 k13 u
1 k23 u
2 ) ,
k33

(6.3)

(6.4)

and the result is used to remove u


3 from the remaining two equations of (6.3), leading to:




2
k13
k23
k11 k33
u
1 +
k12 k13
u
2 = r1 kk13
r3
k33
33
(6.5)




2
k23
k13 k23
k23
k12 k33 u
1 +
k22 k33 u
2 = r2 k33 r3
Next, the terms are evaluated by inserting the numbers given in (6.2). Then the equations
are written in matrix form again and it appears that the familiar equations of the linear
element are reproduced. This should not be taken as a disappointment, as an advantage
of the exercise appears in the post-processing step, when evaluating the internal element
quadratic displacement and linear strain fields according to (6.4). First one obtains the
node displacement u
3 ,
1
L2 p
u
3 = (
u1 + u
2 ) + e ,
(6.6)
2
8AE
from the third equation of (6.3); then the quadratic displacement field u() follows from
the shape functions (6.1):
u() =

L2 p
1
(
u1 + u
2 ) + (
u2 u
1 ) + e
1 2 .
2
2
8AE

(6.7)

2
1
L2 p
=
(
u2 u
1 ) e .
L4
Le
4AE

(6.8)

The linear strain field is


 = u,xi

The new post-processing step makes the element more efficient, as it can now describe the
true solution under constant line load.
c ETH Z

urich IMES-ST, March 26, 2014

6.2 Sample problem quadrilateral element

6.2

47

Sample problem quadrilateral element

This material has partially been taken from the lecture which M. Jabareen held on occasion
of his sabbatical during the fall term 2009 within Strukturanalyse mit FEM [7]. First, the
behavior of the bilinear quadrilateral finite element is scrutinized regarding
1.
2.
3.
4.

patch test
incompressible materials
bending
mesh distortion

Secondly, it will be shown how its performance can be increased by adding incompatible
displacement modes to the bilinear shape functions.

6.2.1

Bilinear quadrilateral element behavior

The element is explained in Section 5.5 and the considerations laid out in Section 5.6 made
clear that it maps bending far better than the linear triangular finite element. Closer
investigations, however, reveal limitations regarding the mapping of nearly incompressible
materials ( 0.5) under plane strain or pure bending. The original bilinear element is
identical with the type plane42 of the FEM software ANSYS Classic [8] with the option
extra displacement shapes: excluded.
Patch Test
The sketch in Fig. 6.3 shows a rectangular domain subject to essential boundary conditions
so that a uni-axial and homogeneous plane strain state x 6= 0, y = 0, and xy = 0 must
result. Table 6.1 lists the node coordinates of the finite element mesh. The figure indicates
node
x

1
0.00
0.00

2
2.00
0.00

3
2.00
1.00

4
0.00
1.00

5
0.65
0.33

6
1.48
0.30

7
1.45
0.77

8
0.53
0.71

Table 6.1: Node coordinates for the patch test. Source: [7]
that, in spite of the element distortions, the homogeneous strain state is mapped exactly.
4

5
1

3
7

6
2

Figure 6.3: Patch test: Geometry, mesh, boundary conditions and displacement field ux

c ETH Z

urich IMES-ST, March 26, 2014

48

Element technology

Nearly incompressible material and volume locking effect


Cooks membrane problem is also visited with the element type plane42. For letting the
mechanically inconsistent volume locking effect appear it is necessary to choose with the
option element behavior: plane strain that z = 0. As Fig. 6.4 testifies, the displace2

1 .0

1 .0

0 .8

0 .8

0 .6

0 .6

4 4
4 n x 2 n

0 .4

u 2 (n u )/u

4 4

u 1 (n u )/u

(n u = 0 .4 9 )

1 6

(n u = 0 .4 9 )

f= 1 N /m m

0 .2
0 .0

0 .4
0 .2
0 .0

0 .0

0 .2

0 .4

0 .6

0 .8

1 .0

0 .0

0 .2

(n u -0 .4 9 )/(0 .4 9 9 7 -0 .4 9 )

4 8

0 .4

0 .6

0 .8

1 .0

(n u -0 .4 9 )/(0 .4 9 9 7 -0 .4 9 )

Figure 6.4: Incompressibily: Cooks membrane problem


ments tend to zero as Poissons ratio tends to 0.5. This contradicts reality because
the material must remain able to deform in modes which do not change the total domain
volume. The smallness of the displacements are due to the volume locking effect of the
bilinear quadrilateral finite element, which will be proven on the simple situation depicted
in Fig. 6.5. Let the local coordinates and coincide, for simplicity, with the global
reference coordinates. Nodes 1, 2, and 4 are fixed in space whereas node 3 may be displaced with u
3 and v3 . The figure also shows the displacement fields depending on the

v
4

u
3

u 1 = v 1 = u 2 = v 2 = u 4 = v 4 = 0

u = F
2

1
2
L

v = F

u 3 =
3

v 3 =
1

(1

+ x

) (1

+ h

(1

+ x

) (1

+ h

)
)

Figure 6.5: Volume locking effect situation


node displacements. The element area changes with the node displacements as depicted
in Fig. 6.6:
1
A(
u3 , v3 ) = (2L1 L2 + L2 u
3 + L1 v3 )
(6.9)
2
The fact, that under plane strain (33 = 0) and the requested volume invariability ( = 0.5)
the area content must remain constant,
A(
u3 , v3 ) = L1 L2 ,

(6.10)

couples the two node displacements:


v3 =
c ETH Z

urich IMES-ST, March 26, 2014

L2
u
3 .
L1

(6.11)

6.2 Sample problem quadrilateral element

A ( u 3 ) = L
2

(L
1

+
1

u
3

49

A ( v 3 ) = L
1

(L

+
2

1
2

v
3

Figure 6.6: Area contents due to node displacements


On the other hand the volume invariability implies that the trace of the strain tensor must
vanish:
tr() = xx + yy + zz = 0 .
(6.12)
The node displacements create the strain fields:
xx = u,x = u, ,x =

1
(1 + ) u
3 ,
2L1

yy =

1
(1 + ) v3
2L2

(6.13)

Substituting of the strain fields into the conditions (6.12) and observing the constraint
(6.11) obtains
1
tr() =
( ) u
3 = 0 .
(6.14)
2L1
This condition can be fulfilled for arbitrary values of and only if u
3 = 0 which proves
the volume locking effect of the bilinear quadrilateral finite element.
Thin structures bending and convergence behavior
The bilinear shape functions of the four-noded quadrilateral finite element permit mapping
of linear bending strain within its domain, which is in contrast to the linear triangular
element. Nevertheless, Fig. 6.7 shows that the accuracy of the mapping of a deflection
P

1 0 n n
H

F e h le r [% ]

-1 0

E = 1 5 0 0

N
m m

L = 1 0 0 m m ,

n = 0 .2 5
,

-2 0
-3 0
-4 0
0

= 1 0 m m

P = 0 .0 1 N

1 0

Figure 6.7: Bending of a cantilever beam and mesh dependent mapping error
line depends much on the quadrilateral element mesh density. In the following it is investigated what properties the element should posses in order to be able to map bending
well. The investigation starts with imposing, in Fig. 6.8, an admissible stress field within
an area element underlying pure bending caused by external bending moments so that no
transverse force arises. The equations in the figure indicate that the bending stress x
be linear through the thickness and the two other stress components, y and xy , vanish.
c ETH Z

urich IMES-ST, March 26, 2014

50

Element technology

= a y ,

s
x

= C

1 1

s
y

+ C
x

= t

1 2

1 2

= 0

a
=

y ,

x y

= C

= C

1 2

3 3

+ C

x y

2 2

e
y

= 0

= 0

x y

= y

1 1

1 1

x
s

1 2

= C

1 1

2
1 2
2 2

2 2

Figure 6.8: Illustration to pure bending


The constitutive equations obtain that bending strain is also linear through the thickness
an that a through-the-thickness strain arises due to Poissons effect. The shearing strain
vanishes with the shearing stress.
A displacement field is obtained by integrating the strain fields. First the direct strains
are integrated:
Z

u = x (y)dx + C1 (y) = xy + C1 (y)


(6.15)
C11
Z
C12 y 2
+ C2 (x)
(6.16)
v = y (y)dy + C2 (x) =
C22 C11 2
where the functions C1 (y) and C2 (x) must meet the requirement that the displacement
fields u and v must not create shear strain (compatibility condition):

xy = u,y +v,x = x + C1 (y),y +C2 (x),x = 0


C11

(6.17)

Since the functions C1 (y) and C2 (x) are separable with respect to their spatial variables,
C1 (y) = C1 must be a constant and after integration it follows that
x2
C2 (x) =
C1 x + C3 (y) .
C11 2
Inserting the latter results into (6.15) and (6.16) obtains the displacement fields
 2


x
C12 y 2
u(x, y) = xy + C1 ,
v(x, y) =
+
C1 x + C3 (y)
C22 2
C11
C11 2

(6.18)

(6.19)

The comparison with the bilinear element shape functions of our present quadrilateral
finite element shows that this cannot map pure bending. The result of a simulation with
ANSYS corroborates this, see the shear strain distribution shown in Fig. 6.9,
Thin structures bending and distorted elements
Under a homogeneous strain distribution the element passes the patch test as the elements distortion does not impair results quality, see Section 6.1. On the other hand, the
observed weakness regarding pure bending lets expect an effect of mesh distortion on calculated deflections. Fig. 6.10 shows the mesh distortion controlled by the parameter a.
With increasing mesh distortion the erroneous results, obtained with low mesh density,
deteriorate even further.
c ETH Z

urich IMES-ST, March 26, 2014

6.2 Sample problem quadrilateral element

51

Figure 6.9: Unphysical connection of pure bending with shear by the bilinear element

L = 1 0 0 m m
H = 1 0 m m
P = 0 .0 1 N

E = 1 5 0 0 M P a
n = 0 .2 5
P

E r r o r d e fle c tio n [% ]

0
-2 0
-4 0
-6 0
-8 0
-1 0 0

0 .0 0

0 .2 5

0 .5 0

0 .7 5

1 .0 0

a /2 H

Figure 6.10: Influence of mesh distortion on deflection error

6.2.2

Additional incompatible element displacement modes

The weakness of the bi-linear four-node element, exposed in Sections 6.2.1 and 6.2.1,
can be traced back to the lack of quadratic terms in the element shape functions. For
simulation of mechanical situations dominated by bending the problem could be solved by
simply using higher-order elements, for instance the eight-node Serendipity type element
presented earlier, and accepting the higher number of degrees-of-freedom and the larger
bandwidth.
Much more interesting is the question whether the bilinear element could be improved so
that, in spite of having only four nodes, it can map bending better. Fig. 6.11 indicates
the affirmative answer: Add to the bi-linear shape functions the quadratic displacement
1 and
2:
modes


1 = 1 2 ,
2 = 1 2

(6.20)
These vanish at element nodes and are therefore connected with abstract weight factors
:
T
+
u = T u
(6.21)
Consequently, the displacement solution is still continuous across nodes but along element
edges gaps or overlaps may occur, hence the name incompatible modes.
We stick with the convention


T = u
1 v1 u
2 v2 u
3 v3 u
4 v4
u
(6.22)
and arrange the weight factors of the incompatible modes:


T =
u1
v1
u2
v2

(6.23)
c ETH Z

urich IMES-ST, March 26, 2014

52

Element technology

F
1

F
3

F
2

Figure 6.11: Bi-linear shape functions (Fig. 5.11) and incompatible quadratic modes
which is
Their contribution to the material deformation is considered with the matrix B,
calculated analog to the familiar matrix B:

 = B
u+B

(6.24)

For deriving the stiffness matrix of the enhanced element one considers the total deformation energy U :
Z
1
T Cd
U =
2
Z
T

1

C B
d
=
B
u+B
u+B
(6.25)
2
and to
delivers the system of equations:
Taking the variations with respect to u
Z


d = r
BT CB
u + BT CB


T CB
T CB

d = 0
B
u+B

(6.26)

Carrying out the integration rule obtains stiffness matrices and therefore (6.26) can also
be written in the more concise form:
"
#(
) ( )

Kuu Ku
u
r
=
(6.27)

Ku K

0
of the incompatible modes,
The inner degrees-of-freedom, which are the weighting factors
are eliminated by first resolving the second system of equations for them,
= K1

Ku u
and then substituting the result in the first system of equations:

= r
Kuu Ku K1
Ku u
c ETH Z

urich IMES-ST, March 26, 2014

(6.28)

(6.29)

6.2 Sample problem quadrilateral element

53

Enhanced four-node element patch test


On the basis of the numerical results, let it be accepted without analytical investigation
that the simple four-node element passes the homogeneous-strain patch test described in
Section 6.2.1. The results of the simple element serve as a reference solution which the enhanced element must find as well. The specified displacements
u0 lead to the homogeneous
state of strains 0 and the constitutive equations find the homogeneous state of stresses
0:
0 0 = B
u
u0 0 = C0 = CB
u0
(6.30)
The enhanced element passes the homogeneous-strain patch test if the quadratic modes
do not contribute to the displacements because their derivatives yield linear strain distributions. Therefore, the weighting factors must vanish. Because of (6.26) it must then
also hold that
=0
(6.31)
Ku u
This result together with (6.30) identifies the condition:
Z
Z
T CB
T 0 de = 0
B
B
ude =
e

T de = 0
B

(6.32)

helps assessing whether condition (6.32) is satisfied:


Expanding the matrix B

y,
0
y,
0

= 2 0
x,
0
x,
B

Det(J0 )
x, y, x, y,

(6.33)

The Jacobian connected with homogeneous strain, J0 , must be a constant and then all
are linear. For integrating linear polynomials the one-point Gauss rule suffices.
entries in B
At the center, = = 0, the linear functions vanish so that condition (6.32) is satisfied
and the enhanced element satisfies the homogeneous-strain patch test as well.
Enhanced four-node element and material incompressibility
We revisit the problem analysis in Section 6.2.1 and note that the incompatible displacement modes can not change the area of the element domain. Therefore the equations (6.10)
through (6.12) continue to hold in the presence of the incompatible modes. However, the
latter have an influence on the strains and instead of (6.13) it holds that:
xx =

1
[(1 + ) u
3 2
u1 ] ,
2L1

yy =

1
[(1 + ) u
3 2
v2 ]
2L2

(6.34)

The condition that, for an incompressible material, the trace must vanish (6.12) gives here:
tr() =

1
1
1
( ) u
3

u1

v2 = 0
2L1
L1
L2

(6.35)

Comparing the coefficients shows that the trace can vanish for arbitrary values of and
, or anywhere in the element domain, if
1

u1 = u
3
2

v2 =

L2
u
3
2L1

(6.36)

As a result, the incompatible displacement modes mitigate the volume locking effect of
the bi-linear quadrilateral finite element as illustrated with Fig. 6.12.
c ETH Z

urich IMES-ST, March 26, 2014

54

Element technology

1 .0

0 .8

0 .8

4 4
4 n x 2 n

u 1 (n u )/u

4 4

(n u = 0 .4 9 )

1 .0

0 .6
0 .4

0 .6

d is p la c e m e n t fo r m u la tio n

d is p la c e m e n t fo r m u la tio n

(n u = 0 .4 9 )

1 6

u 2 (n u )/u

f= 1 N /m m

in c o m p a tib le m o d e s

0 .2

0 .4

in c o m p a tib le m o d e s

0 .2
0 .0

0 .0
0 .0

0 .2

4 8

0 .4

0 .6

0 .8

0 .0

1 .0

0 .2

0 .4

0 .6

0 .8

1 .0

(n u -0 .4 9 )/(0 .4 9 9 7 -0 .4 9 )

(n u -0 .4 9 )/(0 .4 9 9 7 -0 .4 9 )

Figure 6.12: Cooks Problem considered with basic and enhanced quadrilateral elements
Enhanced four-node element and pure bending
Referring back to Section 6.2.1 a state of pure-bending is imposed on the finite element
by specifying the displacements u
1 = u
3 =
u and u
2 = u
4 = u
. This specifies the general
displacement field u(, ):
u(, ) = u
+ (1 2 )
u1 + (1 2 )
u2

(6.37)

We allow that the bending deformation can effect the node displacements along y so that
v1 = v2 = v3 = v4 = v which gives the displacement field:
v(, ) = v + (1 2 )
v1 + (1 2 )
v2

(6.38)

Because of the symmetry of geometry as well as loading it is obvious to choose:


u(0, 0) = 0 = (1 2 )
u1 + (1 2 )
u2

u1 =
u2 = 0

(6.39)

Along the other direction rigid-body motion is suppressed with specifying thet v(1, 0) =
v(1, 0) = 0:
v(1, 0) = 0 = v +
v2

v2 =
v
(6.40)
Then the displacement fields are:
u(, ) = u

v(, ) = (1 2 )
v1 + 2 v

(6.41)

c
b

Figure 6.13: Pure bending with incompatible modes: Bending strain x (a), transverse
strain y due to Poissons ratio (b), and vanishing shear strain xy (c).
By identifying x = and y = , comparison of coefficients

u
= ,
C11

C1 = 0

c ETH Z

urich IMES-ST, March 26, 2014

v1 =

,
2C11

C11
v =
,
2C11 C22

C3 =
v1 ,

(6.42)

6.2 Sample problem quadrilateral element

55

obtains that the two displacement fields (6.19) and (6.41) can be consistently transformed
into each other. Particularly it is easily shown that the shear strain xy now vanishes as
illustrated in Fig. 6.13. Fig. 6.14 shows that the enhanced element can map the structural

P
0

F e h le r [% ]

1 0 n n
H

E = 1 5 0 0

N
m m

n = 0 .2 5
,

L = 1 0 0 m m ,

= 1 0 m m

-1 0
d is p la c e m e n t fo r m u la tio n

-2 0

w ith in c o m p a tib le m o d e s

-3 0
-4 0
0

P = 0 .0 1 N

1 0

Figure 6.14: Cantilever beam bending and influence of mesh density on error
response of a cantilever beam even with a rough mesh.
Bending, element distortion, and incompatible modes
Fig. 6.15 compares the influence of mesh distortion on the deflections as predicted by the

L = 1 0 0 m m
H = 1 0 m m
P = 0 .0 1 N

E = 1 5 0 0 M P a
n = 0 .2 5
P

E r r o r d e fle c tio n [% ]

0
-2 0
w ith in c o m p a tib le m o d e s

-4 0
-6 0
-8 0
-1 0 0

0 .0 0

a
L

0 .2 5

0 .5 0

0 .7 5

1 .0 0

a /2 H

Figure 6.15: Mesh distortion influence on simulated deflection error


base and enhanced finite elements. In contrast to the base element the enhanced element
obtains good results nearly independent of mesh distortion.

c ETH Z

urich IMES-ST, March 26, 2014

56

c ETH Z

urich IMES-ST, March 26, 2014

Element technology

Chapter 7

Numerical preprocessing
This chapter discusses aspects of the automated pre-processing.

7.1

Element matrices size

The sketch on the upper left of Fig. 7.1 serves as a legend for the color coding of the
here considered matrices whose dyadic products give the integrand of the element stiffness
matrices. The matrices B are shown in yellow and the matrices C, expressing the constiB
C B
C

2 2 = 4

C B

6 6 = 3 6

1 2 1 2 = 1 4 4

8 8 = 6 4

1 6 1 6 = 2 5 6

Figure 7.1: Structure and size of matrix BT CB for various elements


tutive law, in purple. The dark-green marked product CB of the two matrices presents
stresses and the dyadic product BT CB rendered in light-green corresponds with the specific deformation energy.
The remaining sketches illustrate the sizes of the matrices of specific element types. The
simple truss element yields the smallest matrices: The integrated constitutive law EA is
only a number and the matrix B contains only two entries. The integrand of the stiffness
matrix, as well as the stiffness matrix itself, contains only 2 2 entries.
The most complex element considered in class is the two-dimensional Serendipity element
with quadratic shape functions for analyzing plane state problems. Its stiffness matrix has
16 16 rows and columns or 256 entries. Its higher power in mapping geometry as well as
solutions complexity is penalized with much higher computational effort, if compared with
c ETH Z

urich IMES-ST, March 26, 2014

58

Numerical preprocessing

the bi-linear four-node element. Fig. 7.2 illustrates the operations for forming the matrix
u = u~

= L u = L u~ = B u~

u~

u~

u
u

B
L

Figure 7.2: Forming of the matrix B of the quadratic Serendipity Element


B for the complex element, which contribute significantly to the numerical pre-processing
effort.

7.2

Gauss quadrature rule

In order to obtain the numerical system of equations the integrations over domain and
boundary must be carried out. Very simple geometries and shape functions, occurring
for instance with the truss, beam, or linear triangular elements, allow integration by
hand and inserting the integrated results into a computer program which contributes
to its numerical efficiency. However, in general an exact integration is not feasible and a
numerical integration, or quadrature, rule must be used. Two-dimensional elements are
often most efficiently integrated by a Gauss quadrature rule:
Z

F(, )dd =
1

NG X
NG
X

wi wj F (i , j ).

(7.1)

i=1 j=1

Chapters 3 through 5 explain that a reference finite element is described and evaluated
in terms of normalized local coordinates. The integration over the normalized domain is
replaced by a sum of weighted function values at specified locations, the Gauss points. The
Gauss quadrature exactly integrates a polynomial function of given order with a minimum
number of sampling points. Fig.7.3 provides vivid pictures for understanding why the onepoint rule exactly integrates linear polynomials and the two-point rule suffices for cubic
polynomials.

-1

0 .0

-1

-0 .5 7 7 4

0 .5 7 7 4

-1

-0 .5 7 7 4

0 .5 7 7 4

Figure 7.3: Vivid pictures for the one- and two-point Gauss quadrature rules

c ETH Z

urich IMES-ST, March 26, 2014

7.2 Gauss quadrature rule

7.2.1

59

Polynomial order and quadrature rule

This section shows how to derive the coordinates and weights for exact integration of a
polynomial of a given order. It is most convenient to consider a function over a onedimensional and normalized domain 1 1. The numerical rule must obtain the
exact integral:
Z 1
NG
X
F ()d =
wi F (i ).
(7.2)
1

i=1

The chosen domain symmetry with respect to its coordinate has the advantage that the
integrals of all terms of F with odd powers vanish. Therefore it is sufficient to consider
the even polynomial powers only:
F =

m
X

a2k2 2k2 .

(7.3)

k=1

The expression for F is now substituted in (7.2) to obtain:


NG
m
X
X
a2k2
=
wi
F ()d = 2
2k 1
1

i=1

k=1

m
X

!
a2k2

2k2

NG
X

wi F (i ) .

(7.4)

i=1

k=1

The samples illustrated in Fig. 7.3 indicate that the number of Gauss points must be equal
to the highest occurring even power, or NG = m. It is useful to exchange the sequence of
the summation symbols on the right-hand side of (7.4):
!
m
m
m
X
X
X
a2k2
(7.5)
2
=
a2k2
wi 2k2 .
2k 1
k=1

i=1

k=1

The resulting equation must be satisfied for each parameter a2k2 , giving m equations for
finding the m supporting points i and m weights wi :
m

X
2
=
wi 2k2 .
2k 1

(7.6)

i=1

Domain symmetry provides the necessary m additional conditions:


- supporting points with non-zero coordinates have negative mirror partners
- the weights of these pairs are equal
Sample case: derivation of the three-point rule
Formal processing of (7.6) yields, because of m = 3, three conditional equations:
2

= w1

2/3 =

w1 12

+ w2
+

w2 22

+ w3
+ w3 32

(7.7)

2/5 = w1 14 + w2 42 + w3 42
Because of domain symmetry it must hold that 3 = 1 , 2 = 0 und w3 = w1 . Therefore,
the last two equations of (7.7) carry only 1 and w1 :
2/3 = 2w1 12
2/5 = 2w1 14

(7.8)

c ETH Z

urich IMES-ST, March 26, 2014

60

Numerical preprocessing

Resolving the first of the two equations in (7.8) for w1 yields:


w1 =

1
312

(7.9)

Substituting the result into the second equation in (7.8) obtains the Gauss points 1 and
3 ,
r
r
3
3
1 =
, 3 =
,
(7.10)
5
5
and substituting these in turn in (7.9) obtains the value of the Gauss weight w1 :
w1 = w3 =

5
.
9

(7.11)

With this follows from the first of the equations in (7.7) the Gauss weight w2 with
w2 =

7.2.2

8
.
9

(7.12)

Table and two-dimensional point grating

Table 7.1 gives the Gauss-point coordinates and weights for Gauss quadrature of various
rule
1 point

point
1

2 points

coordinate
0

numerical value
0.0000000000

q
13
q

-0.5773502692

1.0000000000

1
3

0.5773502692

1.0000000000

q
35

-0.7745966692

0.5555555556

0.0000000000

5
9
8
9

3
5

0.7745966692

5
9

0.5555555556

-0.8611363116

18 30
36

0.3478548452

-0.3399810436

18+ 30
36

0.6521451548

0.3399810436

18+ 30
36

0.6521451548

0.8611363116

18 30
36

0.3478548452

2
3 points

1
2

3
r

4 points

1
2
3
4

q
3+2 65

7
r q
32 65

7
r q
32 65
7
r q
3+2 65
7

weight
2

numerical value
2.0000000000

order
1
3

0.8888888889

Table 7.1: Coordinates and weights of Gauss quadrature


polynomial orders. The points placement in two-dimensional element domains is visualized
with Fig. 7.4. Here, the total number of Gauss points is the square of those within the
one-dimensional domain.

c ETH Z

urich IMES-ST, March 26, 2014

7.3 Selection of the quadrature rule

E le m e n t T y p e

P o ly n o m ia l

lin e a r

61

e x a c t to

h =

0 .5 7 7 ...

3
x
h = - 0 .5 7 7 ...
h

q u a d r a tic

h =

0 .7 7 4 ...

h =

0 .0
x

h = - 0 .7 7 4 ...

c u b ic

h =

0 .8 6 1 ...

h =

0 .3 3 9 ...

h = - 0 .3 3 9 ...
h = - 0 .8 6 1 ...

Figure 7.4: Gauss points in two-dimensional quadrilateral finite reference elements

7.3

Selection of the quadrature rule

The number of Gauss points must be chosen appropriately: High enough for correct integration but not higher, as to not create unnecessary computational effort.

7.3.1

Integration of the stiffness matrix

The number of Gauss points, necessary for a given polynomial order, can be looked up
in Table 7.1. Considering undistorted, particularly rectangular elements, the polynomial
order of the specific deformation energy term can be derived from the order of the general
polynomial underlying the element shape functions. Fig. 7.5 illustrates the approach: First
M a x im a le P o te n z in

P o ly n o m a n s a tz :

x , h

x h

B - M a tr ix

1 , 1

x , h

In te g ra n d

:
:

1 , 1

, h

, h

x 2h , x h

, h

x , h
2

2
2

, h
, h

Figure 7.5: Highest polynomial powers occurring in matrix B


one finds the highest existing powers occurring in the shape functions. Without having
to consider details of the element formulation one obtains the highest powers occurring in
the matrices B simply with the derivatives of the highest powers occurring in the shape
functions. The dyadic product BT B contains, naturally, the sums of the highest powers
c ETH Z

urich IMES-ST, March 26, 2014

62

Numerical preprocessing

(squares of the polynomial terms) and these determine the Gauss quadrature rule.
According to Table 7.1 the linear triangular element requires the one-point rule only, the
bi-linear quadrilateral and quadratic triangular elements the two-point rule, respectively,
and the quadratic Serendipity element the three-point rule. It appears that the respective
rules can integrate polynomials one order higher than the specific deformation energy terms
contain. So to say there is a buffer which is welcome as it can reduce the integration errors
arising with element shape distortion. Shape distortion can render the specific energy
terms rational functions which topic is discussed in Section 7.4.

7.3.2

Integration of the right-hand side

When integrating the boundary terms it may be assumed that the higher polynomial order
of the specific energy terms determines the integration rule. The remaining question is:
How complex may the stress distributions along the element boundary be which can be
exactly integrated by the given rule?
Consider for instance the boundary term (3.21) of a quadrilateral element of a rectangular
shape. Its shape functions (5.25) change linearly along the local coordinates or . The
two-point Gauss rule necessary for this element integrates exactly polynomials up to the
third order. Thus, the kinematically equivalent nodal forces will be integrated exactly if
the external stress distribution is not more complex than a second-order polynomial.

7.4

Element shape distortion and integration error

The discussion in Section 7.3.1 on numerical integration and finding the appropriate
quadrature rules presumes undistorted elements. Then, the elements of the Jacobian
matrix (3.30) are homogeneous within the element domain.
Generally, element distortion causes inhomogeneity of the Jacobian so that its inverse
contains rational functions (polynomials divided by other polynomials). Only the linear
triangular element is not prone to integration error which explains why it can be integrated by hand. The integration error is in the following studied on the bilinear quadrilateral element and Fig. 7.6 illustrates the influence of a parameter within the range
1 + D ,1 + D

0 , 1

0 , 1

1 + D ,1 + D

0 , 1

1 + D ,1 + D
D = 1

0 , 0

D = 0

1 , 0

0 , 0

D = -0 .5

1 , 0

0 , 0

Figure 7.6: Bilinear quadrilateral element with distortion parameter

c ETH Z

urich IMES-ST, March 26, 2014

1 , 0

7.4 Element shape distortion and integration error

63

and y
are
0.5 < . The global coordinates of the nodal points x

0
0

=
=
x
,
y
.

1
+

1
+

(7.13)

The derivatives of the shape functions (5.25) with respect to the local element coordinates
and are

(1

)
(1

(1 )
(1 + )
1
1
,
, =
.
(7.14)
, =

4
4
(1
+
)
(1
+
)

(1 + )
(1 )

With these one can write the Jacobian matrix (3.30) as a function of the distortion parameter ,
"
# "
#
(1 + )
x, y,
1 2 + (1 + )
=
,
(7.15)
J=
4
(1 + ) 2 + (1 + )
x, y,
and obviously = 0 obtains the expected result of a diagonal matrix with constant entries.
The determinant of the Jacobian (7.15) is
J=

1
+ (2 + + ) ;
4
8

(7.16)

it becomes zero at = = 1 if = 0.5 which is illustrated at the right in Fig. 7.6. In


this case the inverse of the Jacobian,
"
# "
#
2
+
(1
+
)
(1
+
)
,
,
x
x
1
=
J1 =
,
(7.17)
1+
(1 + ) 2 + (1 + )
,y ,y
2 (2 + + )
becomes singular. The matrix B connects the node displacements with the element strain
distributions. It contains the global derivatives ,x and ,y of the shape functions in
terms of the local coordinates and . Presently it suffices to consider the shape function
1 only. Its derivative with respect to xis
1 ,x = 2

1 + ( )
1 + 0.5(2 + + )

(7.18)

and it is interesting to study the distribution of its square value across the element domain,
since this value has to be integrated with the the two-point Gauss rule. The distribution is
presented in Fig. 7.7 along the line = 0.5773502692, and the various curves correspond
with respective values of the distortion parameter as indicated in the figure. All curves
are fitted with quadratic polynomials and the values R give the correlations. Severe
deviation from unity, namely with R = 0.6627, arises only for = 0.5 where the
quadrilateral deteriorates to a triangle.
Since the two-point Gauss rule integrates polynomials up to the third order exactly, the
actual integration errors are much smaller than those presented in Fig. 7.7. The deviations
of the correlations from unity of quadratic and cubic fits are presented in Fig. 7.8. The
result justifies the conclusion that the bilinear finite element integrates quite accurately
over a wide range of reference shape distortion.
c ETH Z

urich IMES-ST, March 26, 2014

64

Numerical preprocessing

3 5
D

3 0
D
D

In te g r a n d

2 5
D

2 0

(F

1 5

,x(D = -0 .5 ))

D
2

P o

1 0

P o

P o
P o
0

-5 -1

-0 .5

0 .5

E ta

P o
P o
P o

= -0
= -0
= 0
= 0
= 0
= 0
= 1
ly .
ly .
ly .
ly .
ly .
ly .
ly .

.5 0
.2 5
.0 0
.2 5
.5 0
.7 5
.0 0
(D = -0
(D = -0
(D = 0
(D = 0
(D = 0
(D = 0
(D = 1

.5 0
.2 5
.0 0
.2 5
.5 0
.7 5
.0 0

R
)

)
)

)
)
)

= 0 .6
= 0 .9
= 1
= 1
= 0 .9
= 0 .9
= 0 .9

6 2 7
9 9 7
9 9 8
9 9 6
9 9 2

Figure 7.7: Distribution of integrand (1 ,x )2 and square fit

deviation from polynomials

1.0E-03
fit to quadratic polynomial

8.0E-04

fit to cubic polynomial

6.0E-04
4.0E-04
2.0E-04
0.0E+00
-0.25

0.25

0.5

0.75

element distortion measure

Figure 7.8: Fit of the intagrand to quadratic and cubic polynomials

7.5

Pre-processing program parts

The pre-processing serves the assembly of the numerical system of equations whose solution
simulates the structural response to specified boundary conditions.

7.5.1

Assembly of the global stiffness matrix

The system equations are assembled from the element equations within a loop on all finite
elements. A sample FORTRAN subroutine is given in Fig. 7.9. To save space, and because
only the essential structure of the program is here of interest, the common statements for
transferring data from one program unit to others have been removed.
The loop on elements is marked with 200. With in the nested loop 210 on the element
nodes the global coordinates of these are collected in XEL and Y EL. The integer array
ICONEL relates locally to globally numbered degrees-of-freedom.
Once the element geometry is known, the element stiffness matrix ELM(16,16) is provided
by the subroutine ONSTIFF and then added to the system matrix SYSMAT within the
loop 220. The system matrix is band storage form for symmetric matrices. Therefore
its number of rows equals the number of degrees-of-freedom NDGF and its number of
columns equals the half-band width NHBW. The matrix dimensions are determined whilst
c ETH Z

urich IMES-ST, March 26, 2014

7.5 Pre-processing program parts

65

S U B R O U T I N E O N A S S M B L
I M P L I C I T R E A L * 8 ( A - H , O - Z )
I N C L U D E ' S I Z E . F O R '
C

D I M
D I M
D I M
D I M
D I M
C

E N S
E N S
E N S
E N S
E N S

I O N
I O N
I O N
I O N
I O N

D O

1 1 0
1 0 0
C

1 0 0 I
R H S ( I
D O 1 1
S Y
C O N T I
C O N T I N U E

N D G F
N H B W
C

P O I
I N B
S Y S
X E L
E L M

= 1 , N
) =
0 J =
S M A T
N U E

D O
C

2 1 1
2 1 0

D O

2 2 1
2 2 0
2 0 0

O N

2 1
X E
R E
I L
I S
D O

0
L (
L (
O
Y
2

I
C O N T
C O N T I N U

- - - - -

D G F
0 . 0 D 0
1 , N H B W
( I , J ) =

N T S
F N B
, N N
)
C O N

, 2 ) , I C O N ( N N E L , 9 ) , E L A S T I C I T Y ( 4 , 2 )
C ( N N N B C ) , I E B C ( N N E B C ) , D E B C ( N N E B C )
H B W ) , R H S ( N N D G F )
E L ( 1 6 )

0 . 0 D 0

T H E

N U M B E R

O F

E L E M E N T S

K E L = 1 , N E L

C A L L
C

P O I
C ) ,
D G F
L ( 8
) , I

2 0 0
D O

( N N
N N B
( N N
, R E
, 1 6

0
=

- - - - - L O O P
C

N T S
C ( N
M A T
( 8 )
( 1 6

I 8
I 8
I 8
=
=
1 1
C O
I N
E

= 1
) =
) =
( I
( I
I
N E
U E

, 8
P O
P O
8 C O
2 =
L (

I N
I N
1 )
N (
1 ,
I L

T S
T S
* 2
K E
2
O +

( I C O N ( K E L , I 8 ) , 1 )
( I C O N ( K E L , I 8 ) , 2 )
L , I 8 ) - 1 ) * 2
I 2 )

I S Y

I 2

O N S T I F F

A S S E M B L E

2 2 0 I
I S
=
D O 2 2
J S
J B
I F
I F
N H
N D
S Y
C O N T I
C O N T I N U E
C O N T I N U E
R E T U R N
E N D

1 6 = 1
I C O
1 J 1
=
=
( J B
( E L M
B W =
G F =
S M A T
N U E

S Y S T E M
, 1 6
N E L
6 = 1
I C O
J S
( I 1
M A
M A
( I S

O F

E Q U A T I O N S

( I 1 6 )
, 1 6
N E L ( J 1 6 )
- I S + 1
. L
6 , J 1 6 ) . E
X 0 ( N H B W ,
X 0 ( N D G F ,
, J B ) = S

I N

B A N D

S T O R A G E

M O D E

T . 1
) G O T O 2 2 1
Q . 0 . 0 D 0 ) G O T O 2 2 1
J B )
I S )
Y S M A T ( I S , J B ) + E L M ( I 1 6 , J 1 6 )

Figure 7.9: Subroutine for assembly of the global stiffness matrix


assembling the system stiffness matrix.

7.5.2

Calculation of element stiffness matrices

The element stiffness matrices are calculated by the subroutine ONSTIFF shown in Fig.
7.10. The routine considers the quadratic Serendipity element for analyzing problems with
rotational symmetry. It uses three-point Gauss quadrature and the arrays GP3(3) and
GW3(3) contain the Gauss-point coordinates and weights. The element stiffness matrix
is initialized with loop 100.
The integration ensues within two nested loops. The outer loop 2000 runs on the three
values of the local coordinate and the inner, 2100, on those of so that all nine Gauss
points are visited. At each of these points the matrix B as well as the determinant J of the
c ETH Z

urich IMES-ST, March 26, 2014

66

Numerical preprocessing

S U B R O U T I N E O N S T I F F
I M P L I C I T R E A L * 8 ( A - H , O - Z )
C

D I M E N S I O N
D I M E N S I O N
D I M E N S I O N
C

D A T
0 .
D A T
1 0 .

N G 3 , ( G
8 8 7 2 9 8 3 3
A
( G
2 7 7 7 7 7 7 7

P I

1
C
C

1 0 0 I
D O 1 1
E L
C O N T I
C O N T I N U E

D O

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

1
0
0

P 3 (
4 6 D
W 3 (
7 8 D

I ) , I = 1 , 3 ) / 3 , 0 . 1 1 2 7 0 1 6 6 5 4 D 0 , 0 . 5 D 0
,
0 /
I ) , I = 1 , 3 ) /
0 . 2 7 7 7 7 7 7 7 7 8 D 0 , 0 . 4 4 4 4 4 4 4 4 4 4 D 0 ,
0 /

D A T A N ( 1 . 0 D 0 ) * 4 . 0 D 0

D O
1 1 0
1 0 0

E L M ( 1 6 , 1 6 ) , X E L ( 8 ) , Y E L ( 8 )
C E L ( 3 , 3 ) , F F ( 1 6 , 1 6 ) , W O R K ( 1 6 , 3 )
B M A T _ L ( 1 6 , 3 ) , B M A T _ R ( 3 , 1 6 ) , G P 3 ( 3 ) , G W 3 ( 3 )

= 1 , 1 6
0 J = 1 , 1 6
M ( I , J ) = 0 . 0 D 0
N U E

2 0 0 0
D O 2 1
C A
C A
C A
C A
D O

I G = 1 ,
0 0 , J G
L L O N
L L O N
L L M A
L L M A
2 1 1 0
D O 2
E
C O N T
C O N T I N U
C O N T I N U E
C O N T I N U E

N G
= 1
S H
C M
T M
T M
I
1 1
L M
I N
E

3
, N
A P
A T
L D
L D
= 1
1
( I
U E

G 3
E (
( R
0 (
0 (
, 1
J =
, J
6

G P 3 ( I G ) , G P 3 ( J G ) )

)
B M A T _ L , 1 , 1 6 , 1 , 3 , C E L
, 1 , 3 , W O R K )
W O R K ,
1 , 1 6 , 1 , 3 , B M A T _ R , 1 , 1 6 ,
F F )
1 , 1 6
) = E L M ( I , J ) + F F ( I , J ) * D J * G W 3 ( I G ) * G W 3 ( J G )

R E T U R N
E N D

Figure 7.10: Subroutine for calculating element stiffness matrices


Jacobian matrix J are calculated by the called subroutine ONSHAPE. The first step then
forms the matrix product BT C and the second one the integrand BT CB. The innermost
loop 2110 runs on the degrees-of-freedom, multiplies the contributions of the individual
Gauss points with the area increments dxdy = Jdd, and adds the results to the stiffness
matrix ELM.

7.5.3

Calculation of matrix B and Jacobi-matrix determinant

The matrices B along with the necessary Jacobian transformations are provided by the
subroutine ONSHAPE shown in Fig. 7.11. First, the derivatives , and , are calculated
and written onto the arrays RNX and RNY. Next, the operations (3.30) are carried out,
the entries of the Jacobian assigned to the variables EL11, EL12, EL21 and EL22, the
c ETH Z

urich IMES-ST, March 26, 2014

7.6 Pre-processing numerical effort

S U B R O U T I N E
C

R N X
R N X
R N X
R N X
R N X
R N X
R N X
R N X

R N
R N
R N
R N
R N
R N
R N
R N

C
C

2 1 0
C
C
C

R (
R (
R (
R (
R (
R (
R (
R (

=
=
=
=
=

- 1 .
4 .
4 .
4 .
4 .
=

0 D
0 D
0 D
0 D
0 D

1 )
2 )
3 )
4 )
5 )
6 )
7 )
8 )

=
=
=

=
=

0 D
0 D
0 D
0 D

- - - - - D E T E R M I N A N T

O F

T H E

D J

0 D
0 D
0 D
0 D
= 1
1 1
1 2
2 1
2 2

E L 1 1

E L
= E L
= - E L
= - E L

2 0 0 I 8
I 1 6 =
D O 2 2 0
B M A
B M A
B M A
B M A
C O N T I N
C O N T I N U E

0
, 8
+
+
+

E L 2 2

/ D
/ D
/ D
/ D

= 1
( I
J
T _
T _
T _
T _
U E

0
-

W I T H
+
+

3 . 0 D 0 * R
4 . 0 D 0 * X
Y )
* ( 1 .
Y )
Y ) * ( 1 .
+

5 .
1 .
3 .
1 .
X
X
X

J A C O B I A N

* X
* X
* X
* X

T O

A X I A L

4 . 0
4 . 0
4 . 0
4 . 0

D 0 *
D 0 *
D 0 *
D 0 *

0 D 0

0 D 0

R E S P E C T
0 D 0
0 D 0
0 D 0
0 D 0
) * (
)
* (
)

B M A T _ R ( 3 , 1 6 ) ,

X
X

2 .
2 .
2 .
2 .

0 D
0 D
0 D
0 D

2 .
2 .
2 .
2 .
)

0 D 0
0 D 0
0 D 0
0 D 0

0 *
0 *
0 *
0 *

R *
R *
R *
Y *

R
R

2 . 0 D 0 * X

R A D I U S

4 . 0
4 . 0
4 . 0
+ 4 . 0
1 . 0 D 0 -

R N R ( 8

D I R E C T I O N

Y *
Y *
Y *
Y *

2 . 0 D 0 * X

T O

R N X ( 8 ) ,

D 0 *
D 0 *
D 0 *
D 0 *
2 .

1 . 0 D 0

Y * X
Y * X
Y * X
Y * X
0 D 0 *

+
Y

+
-

2 . 0 D 0 * Y

* X *
* X *
* X *
* X *
X

X
X

M A T R I X

0
0

2 2
1 1
1 2
2 1

R E S P E C T

4 . 0 D 0 * X

0
* ( 1 . 0 D 0
0 * X * ( 1 . 0 D 0
0 * X
0 * X * ( - 1 . 0 D 0

) * R
) * R
) * R
) * R

0 .
0 .
0 .
0 .
I 8
E L
E L
E L
E L

4 . 0 D 0 * Y
4 . 0 D 0 * Y

( I 8
( I 8
( I 8
( I 8

D O

2 2 0
2 0 0

X E L
R E L
X E L
R E L

1 1 =
1 2 =
2 1 =
2 2 =
2 1 0
1 1 =
1 2 =
2 1 =
2 2 =

- - - - - D E F I N E

T H E

E L
E L
E L
E L
D O
E L
E L
E L
E L

W I T H

D E R I V A T I V E S

4 .
4 .
4 .
4 .

B M A T _ L ( 1 6 , 3 ) ,

5 . 0 D 0 * Y
1 . 0 D 0 * Y

0 * Y * ( - 1 . 0 D
0 * Y
0 * Y * ( 1 . 0 D
0 *
( 1 . 0 D

- 3 . 0 D 0
- 1 . 0 D 0

=
=

O F

F X X
F Y Y
F Y X
F X Y
C

Y E L ( 8 ) ,

D E R I V A T I V E S

- 3 . 0 D 0

- - - - - F A C T O R S

- - - - - E L E M E N T S

( 1 )
( 2 )
( 3 )
( 4 )
( 5 )
( 6 )
( 7 )
( 8 )

- - - - - P A R T I A L

X E L ( 8 ) ,

- - - - - P A R T I A L
C

C
C

O N S H A P E ( X , Y )

D I M E N S I O N

67

F O R
J

E L 1 2

N X (
N X (
N R (
N R (

I 8 )
I 8 )
I 8 )
I 8 )

J A C O B I A N
*

E L 2 1

T R A N S F O R M A T I O N

I N T O

L O C A L

C O O R D I N A T E S

J
J
J
M A T R I X

, 8
8 - 1
= 1
L ( I
L ( I
R ( J
R ( J

) *
, 2
1 6
1 6
+ 1
+ 2

O F

I N T E R P O L A T I O N

F U N C T I O N S

2
+ J ,
+ J ,
, I 1
, I 1

J +
J +
6 +
6 +

1 )
2 )
J )
J )

=
=

=
=

R N
R N
R N
R N

X (
X (
X (
X (

I 8
I 8
I 8
I 8

)
)

*
)

*
*

F X
F X
F X
F X

+
Y

Y
X

+
+

R N
R N
R N
R N

R (
R (
R (
R (

I 8
I 8
I 8
I 8

)
)
)

*
*

*
*

F Y
F Y
F Y
F Y

X
Y
Y

R E T U R N
E N D

Figure 7.11: Subroutine for calculating the matrix B B-Matrix and the Jacobian matrix
entries of the inverse Jacobian calculated and assigned to the variables FXX, FXY, FYX
and FYY. These are used to calculate the matrix B entries.

7.6

Pre-processing numerical effort

This section compares the numerical effort to be invested in the bilinear and quadratic
quadrilateral elements. The numbers of calculation steps of the individual operations N1
through N8 is estimated for both element types and recorded in Table 7.2. The total
number of operations is estimated with the formula
NF LOP S = N1 [N2 + 2N3 N4 + N5 + N6 + N 7 + N8 ]

(7.19)

c ETH Z

urich IMES-ST, March 26, 2014

68

Numerical preprocessing

factor
N1
N2
N3
N4
N5
N6
N7
N8

effort
integration points
stiffness matrix entries
shape functions
terms in each shape function
B-matrix entries
operations for calculating Jacobian
operations BT C
operations BT CB
total operations

number lin
4
64
4
4
16
64
72
192
1760

number quad
9
256
8
8
32
256
144
768
14256

Table 7.2: Numerical effort for calculation of one element stiffness matrix
and the results listed in the last row of the table indicates that the quadratic element
requires an effort eight times higher than the bilinear element.

c ETH Z

urich IMES-ST, March 26, 2014

Chapter 8

Degree-of-freedom coupling
Structural analysis often has to consider essential boundary conditions which cannot be
mapped by the method for implementing individual degrees-of-freedom explained in Section 4.6. In particular, linear couplings
ij u
j + i = 0

(8.1)

can occur which are exemplified with Fig. 8.1. Such couplings must also be modeled for
solving contact problems in order to prevent penetration of two or more distinct bodies.
Also in the case of describing rigid-body motions in context of solving harmonic vibration
problems with vector iteration, see Section 15.1.4, degrees-of-freedom are linearly coupled.
The couplings allow to eliminate degrees-of-freedom from the total set u
as the eliminated

ones can be expressed in terms of the remaining unknowns u


and the specified geometric
boundary conditions u
:
u
= Z
u + u

(8.2)
The constraining matrix Z contains the linear equations (8.1) for eliminating the coupled
degrees-of-freedom as well as the unit matrix I for assigning the remaining unknowns. The
number of its rows corresponds with the full set of degrees-of-freedom and the number of its
columns is less by the number of coupling equations and geometric boundary conditions.
The relations (8.2) are substituted into the discrete form of the total potential energy:
=



1
+u
) u
+u
u
T ZT K (Z
u + u
T ZT r
2

(8.3)

Taking the variation of with respect to the solution parameters is the same as deriving
the terms in (8.3) with respect to the independent unknowns u
. Requiring that the variation must vanish provides the numerical system of equations with a symmetric coefficient
matrix:
ZT KZ
u = ZT (r K
u) .
(8.4)
After resolving (8.4) for the reduced set of independent unknowns one obtains the dependent unknowns simply with (8.2). The following sample problem shall provide an idea of
how to construct the constraining matrix Z from (8.1).

8.1

Sample problem slanted support

The structure sketched in Fig. 8.1 underlies the geometric boundary conditions:
u
1 = 0

w
1 = w

1 = 0

u
2 + w
2 = 0

(8.5)

c ETH Z

urich IMES-ST, March 26, 2014

70

Degree-of-freedom coupling

2
1

x , u
z , w

Figure 8.1: Boundary conditions involving degree-of-freedom coupling


The condition at the right in (8.5) couples the displacements u
2 and w
2 , thus modeling

the slanted gliding support oriented at an angle of 45 . Of the total set of six degreesof-freedom only two remain independent: The rotation 2 and the deflection w
2 , if the
displacement u
2 is selected to depend on the deflection. The set of all degrees-of-freedom
u
are now expressed in terms of the remaining independent u
according to (8.2) by using
a constraining matrix C:

u
1
0 0
0

0 0

w
1
w




1
0
0
0
2

=
+
(8.6)
1 0 2
u
2
0

w

1 0
0

0 1
0
2
The first three equations describe the geometric boundary conditions for individual
degrees-of-freedom, the fourth expresses the dependence of u
2 from w
2 , and the last two
equations imply that w
2 and 2 remain independent unknowns. If the stiffness matrix K
of the combined beam-rod element is given with

1 0 0 1 0 0
0
0
0
0 0
0
0 0 0
0
0 0 0
6
3L 0 6 3L

2
2

EI 0 3L 2L
0 0 0
0 3L L EA 0 0 0
, (8.7)
+
K= 3

1 0 0
0
0
0 0
0
L 0

L 1 0 0
0 0 0
0 6
0 0 0
3L 0 6
3L
0 0 0
0 0 0
0 3L L2 0 3L 2L2
the reduced system (8.4) is written here with:





w
2
+ 6 3L
6
=

w,

3L 2L2
3L
2

AL2
2I

It is easily resolved by hand and yields the displacements:









w

3w

w
2
L
2L2 3L
6
=
=
3L

2
(2 + 3)L2 3L + 6
(2 + 3)L
The complete displacement solution follows simply from (8.2):

(2
+
3)L

0
u
=
.
3L

(2 + 3)L

3L

c ETH Z

urich IMES-ST, March 26, 2014

(8.8)

(8.9)

(8.10)

Chapter 9

Solution of the system equations


The FEM pre-processing step creates the system equations
Ku = f

(9.1)

in the main-processing step. The


which are solved for the unknown primary solution u
solution step is not at the core of the finite-element method; however the choice of solution
methods may influence the structural analysis efficiency. Therefore it is important to
understand the influence of stiffness matrix size and population structure on the best
choice of the solution algorithm.

9.1

Band width minimization

All nonzero stiffness-matrix entries lie within a band enclosing the main diagonal and
some number of secondary diagonals. In case of symmetric laminates the half-bandwidth
is often considered; the number includes the main diagonal plus the count of secondary
diagonals to one side of it. The bandwidth of a stiffness matrix depends on the global node
numbering; more specifically, the band width is determined with the maximum difference
of the global node numbers connected by any one of the finite elements, multiplied by
the number of degrees-of-freedom per node. After assembly the bandwidth can be greater
than the necessary minimum. The numerical effort of direct solution algorithms increases
quadratically with the bandwidth. For illustration consider the simplistic structure shown
in Fig. 9.1 where the nodes are numbered after different schemes. The corresponding
2 2

2 1

2 0

1 9

1 8

1 7

1 6

1 5

1 4

1 3

1 2

1 0

1 1

1 2

1 3

1 4

1 5

1 6

1 7

1 8

1 9

2 0

2 1

2 2

1 0

1 1

1 0

1 2

1 4

1 6

1 8

2 0

2 2

1 1

1 3

1 5

1 7

1 9

2 1

2
1

6
5

Figure 9.1: Node numbering schemes and stiffness matrix population structure
respective matrix population structures are shown to the right. The upper row illustrates
c ETH Z

urich IMES-ST, March 26, 2014

72

Solution of the system equations

the worst case since the lowest and the highest node numbers are assigned to the same
finite element. This element determines the half-bandwidth NHBW = 22. The scheme
shown in the middle row is an improvement as the numbers increase along the direction
with several elements. This gives a half-bandwidth of NHBW = 13 in every single elements.
An optimum numbering scheme is shown in the lower row, where the node numbers increase along the direction with only one element. The scheme obtains the smallest possible
half-bandwidth NHBW = 4.
The present considerations is completed with sketching the three matrices population
structures in band storage in Fig. 9.2. The band storage mode saves much memory, par-

Figure 9.2: Symmetric matrices in band storage mode


ticularly if the systems of equations are large and the band structure pronounced. For the
here considered symmetric matrices the main diagonal is placed in the first column of the
band-storage mode. The difference in bandwidth between the worst and optimum cases
illustrated in the figure causes the numerical solution effort to differ by a factor of about
30.
The definition of a profile is based on a more refined characterization of the matrix structure: It is the average of the bandwidths of each individual row:
Nequ
1 X
P =
NHBW (k).
Nequ

(9.2)

k=1

Utilization of the profile can accelerate solution processes as the values, related to the
present sample problems, listed in Table 9.1 indicate. Commercial FEM programs include
strukture
bandwidth
profile

Fig. 9.2 left


22
7.4

Fig. 9.2 center


13
7.9

Abb. 9.2 right


4
3.3

Table 9.1: Bandwidths and profiles of the population structures shown in Fig. 9.2
algorithms for minimizing the bandwidth [9] or the profile [10, 11]. These reorder the
global node numbering.
The numerical solution methods for linear systems of equations fall into two direct and
the iterative ones. The direct ones are most efficient for smaller systems with pronounced
band structure. The iterative methods can be more efficient than direct ones for very large
systems and when a pronounced band structure does not exist.

c ETH Z

urich IMES-ST, March 26, 2014

9.2 Direct solution methods

9.2

73

Direct solution methods

The principle of the direct solution methods is a triangular decomposition of the nonsingular matrix K of the system of equations Ku = f . The decomposition presents K as
a product of two triangular matrices, LR = K as Fig. 9.3 illustrates. However, if the
K = L R
R
K
R
L

u = r
u = y

Figure 9.3: Triangular decomposition principle


values of the main-diagonal entries differ very largely the accuracy of the decomposition
can be increased if division by small numbers is avoided through exchange of rows. Then
the decomposition is given with a permutation matrix P, namely LR = PK. The decomposition transforms the original system of equations Ku = f to an equivalent system
Ru = y. The question of how to achieve such a decomposition will be addressed later and
for now one can consider that
Ku = (LR)u = L(Ru) = f

(9.3)

and solve the system for the substitution y = Ru


Ly = f .

(9.4)

This is achieved with the process of forward substitution:

y1 =

f1
,
l11

yi =

1
fi
lii

i1
X

lij yj ,

i = 1, 2, , N.

(9.5)

j=1

Subsequently one can obtain the desired solution u from Ru = y by the process of back
substitution:

N
X
yN
1
uN =
,
ui =
yi
rij uj ,
i = N 1, 2, , 1.
(9.6)
uN N
rii
j=i+1

Fig. 9.4 illustrates both processes. The decomposition process explained in the following
sections differ from each other by drawing advantage from the particular properties of
matrices K to achieve computing time savings. It is important to note that that the right
and left (or upper and lower) triangular decompositions R and L of a banded matrix K
are always within the bandwidth of K.
c ETH Z

urich IMES-ST, March 26, 2014

74

Solution of the system equations

b a c k s u b s titu tio n

fo r w a r d s u b s titu tio n
K u

(L R )u

L (R u )
y = f
L

L y

R u

u = y

f
f
y

Figure 9.4: Principle direct solution

9.2.1

Gauss elimination

Gauss elimination applies to a system of equations

k11 k12 k1n


u1
k21 k22 k2n
u2

..

..
..
..
..
.
.
.
.
.

kn1 kn2 knn


un
the elementary row and column operations
to a right triangular matrix R:

r11 r12 r1n


0 r22 r2n

..
..
..
..
.
.
.
.
0

rnn

f1

f2

=
.

..

fn

(9.7)

indicated in Fig. 9.5 to transform matrix K

y1

y2

=
.
..

un
yn
u1
u2
..
.

(9.8)

Since, as a matter of course, the operations are also applied to the right-hand side the sys-

f
R

Figure 9.5: Gauss Elimination at a fully populated matrix


tem of equations Ru = y is created. The last equation of this system is easily resolved for
un and by successive substitution of the respective previous results the complete solution
is found step-by-step after the formula:

k
X
1
ui =
yi
rij uj .
(9.9)
rii
j=i+1

c ETH Z

urich IMES-ST, March 26, 2014

9.2 Direct solution methods

75

The number of floating-point operations (flops) for a fully populated matrix is n3 /3+O(n2 ).
For a banded matrix the effort is reduced and Fig. 9.6 illustrates that the right triangular

y
R

Figure 9.6: Gauss Elimination at a banded matrix


matrix remains within the half-bandwidth. The number of floating-point operations is
then given with n(2p2 + 0.5p 1) 0.5p + np.

9.2.2

Cholesky decomposition

The linear FEM usually creates stiffness matrices K which are both symmetric (K = KT )
and positive definite, or uT Ku > 0 for any non-zero array u. Then, Cholesky decomposition can be applied with advantage as it requires with n3 /6 + O(n2 ) asymptotically only
one half of the floating-point operations than Gauss elimination. There are two variants
of Cholesky decomposition: The first one is K = BT B and the second is K = BT DB.
The second one is numerically preferable as it avoids calculation of square roots.
First Cholesky decomposition variant
The method includes the following steps:
1. decompose K = BT B. For each j = 1, n in increments of 1
(a)
bjj

v
u
j1
u
X
= tkjj
b2ij
i=1

(b) for each k = j + 1, n in increments of 1


bjk

1
=
bjj

kjk

j1
X

!
bik bij

i=1

Fig. 9.7 imparts a visual picture of the process.


2. forward substitution f = BT b. For each j = 1, n in increments of 1
!
j1
X
1
bj = fj
bij bi
bjj
i=1

3. back substitution Bu = b
c ETH Z

urich IMES-ST, March 26, 2014

76

Solution of the system equations

(a)
bn
bnn

un =

(b) for each i = n 1, 1 in Increments of 1


1
ui =
bii

n
X

bi

!
bik uk

k=i+1

The determinant of K can now be calculated conveniently with


|K| = |BT ||B| = (b11 b22 bnn )2 .

1 1

1 1

1 2

1 3

k
k

2 2

1 2

2 3

1 3

2 3
3 3

= b

b

1 3

2 3

3 3

2 3

3 3

b
k

2 2

1 3

2 2

1 1
1 2

1 2

1 1

= b

2
1 1

k
k

1 1

1 1

1
=

1 2

2 2

1 2

1 1

2 2

- b

2
1 2

b
b

2 3

1 2

= b 11b

2 2

= b

1 2

1 2

+ b

2 2

k
k

1 3

= b 11b

2 3

= b

3 3

= b

1 2
2

1 3

3 3

1
=
=

1
=

1 3

1 3

1 1

(k

2 3

- b

1 2

3 3

- b

2
1 3

- b

1 3

2 2

2
2 3

1 3

+ b

1 3

+ b

2 3

2 2

+ b

2 3
2
3 3

Figure 9.7: Elucidation of the first Cholesky decomposition variant

Second Cholesky decomposition variant


The matrices and vectors have the following structures:

d1
1 b12

d2
1 b23

.
..
..
..
D=
,B =
.
.

dn1
1
dn

b12
b2n
..
.
bn1,n
1

,b =

z1

z2

,z =
.
..

bn
zn
b1
b2
..
.

Perform the following steps


1. decompose K = BT DB. for each j = 1, n in increments of 1
(a) for each i = 1, j 1 in increments of 1
i. h = kij
ii. bij = h/di
iii. for each k = i + 1, j in increments of 1 kkj is replaced by kkj hbik
(b) dj = ajj
Fig. 9.8 imparts a visual picture of the process.
2. forward substitution f = BT z, Db = z. for each j = 1, n in increments of 1
(a) zj = fj
c ETH Z

urich IMES-ST, March 26, 2014

9.3 Iterative solution methods

77

(b) for each i = 1, j 1 in increments of 1 zj := zj bij zi


(c) bj = zj /dj
3. back substitution Bu = b. for each j = n, 1 in increments of 1
(a) uj = bj
(b) for eachi = j + 1, n in increments of 1 uj := uj bji ui
The determinant of K is calculated with
|K| = |BT ||D||B| = |D| = (d1 d2 dn ) .

0
1

0
k

1 1

1 2

1 3

k
k

1 2
2 2

2 3

1 3

2 3
3 3

= b

b

1 3

0
1

1 2

1 1

d 3
0

= d

2 3

k
k

1 2
2 2

= d 1b
= d 1b

b
0

d 1b

0
2

0
d 1

0
d

1 2

1 2

d 2b
2

d
k

d 1b

1 2

+ d

1 3

1 2
2

1 2

k
k

1 1

= d 1b

1 3

2 3

= d 1b

1 2
2

1 3

1 2

= d 1b

= k
1

2 3

1 3

3 3

1 3

=
= k

1 2

1 3

2 2

- d 1b

1 2

2 3

1 3

+ d 2b

+ d 2b

2 3

k
=

1 3

=
= k

(k

2 3

- d 1b

1 2

1 3

2
3 3

- d 1b

2
1 3

- d 2b

2
2 3

2 3

+ d
3

Figure 9.8: Elucidation of the second Cholesky decomposition variant

9.3

Iterative solution methods

Iterative methods avoid the triangular decomposition process. Proofs of convergence of


some iterative solution methods are based on eigenvector and eigenvalue analyzes of a
matrix B. An eigenvector v of B is a non-zero vector which does not rotate if B is applied
to it: Bv = v. The scalar is an eigen value of B. If matrix B is applied many times
to an eigenvector v, Bi v = i v, the resulting vector i v will tend to zero if || < 1, or
else tend to infinity. Eigenvectors are linearly independent of each other and can thus be
used as a basis to express any vector x existing in their vector space in terms of a linear
combination: x = 1 v1 + 2 v2 + + n vn (expansion theorem). Repeated application
of B on x yields Bi x = 1 i1 v1 + 2 i2 v2 + + n in vn . if all eigenvalues are less than
one Bi x converges against zero. This explains the significance of the convergence radius
(B),
(B) = max|i |,
(9.10)
of a B for numerical analyzes.

9.3.1

Jacobi iteration

Stiffness matrices K created by the FEM satisfy the assumption that all main-diagonal
entries kjj are non-zero. Then, K is partitioned into a diagonal matrix D, which just
c ETH Z

urich IMES-ST, March 26, 2014

78

Solution of the system equations

contains the main diagonal of K, and a matrix E containing all remaining entries of K.
So it holds that K = D + E and from this one develops
Kx = r
Dx = Ex + r

(9.11)

x = D1 Ex + D1 r
x = Bx + z

wobei

B = D1 E,

z = D1 r.

The diagonal matrix D is easily and with little numerical work inverted. Fig. 9.9 illustrates
the operations on a sample matrix structure. A solution method must improve a starting
1

1
2

1
2

3
4

5
6

2
3

4
5

1
2

3
4

4
5

6
7

5
6

7
8

7
8

9
1 0

9
1 0

1 1
1 2

1 1
1 2

1 3
1 4

1 0
1 1

1 2
1 3

1 2
1 3

1 4
1 5

1 4
1 5

1 6

8
9

1 0
1 1

7
8

1 5

1 6

1 6

D
E

-1

=
B

Figure 9.9: Forming of the matrix B for Jacobi iteration


vector x which does not satisfy (9.11); the Jacobi method applies the recursion rule
x(+1) = Bx() + z

(9.12)

where it must be shown why or under what circumstances the sequence will converge. To
this end each iteration result is expressed as a sum of the exact solution x and the error
vector e() = x() x. Then it follows from (9.12) that
x(+1) = Bx() + z
= B(x + e() ) + z
= Bx + z + Be()

(9.13)

= x + Be()
e(+1) = Be() .
If the error is expressed in terms of a linear combination of eigenvectors it becomes,
together with the considerations in Section 9.3, obvious that it will, independent of the
starting vector, tend to zero if (B) < 1 and this is the case if

n
X
kij


max
kii < 1
1jn

column sum criterion

(9.14)

row sum criterion

(9.15)

i=1
i6=j

or


n
X
kij


max
kii < 1
1in
j=1
j6=i

applies.
c ETH Z

urich IMES-ST, March 26, 2014

9.3 Iterative solution methods

9.3.2

79

Gauss-Seidel iteration

The Gauss-Seidel iteration method improves the Jacobi iteration with an up-date algo(+1)
rithm. After the first entry of the new vector x1
has been calculated, it appears rational
(+1)
to overwrite with it that of the previous vector prior to calculating the second entry x2
as Fig. 9.10 illustrates. Continuing the up-date step for the calculation of all following

1
2
3
4
5
6
7
8
9
1 0
1 1
1 2

( i+ 1 )

x
B

(i)

w ith u p d a te s
+

Figure 9.10: Gauss-Seidel modification of the Jacobi method


entries gives the iteration rule
(+1)
xi

i1
n
X
X
bi
aik (+1)
aik ()
=

xk

x
aii
aii
aii k
k=1

(9.16)

k=i+1

It is called Gauss-Seidel or sequential method and generally converges faster than the
Jacobi method.

9.3.3

Method of conjugate gradients

The method of conjugate gradients finds the minimum of a continuous differentiable and
convex function. It can be used to solve linear systems of equations without decomposing
the matrix of coefficients. The solution of the linear problem

minimizes the functional

K
u=r

(3.21)

1 T
K
T r
=U W = u
uu
2

(3.2)

whereupon its gradient

= K
ur
(9.17)

u
vanishes. Fig. 9.11 gives a visual impression. Compared with direct solution, the method
of conjugate gradients reduces the space and time complexity to O(m) where m is the
number of non-zero entries in the stiffness matrix. The method guarantees finding of the
exact solution of a system of equations with n unknowns in n iterations. It is explained
with basic notions, derivations, visualization, proof of convergence, and conditioning by
0
Shewchuk [12]. Presently only the algorithm is summarized. First, a starting vector u
must be given. It must satisfy the geometric boundary conditions and therefore it is
(
u) =

c ETH Z

urich IMES-ST, March 26, 2014

80

Solution of the system equations

Figure 9.11: Quadratic functional


convenient to reduce the system of equations so that the specified displacements do not
appear in it (see Chapter 8). For the rest, the starting vector can be arbitrary even though
an existing good estimate may save many iterations and thus significantly contribute to
the solution efficiency.
In order to find the minimum of better solutions along search directions s are determined
successively,
(i+1) = u
(i) + (i) s(i) ,
u
(9.18)
where the step length (i) must be adjusted so that the derivative of with respect to
(i) vanishes:
(
u())

= 0

u

u
= T s(i)


(i)
(i) + (i) si r s(i)
= K u
=


i
(i) + K(i) s(i) s(i)

(9.19)

(i) =

(i)T si
s(i)T Ks(i)

The thus determined step length is substituted into 9.18 to obtain the improved solution
(i+1) . There, a new search direction is found in terms of a linear combination of
estimate u
the negative of the total-potential-energy gradient and the previous search direction. The
method of conjugate gradients guarantees finding of the exact solution within N iterations,
or line searches, where the N is the number of unknowns, if the search directions are linearly
independent of each other with respect to the matrix of coefficients K, or s(i)T Ks(i+1) = 0,
where Fig. 9.12 gives a visual impression and proof is given in [12]. The factor for the
linear combination is found with


s(i)T K(i+1)
s(i)T Ks(i+1) = s(i)T K (i+1) + s(i) = 0 =
s(i)T Ks(i)

(9.20)

The matrix K appearing in the denominator of the result in (9.20) can be eliminated, and
thus the evaluation of made more efficient numerically, by exploiting the fact that the gradients, obtained at each line search, are all orthogonal to each other, or (i+1)T (i) = 0.
The line search along s(i) obtains:


(i)T si
(i) + (i) s(i) r = (i) (i)T
Ks(i)
(i+1) = K u
s Ks(i)
c ETH Z

urich IMES-ST, March 26, 2014

(9.21)

9.3 Iterative solution methods

81

s
r

= s

(1 )

(0 )

(1 )

(1 )

s
x

(1 )

(0 )

(1 )

(1 )

Figure 9.12: Orthogonal (left) and conjugate (right) search directions


The scalar product of the new gradient with itself gives
(i+1)T (i+1) = (i+1)T (i)

(i)T si
(i+1)T Ks(i)
s(i)T Ks(i)

(9.22)

Because of the orthogonality of the gradients and after rearranging it is obtained that:
(i+1)T Ks(i)
(i+1)T (i+1)
=

.
s(i)T Ks(i)
(i)T si

(9.23)

Substituting the result into the expression for the factor in (9.20) eliminates the stiffness
matrix K in there. Furthermore, the previous search direction s(i1) and the present
gradient (i) are always orthogonal to each other so that
(i)T s(i) = (i)T (i) + (i)T s(i1) = (i)T (i)

(9.24)

A summary of the method of conjugate gradients is


(0)

= K
u0 r

(i)

s(0)

(i)T (i)
s(i)T Ks(i)

(i+1) = u
(i) + (i) s(i)
u
(9.25)

(i+1) = (i) (i) Ks(i)


s(i+1)

= (0)

(i+1) =

(i+1)T (i+1)
(i)T (i)

= r K
ui

Preconditioning of the system of equations


The deviation of the contour lines from a circular shape reflects in the eigenvalues of the
matrix K. If all contour lines are circles as the sketch to the left in Fig. 9.12 indicates one
obtains multiple eigenvalues 1 = 2 = = n and the condition number ,
=

max
,
min

(9.26)

is equal to one. It is intuitively clear that a system of equations with = 1 is exactly


solved with in only one iteration step. The worse a system of equations is conditioned, i.e.
the more the contour lines deviate form circles as the sketch to the right in Fig. 9.12 shows,
the greater the condition number becomes and the more iterations up to the maximum
number n are necessary. Preconditioning manipulates a given system of equations so that
c ETH Z

urich IMES-ST, March 26, 2014

82

Solution of the system equations

the condition number becomes as small as possible. The problem is to find a positive
definite matrix M which approximates the stiffness matrix K. Then the original system
of equations K
u = r is indirectly solved by solving instead the preconditioned system
M1 K
u = M1 r .

(9.27)

If (M1 K)  (K), (9.27) converges faster than the original system. Thereby it must be
noted that the matrix M1 K is generally non-symmetric even though both M and K are
symmetric. Then one can resort to the untransformed conjugate gradient method which is
also described in [12].
The remaining problem, of finding of a good preconditioner M without investing too much
computational effort in it, is an on-going research topic. A simple but not overly efficient
preconditioner is given with the main diagonal of K.
An approach in context of simulating sequential processes such as damage accumulation
in composite material laminates is explained in [2].

9.4

Vector iteration for eigenvalue problems

The method of vector iteration with matrix deflation [13] finds eigenvectors and eigenvalues
of an eigenvalue problem such as arises from harmonic vibrations:


9.4.1


K 2M u
= 0.

(9.28)

Vector iteration

(r) of a system possess the orthogonality


The method utilizes the fact that the eigenvectors u
property

= 0 s 6= r
(r)T (s)

,
r, s = 1, 2, ..., n
(9.29)
u
u
6= 0 s = r
wherefore they are linearly independent and can be used as a basis to express any vector
in the same vector space:
= c1 u
1 + c2 u
(2) + ... + cn u
(n) 6= 0 .
u

(9.30)

The sought eigenvectors satisfy the standard eigenvalue problem


(r) ,
D
u(r) = r u

D = K1 M,

r =

1
,
r2

r = 1, 2, ..., n.

(9.31)

(r) onto itself save a constant factor r =


The dynamic matrix D maps an eigenvector u
0 with D creates a significantly different,
1/r2 . Multiplication of an arbitrary vector u
1 . The operation can be analyzed with the help of the expansion theorem
rotated, vector u
(9.30):
n
n
X
X
r (r)
1 = D

u
u0 =
cr D
u(r) = 1
cr u
(9.32)
1
r=1

r=1

The factors r /1 of the respective eigenvectors drive the method because the eigenvalues
r are monotonically increasing 1 2 n . With increasing r, indicating increasing eigenvector complexity and circular eigen frequency , the fractions r /1 decrease so
c ETH Z

urich IMES-ST, March 26, 2014

9.4 Vector iteration for eigenvalue problems

83

that with repeated application of D the contributions of the higher eigenvectors become
ever smaller. This defines the iteration rule

n
n 
X
1 p
1 X p
1
r p
(r)
1
0 = p1
(r)
p =
= c1 u
+
D
up1 =
D u
cr u
(9.33)
u
r cr u
1
1

1
1 r=1
r=2
By choosing the integer p sufficiently great the first term of the sequence (9.33) becomes
dominant and therefore the first eigenvalue 1 is called dominant eigenvalue. It follows
lim

1
1
p = lim p1 Dp1 u
1 = c1 u
(1) .
u
p
1
1

(9.34)

(1) is obtained with sufficient accuracy, the correIf, in the limit, the first eigenvector u
sponding circular eigenfrequency is calculated from (9.31):
1
u

2
1 =
.
(9.35)
|D
u1 |
It remains the problem of finding the more complex eigenvectors and higher eigenvalues.

9.4.2

Matrix deflation

For finding the higher eigenvectors and eigenvalues the dynamic matrix D must be manipulated so that the iteration (9.34) convergences against the higher eigenvectors. The
manipulation is called matrix deflation; It forms a new dynamic matrix D(2)
(1) u
(1)T M.
D(2) = D 1 u

(9.36)

(1) is normalized to satisfy u(1)T Mu(1) = 1. The modified


Here the first eigenvector u
(2)
matrix D has the same eigenvalues as D(1) with the exception that 1 has been removed.
This is illuminated by multiplying (9.36) with an arbitrary vector u0 :
0 =
D(2) u

n
X

(r) =
cr D(2) u

r=1

n
X

(1)
cr D
u(r) 1 u

r=1

Because of the orthogonality property of the



= 0 s 6= r
(r)T
(s)
,
u
Mu
6= 0 s = r

n
X

(1)T M
cr u
u(r) .

(9.37)

r=1

r, s = 1, 2, ..., n

(9.38)

of the eigenvectors (9.37) reduces to


D(2) u1 =

n
X
r=1

(r) =
cr D(2) u

n
X

cr D
u(r) .

(9.39)

r=2

The dynamic matrix D(2) is called deflated matrix corresponding with the second eigenvalue. Since 2 is the dominant eigenvalue of D(2) the deflation process can be continued
(3) and eigenvalue 3 with
for obtaining the third eigenvector with u
D(3) = D(2) 2 u(2) u(2)T M .

(9.40)

The process is generalized with


D(s) = D(s1) s1 u(s1) u(s1)T M

s = 2, 3, ..., n .

(9.41)

c ETH Z

urich IMES-ST, March 26, 2014

84

Solution of the system equations

As an addendum it remains to understand the orthogonality property (9.38). Consider


two different solutions of the eigenvalue problem (9.28) and write these in the form
K
u(r) = r2 M
u(r) ,

(9.42)

K
u(s) = s2 M
u(s) .

(9.43)

(s)T and (9.43) with u


(r)T to obtain
Now multiply (9.42) with u
(s)T K
(s)T M
u
u(r) = r2 u
u(r) ,

(9.44)

(r)T K
(r)T M
u
u(s) = s2 u
u(s) .

(9.45)

Finally transpose (9.45), subtract the result from (9.44) and obtain
 (s)T

M
u(r) = 0.
r2 s2 u

(9.46)

The fact that in general the eigenfrequencies differ from each other obtains the orthogonality property (9.38).

c ETH Z

urich IMES-ST, March 26, 2014

Chapter 10

Substructure technique
When developing a new structural design it can be that part of the structure is already
fixed while another part of the same structure is still subject to iterative design improvement. Both parts are elastically coupled so that the structural behavior of both parts
can not be analyzed separately. The finite elements shown in Fig. 10.1, rendered in pink,

b b

a b

a a

1 2

1 4

1 6

1 8

2 0

2 4

2 8

3 2

1 1

1 3

1 5

1 7

1 9

2 3

2 7

3 1

4
6

1 0

2 2

2 6

3 0

2 1

2 5

2 9

1
3

Figure 10.1: Partitioning of a mesh for sub-structure technique

model the part subject to changes. They contribute to the stiffness matrix entries related
to degrees-of-freedom on all nodes marked with color red. The elements shown with orange and yellow colors model the fixed part of the structure. Therefore the stiffness-matrix
entries relating to the nodes marked with yellow are not changing during the design development. The elements shown in orange connect those with variable and those with fixed
properties.
The substructure technique eliminates the fixed system equations within an antecedent
solution process which can reduce the numerical effort of the structural analyzes of the
sequentially improving design solutions.
Two conceptually different approaches for solving the sub-structure problem will be shown;
both lead to the same results. The one approach follows partitions the system of equations as already indicated with Fig. 10.1, the other one replaces the fixed structure by a
continuous elastic support along the interface between the fixed and variable parts of the
structure.
Both approaches are explained by using the academic sample problem shown if Fig. 10.1.
The chapter concludes with application of the continuous elastic support model to a loadintroduction optimization problem.
c ETH Z

urich IMES-ST, March 26, 2014

86

Substructure technique

10.1

Partitioning of the system of equations

a
The system of equations K
u = r is partitioned with respect to those degrees-of-freedom u
whose stiffness-matrix coefficients Kaa may change with design changes, and those which
b:
are fixed, namely u


 

Kaa Kab
u
a
ra

=
,
Kba = Kab T .
(10.1)
Kba Kbb
u
b
rb
The coefficients Kab correspond with the coupling stiffness entries of the finite elements
which connect the fix and the variable parts of the mesh and which are shown in with an
orange color in Fig. 10.1.
The subsystem appearing in the second row of the partitioned system (10.1) is resolved
b of the fixed part with the objective of eliminating them so
for the degrees-of-freedom u
that they do not appear explicitly in further solution processes:
u
b = K1
a ) .
bb (ra Kba u

(10.2)

Substituting the result into the first row of (10.1) gives a reduced system for the degreesof-freedom u
a of the variable part only:

Kaa Kab K1
a = ra Kab K1
(10.3)
bb Kba u
bb rb .
Since the sub-matrix Kbb is not affected by the structural changes, it must be inverted
only once and K1
bb can be stored in memory for further use in future iterations. However,
sub-matrix Kaa must be newly assembled in each iteration and the matrix multiplications
illustrated with Fig. 10.2 must be performed:
= Kab K1 ,
K
bb

b,
r = ra Kr

= Kaa KK
ba .
K

(10.4)

If the reduced system of equations


ua = r
K

(10.5)

is much smaller than the original one, or nDGF a << nDGF b , the computing time savings
can be quite significant.
K

a a

b a

a b

b b

= K

a b

K
b

b a

b b

a b

b b

K r
b

= K K

b a

Figure 10.2: Partitioning of a structure for sub-structure technique


For a close look at the efficiency of the sub-structure technique consider the situation
shown in Fig. 10.1. The mesh nodes are on purpose not numbered to give the smallest
bandwidth for the total system but rather to disentangle the fixed and the variable parts
of the mesh as much as possible as Fig. 10.3 illustrates. If each node carries only one
degree-of-freedom the mesh carries nDGF = 32 degrees-of-freedom with a half-bandwidth
nHBW = 14. Partitioning obtains nDGF a = 10 and nHBW a = 4 as well as nDGF b = 22
und nHBW b = 6.
c ETH Z

urich IMES-ST, March 26, 2014

10.1 Partitioning of the system of equations

87

1
2

2
3

3
4

4
5

5
6

6
7

7
8

8
9

a b

9
1 0

1 0
1 1

1 1
1 2

1 2
1 3

1 3
1 4

=
K

1 5
1 6
1 7
1 8
1 9

a a

1 4

1 5
1 6
1 7
1 8
1 9

2 0

b b

2 0
2 1

2 1
2 2

2 2
2 3

2 3
2 4

2 4
2 5

2 5

2 6
2 7
2 8
2 9
3 0

2 6
2 7

b a

2 8
2 9
3 0

3 1

3 1
3 2

3 2

Figure 10.3: Partitioning of the system of equations and matrix operations


First the stiffness matrix Kbb must be inverted. This requires, with Gauss elimination and
back-substitutions, 4529 floating-point operations (flops). One flop includes a multiplica = Kab Kbb 1 requires,
tion and an addition. The forming of the unsymmetric matrix K
2
without utilizing the existing banded structure, nDGF a nDGF b = 4840 flops. Its population structure is shown in Fig. 10.4. Its product with Kba requires n2DGF a nDGF b = 2200

a b

- 1
b b

1 1
1 2
1 3
1 4
1 5
1 6
1 7
1 8
1 9
2 0
2 1
2 2
2 3
2 4
2 5
2 6
2 7
2 8
2 9
3 0
3 1
3 2

= Kab Kbb 1
Figure 10.4: Structure of matrix K
flops and obtains the symmetric matrix shown in Fig. 10.5. The operations for creating

b a

=
K

a b

- 1
b b

b a

Figure 10.5: Structure of matrix Kab Kbb 1 Kba


Kab Kbb 1 Kba must be performed only once and require the initial investment of 6040
= Kaa Kab K1 Kba ,
flops. Subtracting this matrix from Kaa obtains the matrix K
bb
illustrated in Fig. 10.6, of the reduced system of equations for degrees-of-freedom ua . The
has the bandwidth n
matrix K
= 7.
HBW K
The numerical size and resulting numerical effort of the original and reduced systems of
equations are summarized in Table 10.1. For the original system the optimum half bandwidth nHBW = 6 is used. With this sample problem the partitioning with its initial
effort of 10569 flops gives an efficiency advantage only if at least seven design solutions are
c ETH Z

urich IMES-ST, March 26, 2014

88

Substructure technique

1
2
2

3
3

4
4

5
5

6
6

7
7

8
8

9
9
1 0

a a

a b

- 1
b b

b a

1 0

Figure 10.6: Structure of matrix K


analyzed with the samller system. With larger systems occurring in practice numerical
savings can be much more significant.
structure
original
reduced

NDGF
32
10

NHBW
6
7

NF LOP S
2557
1046

Table 10.1: Numerical size and flops without and with sub-structure technique

10.2

Method of continuous elastic foundation

Another approach for eliminating degrees-of-freedom, so that these do not appear any more
in sequential solution procedures, is to replace of the respective mesh part by a continuous
elastic support. Fig. 10.7 indicates the support with hatched lines. The elastic support
K

E B

a a

4
2
1

6
3

8
7

1 0

Figure 10.7: Replacement of fixed substructure in Fig. 10.1 by elastic foundation


must exhibit structural properties which have the same effect on the degrees-of-freedom on
the nodes along the interface as the mesh part to be eliminated. Within a solution process
with several load cases the response of the fixed mesh part to unit forces indicated in Fig.
10.8 are calculated. are calculated. Each of the responses is in terms of the displacements

1 2

1 4

1 6

1 8

2 0

2 4

2 8

3 2

1 1

1 3

1 5

1 7

1 9

2 3

2 7

3 1

4
6

1 0

2 2

2 6

3 0

2 1

2 5

2 9

Figure 10.8: Application of unit forces to the mesh part to be eliminated


c ETH Z

urich IMES-ST, March 26, 2014

10.3 Substructure technique in structural optimization

89

uj (fi ) of all boundary nodes. They are used to assembly a compliance matrix SEB :
(SEB )ji = uj (fi ).

(10.6)

Naturally the matrix is fully populated because of the elastic coupling all nodes will displace even only one of them is subject to a force. The inverse is the stiffness matrix
representing the continuous elastic support,
KEB = SEB 1 ;

(10.7)

it must be stored for future use. It reproduces the influence of the eliminated part of the
mesh on the variable mesh of the structural part to be optimized and maps the influence
of the geometric boundary conditions on the eliminated structure. Therefore KEB is
solvable, which was silently assumed with (10.6). The structural analysis of the part to be
optimized requires only the solution of
ua = (Kaa + KEB ) u
K
a = ra .

(10.8)

For the sample problem given in Fig. 10.1 and if using the same node numbering one
obtains the same matrix population structures as Fig. 10.6 illustrates.
It is important to realize that the continuous elastic foundation forms a super element
coupling all degrees-of-freedom on the interface between the two partial structures or
meshes. The reduction of the number of degrees-of-freedom is penalized with an increase
of the bandwidth, which may be critical as the number of flops scale with the square of it.

10.3

Substructure technique in structural optimization

An example for a structural model, on which the method of continuous elastic support can
be applied with advantage, stems from an optimization problem for a bonded load introduction element, the so-called onsert. The onsert is axisymmetric with a threaded center
hole and is bonded to the face sheet of a sandwich structure. A study on onsert structural
strength [14, 15] considers the load case of an axisymmetric force, i.e. perpendicular to
the sandwich plane, only. The sandwich is supported at its circumference. The force that
the bond layer can transmit is to be maximized by an optimum onsert shape, see the mesh
plots in Fig. 10.9. The design changes concern the onsert only and all other parts of the

Figure 10.9: Structural model with onsert of the initial and optimized shapes
structure remain the same, which was seductive to try out the continuous elastic support
method. The results in Table 10.2 show, that, for the chosen mesh, the advantage of a
smaller number of unknowns and the disadvantage of a larger bandwidth almost cancel out
each other so that the numerical efforts of the original and reduced systems of equations
require nearly the same number of flops for resolving them.

c ETH Z

urich IMES-ST, March 26, 2014

90

Substructure technique

Struktur
original
fixed part, sandwich and bond layer
variable part, onsert

NDGF
10866
7650
3395

NHBW
154
82
274

NF LOP S
2690 4110 604
140 0090 700
2610 3170 636

Table 10.2: Efficiency of the sub-structure technique with the model of a bonded structure

c ETH Z

urich IMES-ST, March 26, 2014

Chapter 11

Post-processing
The solution of the system equations, explained in Chapter 10, obtains the primary solution of the node degrees-of-freedom. In case of a static structural analysis, they are
displacements and, if one thinks of beams, plates, or shells, rotations. The primary solution answers the question of how a structure deforms under applied forces or enforced
displacements but does not quantify material strains or stresses. The distributions of
strains and stresses are obtained by the post-processing step. It identifies the node displacements of each finite element where the connectivity matrix solves the problem of
mapping global degrees-of-freedom onto the local ones. The element node displacements
and the element shape functions provide with (3.17) the displacement fields within the
respective finite elements.
Note that the displacement field is continuous, at least for finite elements in standard
displacement formulation, across the element interfaces. However, because of the displacement fields being defined on element level, strains and stresses generally jump at
the interfaces. Since the true solution stresses are continuous as long as the material is
homogeneous, the question arises at which points within the finite element domain the
simulated stresses agree best with those of a true solution.

11.1

Calculation of strains

by
The strains  are determined with the element degrees-of-freedom u
= B
u.
 = DT u

(11.1)

Formally (11.1) can be evaluated anywhere within the element domain. However, the
results agree better with a true solution at certain points than at others.
For introducing the problem consider the following sample illustrated with Fig. 11.1 which
could be a drill gear loaded by its own weight only. Gravity creates a body force f = g.
The structural model truss suffices for the modeling. Then the homogeneous body force f
is simply multiplied with the constant cross-sectional area A to obtain the homogeneous
line load q = f A. The integral of it over length determines the internal force N . The upperend suspension must carry the total weight while the lower end carries no load. Because
of the cross-sectional area being a constant the strain is proportional to the internal force.
It has a maximum value at the upper suspension and decreases linearly to vanish at the
free end. Consequently the compatible displacement field must be a quadratic function.
In contrast, the FEM solution foresees only a linear displacement field within each truss
element and, if the structure is modeled with only two elements, one obtains the crude
c ETH Z

urich IMES-ST, March 26, 2014

92

Post-processing

N o d e

E , A , r

G a u s s P o in t

T r u e S tr a in

T r u e D is p la c e m e n t

F E M

F E M

S tr a in

D is p la c e m e n t

Figure 11.1: Truss model of a drill gear under gravity load


approximations of displacement and strain fields as they are also drawn in Fig. 11.1. Here
the displacement approximation errors vanish at the nodes and more problematic appears
the strain representation, which can only be constant within each finite element. At least,
they agree with the true strain distribution at the element center points. These points are
therefore optimum points for strain simulation.
A method for determining such optimum points for strain (or stress) calculation is due to
Barlow [16]. Consider the approximation u over a one-dimensional element with a linear
polynomial depending on the local coordinate 1 1,

 

1
1
T
, =
u = p , p =
(11.2)

2
and assume that a true solution v exists whose distribution follows the, closest more
complex, quadratic polynomial:

1
1

2
.
(11.3)
v = qT , q =
, =

2
3

First let the approximation u coincide with the true solution v at the nodes = 1 and
= 1:
= 1 1 2 = 1 2 + 3
.
(11.4)
=
1 1 + 2 = 1 + 2 + 3
The same equations read in matrix notation


1 1
1
1



1
2


=

 1
1 1 1

,
1
1 1 2
3

(11.5)

or short
P = Q.
c ETH Z

urich IMES-ST, March 26, 2014

(11.6)

11.1 Calculation of strains

93

There the matrices P and Q contain


values of , the distributions p and q:
T
p (1 )

..
P=
.
pT (n )

in their rows, corresponding with the respective

qT (1 )

..
Q=
.
.
T
q (n )

(11.7)

In order to express the parameters in terms of , equation (11.6) is multiplied with the
inverse P1 of the square matrix P:
= P1 Q.

(11.8)

The final step follows from answering the question at which points the derivative of the
true distribution agrees with that of its approximation. The derivatives of both are:
u, = pT ,

v, = qT , .

(11.9)

Equating both derivatives,


pT , = qT , ,

(11.10)

use of (11.8), factoring out of the common factor and forming of the transposed deliver
a system of equations for the position at which both derivatives have the same values:
QT PT p, = q, .

(11.11)

Returning to the concrete sample, one finds:


PT =

1
2

1 1
1
1

1 0
= 0 1 .
1 0


,

QT PT

Substituting the result (11.12) in (11.11) gives the equations


0 0
1
1
,
=


2
0

(11.12)

(11.13)

from which the optimum point = 0 at the center of the linear element is found, as Fig.
11.2 illustrates.
T r u e D is p la c e m e n t

T r u e S tr a in

F E M

F E M

D is p la c e m e n t

S tr a in

Figure 11.2: Optimal strain point within a linear finite element


The general scheme is now applied to the slightly more complex problem of finding the
optimal strain points of a quadratic finite element. Consider the approximation p and
assumed true solution q together with their respective derivatives:

1
0

1
0

1
p=
, p, =
and q =
,
q,
=
.
(11.14)

2
2
2

2
3

3
c ETH Z

urich IMES-ST, March 26, 2014

94

Post-processing

With the node coordinates 1 = 1, 2 = 0, and 3 = 1 one obtains


matrices P, Q and the product QT PT :

1 0
1 1 1
1 1 1 1

0 1
P = 1 0 0 , Q = 1 0 0 0 , QT PT =
0 0
1 1 1
1 1 1 1
0 1

with (11.7) the

0
0
.
1
0

Inserting of these matrices in (11.11) gives the equations

0
0

1
1
=
.
2
2

2
1
3

(11.15)

(11.16)

The desired result is found in


the fourth of these equations. One obtains from there the
two optimal points = 1/ 3 oder = 0.577350269. With the strains, calculated at
these points, one can according to Fig. 11.3 extrapolate a linear approximation of the
T r u e S tr a in
F E M

S tr a in

Figure 11.3: Optimal strain points within a quadratic finite element


strain distribution within the finite element. It is interesting to note that all points are
identical with Gauss points. Fig. 11.4 shows the points arrangements for integrating its
G a u s s Q u a d r a tu r e S u p p o r tin g P o in ts

O p tim u m

S tr e s s P o in ts

Figure 11.4: Integrations- und optimale Dehnungspunkte im quadratischen Element


stiffness matrix and for evaluating the strains.

c ETH Z

urich IMES-ST, March 26, 2014

Chapter 12

Reduced modelling techniques


Some structural situations, quite complex at first sight, contain an inner simplicity the
recognition and exploitation of which can yield drastic reductions of modelling effort and
numerical cost. The situations with inner simplicity include geometric periodicity in combination with globally homogeneous loading situations and generalized plane-strain states.
The geometry of periodic structures consists of the repetition of one geometric element
along one, two, or three spatial directions. If the deformation field in any one geometric
element is the same as in all others, the deformation of the whole structure follows the
same periodicity as the geometry. The geometric element may be represented by a unit-cell
model which can be smaller than the geometric element, if the latter contains symmetries.
Then, the modelling and structural analysis of a unit-cell model will be representative
for all unit cells or geometric elements and thus the whole structure, and the results can
be used to extract the homogenized response of the structure to loads. This can strictly
only be true if the structure is of infinite extension. Boundaries of the periodic structure
will generally induce edge effects perturbing the homogeneity of the averaged inner forces.
Unit-cell simulations will obtain the inner solution to a problem with finite domain and
ignore the edge effects. In order to make a unit-cell model actually behave as if it was
a part of the periodic structure, symmetry and periodicity conditions must be analyzed
and implemented via essential boundary conditions and degree-of-freedom couplings as
outlined in Section 12.1. A sample problem is presented in Section 12.1.1.
The student is familiar with the quite limiting definition of a plane-strain state where the
strains with subscript referring to the out-of-plane direction vanish, e.g. z = yz = xz =
0. A more general situation allows all strain components to exist and requires only that
they remain constant along one direction. The simplifications of the basic elasticity equations are outlined in Section 12.2. Test specimens made of multidirectional laminates can
deform quite complex and suffer from severe edge effects; however the deformations and
stresses remain constant along the longitudinal, or loading, direction so that the deformation in any one cross-section is the same as in all other cross-sections. The development of
a two-dimensional model and dedicated finite element formulation is explained in 12.1.1.
Calculation of the homogenized substitute plate properties of a periodically corrugated
sheet includes both of the inner simplicities: periodicity along the wavy direction and
generalized plane strain along the other. The development of a unit-cell model of it with
dedicated finite elements is explained in Section 12.3.
c ETH Z

urich IMES-ST, March 26, 2014

96

Reduced modelling techniques

12.1

Periodic systems

For a continuum presentation of a periodic structure the mechanical unit cell should be
defined so that it fits its adjacent neighbors without gaps. Fig. 12.1 illustrates an example

y
L
y

L
x

Figure 12.1: Illustration to periodicity


where the unit cells in the undeformed reference configuration are rectangles. It also
indicates that under homogeneous overall deformation, here shear, the unit cells must
still fit into each other continuously even though the contours of their boundaries may
take on complex shapes. Obviously this implies that the respective opposite edges of the
quadrilateral unit cells must undergo the same change of shape. The total deformation
is a superposition of the change-of-shape and a rigid-body component, which is dictated
by the overall deformation. The overall strains distribution must be homogeneous with a
compatible displacement field which can at most be linear:
u
x = u0x + xC11 + yC12 + zC13
u
y = u0y + xC21 + yC22 + zC23

(12.1)

= u0z + xC31 + yC32 + zC33

u
z

where the bar indicates the part of the displacement field corresponding to the homogeneous strains and where u0i denote rigid-body motion of the unit cell. For the unit cell
indicated in Fig. 12.1 it is chosen that u0x = u0y = 0. The parameters Cij we identifiy with
the components of the homogeneous parts of the strains:
u
x = u0x + xhx

h
+ yxy
xy

u
y

= u0y

h (1 ) + yh
+ xxy
xy
y

u
z

= u0z

h
+ xzx
zx

h (1 )
+ zzx
zx
h
+ zyz
yz

(12.2)

h (1 ) + zh
+ yyz
yz
z

h
The meaning of the factors ij is illustrated with Fig. 12.1, where the shearing strain xy
comes from the linear increase of ux along y which is expressed with xy = 1. Loading
is applied by specifying values to selected homogeneous strain components, for instance
h =
xy
xy , xy = 1.
The elasticity problem, solved by use of the finite element method, yields the complete
corresponding with
displacement field u including the periodic deviations from the part u
the homogeneous parts of the strain fields. The periodic solution must have the properties:

u(Lx , y, z) = u(0, y, z) +
ux
u(x, Ly , z) = u(x, 0, z) +
uy
u(x, y, Lz ) = u(x, y, 0) +

uz

where
ui are constants which are either specified or solved.
c ETH Z

urich IMES-ST, March 26, 2014

(12.3)

12.1 Periodic systems

12.1.1

97

Sandwich structure with honeycomb core

A typical sandwich structure, such as shown in Fig. 12.2, consists of two face sheets
and a core much thicker than those. Typically, the face sheets are made from materials

Figure 12.2: Honeycomb with thin GFRP weaving face sheets and Nomex honeycomb core
with high stiffness and strength whereas the core is made from some low-density material
with lower stiffness and strength. The low density is provided with either randomly (i.e.
foam) or periodically (i.e. honeycomb) structured materials. Such a sandwich structure
exhibits very high mass-specific stiffness and strength in bending, which is explained with
the parallel-axes theorem. As Fig. 12.3 The name honeycomb stems from the fact that,
approximately, testifies, the periodic structure of such material consists of approximate

w
L

Figure 12.3: Nomex honeycomb core material


hexagons, hence the name honeycomb. Fig. 12.4 illustrates the manufacturing process

Figure 12.4: Expansion process for making honeycomb structured materials and geometric
idealization
steps of honeycomb materials. Because of it the L walls, called nodes in the sketch on the
right in Fig. 12.4, consist of two base sheets plus a layer of bonding material inbetween
c ETH Z

urich IMES-ST, March 26, 2014

98

Reduced modelling techniques

them.
For the analysis of large structures containing sandwich plates it is convenient and computational efficient to use homogenized stiffness properties of a substitute plate. Another
objective might be the analysis of the deformations resulting from the interplay between
facesheets and core driven by humidity changes [17]. Both objectives can be reached by
using a unit-cell approach. Even though the periodic pattern is composed of hexagons, a
unit cell is more conveniently chosen to have a brick shape as suggested with Fig. 12.5
and justified with geometric symmetries. Also, if the sandwich is symmetric, its midplane
y
x
y

x
-

L
W

L
-

x
-

y
-

Figure 12.5: Nomex honeycomb core material


is also a plane of symmetry. As part of an exercise it is left to the student to formulate the
boundary and periodicity conditions for a mesh of finite elements mapping the unit cell.

c ETH Z

urich IMES-ST, March 26, 2014

12.2 Generalized plane-strain state

12.2

99

Generalized plane-strain state

Students of mechanical engineering or mechanics are familiar with the concepts of plane
stress and plane strain. For the latter, it is typically assumed that the direct strain component perpendicular to the modelling plane be zero. In terms of mathematical modelling,
the thus defined plane strain and plane stress models differ from each other only in the
definition of the entries of the stiffness matrix relating to each other the in-plane stress
and strain components, see Appendix A. The plane stress models underly the theories of
thin plates and shells whereas the specific plane stress models are useful for the analysis
of dams, for instance.
A generalized state of plane strain requires only that, along a generator (say: x) of a structure whose geometry remains constant with respect to the generator direction, all strains
remain constant. Global deformation fields compatible with the generalized plane strain
assumption are illustrated in Fig. 12.6. The indexing of the global deformation modes is
x

k
y

j
z

j
y

j
x

Figure 12.6: Generalized plane-strain compatible global deformation modes


meant for compact bodies: 0x is a homogeneous strain field, y and z are circular-bending
curvatures about the y and z axes, respectively, and x is the twist around the x axis.
From the figure the contributions of each deformation mode can be found:
u
x = x0x + zxy
u
y

xyz
1 2
2 x z

u
z

1 2
2 x y

yz y
zx x

yz z
+ zx z

(12.4)

+ xy x xy y

The global deformation modes contribute to the strain distributions:


x
y
z
yz
zx
xy

=
=
=
=
=
=

0x
uy ,y
uz ,z
uy ,z + uz ,y
ux ,z
ux ,y

yz

+ zy

+ x( z y )
+ y( x z )

+ z( y x )
(12.5)
The shearing strain yz contains a term which depends linearly on the generator x. Since
the generalized plane strain assumption requires that all strains must be constant along
the generator, it must be that y z = 0. The twist components as sketched in Fig. 12.6
can not exist independently of each other:
y x = xy ,

z x = xz ,

xz = xy

(12.6)

c ETH Z

urich IMES-ST, March 26, 2014

100

Reduced modelling techniques

and then the displacement fields corresponding to the global deformation modes become:
u
x = x0x + zxy
u
y

u
z

xyz
1 2
2 x z

1 2
2 x y

+ zxxy

(12.7)

xyxz

and the strains are:


x
y
z
yz
zx
xy

=
=
=
=
=
=

0x

yz

uy ,y
uz ,z
uy ,z + uz ,y
ux ,z
ux ,y

+ zy
(12.8)

yxz
+ zxy

The results can be abbreviated with


(x, y, z) ,
u(x, y, z) = u (y, z) + u

(y, z) = Lu (y, z) + (y, z)

(12.9)

The differential operator L here is:

0
()
y

L=

()
z

0
()
z

()
y

()
z

()
y

(12.10)

The same differential operator, applied to the stress components written in Voight notation
to describe the equlibrium equations, expresses that the stresses do not vary along the
generator x:
yx ,y + zx ,z = 0
LT = 0

y ,y

+ zy ,z

= 0

yz ,y

= 0

z ,z

(12.11)

Now the weak variational form can be written specific to the generalized plane-strain
problem:
Z
Z
Z
T
T T

=
(u ) L C (Lu + ) d
(u) f d
(u)T d
=0
(12.12)

The discretization of it gives the numerical system of equations:


XZ
XZ
XZ
XZ
T

T
T
B CBd
u =
B d +
f d +
NE

c ETH Z

urich IMES-ST, March 26, 2014

NE

NE

NE

T d

(12.13)

12.2 Generalized plane-strain state

12.2.1

101

Multidirectional FRP laminate test simulation

Fiber reinforced plastics (FRP) are composite materials with high degree of anisotropy. In
the direction parallel to the reinforcing fibers, mass-specific stiffness and strength values
are high. Highest performance as well as anisotropy are reached when the reinforcing fibers
are made of carbon and aligned unidirectional. The elastic coupling of the layers depends
on the angular position of the fibers with respect to the direction of applied strain, which
is illustrated in Fig. 12.7. The sheet materials are typically used as multidirectional lami-

Figure 12.7: Anisotropic coupling effects in unidirectional composites


nates. Because of mutal constraint, the individual layers within a laminate can not freely
exercise their elastic coupling behavior. Thus, individual layers of a laminate subjected
to a uniaxial stress testing situation as illustrated in Fig. 12.8, will suffer internal stresses
not only along the loading direction x but also the width direction y. At the free edges

F
x , u
x

z , u
z

y , u
y

Figure 12.8: Test specimen subject to unidirectional stress


y = b/2 the internals stresses y and xy are subjected to the natural stress-free boundary condition. Because of the equilibrium conditions, the existence of stress gradients y ,y
xy ,y causes interlaminar stress yz , xz , and z to come into existence at the free edges.
All of this is named free-edge effects. Free-edge effects influence the results of stiffness and
strength measurements and also the performance of laminated structures in general.
Simulation of the test situation of a multidirectional laminate with finite volume elements
may quickly lead to large numbers of degrees-of-freedom. At sufficiently large distances
from the clamps, the deformation state of any one cross-section must be the same as those
of all other cross-sections: There is a generalized plane strain state with the loading direction as generator. The circumstance allows analysis of edge effects with a two-dimensional
mesh and dedicated finite elements where three displacement degrees-of-freedom are defined on each nodal point.

c ETH Z

urich IMES-ST, March 26, 2014

102

12.3

Reduced modelling techniques

Corrugated sheets

Corrugated sheets are probably best known as inexpensive roofing materials. Airplane
alternative design explored in past times has relied on the highly increased bending stiffness
effected by the corrugation for introducing the necessary fuselage and wing structural
stability. Corrugation lends anisotropy to plates which is much higher even than that
of fiber-reinforced materials. Particularly, membrane stiffness along the corrugation is
much reduced or enduracble strains are much increased. Corrugated sheets offer candidate
solutions for flexible skins needed for morphing wing design, where a shape of change calls
for high flexibility along the cord direction and light-weight design criteria for high stiffness
along the span direction.
Fig. 12.9 obviates that the meshing of a corrugated sheet must lead to very large numbers

Figure 12.9: Meshing of a corrugated sheet


of degrees-of-freedom if it contains many periods of the corrugation pattern. On a large
scale, the structural properties of a corrugated sheet can be expressed with those of a
homogeneous substitute plate in terms of a combined membrane and plate stiffness matrix
as it appears in the theory of laminated plates. Fig. 12.10 indicates that the entries of
4
3

N
N

M
M
M

~
A

y
x y
x
y
x y

A~

= ~
B

B~

~
A

1 1

1 2

1 6

1 1

1 2

1 6

2 6

1 2
2 2
2 6

B
B

~
~
~

6 6

1 6

2 6

6 6

1 6

1 1

1 2

1 6

1 2
2 2
2 6

2 6

~
B

2 6

D
D

1 6

2 2

~
B

1 2

1 2

~
B

1 1

2 6

~
B

1 6

2 2

~
A

1 2

~
~
~

6 6
1 6
2 6
6 6

e

e
g

k
k

k

x
x

0
x

x
y

x y

Figure 12.10: Deformation load cases for calculating the entries of a substitute ABD matrix
the substitute ABD matrix can be calculated by use of a unit cell and load cases with
the sketched specified deformations. The unit-cell conventions are introduced with Fig.
12.11, where a corrugation pattern is shown which consists of circular segments. The
is calculated column-wise, where the entries of each column
initial-stiffness matrix ABD
correspond to a respective prescribed unit-value deformation component. For instance,
for calculating the first column Ai1 , the deformation vector {1, 0, 0, 0, 0, 0} is specified and
the simulation results are evaluated to obtain the line loads whose values are then equal
to the entries in the first column.
c ETH Z

urich IMES-ST, March 26, 2014

12.3 Corrugated sheets

103

k = 1 /R
u

x 2(y )
c

z (y )

u
z

R s in y

R c o s y

u
y

y (y )
y

y
P

y
0

Figure 12.11: Conventions of the corrugated-sheet unit cell


The solution to the corrugated-sheet reduced modeling problem is based on the unit-cell
approach as well as the utilization of a generalized plane-strain state. Load cases 1, 4,
and 6 specify the strain components 0x , x , and xy , respectively, with x being the out-ofplane direction. Mapping of the unit cell with a finite-element mesh (see Fig. 12.12) poses
s
1

i-1

s
2
3
4

0
1

2
3

li

Figure 12.12: Finite-element mesh of the corrugated sheet


some interesting problems, particularly if the sheet material is a laminate made from
fiber-reinforced composite materials. The finite-element formulation of the generalized
plane-strain problem is contained in the form (12.13). When implementing geometrical
boundary conditions it must be respected that the model be compatible with shell theory:
supports must be defined and line loads must attack at the midplane z = 0. Periodicity
requires that the local deformations along the boundaries u2 = constant, u3 at both
ends be the same. Fig. 12.13(a) illustrates the point by using a somewhat simplified
and illustrative example where the corrugated sheet is degenerated to a flat plate and
subject to load case 1. Because of the Poisson effect and the specified strain 0x the mesh

Figure 12.13: Midplane support and periodicity at load case 1


tends to contract along the y direction. Specifying only uy = 0 at the nodal points in the
midplane leads to a warping of the boundaries which implies discontinuity of the periodic
displacement field as the top of Fig. 12.13(a) shows. In the general case of a corrugated
sheet, as illustrated with Fig. 12.13(b) for the same load case, the ends must be allowed
c ETH Z

urich IMES-ST, March 26, 2014

104

Reduced modelling techniques

to rotate.
A laminate can be unsymmetric so that its midplane does not coincide with an interface
between two adjacent layers. If the finite elements are arranged so that only one element
row is used for each laminate layer, there would be no nodal points on the midplane. The
problem could be mitigated by meshing the central layer containing the midplane so that
nodal points are on the midplane. The other and more consistent method for solving the
problem is to support the mesh at the midplane even though there is no nodal point as Fig.
12.14 illustrates. This can be achieved with constraints on the nodal-point displacements
1

1 2

- 1 /- 1

4
8

1
z
5
9

1 3
5

1 6

1 1

h
9

1 4

1 5

6
2

1 0

1 ,1
2

Figure 12.14: Midplane support and periodicity at load case 1


at the shown boundary edge = 1 of the pertinent finite element, which is here of the
Lagrangian type with 16 nodal points. The displacements at the local coordinate m of
the point on the laminate midplane are
u(y = 0, z = 0) = (1, m )
u

(12.14)

At the element edge = 1 only the shape functions 1 , 2 , 5 , and 9 contribute to


the displacement field. Therefore, fixing the displacement components along x and y at
y = z = 0 gives the conditions:
1 u
x1 + 2 u
x2 + 5 u
x5 + 9 u
x9 = 0,

1 u
y1 + 2 u
y2 + 5 u
y5 + 9 u
y9 = 0

(12.15)

The penalty method minimizes a pseudo-potential constructed from the errors of the
unsatisfied constraints, which competes with the total potential energy for minimization:
= + P ,

T T u

P = P eT e = u

Taking the variation with respect to the solution


stiffness matrix,

1 1 1 2
1 2 2 2
KP = P
1 5 2 5
1 9 2 9

(12.16)

gives a contribution to the


parameters u
1 5
2 5
5 5
5 9

1 9
2 9

5 9
9 9

(12.17)

The nodal-point displacements must be identified with the global degrees-of-freedom numbers by use of the connectivity matrix. The penalty factor P must be chosen; it will
produce good results for practical purposes over a wide value range. Obviously, by letting
P 0 the constraints will not be respected. With increasing P the constraint error will
decrease but it must be kept in mind that, if letting P , the penalty pseudo potential
c ETH Z

urich IMES-ST, March 26, 2014

12.3 Corrugated sheets

105

will numerically dominate the total potential energy so that the elasticity problem will not
be solved correctly.
For a more thorough presentation of the corrugated-sheet set of problems the interested
student is referred to the publication [18].

c ETH Z

urich IMES-ST, March 26, 2014

106

c ETH Z

urich IMES-ST, March 26, 2014

Reduced modelling techniques

Chapter 13

Sequential processes
Under loading a structure will deform so that reference and deformed configurations are
not the same. Points of load attack move which alters the global equilibrium of forces.
The local strains may change the relations between them and the stresses. Even more
complicated can it be if these changes do not only depend on load but also on time.
For many practical engineering problems such changes are negligibly small and the systems
of equations K
u = r are linear as the stiffness matrix K is a constant.
The stiffness matrix is related to the total deformation energy which the structure stores
under given external loading. The deformation energy depends on the geometry of the
structure as well as on the material of which it is made and a minimum complexity of
nonlinear problems is K = K (u, (u, t)). There are several kinds of non-linearity: Material
non-linearity may arise at moderate strains (yield limit) or even moderate displacements
may cause non-linear displacement-strain relations (Fig. 13.1a), geometric non-linearity
may change the shape of a structure significantly even though the strains and stresses
remain small (Fig. 13.1b), and contact problems change, under increasing loading, the
geometric boundary conditions. Further complications, such as the dependence of load of
Q
F

(a )

(b )

(c )

Figure 13.1: Limits of linear modeling


the displacements are easily imagined where Fig. 13.2 gives an illustration.

Figure 13.2: Change of natural boundary conditions due to sagging of a supporting wall
Simulation of non-linear problems solves a sequence of linear problems. A load, triggering
non-linear effects, is applied in small increments. The changes of the structure or of the
relation between strains and stresses, appearing at the respective load steps, are taken into
c ETH Z

urich IMES-ST, March 26, 2014

108

Sequential processes

account with modifications of the stiffness matrix. Fig. 13.3 illustrates an error to which
the linear simulation of a geometrically non-linear problem may lead.
P
c o r r e c t w ith g e o m e tr ic a lly n o n lin e a r s im u la tio n
u
1

u
2

u
3

u
4

w r o n g w ith lin e a r s im u la tio n

Figure 13.3: Large deflection of a slender cantilever beam

13.1

Laminated structures damage progression simulation

The simulation of progressing damage, where an X-ray photograph of a momentary damage


state is shown in Fig. 13.4, in structures made from laminated fiber-reinforced plastics

Figure 13.4: Fatigue damage in a notched CFRP laminate [45, 90, 45, 0]s . Source [19]
as a typical nonlinear problem is here considered as it connects topics of the present and
other lecture classes of the major Lightweight Structures, namely:
-

mechanics of composite materials


structural analysis with FEM
failure criteria for unidirectional composites
damage models
iterative solution of system equations
pre-conditioning of the system matrix
efficiency of various equations solution methods

13.1.1

Damage modeling

Total failure of a multidirectional laminate requires fiber rupture. Before the onset of
fiber rupture matrix failures will occur, having little influence on structural strength under quasi-static loading if the fibers are very much stiffer and stronger than the embedding
matrix. Both damage modes are simplistically illustrated in Fig. 13.5. The first damage occurring in a multidirectional laminate is called first-ply failure. Prediction of the
c ETH Z

urich IMES-ST, March 26, 2014

13.1 Laminated structures damage progression simulation

109

b
a

Figure 13.5: Fiber rupture (a) and matrix failure (b) in a multidirectional laminate
structural strength, i.e. the highest load which can be transferred or critical load factor
on a given load case, requires a progressive damage analysis. It is an iterative repetition
of the search for the first-ply failure with subsequent modeling of the identified damage
increment.
Neither tensile nor shear forces can be transmitted perpendicular to the fracture surfaces.
A very crude damage model reduces Youngss moduli along the respective directions by a
factor of :
tensile fiber failure
compressive fiber failure
tensile matrix failure
compressive matrix failure

13.1.2

: E1 = E1 ,
:
: E2 = E1 ,
:

G12
G12
G12
G12

= G12 ,
= G12 ,
= G12 ,
= G12 ,

12

12

12

12

= 21

= 21

= 21

= 21

=0
=0
=0
=0

(13.1)

Damage states and displacement fields

Each damage event reduces the apparent material stiffness within its vicinity. For a finiteelement model, or mesh, for simulating the structure,
K
u = r,

(13.2)

the vicinity of a new failure event is the finite element within which it is predicted. Modeling of the damage increment will change its stiffness matrix to map the loss of material
stiffness and possibly residual stresses are relaxed. Consequently, these changes affect the
displacement of the whole mesh and with these the local material strains. Thus, each new
damage increment affects the spatial probability distribution of further damage events.
All of these changes are captured with the following equation:


 

K(0) + K u
(0) +
u = r(0) + r .
(13.3)
Magnitudes with superscript (0) relate to the undamaged state of the structure. Removing
the system equations relating to the initial undamaged state of the simulated structure,
K(0) u
(0) = r(0) ,

(13.4)

reduces (13.3) to the form





K(0) + K
u = r K
u(0) .

(13.5)

For further elaboration of (13.5) let it be accepted that the system (13.4) has already been
solved by direct methods to obtain the displacement field u
(0) . This implies invertibility
(0)
of the system matrix K which is the case if at least statically determinate support has
been realized with geometric boundary conditions. Whereas (13.3) calculates the total
c ETH Z

urich IMES-ST, March 26, 2014

110

Sequential processes

displacements corresponding to a given damage state, (13.5) finds the deviations, caused
by the damage, from the initial displacement field found by solving (13.4). The changes
u
evoked by consecutive damage events modify in increments the initial solution calculated
with (13.4).

13.1.3

Preconditioning of the system of equations

If the difference
u between two consecutive displacement fields is small enough the
numerical effort of iterative solution methods can be smaller that that of direct ones
because the existing solution provides an excellent guess for the new solution changed
by the current damage event. In addition, the decomposed form of K(0) , obtained by
a direct solution step, can be used as a pre-conditioner for improving the convergence
of iterative methods to obtain an approximate solution of (13.5). Application of K(0) to
1
(13.5) shows that only those finite elements that were modified to map damage K(0) K
cause deviation from the unit matrix:




1
1
r K
u(0) .
(13.6)
u = K(0)
I + K(0) K
However, the deviation is generally not symmetric even though both matrices, K(0)
K, are symmetric. The left-hand side of (13.6) can be evaluated with


I + K(0)


K
u =
u + L
u

and

(13.7)

where the operation L


u first creates the vector K
u by matrix multiplication on
element level and secondly the back-substitution process is applied on it.

13.1.4

Notched tensile test CFRP specimen model

The notch in the center of a tensile test coupon sketched in Fig. 13.6 creates stress con2 R
y
x

2 W

2 L

Figure 13.6: Test specimen geometry and quarter-model FE mesh


centrations, which will provoke early damage events; under further load increase damage
will spread. The FEM mesh utilized the double symmetry of the test specimen situation.
The clamps enforce the specimen ends to remain straight. Therefore, load is applied in
terms of inhomogeneous essential boundary conditions.
The investigated laminate designs are given in Table 13.1 and the material properties of
the basic unidirectional composites of which the laminates are made are listed in Table
13.2.
c ETH Z

urich IMES-ST, March 26, 2014

13.1 Laminated structures damage progression simulation

No.
1
2
3
4
5

111

Table 13.1: Laminates


name
stacking sequence
unidirectional transverse
[90]
unidirectional longitudinal
[0]
symmetric cross-ply
[0, 90]s
quasi-isotropic
[0, 45, 90]s
bi-directional
[45]s

Table 13.2: UD T300 material data (DORNIER SYSTEM)


property
unit
1/23
2/13
3/12
Youngs moduli
M P a 135000 10000 10000
shear moduli
MPa
3846
5000
5000
Poissons ratios

0.3
0.27
0.27
1
6
6
thermal expansion coefficients K
0.6
30.0
30.06

Figure 13.7: Damage pattern unzip action after 15 damage events

13.1.5

Simulated damage states

The simplest damage pattern appears with the unidirectional reinforcement transverse to
the loading direction, [90]. Starting from the notch edge, the laminate is simply torn into
two symmetric halves and the damage spread is similar to opening a zipper. If all fibers
are parallel to the loading direction, [0], the damage propagation is determined by other
damage mechanisms. Because of the low shear strength, if compared to the fiber tensile

a
b

Figure 13.8: Damage patterns simulated for various laminates


strength, of the material the shear stresses at the notch edge will trigger failure there and
create the inter-fiber failure aligned almost tangential to the notch edge, see Fig. 13.8a).
Similar trends appear with the cross-ply laminate [0, 90]s , see Fig. 13.8b), because here the
shear strength is as low as that of the uni-directional laminate. Within the quasi-isotropic
laminate [0, 45, 90]s the damage is spatially more confined which is due to the arresting
effect of the multidirectional fiber arrangement, see Fig. 13.8c). Within the bidirectional
laminate [45]s the cracks tend to spread along the 45 degree direction, see Fig. 13.8d).
c ETH Z

urich IMES-ST, March 26, 2014

112

13.1.6

Sequential processes

Conjugated gradient method solution efficiency

The solution efficiency has been studied with the unzip action damage propagation seen
with the 90 degree unidirectional laminate illustrated in Fig. 13.7, because this propagation
pattern remains the same for varying degrees of solution accuracy. The solution quality as
well as solution times of the conjugated gradient method (CG) have been compared with
those of a direct solution method (DS), namely Gauss elimination and back substitution.
A measure of the CG solution quality is given by the load factors predicted for triggering
the sequence of damage events, where the values obtained with direct solution are plotted
in Fig. 13.9 and those of the first five iterations with CG are shown in Fig. 13.10. The

Figure 13.9: Failure-load factors reference values after DS. Source: [2]
sequence of approximations converges with the numbers of iterations as shown in the plot
to the left of Fig. 13.10. The deviation of the approximations from the direct solution
results is measured with
25
X
s=
( DS )2
(13.8)
1

where a mesh with 25 finite elements between the specimens notch and its straight edge
has been chosen. The plot tn the right of Fig. 13.10 gives the rate of convergence, where

five iterations closing in at the exact results

convergence rates over first ten iterations

Figure 13.10: Failure-load factors: unzip action and CG. Source: [2]
the regression line indicates that, on average, every two iterations with CG reduce the
error by one order of magnitude.
The numerical efforts of the DS and CG solution methods expended into the unzip action
have been measured for various mesh sizes, where both the numbers of unknowns as well
as the bandwidth values changed so that the influence of both could be captured. The
mesh parameters, resulting mesh size indices, and computing time for direct and iterative
c ETH Z

urich IMES-ST, March 26, 2014

13.1 Laminated structures damage progression simulation

113

solution methods are listed in the Table printed in Fig. 13.11. For evaluating the influence

Figure 13.11: Mesh parameters, problem sizes, and computing times. Source: [2]
of size parameters on computing times a model is developed. It combines the knowledge
of what processing steps computing times are composed of with existing equations for
calculating numbers of floating-point operations (FLOPS) for the various processing steps.
The computing times decompose into the following parts:
- tGT

triangular matrix decomposition with Gauss elimination

- tBS

back substitution upon matrix decomposition

- tSP

computing of scalar products

- tF P F

next failure event prediction

Known relations are weighted with scaling factors , , , and for fitting measured
computing times:

2
tGT
= NDGF 2NHBW
+ 23 NHBW 1
tBS

NDGF NHBW

tSP

NDGF NIT ER

tF P F

NEL

(13.9)

The computing times of DS and CG depend on the number of failure events NF E and
follow from the sums:
tDS

NF E (tGT + tF P F )
(13.10)

tCG = tGT + (NF E 1) (tBS + tSP + tF P F ) + tF P F


The data listed in Table 13.11 are ordered after increasing computing times and shown in
Fig. 13.12 with open markers, namely direct solution with circles and iterative solution
with squares. The solid symbols show fitted model predictions. The diagram indicates
that solution time increases with problem size stronger in case of direct solution than in
case of conjugated gradients.
c ETH Z

urich IMES-ST, March 26, 2014

114

Sequential processes

Figure 13.12: Computing times measurements and fitted model predictions. Source: [2]
The model is used, together with some simplifications, to study the inherent systematic
of the difference as is explained in the following. The simplifications come with introducing one model size parameter on which all other parameters depend by some reasonable assumptions. The model-size parameter N is the same as the number of degrees-offreedom NDGF . For large models the number of elements is approximately the same as
the number of nodes. Since the latter carry here two degrees-of-freedom each, it holds that
NEL = 0.5N . The values listed in Table 13.11 justify the assumption that the bandwidth

increases with the square root of the number of degrees-of-freedom,


or NHBW = N . The

number of failure events is also estimated NF E = N . The assumptions give a formula

tCG
1 (2 + ) N + (1.5 + 0.5) N

=
(13.11)
tDS
N
2N + 1.5 N + 0.5
for the computing time ratio as a function of problem size which is illustrated in the
diagram shown in Fig. 13.13. It indicates that CG computing times asymptotically scale

Figure 13.13: Scaling of CG computing time relative to those of DS. Source: [2]

less than those of direct solution by a factor of N . Due to the excellent preconditioning
and starting vectors the break-even point is on the order of N 1000.

c ETH Z

urich IMES-ST, March 26, 2014

Chapter 14

Finite beam elements


The bending problem adds interesting aspects to the FEM: The shear rigid beam finite
element transfers the deflection line as well as its derivative continuous over nodes, thus
possessing C 1 continuity. On the other hand, the shear compliant element derivation offers
an opportunity to study a locking effect and its mitigation with a reduced integration
technique. The following Sections 14.1 through 14.5 recall beam theory.

14.1

Equilibrium at a small beam element

The structural element beam differs from that of the rod by having its load transverse to
its axis rather than longitudinal. For geometric reasons the transverse forces cannot exist
without being accompanied by internal bending moments. Fig. 14.1 gives the equations
M

M + d M

Q + d Q

d x
y

Q + d Q

d x
y

x
z

x
z

F : - Q
: - M

+ q d x = 0

Q = - q

+ d M ) - Q d x = 0

M = Q

+ (Q + d Q )
+ (M

Q + q = 0 ,

M - Q = 0

Figure 14.1: Equilibrium of forces and of moments at small beam element


0

of equilibrium under the influence of a line load q. The superscript prescribes taking the
derivative with respect to the beam axis coordinate x.

14.2

Kinematics of bending and shear

The kinematic relations of the shear-compliant beam are depicted in Fig. 14.2. The
first-order shear theory after Mindlin [20] is based on the unphysical assumption that
cross-sections remain plane under shear deformation [21].
c ETH Z

urich IMES-ST, March 26, 2014

116

Finite beam elements

b = 0
z

d e fle c tio n lin e s lo p e


c r o s s - s e c tio n a l r o ta tio n
s h e a r a n g le
g
b

g = w '

w '
k in e m a tic s :

w '= g - b

g = 0
w

g = w '+ b
u ( x , z ) = zb ( x )

b = - w '

e ( x , z ) = z b '( x ) = z k

u (z )= b z

Figure 14.2: Kinematic relations at an infinitesimal beam element

14.3

Integrated constitutive laws

Structural mechanical models relate stress resultants such as a longitudinal force N (truss),
torsional moment T (rod), transverse force Q (beam) or bending moment M (beam) to the
conjugate deformation measures to obtain structural element stiffness properties, which are
here dubbed integrated constitutive laws. The resistance against first-order-theory shear
deformation is derived with Fig. 14.3 and that against bending deformation in Fig. 14.4.

= b

= G b

t
2

t d z
2

= g G b
= g G b
= g G b t
Q

gd z
t
t

d z

[t

- (- t)

= G A g

b = 0
w
g = w '

Figure 14.3: Transverse force, shear stress and shear stiffness after first-order theory

14.4

Shear-stress distribution and shear correction factor

The shear stress equilibrates the transverse force but must vanish on the upper and lower
beam surfaces to satisfy the natural boundary condition there. Assuming cross-sections to
remain plane, or (z) = const, contradicts the natural boundary conditions wherefore the
value GA misrepresents the true shear stiffness. A shear correction factor arises from comc ETH Z

urich IMES-ST, March 26, 2014

14.4 Shear-stress distribution and shear correction factor

= b
M
=

E b
= k E b

117

s (z )zd z
t

e (z )zd z
t

z 2d z

3
3
1 t
t
= k E b - -
3 2
2
3
B t
= k E
1 2
= E Ik

- b

u (z )= b z

Figure 14.4: Bending moment, curvature, and bending stiffness


paring the unrealistic shear stress distribution with one that satisfies the natural boundary
conditions. The latter distribution is parabolic and Fig. 14.5 shows how it is constructed
e q u ilib r iu m

o f m o m e n ts

lin e a r b e n d in g s tr a in d is tr ib u tio n

M = Q = E Ik
Q

e ( x , z ) = z k ( x ) = z

H o o k e 's la w

s = E e

lo c a l e q u ilib r iu m

t ,z = - s

in te g r a tio n th r o u g h th ic k n e s s

t (z ) = -

s h e a r s tr e s s d is tr ib u tio n

t (z ) = -

E I

s d z =

Q
I

z
2

x z

z d z

t

2
2

Q
2 I

Figure 14.5: Derivation of a parabolic shear stress distribution


by integrating the local equilibrium condition with the bending stress gradient through
the thickness.

c ETH Z

urich IMES-ST, March 26, 2014

118

Finite beam elements

A shear correction factor is found by following the thought that the integral of the
specific deformation energy u over the cross-sectional area must be the same for both the
homogeneous and the parabolic distributions. Fig. 14.6 explains the integration of the

x z

u d zd x =
t

t
2

b Q
=
8 G I

t ( z )g ( z )d zd x =
t

t (z )2 d zd x

b Q
t
- d zd x =
8 G I
2
2

2 G

b Q 2 z 5 2 z
=

8 G I 2 5
3

t
t
+ z
2
2
2

z
t

t
t
- 2 z + d zd x
2
2
2

2 b Q
d x =
1 5 G I

2
2

t
d x
2

Figure 14.6: Deformation energy of the parabolic shear-stress distribution


parabolic shear-stress distribution and Fig. 14.7 explains that of the constant distribution.

x z

u d zd x =
t
1

t g fA d x =

2 G

Q

2 G A
1

fA d x =

fA d x = f

d x

2 G A

Figure 14.7: Deformation energy of the constant shear-stress distribution


Quantities with respect to the constant distributions are marked with an upper bar. The
shear-correction factor f , already introduced with Fig. 14.7, is determined by equating
the two different deformation-energy expressions as Fig. 14.8 explains.

x z

2 b Q
u d x = f
=
t
2
2 G A 1 5 G I
!

4 A b t
f =

1 5 I 2 2

2
2

t

2
4 tb

b t3
1 5
1 2

= 2

u d x

5 7 6
=
= 1 .2
3 2
4 8 0

Q = G A sg

g = f
A
s

t
G

= fA

Figure 14.8: Equating the deformation energies for finding the shear correction factor

c ETH Z

urich IMES-ST, March 26, 2014

14.5 Differential equations

14.5

119

Differential equations

The basic equations of the mechanical structural element beam must be brought together to
a meaningful differential equation. The representation shown in Fig. Abb. 14.9 eliminates
the shear strain from the equilibrium of the forces as well as from that of the moments.
G A
s

(w

+ b ) + q = 0

fo r c e e q u ilib r iu m

Q + q = 0

g = w + b

Q = G A sg

E I b + q = 0

c o n s titu tiv e la w

M - Q = 0
E I w - q = 0

k in e m a tic s

w = g - b

= E Ik = - E Ib
M

m o m e n t e q u ilib r iu m

- E I b - G A

(w
s

+ b

= 0

Figure 14.9: Beam differential equations

14.6

Shear-rigid beam finite element

In the case of a shear rigid beam (thickness is small if compared with the length of he
beam structure) one proceeds from the differential equation EIw0000 q = 0 it is agreed
that = 0.

14.6.1

Weak variational form of the shear-rigid beam problem

In addition to the standard procedure the unknown displacement solution appears in its
fourth spatial derivative. Fig. 14.10 demonstrates that twice partial integration leads to
p r in c ip le o f v ir tu a l d is p la c e m e n ts :
tw ic e p a r tia l in te g r a tio n :

w e a k v a r ia tio n a l fo r m :

d P
L

L
e

L
e

d w

(E

I w - q ) d x = 0
L

d w E I w d x = d w E I w
0

= d w E I w
0

d w E I w d x =

L
e

L
0

d w E I w d x
L

- d w E I w

d w q d x - d w E I w

L
0

L
0

d w E I w d x

+ d w E I w

L
e

Figure 14.10: Derivation of the weak variation form for the shear-rigid beam
the weak variational form, where both the virtual displacements as well as the unknown
displacement solution appear in their second derivatives. This creates two boundary terms
the mechanical interpretation of which is obtained with the help of beam theory.

c ETH Z

urich IMES-ST, March 26, 2014

120

Finite beam elements

14.6.2

Boundary terms interpretation

With the help of the beam theory equations one identifies in Fig. 14.11 the boundary terms:
w e a k v a r ia tio n a l fo r m
L

d w E I w d x =

b o u n d a ry te rm

d w q d x - d w E I w
e

+ d w E I w

L
e

in te r p r e ta tio n

w = - k ,

- E Ik = - M

k in e m a tic s

c o n s titu tiv e e q .

E I w = - M ,

= - d w M

L
e

s p e c ifie d b e n d in g m o m e n ts

E I w = - M = - Q

c o n s titu tiv e e q .

d w E I w

d w E I w

L
e

= - d w Q

L
e

s p e c ifie d fo r c e s

e q u ilib r iu m

Figure 14.11: Boundary terms interpretation


They signify specified bending moments and transverse forces, which form conjugated pairs
together with the rotation and deflections at the same parts of the boundary, respectively.

14.6.3

Conclusions for the discretization step

The shape functions appear in their second derivatives and therefore must be differentiable
twice. The boundary terms include deflections and rotations as independent degrees-ofw e a k v a r. fo rm

w e a k v a r ia tio n a l fo r m

w ith b o u n d . te r m s in te r p r e te d :

d w E I w d x =
e

d w q d x + d w Q

- d w M

L
e

c h a r a c te r is tic s :

- b o u n d a r y te r m s im c lu d e p r im a r y v a r ia b le a n d d e r iv a tiv e
- s h a p e fu n c tio n s m u s t b e d iffe r e n tia b le tw ic e

d is c r e tiz a tio n : S tiffn e s s m a tr ix a n d r ig h t-h a n d s id e :

F E I F T d x ,

r =

F q d x + F Q

L
e

- F M

L
0

c o n c lu s io n :
tw o -n o d e e le m e n t w ith H e r m ite p o ly n o m ia l

Figure 14.12: Weak variational form properties and conclusions


freedom. An element with two nodes therefore carries four degrees-of-freedom, corresponding with a third-order polynomial.

14.6.4

Polynomial complexity and Hermite shape functions

The discretization of the variable w of the shear-rigid beam finite element with two nodes,
shown in Fig. 14.13, is:

w() =

c ETH Z

urich IMES-ST, March 26, 2014

1 2 3

w
1

1
4
.

w

(14.1)

14.6 Shear-rigid beam finite element

121

The shape functions are to be connected with the general cubic polynomial in terms of
the global coordinate [1]:
w(x) = 0 + 1 x + 2 x2 + 3 x3 .

(14.2)

The shape functions must satisfy the end conditions:


x
b
1

x
0

w
1

L
y

b
2

b
1

- 1

w
2

b
2

2
x

Figure 14.13: Degrees-of-freedom of the two-node beam finite element


w(x0 ) = w1 w(x1 ) = w2
.
(x0 ) = 1 (x1 ) = 2

(14.3)

Next, the parameters i are expressed in terms of the degrees-of-freedom so that the
continuity conditions in terms of of the global coordinate are automatically satisfied:
w1 =
1

0 +

1 x0 +

2 x20 + 3 x30

= w0 (x0 ) = 1 22 x0 33 x20

w2 =
2

w(x0 ) =

w(x1 ) =

0 +

1 x1 +

2 x21 + 3 x31

(14.4)

= w0 (x1 ) = 1 22 x1 33 x21

The same equations are written in matrix form:

2
3

w
1
x
x
x

1
0
0
0
0

1 0 1 2x0 3x2 1

1 x1
2
x21
x31

w2

2
0
1
2x
3x

2
1
3
1

(14.5)

Now the equations can be resolved for the parameters i by inverting the matrix of coefficients in (14.5). Substituting the result into (14.2) obtains:
2

3

xx0
0
+
2
1 = 1 3 xx
Le
Le

2
0
2 =
(x x0 ) 1 xx
Le

2

3
(14.6)
xx0
0
3 =
3 xx

2
Le
Le


2
xx0
0
4 = (x x0 ) xx

Le
Le
The definitions of the global and the local normalized element coordinates given in Fig.
14.13 obtain the transformation:
x x0
1+
=
.
Le
2

(14.7)

c ETH Z

urich IMES-ST, March 26, 2014

122

Finite beam elements

Substituting these into (14.6) obtains the shape functions introduced with (14.1).
1
4 (2

1 =
2 =
3 =
4 =

3 + 3 )

Le
2
3
8 (1 + + )
1
3
4 (2 + 3 )
Le
2
3
8 (1 + )

(14.8)

Notice that these, in contrast to the previously considered isoparametric elements, must
consider the actual length of the beam finite element within the mesh. The shape functions
(14.8) are plotted in Fig. 14.14. The beam finite element is C 1 continuous. The virtual

1.0
0.8
P1

phi(i)

0.6

P2
P3

0.4

P4

0.2
0.0
-1.0

-0.8

-0.6

-0.4

-0.2

0.0

0.2

0.4

0.6

0.8

1.0

-0.2
-0.4
local beam element coordinate

Figure 14.14: Allgemeiner Polynomansatz


deformation energy of the weak variational form requires the second derivative of the shape
functions with respect to the global coordinate x. Same as for the truss element, it holds
that d/dx = 2/Le and with that one obtains for the first and second derivatives:

6 + 6 2
6

L
+
2L

3L

3L

e
e
e
e
e
1
1
,x =
, ,xx = 2
.
(14.9)

4Le
Le
6 6 2
6

L 2L 3L 2

L 3L

e
e
e
e
e
The formal steps of creating the dyadic product and integrating it over the element domain,
where dx = Le /2d, obtain the stiffness matrix K. Analog steps yield the right-hand side.
The element equations are:

w
1
6
3Le 6 3Le
6
Q1

qL

2
2EI
1
L2e
Le
M1
e
3Le 2Le 3Le

+
(14.10)
=
3Le
6
3Le
w
2
6
Q2

L3e 6
12

3Le
L2e
3Le 2L2e
Le
M2
2
The shear-rigid beam element implies that = 0 and therefore the cross-section rotation
is always the negative of the deflection line derivative = w0 .
c ETH Z

urich IMES-ST, March 26, 2014

14.7 Shear-compliant beam finite element

14.7

123

Shear-compliant beam finite element

The shear-compliant beam element arises from removing the constraint of vanishing shear
strain . Then, the direct coupling w0 = must be replaced with the kinematic relation
w0 = . Thus, the shear-compliant beam finite element is only C 0 continuous.

14.7.1

Penalty method and shear-compliant beam weak variational form

We follow a method invented by Zienkiewicz [6], where the shear-compliant beam functional is developed from the shear-rigid beam functional. The total potential energy of the
shear-rigid beam is easily obtained from the weak variational form given in Fig. 14.12:
L
L
Z
Z L



1 L
00 2
0


=
EI w
dx
wqdx wQ + w M
(14.11)
2 0
0
0
0
The positive sign of the rightmost term takes into account that the rotation, which is
conjugated to the bending moment, is the negative of the derivative of w.
The decoupling of the rotation from the derivative w0 requires rewriting of (14.11):
L
L
Z
Z L



1 L
0 2


=
EI dx
wqdx wQ M
(14.12)
2 0
0
0
0
The kinematic relation of shear angle , rotation , and slope w0 is now formally introduced
with a variational principle by way of a penalty function:
Z L
2

=+
w0 + dx.
(14.13)
0

It appears that a mechanically meaningful interpretation of the penalty factor is


1
= GAs
(14.14)
2
as the additional term then represents the deformation energy related to the shear strain.
Variation of (14.13) obtains the weak variational form of the shear-compliant beam.
L
L
Z L
Z L


 0


0
0
0


EI + w + GAs w + dx =
wqdx + wQ + M
(14.15)
0

The form suggests to introduce a partitioned matrix form of the integrated constitutive
laws:
"
#
EI
0
C=
.
(14.16)
0 GAs
The according structures of the remaining mechanical quantities are:
)
( )
(
)
( )
(
q
w

M
.
, =
, =
, q=
u=

Q
0
Together with the differential operator L
"
L=

( )0

( )0

#
(14.18)

one finds the compact representation of the weak variational form 14.15:
L
Z L
Z L

T
T
T
(Lu) C (Lu) dx =
u qdx + u
.
0

(14.17)

(14.19)

c ETH Z

urich IMES-ST, March 26, 2014

124

14.7.2

Finite beam elements

Shear-compliant beam weak variational form discretization

The primary variables appearing in the weak variational form of the shear-compliant beam
appear in their first derivatives at most. Therefore linear shape functions (4.12) introduced
in Section 4.2 are sufficient. The discretization of the displacement and rotation fields is
described with:

w
1

"
#
(
)

0
1
+

0
w

1
1
T
u= u

=
.
(14.20)
2
0
1
0
1+

w
2

2
With the matrix B
1
B = LT =
2Le

"

2 Le (1 ) 2 Le (1 + )

#
(14.21)

and the partitioned constitutive law one manages the partitioning of the element stiffness
matrix into a bending and a shearing part: 2
Z 1
Z 1
Le
Le
T
BT Cs B d.
(14.22)
K = Kb + Ks =
B Cb B d +
2
2
1
1
Performing the multiplications and integrations obtains the matrices:

0
0
0
0
6
3Le 6 3Le

2
L2e
1
0 1
EI
GAs

0
3Le 2Le 3Le

Kb =
. (14.23)

, Ks =
Le 0
6Le 6
3Le
6
3Le
0
0
0

2
2
3Le 2Le
3Le
Le
0 1
0
1
The right hand side becomes:

0
qLe
r=
+

2
1

1
Q

1
M
2

2
M

(14.24)

Please note that the bending and the shear stiffness matrices are added like for a system
of springs in-parallel. Obviously, this is not physical, as in reality the displacement due to
the shear compliance must be added to the displacements which occur due to the bending
compliance. Thus, in reality we have a system of springs in sequence. The circumstance,
that the shear-compliance finite element as has been derived here is physically wrong leads
to the locking effect which occurs if the shear-compliant beam finite element is used to
model thin beam situations, where shear influence must be negligible.

c ETH Z

urich IMES-ST, March 26, 2014

14.7 Shear-compliant beam finite element

14.7.3

125

Shear-compliant beam finite element locking effect

The shear-compliant beam functional leads to the form



(Kb + Ks ) u
= EIK0b + GAs K0s u
=r,

(14.25)

The shear-compliant beam element is meant for modeling of compact beams, where the
thickness is not very much smaller than the length. If using it for thin beams, or EI <<
GAs , the system of equations degenerate to
GAs K0s u
=r.

(14.26)

Simulations then exhibit the locking effect, as the calculated displacements are much too
small. For illustration of it consider with Fig. 14.15 a cantilever beam which is loaded
r e d u c e d in te g r a te d s h e a r s tiffn e s s m a tr ix

s h e a r - r ig id b e a m
= 1

= 2

= 4
e

= 8
e

1 .0

A n te il a n d e r e x a k te n D u r c h s e n k u n g

= 1 6

s h e a r - c o m p lia n t b e a m

0 .9
0 .8
0 .7
0 .6
0 .5
0 .4

fu lly in te g r a te d s h e a r s tiffn e s s m a tr ix

0 .3
0 .2
0 .1
0 .0
1

1 6

A n z a h l E le m e n te

Figure 14.15: Shear-compliant beam element locking effect


at its free end with a transverse point force. Mesh density influence is investigated by
subdividing the length of the beam with 1, 2, 4, 8, or 16 finite elements and comparing
the obtained results with exact solutions. Also, the two cases of a shear-rigid and a
shear-compliant beam are considered. The results obtained by using the shear-compliant
beam finite element are shown with the two lower curves. While the exact result of the
shear-compliant beam situation is approximated with increasing mesh density, the element
locks in case of the shear-rigid beam situation. The two upper curves are obtained after
modifying the present finite element and indicate a much improved aptitude for both beam
situations.

14.7.4

Locking effect mitigation by reduced integration

For thin-beam situations, the displacements calculated with the degenerated stiffness matrix in (14.26) vanish as long as the system matrix Ks is not singular. Recall that system
stiffness matrices are principally composed of singular element matrices and therefore remain singular as well until essential boundary conditions, equivalent to a at least statically
c ETH Z

urich IMES-ST, March 26, 2014

126

Finite beam elements

determinate support, have been implemented. To solve the problem of the locking effect it
must be achieved that the system matrix Ks remains singular even after implementation
of the essential boundary conditions.
The desired effect is achieved by not fully integrating the shear stiffness matrix. The
result shown in (14.23) is obtained by analytical integration or, equivalently, with Gauss
quadrature. As quadratic terms appear in the matrix B, an exact integration requires
the two-point Gauss rule. A reduced integration, i.e. by using the one-point Gauss rule,
reduces the matrix rank from 2 to 1, i.e. the integrated matrix contains only one linearly
independent row. The matrix of rank 1, Ks (1) , remains singular even after implementation
of the boundary conditions:

4
2Le 4 2Le

2
2

2Le
L
2L
L
e
GA
e
e

s
Ks (1) =
(14.27)
.

4Le 4
2L
4
2L
e
e

2Le
L2e
2Le
L2e
The following analytical inspection by Hughes [22] glaringly illuminates the locking effect
and its mitigation. Consider the cantilever beam situation, illustrated with Fig. 14.16,
which is loaded at its free end with transverse force or bending moment and presented with
M ,b
L

Q , w

Figure 14.16: Cantilever beam analyzed with one single finite element
only one single finite element. Use of the fully integrated stiffness matrix Ks (2) obtains
the system equations
"
"
#
#! (
) (
)
6 3Le
w
Q
EI 0 0
GAs
+
=
.
(14.28)
Le 0 1
6Le
3Le 2L2e

M
Obviously also shear stiffness matrix has rank 2 and is regular. Resolution of the two
equations for the primary variables gives
w=

L
3EI + L2 GA
Q+
M,
3GA
2

2M + LQ
,
2

EI
GAL
+
.
L
12

(14.29)

As the beam becomes very thin the bending stiffness becomes very small if compared to
the shear stiffness, or EI << GA, and expects that the shear-rigid situation should be
dominated by bending. The numerical system of equations behaves in the opposite and
yields in the limiting case,
w=

4LQ + 6M
,
GA

12M + 6LQ
,
GAL

the result that deflection and rotation depend on shear stiffness only.
The reduced-integrated stiffness matrix Ks (1) obtains the system equations
"
#
"
#! (
) (
)
4 2L
w
Q
EI 0 0
GAs
+
=
.
Le 0 1
4L
2L L2

M
c ETH Z

urich IMES-ST, March 26, 2014

(14.30)

(14.31)

14.7 Shear-compliant beam finite element

127

Here, the shear-stiffness matrix has rank 1 and is singular. Resolution for the primary
variables presently gives

 3
L2
L
L
2LM L2 Q
Q
w=
+
M,
=
(14.32)
4EI
GA
2EI
2EI
and it becomes apparent that the rotation always depends on the bending stiffness EI
only. In the limiting case of a very thin beam also the deflection
w=

L3
L2
Q
M
4EI
2EI

(14.33)

depends on bending stiffness only, which is a mechanically meaningful, or correct, result.


Let it be remarked that thin-beam situations are analyzed more efficiently if using the
shear-rigid beam finite element: The problem underlying the results shown in Fig. 14.15
are solved exactly with only one element.

c ETH Z

urich IMES-ST, March 26, 2014

128

c ETH Z

urich IMES-ST, March 26, 2014

Finite beam elements

Chapter 15

Eigenvalue problems
Calculation of eigenfrequencies of harmonic vibrations and critical buckling loads of instability problems together with the eigenmodes belong to the eigenvalue problems and are
here explained.

15.1

Harmonic Vibrations

15.1.1

Equations of motion

Dynamic problems include forces arising from acceleration and the inertia of mass which
must be considered with the equilibrium of forces at an infinitesimal small volume element.
Adding the dynamic forces to the equilibrium conditions, see Fig.3.1, gives the equations
of motion:
ij ,j +fi =
ui
(15.1)
or
LT CLu + f =
u.

15.1.2

(15.2)

Weak variational form and system of equations

The weak variational form including mass inertia effects is


Z
Z
Z
Z

T T
T
T
u L C (Lu) d +
u
ud =
u f d + uT
d .

Its discrete form on element level


Z
Z
Z
Z
(t) =
BT CBd
u(t) +
T du
f (t)d +
(t)d .

(15.3)

(15.4)

obtains the numerical element equations. The second integral on the left-hand side of
(15.4) is called mass matrix M and with evaluation of the integrals and assembly of all
element contributions one obtains the numerical system of equations:
(t) = r(t).
K
u(t) + Mu

15.1.3

(15.5)

Relation between amplitudes and accelerations

Harmonic vibrations exist in the absence of energy dissipation. If the amplitudes are
much smaller than the size of considered structure, the kinetic and deformation energies
c ETH Z

urich IMES-ST, March 26, 2014

130

Eigenvalue problems

are transformed into each other in sinusoidal time dependence. The time dependence is
written for the displacements as
(t) = u
cos(t)
u
(15.6)
with the angular eigenfrequency . Taking derivatives with respect to time t obtains the
relation between displacement amplitude and acceleration:
= 2 u
.
u

(15.7)

The harmonic vibrations problem is written as:



=0.
K 2M u

15.1.4

(15.8)

Further aspects of solving an harmonic vibrations problem

The vector-iteration solution method explained in Section 9.4 solves the harmonic vibrations problem. Building the dynamic matrix D requires inverting the stiffness matrix K
which must then be regular. When simulating the vibration properties of unsupported
systems, the problem appears that the stiffness matrix is singular. The difficulty is mitigated by modifying the system on the basis of the following thoughts.
Rigid-body motions are free of deformation energy:
1 T
K
u
u=0
2

or

K
u = 0.

(15.9)

On the other hand the product of the mass matrix and the rigid-body-motion displacement
fields do not give a null vector:
M
u 6= 0.
(15.10)
Satisfying of the eigenvalue problem (15.8) now implies vanishing of the eigenfrequencies.
A number of c independent rigid-body motions must give a c fold zero eigenvalue:
12 = 22 = . . . =c2 = 0.

(15.11)

A system with N degrees-of-freedom then retains n = N c eigenvectors with deformation


energy. With known rigid-body-displacement fields constraining equations
N = C
u
un

(15.12)

can be set up which transform the original problem with N eigenvectors into a smaller system with n eigenvectors which all carry deformation energy and where the stiffness matrix
is regular. The constraining equations are contained in a constraining matrix C and Fig.
15.1 explains its structure. The starting point of constructing the constraining matrix is
N = C
the unit matrix I which maps any vector onto itself with the operation u
uN . The
N are expressed in terms of a
unit matrix is then modified so that the first c entries of u
linear combination of the remaining n entries. Fig. 15.1 indicates that the replacement fills
the first c rows, starting with the column c + 1, with the constraining equations. The first
c columns contain zeros and omitting these obtains the matrix C whose first c columns
contain the constraining equations while the remaining equations contain the unit matrix
I with n n rows and columns.
The needed constraining equations follow from the orthogonality property of the eigenvectors with respect to the mass matrix,
(r)T M
u
u(s) = 0, r 6= s,
c ETH Z

urich IMES-ST, March 26, 2014

(15.13)

15.1 Harmonic Vibrations

131

c
c

...
0

...
0

...
0

...

...

...

...

...

...

...

...

...

...

...
1

...
0

...
0

...
0

...

...
0

...
0

...
0

...
0

...

...

...

...

...

...

...

...

...

...
0

...

...

N
0

u n it m a tr ix

n
n

1 ,c + 2

...

...

...

c ,c + 2

...

...
0

0
1

...

...

...

...

1 ,c + 1

...
c

c ,c + 1

1 ,c + 2

...

...

...

c ,c + 2

...

...

0
1

...
0

...

...

...

...

...

...

...
1

...
1

1 ,N

...

1 ,c + 1

...
c

c ,N

c ,c + 1

im p le m e n ta tio n o f c o n s tr a in ts

1 ,N

...
c ,N

c o n s tr a in in g m a tr ix

Figure 15.1: Constraining matrix structure


which allows to express all N eigenvector entries in terms of the remaining n = N c
(r) , 1 r c are known and form the
entries. The rigid-body-displacement arrays u
N c . It is multiplied with the mass matrix M where the operation
columns of a matrix U
is partitioned as Fig. 15.2 illustrates. The orthogonality property (15.13) can now be
M
M

c c

~
T

c c

~
T

n c

n c

c c

c n

n n

c n

Figure 15.2: Rigid-body-motion constraint


written as
(r)T M
(s)
(s)
u
u(s) = Rcc u
c + Rnc u
n =0

(15.14)

(s) in
and provides upon inversion of the matrix Rcc the first c entries of any eigenvector u
terms of its remaining entries,
1
(s)
(s)
u
c = Rcc Rnc u
n ,

(15.15)

which relations are the desired constraining equations. With the help of the constraining
matrix C, whose population structure is indicated in Fig. 15.1, one obtains the smaller
matrices
K0 = CT KC,
M0 = CT MC,
(15.16)
which form the reduced eigenvalue problem

= 0.
K0 2 M0 u

(15.17)

with a regular stiffness matrix K0 .


c ETH Z

urich IMES-ST, March 26, 2014

132

Eigenvalue problems

15.2

Instability and FEM

The approach for calculating critical buckling loads with FEM is here demonstrated by
referring to an Euler buckling case. Fig. 15.3 illuminates the influence of the longitudinal
P
a s s u m p tio n s :
- n o lo n g itu d in a l c o m p lia n c e
- fo r c e r e m a in s v e r tic a l
- n o fu rth e r e x te rn a l fo rc e s

w d x
w

d x

e q u ilib r iu m

P
P

[
=

a t d e fle c te d lin e e le m e n t:

E I w 2 - q w - P
2

= w d x P

d M

d P

[d
=

M = w P

(
1

w 2

)]

d x

w E I w - d w q - P ( d w w )

d x

Figure 15.3: Equilibrium in the deflected configuration and energy functional


force P upon the equilibrium of a infinitesimal small line element in the deflected configuration of a rod. The deflection differential w0 dx causes a bending moment w0 dxP and
both quantities contribute, as a conjugated pair, external work to the energy functional.
Assuming that the rod is rigid in its longitudinal direction and no transverse force is applied, there is no other contribution. Fig. 15.4 summarizes the derivation steps from the
L

[d

w E I w - P ( d w w ) ] d x = 0

(d

u~
T

E I F F T u~ - P d u~

E I B
K

B
K

u~ - P B

B
N

u~
N

K u~ - P N u~ = 0 ,

(K

- P N

) u~

= 0

K
=

d x = 0

d x ,

N =

d x = 0

E I B
0

F F T u~
T

B
K

B
N

T
N

d x

e ig e n v a lu e p r o b le m !

Figure 15.4: Discretization step numerical eigenvalue problem


weak variational form to the numerical eigenvalue problem. It appears that the critical
load, or more generally a critical factor on a given load case, takes the role of eigenvalue.
The equilibrium at the deflected system gives the so-called geometric stiffness matrix N.
The used sample problem is worked out concretely by referring to the shear-rigid beam
finite element introduced in Section 14.6 and Fig. 15.5 summarizes all needed matrices.
Let the rod be modeled with one single finite element only. Because of the clamping at
the lower end there remain only two degrees-of-freedom, namely the deflection and the
rotation at the free end. The eigenvalue problem and its solution by vector iteration, see
c ETH Z

urich IMES-ST, March 26, 2014

15.2 Instability and FEM

- 6 x
1 - L + L x
=

8 4 + 6 x
L + L x

+ L x
- L x

+ 2 x
- L x
- 2 x
- L x

133

- 6

1 L + 2 L x
= F =

4 L 6
L - 2 L x
N

+ 6 x
- 3 L x
- 6 x
- 3 L x

,
2
2

6 x
- 3 L x

- 6 x
- 3 L x

= F =
K

- L

1 .0

3 6
- 3 L

- 3 6
- 3 L

0 .8

p h i(i)

P 2
P 3

0 .4

P 4

0 .0
-0 .2

-0 .8

-0 .6

-0 .4

-0 .2

0 .0

0 .2

0 .4

0 .6

N =

0 .2

-1 .0

- 3 L e - 3 6 4 L 2e 3 L e 3 L e 3 6
3
- L 2e 3 L e 4

P 1

0 .6

0 .8

1 5 L

1 .0

-0 .4
L o k a le B a lk e n K o o r d in a te

3 L e
L 2e
L e
L 2e

6
- 3 L
2 E I
=

L 3e - 6
- 3 L

- 3 L
2 L 2e
3 L e
L 2e
e

- 6 - 3 L
3 L e L 2e
6
3 L e
3 L e 2 L 2e

Figure 15.5: Shape functions and matrices of the shear-rigid beam finite element
Section 9.4, are summarized in Fig. 15.6. The inverse stiffness matrix K1 and dynamic
2 E I 6
3
L
e 3 L

3 L e
2
- P
2 L 2e
1 5 L
e

3 6
3 L

w

b

3 L e
4 L 2e

(I

u~

(r )

= 0

u~

(r )

= 0

- 1

N u~

(r )

- P rN
- 1

- P rK
K

s o lu tio n w ith v e c to r ite r a tio n ,


s a m e a s fo r h a r m o n ic v ib r a tio n s :

D u~

(r )

1 ~
u

(r )

r
( r )

= l

u~

(r )

Figure 15.6: Sample eigenvalue problem and solution with vector iteration
matrix D as well as chosen numbers for numerical evaluation are shown in Fig. 15.7. The
- 1

L
e

2 L

6 E I - 3 L

2
e

- 3 L e

E = 7 0 '0 0 0 M P a ,

A = 1 0 0 m m

D =

1 5 E I - 3 0 L
4

I = 8 3 3 .3 m m
,

2 1 L

2
e

- 2 L 3e

5 L 2e

L = 1 0 0 0 m m
,

Figure 15.7: Dynamic matrix and model data


eigenform and critical load predicted with the data are shown in Fig. 15.8 which also il1 .0

1 6 0

0 .8

K n ic k la s t u n d D iffe r e n z

(1 )

E ig e n v e k to r k o m p o n e n te n

1 8 0

0 .9

0 .9 9 9 9 9 8 8
=

0 .0 0 1 5 6 7 8

0 .7
0 .6
0 .5
0 .4
0 .3
0 .2
0 .1

1 4 0

(1 )

= 1 4 5 .0 1 N

1 2 0
1 0 0
8 0
6 0
4 0
2 0

0 .0
1

Ite r a tio n

1 0

0
1

1 0

Ite r a tio n

Figure 15.8: Vector iteration results and convergence


lustrates the convergence of the solution process. The eigenvectors are normalized to a
unit length. For comparison the same problem is solved by using ten elements instead of
c ETH Z

urich IMES-ST, March 26, 2014

134

Eigenvalue problems

only one and the computations are performed by a commercial FEM program According
to the numbers given in Fig. 15.9 the commercial code obtains identical results as the

Figure 15.9: Verification with ANSYS


sample problem if only one element is used. The mesh with ten elements obtains a critical
buckling load only slightly lower.

15.3

Critical buckling load and natural frequency

Fig. 15.10 illustrates an interesting relation between critical buckling load factor and
c o n s id e r s tr e s s s tiffn e s s m a tr ix N
w ith lo n g itu d in a l fo r c e P

(K

- P N

u~ = w
2

M u~

lin e a r e x tr a p o la tio n o f th e s q u a r e o f
th e e ig e n fr e q u e n c y to a z e r o v a lu e
2

= w

(P )

P 1w

2
2
2

+ P 2w
- w 12
1

1 .0

0 .9
0 .8

Q u a d r ie r te n o r m ie r te
E ig e n fr e q u e n z

N o r m ie r te E ig e n fr e q u e n z

1 .0

c r it

0 .7
0 .6
0 .5
0 .4
0 .3
0 .2
0 .1
0 .0

0 .9
0 .8
0 .7
0 .6
0 .5
0 .4
0 .3
0 .2
0 .1
0 .0

A u f K n ic k la s t b e z o g e n e L n g s k r a ft

8
1

A u f K n ic k la s t b e z o g e n e L n g s k r a ft

Figure 15.10: Knickgefahr und Abfall der Grundfrequenz


natural frequency of the loaded structure. Knowing the eigenfrequency of a structure in
its reference configuration and measuring the eigenfrequency under load allows to predict
the critical load factor.

c ETH Z

urich IMES-ST, March 26, 2014

List of Figures
1.1
1.2

Steps of an FEM analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . .


Elements of various spatial and shape-function complexities . . . . . . . . .

2.1
2.2
2.3
2.4
2.5
2.6
2.7
2.8
2.9
2.10

Truss Viaduct . . . . . . . . . . . . . . . .
Trusses and connections . . . . . . . . . .
structural member truss . . . . . . . . . .
Truss and FEM model . . . . . . . . . . .
Mapping truss onto finite element . . . . .
Lineloads and nodal forces . . . . . . . . .
Truss finite element equations . . . . . . .
Beam element conventions . . . . . . . . .
Beam boundary conditions . . . . . . . .
Element equations of the shear-rigid beam

3.1
3.2
3.3
3.4
3.5
3.6
3.7
3.8

Basic equations of elasticity in tensor


Symbols of matrix-operator notation
General solid-body problem . . . . .
Total potential energy . . . . . . . .
Infinitesimal tetrahedral element . .
Summary energy methods . . . . . .
Boundary conditions . . . . . . . . .
Volume element with 8 nodes . . . .

4.1
4.2
4.3
4.4

Linear truss element shape functions .


Simplistic truss frame work . . . . . .
Statically determinate truss frame . .
Directional transformation at the truss

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

5.1
5.2
5.3
5.4
5.5
5.6
5.7
5.8
5.9
5.10
5.11

Planar four-node element . . . . . . . . . . . . . . . . . . . . .


Test situation with notched specimen . . . . . . . . . . . . . . .
Symbolic notations and expansions . . . . . . . . . . . . . . . .
Pasquals triangle and quadrilateral elements . . . . . . . . . .
Shape functions of the linear triangular element . . . . . . . . .
Linear triangular element node coordinates and displacements .
Vector line element in local coordinates . . . . . . . . . . . . .
Shape functions along element edges . . . . . . . . . . . . . . .
Linear triangular element kinematically equivalent nodal forces
Bilinear quadrilateral element conventions . . . . . . . . . . . .
Bilinear quadrilateral element shape functions . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

and in
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

8
8
8
9
9
10
10
11
12
13

matrix-operator notations
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

15
16
16
17
19
20
23
24

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

28
30
30
31

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

33
34
35
36
37
37
39
40
40
41
41

. . . . .
. . . . .
. . . . .
element

.
.
.
.
.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

2
3

c ETH Z

urich IMES-ST, March 26, 2014

136

LIST OF FIGURES

5.12
5.13
5.14
5.15
5.16

Results of various models for simulating a cantilever beam loading situation


Ansatzfunktionen Drei- und Viereckelement . . . . . . . . . . . . . . . . . .
Quarter-circle approximation with triangular and quadrilateral elements . .
Circular domain approximation with bilinear and curvilinear elements . . .
Four-node Serendipity element shape functions illustration . . . . . . . . . .

42
42
43
43
44

6.1
6.2
6.3
6.4
6.5
6.6
6.7
6.8
6.9
6.10
6.11
6.12
6.13
6.14
6.15

Quadrilateral element construction, after [6] . . . . . . . . . . . . . . . . . .


Quadratic truss-element shape functions . . . . . . . . . . . . . . . . . . . .
Patch test: Geometry, mesh, boundary conditions and displacement field ux
Incompressibily: Cooks membrane problem . . . . . . . . . . . . . . . . . .
Volume locking effect situation . . . . . . . . . . . . . . . . . . . . . . . . .
Area contents due to node displacements . . . . . . . . . . . . . . . . . . . .
Bending of a cantilever beam and mesh dependent mapping error . . . . . .
Illustration to pure bending . . . . . . . . . . . . . . . . . . . . . . . . . . .
Unphysical connection of pure bending with shear by the bilinear element .
Influence of mesh distortion on deflection error . . . . . . . . . . . . . . . .
Bi-linear shape functions (Fig. 5.11) and incompatible quadratic modes . .
Cooks Problem considered with basic and enhanced quadrilateral elements
Pure bending with incompatible modes . . . . . . . . . . . . . . . . . . . . .
Cantilever beam bending and influence of mesh density on error . . . . . . .
Mesh distortion influence on simulated deflection error . . . . . . . . . . . .

45
46
47
48
48
49
49
50
51
51
52
54
54
55
55

7.1
7.2
7.3
7.4
7.5
7.6
7.7
7.8
7.9
7.10
7.11

Structure and size of matrix BT CB for various elements . . . . . . . . . . .


Forming of the matrix B of the quadratic Serendipity Element . . . . . . .
Vivid pictures for the one- and two-point Gauss quadrature rules . . . . . .
Gauss points in two-dimensional quadrilateral finite reference elements . . .
Highest polynomial powers occurring in matrix B . . . . . . . . . . . . . . .
Bilinear quadrilateral element with distortion parameter . . . . . . . . .
Distribution of integrand (1 ,x )2 and square fit . . . . . . . . . . . . . . . .
Fit of the intagrand to quadratic and cubic polynomials . . . . . . . . . . .
Subroutine for assembly of the global stiffness matrix . . . . . . . . . . . . .
Subroutine for calculating element stiffness matrices . . . . . . . . . . . . .
Subroutine for calculating the matrix B B-Matrix and the Jacobian matrix

57
58
58
61
61
62
64
64
65
66
67

8.1

Boundary conditions involving degree-of-freedom coupling . . . . . . . . . . 70

9.1
9.2
9.3
9.4
9.5
9.6
9.7
9.8
9.9
9.10
9.11
9.12

Node numbering schemes and stiffness matrix population structure


Symmetric matrices in band storage mode . . . . . . . . . . . . . .
Triangular decomposition principle . . . . . . . . . . . . . . . . . .
Principle direct solution . . . . . . . . . . . . . . . . . . . . . . . .
Gauss Elimination at a fully populated matrix . . . . . . . . . . .
Gauss Elimination at a banded matrix . . . . . . . . . . . . . . . .
Elucidation of the first Cholesky decomposition variant . . . . . .
Elucidation of the second Cholesky decomposition variant . . . . .
Forming of the matrix B for Jacobi iteration . . . . . . . . . . . .
Gauss-Seidel modification of the Jacobi method . . . . . . . . . . .
Quadratic functional . . . . . . . . . . . . . . . . . . . . . . . . . .
Orthogonal (left) and conjugate (right) search directions . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

71
72
73
74
74
75
76
77
78
79
80
81

10.1 Partitioning of a mesh for sub-structure technique . . . . . . . . . . . . . . 85


c ETH Z

urich IMES-ST, March 26, 2014

LIST OF FIGURES

137

10.2
10.3
10.4
10.5
10.6
10.7
10.8
10.9

Partitioning of a structure for sub-structure technique . . . . . . . .


Partitioning of the system of equations and matrix operations . . . .
= Kab Kbb 1 . . . . . . . . . . . . . . . . . .
Structure of matrix K
Structure of matrix Kab Kbb 1 Kba . . . . . . . . . . . . . . . . . . .
. . . . . . . . . . . . . . . . . . . . . . . . . .
Structure of matrix K
Replacement of fixed substructure in Fig. 10.1 by elastic foundation
Application of unit forces to the mesh part to be eliminated . . . . .
Structural model with onsert of the initial and optimized shapes . .

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

86
87
87
87
88
88
88
89

11.1
11.2
11.3
11.4

Truss model of a drill gear under gravity load . . . . . . . . .


Optimal strain point within a linear finite element . . . . . .
Optimal strain points within a quadratic finite element . . . .
Integrations- und optimale Dehnungspunkte im quadratischen

. . . . .
. . . . .
. . . . .
Element

.
.
.
.

.
.
.
.

.
.
.
.

92
93
94
94

12.1 Illustration to periodicity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96


12.2 Honeycomb with thin GFRP weaving face sheets and Nomex honeycomb core 97
12.3 Nomex honeycomb core material . . . . . . . . . . . . . . . . . . . . . . . . 97
12.4 Honeycomb expansion process and geometry idealization . . . . . . . . . . . 97
12.5 Nomex honeycomb core material . . . . . . . . . . . . . . . . . . . . . . . . 98
12.6 Generalized plane-strain compatible global deformation modes . . . . . . . 99
12.7 Anisotropic coupling effects in unidirectional composites . . . . . . . . . . . 101
12.8 Test specimen subject to unidirectional stress . . . . . . . . . . . . . . . . . 101
12.9 Meshing of a corrugated sheet . . . . . . . . . . . . . . . . . . . . . . . . . . 102
12.10Deformation load cases for calculating the entries of a substitute ABD matrix102
12.11Conventions of the corrugated-sheet unit cell . . . . . . . . . . . . . . . . . 103
12.12Finite-element mesh of the corrugated sheet . . . . . . . . . . . . . . . . . . 103
12.13Midplane support and periodicity at load case 1 . . . . . . . . . . . . . . . . 103
12.14Midplane support and periodicity at load case 1 . . . . . . . . . . . . . . . . 104
13.1 Limits of linear modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . .
13.2 Change of natural boundary conditions due to sagging of a supporting wall
13.3 Large deflection of a slender cantilever beam . . . . . . . . . . . . . . . . .
13.4 Fatigue damage in a notched CFRP laminate [45, 90, 45, 0]s . Source [19] .
13.5 Fiber rupture (a) and matrix failure (b) in a multidirectional laminate . . .
13.6 Test specimen geometry and quarter-model FE mesh . . . . . . . . . . . . .
13.7 Damage pattern unzip action after 15 damage events . . . . . . . . . . . . .
13.8 Damage patterns simulated for various laminates . . . . . . . . . . . . . . .
13.9 Failure-load factors reference values after DS. Source: [2] . . . . . . . . . . .
13.10Failure-load factors: unzip action and CG. Source: [2] . . . . . . . . . . . .
13.11Mesh parameters, problem sizes, and computing times. Source: [2] . . . . .
13.12Computing times measurements and fitted model predictions. Source: [2] .
13.13Scaling of CG computing time relative to those of DS. Source: [2] . . . . . .

107
107
108
108
109
110
111
111
112
112
113
114
114

14.1
14.2
14.3
14.4
14.5
14.6
14.7

115
116
116
117
117
118
118

Equilibrium of forces and of moments at small beam element . . . . . .


Kinematic relations at an infinitesimal beam element . . . . . . . . . . .
Transverse force, shear stress and shear stiffness after first-order theory
Bending moment, curvature, and bending stiffness . . . . . . . . . . . .
Derivation of a parabolic shear stress distribution . . . . . . . . . . . . .
Deformation energy of the parabolic shear-stress distribution . . . . . .
Deformation energy of the constant shear-stress distribution . . . . . . .

.
.
.
.
.
.
.

.
.
.
.
.
.
.

c ETH Z

urich IMES-ST, March 26, 2014

138

LIST OF FIGURES

14.8 Equating the deformation energies for finding the shear correction factor
14.9 Beam differential equations . . . . . . . . . . . . . . . . . . . . . . . . .
14.10Derivation of the weak variation form for the shear-rigid beam . . . . .
14.11Boundary terms interpretation . . . . . . . . . . . . . . . . . . . . . . .
14.12Weak variational form properties and conclusions . . . . . . . . . . . . .
14.13Degrees-of-freedom of the two-node beam finite element . . . . . . . . .
14.14Allgemeiner Polynomansatz . . . . . . . . . . . . . . . . . . . . . . . . .
14.15Shear-compliant beam element locking effect . . . . . . . . . . . . . . . .
14.16Cantilever beam analyzed with one single finite element . . . . . . . . .

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

118
119
119
120
120
121
122
125
126

15.1 Constraining matrix structure . . . . . . . . . . . . . . . . . . . . . .


15.2 Rigid-body-motion constraint . . . . . . . . . . . . . . . . . . . . . .
15.3 Equilibrium in the deflected configuration and energy functional . .
15.4 Discretization step numerical eigenvalue problem . . . . . . . . . . .
15.5 Shape functions and matrices of the shear-rigid beam finite element
15.6 Sample eigenvalue problem and solution with vector iteration . . . .
15.7 Dynamic matrix and model data . . . . . . . . . . . . . . . . . . . .
15.8 Vector iteration results and convergence . . . . . . . . . . . . . . . .
15.9 Verification with ANSYS . . . . . . . . . . . . . . . . . . . . . . . . .
15.10Knickgefahr und Abfall der Grundfrequenz . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

131
131
132
132
133
133
133
133
134
134

A.1
A.2
A.3
A.4
A.5
A.6

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

143
144
145
146
151
151

Solid body with domain and boundary . . . . . . . . . . . . . .


Stress components at a hexahedron . . . . . . . . . . . . . . . . . . .
Relative motion of two neighboring points within a solid body [23]. .
Shear strain mechanism[23]. . . . . . . . . . . . . . . . . . . . . . . .
Element in axisymmetric solid body and volume element [6]. . . . .
Forces acting upon a volume element of an axisymmetric solid body

c ETH Z

urich IMES-ST, March 26, 2014

List of Tables
3.1

Nodal point coordinates of the element Fig. 3.8 . . . . . . . . . . . . . . . . 24

5.1

Four-node Serendipity element shape functions and derivatives . . . . . . . 44

6.1

Node coordinates for the patch test. Source: [7] . . . . . . . . . . . . . . . . 47

7.1
7.2

Coordinates and weights of Gauss quadrature . . . . . . . . . . . . . . . . . 60


Numerical effort for calculation of one element stiffness matrix . . . . . . . 68

9.1

Bandwidths and profiles of the population structures shown in Fig. 9.2 . . . 72

10.1 Numerical size and flops without and with sub-structure technique . . . . . 88
10.2 Efficiency of the sub-structure technique with the model of a bonded structure 90
13.1 Laminates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
13.2 UD T300 material data (DORNIER SYSTEM) . . . . . . . . . . . . . . . . 111
A.1 Tensor and matrix (Voigt) notations . . . . . . . . . . . . . . . . . . . . . . 147

c ETH Z

urich IMES-ST, March 26, 2014

140

c ETH Z

urich IMES-ST, March 26, 2014

LIST OF TABLES

Bibliography
[1] Reddy, J.N. An Introduction to the Finite Element Method. McGraw-Hill, New York,
1984.
[2] Kress, G., M. Siau, and P. Ermanni. Iterative Solution Methods for Damage Progression Analysis. Composite Structures, 2004. in print.
[3] Reddy, J.N. Applied Functional Analysis and Variational Methods in Engineering.
McGraw-Hill, 1986.
[4] Galerkin, B.G. On Electrical Circuits for the Approximate Solution of the Laplace
Equation. Vestnik Inzh., 19:897908, 1015. in Russian.
[5] G
ateaux, R. Sur la representation des fonctionelles continues.
Academia dei Lincei, 23(1):310315, 1014.

Atti della Reale

[6] Zienkiewicz, O.C. and R.L. Taylor. The Finite Element Method 1. McGraw-Hill,
1989.
[7] Jabareen, M. Finite Element for Linear Elasticity - Element Technology. Zentrum
f
ur Mechanik, ETH Z
urich, Herbstsemester 2008.
[8] . ANSYS Classic Release 12. Canonsburg, PA, USA: ANSYS Inc., 2009.
[9] Cuthill, E. and J. McKee. Reducing the Bandwidth of Sparse Symmetric Matrices.
In Proc. 24th National Conf. ACM, pages 157172, 1969.
[10] Sloan, S.W. An Algorithm for Profile and Wavefront Reduction of Sparse Matrices.
Int. J. Num. Methods Eng., 23:239251, 1986.
[11] Sloan, S.W. A FORTRAN Program for Profile and Wavefront Reduction. Int. J.
Num. Methods Eng., 28:26512679, 1989.
[12] Shewchuk, J.R. An Introduction to the Conjugate Gradient Method Without the
Agonizing Pain. School of Computer Science, Carnegie Mellon University, Pittsburgh,
PA 15213, Aug. 1994.
[13] Meirovitch, L. Elements of Vibration Analysis. McGraw-Hill, 1975.
[14] Kress, G., P. Naeff, and P. Ermanni. Onsert Strength Design. Int. J. Adhesion &
Adhesives, 24:201209, 2004.
[15] Kress, G., D. Fritsche, and P. Ermanni. Failure Criteria and Onsert Shape Optimization. Int. J. Adhesion & Adhesives, in print, 2004.
c ETH Z

urich IMES-ST, March 26, 2014

142

BIBLIOGRAPHY

[16] Barlow, J. Optimal Stress Locations in Finite Element Models. Int. J. Num. Methods
in Engineering, 10:243251, 1976.
[17] Kress, G., M. Winkler. Honeycomb Sandwich Residual Stress Deformation Pattern.
Composite Structures, 89(2):294302, 2009.
[18] Kress, G., M. Winkler. Corrugated Laminate Analysis: A Generalized Plane Strain
Problem. Composite Structures, 93(5):14931504, 2011.
[19] Fatigue Response of Notched Graphite/Epoxy Laminates. American Society for Testing
and Materials, 1985.
[20] Mindlin, R.D. Influence of Rotary Inertia and Shear on Flexural Response of Isotropic
Elastic Plates. J. Appl. Mech., 18(3):138, 1951.
[21] Bathe, K.-J. Finite-Elemente-Methoden. Springer, 1990.
[22] Hughes, T.J., R.L. Taylor, and W. Kanoknukulachai. A Simple and Efficient Finite
Element for Plate Bending. Int. J. Num. Methods in Engineering, 11:15291543, 1968.
[23] Frederick, D. and T.S. Chang. Continuum Mechanics. Scientific Publishers, Cambridge, 1972.
[24] Jones, R.M. Mechanics of Composite Materials. Hemisphere Publishing Corporation,
New York, 1975.

c ETH Z

urich IMES-ST, March 26, 2014

Appendix A

Equations of linear elasticity and


plane problems
A.1
A.1.1

Basic problem of linear elasticity


Equilibrium conditions

Consider with Fig. A.1 an arbitrary domain bounded with . The equilibrium of forces
z
G
n
i

s
i

d W
W

d G

fi
0
y

Figure A.1: Solid body with domain and boundary


dictates that the sum of all forces acting upon the domain and the boundary must vanish.
Consider the stress vector i (x, y, z) acting upon the boundary differential d, the mass
density (x, y, z) of the material contained in the volume differential d, and the body
force fi (x, y, z) acting upon it. Then the equilibrium condition for forces is:
Z
Z
i d +
fi d = 0.
(A.1)

The stress vector i in (A.1) can be expressed with the stress tensor ij and the normal
unit vector ni which is oriented perpendicular to the boundary differential d:
Z
Z
ji nj d +
fi d = 0.
(A.2)

Using the Gauss-Green theorem the boundary integral in (A.2) is transformed into a
domain integral,
Z
Z
ji nj d =
ji ,j d,
(A.3)

c ETH Z

urich IMES-ST, March 26, 2014

144

Equations of linear elasticity and plane problems

where the subscript (,j ) prescribes the respective derivative with respect to x, y, z. So
(A.2) becomes
Z
(ji ,j +fi ) d = 0.
(A.4)

This equation must hold for any subdomain within wherefore the integrand on the
left-hand side of (A.4) must vanish identically:
ji ,j +fi = 0.

(A.5)

The equations hold also for a body in motion if the body force represents the inertia force

u, then one speaks more generally of the equations of motion and all magnitudes depend
on position and time as well. The expanded equations of motion
xx ,x + yx ,y + zx ,z + fx =
ux
xy ,x + yy ,y + zy ,z + fy =
uy
xz ,x + yz ,y + zz ,z + fz =
uz .

(A.6)

are illustrated in Fig. A.2.


s

zx

zy

t
y

s
x

x z

y z

y x

s
y

x y

Figure A.2: Stress components at a hexahedron

A.1.2

Kinematic or strain displacement relations

If a continuum changes its configuration the material in the vicinity of each point is
generally translated and rotated like a rigid body but also subject to straining. The strain
is the part of the motion of two particles relative to each other which can not ascribed to a
pure rigid-body motion. For obtaining a measure of strain consider two neighboring points
before (P, Q) and after (P , Q) a deformation [23]. The points identified with vectors:
P :
Q:
P :
Q:

ai
ai + dai
xi = ai + ui (a, t)
xi + dxi = ai + dai + ui (a + da, t)

(A.7)

The squares of the small distances between both points before and after deformation differ
by
(dx)2 (da)2 = dxi dxi dai dai .
(A.8)
If the positions xi of the particles are given as a function of the initial configuration ai the
vector dxi can be expressed by using the chain rule:


xi
dxi =
daj ,
(A.9)
aj P
c ETH Z

urich IMES-ST, March 26, 2014

A.1 Basic problem of linear elasticity

,t)
+ d a
a
(
u i

d x
i

d a

x i+
d x

145

a i+ d
a

P
a

u i( a , t )

x
i

Figure A.3: Relative motion of two neighboring points within a solid body [23].
whereby the partial derivatives are evaluated at point P . Combination of (A.9) and (A.8)
obtains


xr xr
2
2
(dx) (da) =
ij dai daj .
(A.10)
ai aj
Following the Lagrangian point of view the displacements ui are given as differences between initial and current positions of a point:
ui = xi ai .

(A.11)

This obtains the quadratic strain measure in terms of displacements:


i
h

(a
+
u
)
(a
+
u
)

dai daj
(dx)2 (da)2 = a
r
r
r
r
ij
aj
i
h


i
r
r
=
ri + u
rj + u
ai
aj ij dai daj
h
i
uj
ui
ur ur
= a
+
+
ai
ai aj dai daj .
j

(A.12)

The quantity in brackets is a measure of strain since only if strain exists the expression
(dx)2 (da)2 differs from zero. The Lagrangian strain tensor


uj
1 ui
ur ur
Lij =
+
+
(A.13)
2 aj
ai
ai aj
is invariant with respect to interchanging the indices i and j, or symmetric. The derivation
steps for (A.13) did not imply smallness of displacements or strains. Linear elasticity
assumes displacement gradients being small enough to justify neglect of the quadratic
terms. The familiar linearized strain tensor is written as


uj
1 ui
ij =
+
.
(A.14)
2 xj
xi
Fig. A.4 illustrates the geometric meaning of shear strain.
c ETH Z

urich IMES-ST, March 26, 2014

146

Equations of linear elasticity and plane problems

z
_
R
z
_
R
P
x

_
Q

Q
y
_
P

u n d e fo rm e d

_
x

u ,zd z
u ,xd x
u ,yd y

_
y

d e fo rm e d

Figure A.4: Shear strain mechanism[23].

A.1.3

Generalized Hookes law

The relation between stress tensor ij and strain tensor kl for a linear elastic solid body
is given with a constant fourth-order tensor Cijkl :
ij = Cijkl kl ,

i, j, k, l = 1, 2, 3.

(A.15)

The 81 components of Cijkl are not all linearly independent from each other. The strain
tensor is symmetric and also, in absence of body moments, the stress tensor. Thus, the
subscripts ij as well as kl are commutable and the tensor C can contain 36 independent
components only. From the deformation energy density e
Z
1
e = ij dij = Cijkl kl ij ,
(A.16)
2

the material law is extracted by taking the derivative with the respect to strains and
the elastic constants by taking the second derivative. The sequence of the derivatives is
commutable and therefore the index pairs ij as well as kl can also be exchanged. Thus
the tensor Cijkl is symmetric
Cijkl = Cklij ,
(A.17)
with only 21 independent components which can describe the most general case of an
anisotropic material.

A.1.4

Temperature and moisture strains

Changes of temperature and moisture, T and H, evoke solid-body volume changes


which are described with coefficients of thermal expansion (CTE) ij and swelling coefficients ij . Solid bodies made from anisotropic materials may also suffer changes of shape.
Hindering of the free strains by clamping or due to inhomogeneous material distributions,
such as is the case for a multidirectional laminate made from fiber composite materials,
leads to internal stressing. For capturing such effects in simulations the generalized Hooles
law (A.15) must be appended with the free temperature and moisture strains:
ij = Cijkl (kl kl T kl H) ,
c ETH Z

urich IMES-ST, March 26, 2014

i, j, k, l = 1, 2, 3.

(A.18)

A.1 Basic problem of linear elasticity

147

The systematic lies in distinguishing between mechanical strains M , total strains T , and
free strains R . Mechanical strains are directly related with stresses via the constitutive
equations (A.18). Total strains are connected with the displacement fields by the kinematic
relations (A.14). Free strains due to temperature or moisture changes can be measured
at unconstrained homogeneous solid bodies. All of these strains are related to each other
with:
M = T R .
(A.19)

A.1.5

Displacement formulation of a solid-body problem

The displacement formulation proposes an admissible displacement field as it is continuous and satisfies the essential boundary conditions. A compatible strain field follows from
applying the kinematic relations and from there the constitutive equations give the stress
fields. The latter may not be admissible as they can violate equilibrium conditions and
natural boundary conditions.
The primary solution of a problem in displacement formulation is in terms of the displacements ui , wherefore the basic problem is cast in a displacement differential equation.
First the stresses in the equations of motion (A.5) are substituted by the strains via the
constitutive equations (A.15):
Cijkl kl ,j +fi =
ui .
(A.20)
Secondly the strains are expressed in terms of displacements by use of the kinematic
relations:
1
Cijkl (uk ,lj +ul ,kj ) + fi =
ui .
(A.21)
2
The finite element method for solid-body problems follows the displacement formulation
and uses, at least for static and eigenvalue problems, the principle of the minimum of the
total potential energy. The power of the method, which is the ability to simulate problems
with the most complex domain geometries, rests on the division of the problem domain
into sub domains, the finite elements. The subdivision sets it apart from the Ritz method.
If the finite-element shape functions are fully controlled by node degrees-of-freedom, the
displacement fields are continuous over the mesh. Strains and stresses are generally not
continuous.

A.1.6

Matrix-operator notation

Symmetries of the tensors ij (in absenceof body moments), ij , and Cijkl (path independence of linear problems) permit writing the basic problem of linear elasticity in terms of
matrices and operators. Table A.1 shows how the six independent components of stress
and strain tensors are placed in one-dimensional arrays and the maximum of 27 constants
of Hookes law fill a symmetric matrix with six rows and columns. In a Cartesian system
1
2
3
4
5
6

=
=
=
=
=
=

11
22
33
23
31
12

,
,
,
,
,
,

1
2
3
4
5
6

=
=
=

23
31
12

=
=
=
=
=
=

11
22
33
223
231
212

Table A.1: Tensor and matrix (Voigt) notations

c ETH Z

urich IMES-ST, March 26, 2014

148

Equations of linear elasticity and plane problems

the subscripts 123 can be replaced with xyz. All arrays are defined as follows,

1
1
c11 c12 c13 c14 c15 c16

2
2

21 c22 c23 c24 c25 c26


ux

c31 c32 c33 c34 c35 c36


3
3
uy
=
u=
=
C=
c41 c42 c34 c44 c45 c46


uz
4

c51 c52 c35 c45 c55 c56

5
5

6
6
c61 c62 c36 c46 c56 c66

;,

(A.22)

and the constitutive equations can now be written as


= C.

(A.23)

To determine the Cij from known engineering constants, namely Youngs moduli E, shear
moduli G, and Poissons ratios , it is convenient to write the compliance matrix S for
principal material directions,

1
13
12
0
0
0
E1
E1
E1


23
1

21
0
0
0

E2 E2 E2

31 32
0
0
0
E2
E3

E
(A.24)
S= 3
,
1

0
0
0
0
0
G

23

0
0
0
0
0
G
31

1
0
0
0
0
0
G12
which is then inverted to obtain C. The determinant S of the coupled part relating to an
orthotropic material is
2
2
2
S = S11 S22 S33 S11 S23
S22 S13
S33 S12
+ 2S12 S23 S13 .

(A.25)

With it the entries of the stiffness matrix C can be expressed by those of the compliance
matrix S:
C11 =

2
S22 S33 S23
S

C12 =

S13 S23 S12 S33


S

C13 =

S12 S23 S13 S22


S

C22 =

2
S33 S11 S13
S

C23 =

S12 S13 S23 S11


S

C33 =
C44 =

1
S44

C55 =

1
S55

C66 =

2
S11 S22 S12
S
1
S66 .

(A.26)

Writing the strain-displacement relations concisely requires the symbolic linear, or differential, operator L:

/x
0
0

0
/y
0

0
/z
0
.

 = Lu,
L=
(A.27)

/z /y
0

/z
0
/x

/y /x
c ETH Z

urich IMES-ST, March 26, 2014

A.2 Two-dimensional Problems

149

The same operator L is used to express the equations of motion:


LT s + f =
u.

(A.28)

Finally the displacement differential equation is obtained by combining (A.23), (A.27) and
(A.28):
LT CLu + f =
u.
(A.29)
For taking free strains into account one modifies (A.29) by subtracting the free-strains F
from the total strains T = Lu:

LT C Lu R + f =
u.

A.2

(A.30)

Two-dimensional Problems

Simplified equations exist for describing the mechanical behavior of problems underlying
certain assumptions regarding geometry and loading. Familiar situations include plane
problems and the so-called structure elements such as rod, truss, beam, disk, plate, or
shell.

A.2.1

Plane elasticity problems

Plane problems are formally so closely related to three-dimensional problems that the
corresponding finite elements are often dubbed 2D solid elements. Consider a solid body
whose domain is bounded by two parallel planes separated by a distance h, and let
h be much smaller than the bodys extension along the planes. Let the z direction of a
Cartesian system be perpendicular to the planes and the body be loaded with forces acting
along directions within the xy plane only. Then the out-of-plane stresses are negligibly
small,
z = yz = xz = 0 ,
(A.31)
and one speaks of a plane state of stress.
A plane state of strain follows from a contrary assumption: Namely that the distance of
the bounding planes be much larger than the domain extension along the xy planes. The
special plane strain state explained in many textbooks specifies the strain component z
to be zero. In case of an isotropic material the shear strains zx and yz vanish also.
Both states allow considering the displacements within the xy plane only, and these depend
on the position within this plane only:
u = u(x, y),
v = v(x, y),
w =
0.

(A.32)

Thus, the derivatives of stresses and strains with respect to z can be neglected and the
equations of motion
x ,x +xy ,y +fx = 0
(A.33)
xy ,x +y ,y +fy = 0
as well as the kinematic equations
x = u,x y = v,y xy = u,y +v,x .

(A.34)

c ETH Z

urich IMES-ST, March 26, 2014

150

Equations of linear elasticity and plane problems

simplify accordingly. Both can be expressed with reduced linear operator L:

L=

/x

/y
.
/y /x.
0

(A.35)

The reduced material law is


x = C11 x + C12 y
y = C12 x + C22 y
xy = C66 xy
.

(A.36)

It is only the definition of the cij by which the equations of plane stress and plane strain
states differ from each other.
In case of plane stress consider the compliance matrix and note that that all stress components with a subscript z are zero, (A.31). The respective rows and columns can be
removed form the constitutive equations:

S11 S12 0
1
1
2
2
= S12 S22 0
.

6
0
0 S66
6

(A.37)

Since the matrix (A.37) is coupled in two components only it can be inverted easily by
hand and inserting the definitions of the Sij from (A.24) obtains the reduced stiffness
matrix for plane stress states for anorthotropic material:
C11 =

E1
1 12 21

C12 =

12 E1
1 12 21

C66 = G12 .

(A.38)

In case of plane strain one uses the entries of the stiffness matrix (A.26) which results from
inverting the fully populated compliance matrix (A.24):
C11 =

1 23 32
E2 E3 4

C12 =

21 + 31 23
E2 E3 4

C22 =

1 13 31
E1 E3 4

C66 = G12 ,

(A.39)

with
4=

1 12 21 23 32 31 13 221 32 13
.
E1 E2 E3

(A.40)

Isotropic materials imply the simplifications Ei = E, ij = and G = E/2(1 + ) and the


stiffness matrices for plane stress
C11 = C22 =

E
1 2

C12 =

E
1 2

C66 =

E
2(1 + )

(A.41)

and plane strain


C11 = C22 =

E(1 )
(1 + )(1 2)

states are easily verified.


c ETH Z

urich IMES-ST, March 26, 2014

C12 =

E
(1 + )(1 2)

C66 =

E
2(1 + )

(A.42)

A.2 Two-dimensional Problems

151

y (2 )

g xy(t

x y

e x(s x)
x
q

e q(s q)
x (1 )

e y(s y)

Figure A.5: Element in axisymmetric solid body and volume element [6].

A.2.2

Axisymmetry

Problems of axial symmetry are described in cylindrical coordinates and have in common
with plane problems that derivatives of the state variables exist with respect to two directions only, namely the axial and the radial. Fig. A.5 illustrates the arrangement of
a finite element as it represents the mechanical response of an axisymmetric solid body.
We align the axial direction with y and the radial direction with x. Displacements appear
along these two directions only and (A.32) continues to hold. Since the displacement along
the circumferential direction, , vanishes together with changes of all state variables the
kinematical relations in cylinder coordinates and operator notation are written as:

x

y
=

xy


 = Du

/x

/y

L=

/y /x

1/x
0

u
v

u=


(A.43)

The sketch in Fig. A.6 helps to understand the equilibrium conditions in radial direction
x and in axial direction y:
x:

x ,x

+ xy ,y + x1 (x ) + fx = 0

y : xy ,x +

y ,y

x1 xy

(A.44)

+ fy = 0

Here an orthotropic material with principal material axes parallel to the directions x, y
s

d q d x
2

d x

d q

f
x

= r w

t
x

d q x

y x

d q d x

1
2

(s
x

+ d s
x

) d q ( x + d x )

d q d x

Figure A.6: Forces acting upon a volume element of an axisymmetric solid body
c ETH Z

urich IMES-ST, March 26, 2014

152

Equations of linear elasticity and plane problems

and is assumed:

xy

c11
c12
=
c16

c13

c12 c16 c13


x

c22 c26 c23


y

c26 c66 0

xy
c23 0 c33


(A.45)

No stresses x , and y occur with this material law and the mechanical situation is called
torsion-free axisymmetric case.
.

c ETH Z

urich IMES-ST, March 26, 2014

Вам также может понравиться