You are on page 1of 81

MM2:

Numerical Theory
Lecture Notes
March 2010

Prof. Dr.-Ing. K.-U. Bletzinger


Dipl.-Ing.(FH) Falko Dieringer M.Sc.
Dipl.-Ing. Johannes Linhard

MM2: Numerical Theory

Index

Index
1

Introduction to Form Finding ................................................................................ 1 - 1


1.1
Introduction........................................................................................................... 1 - 2
1.2
Statical problem of form finding .......................................................................... 1 - 3
1.3
Methods for numerical form finding .................................................................... 1 - 7
1.3.1 Reduction of equations..................................................................................... 1 - 7
1.3.2 Force density method and updated reference strategy ..................................... 1 - 8
1.3.3 Dynamic relaxation ......................................................................................... 1 9

Differential Geometry of Surfaces and Continuum Mechanical Basics............. 2 - 1


2.1
Differential geometry............................................................................................ 2 - 1
2.1.1 Description of a point in space......................................................................... 2 - 2
2.1.2 Description of a spatially curved surface in space........................................... 2 - 3
2.2
Continuum mechanical basics .............................................................................. 2 - 6
2.2.1 Deformation gradient F.................................................................................... 2 - 7
2.2.2 Nonlinear strain measures................................................................................ 2 - 8
2.2.3 Stress measures ............................................................................................. 2 10

Plane Stress State and Principal Stresses............................................................. 3 - 1


3.1
Plane stress state ................................................................................................... 3 - 2
3.1.1 Reduced stress tensor ....................................................................................... 3 - 3
3.1.2 Reduced strain tensor ....................................................................................... 3 - 4
3.1.3 Hookean law for the plane stress state ............................................................. 3 - 5
3.2
Principal stresses................................................................................................... 3 - 6
3.2.1 Principal stress determination .......................................................................... 3 - 6
3.2.2 Mohrs circle of stress...................................................................................... 3 - 7

Numerical cutting pattern generation ................................................................... 4 - 1


4.1
Partitioning of the structure into strips ................................................................. 4 - 2
4.1.1 Definition of cutting lines ................................................................................ 4 - 2
4.1.2 Geodesic line calculation during form finding ................................................ 4 - 3
4.1.3 Geodesic line calculation on a found form ...................................................... 4 - 3
4.2
Flattening and compensation of the strips ............................................................ 4 - 4
4.2.1 Simple triangulization technique ..................................................................... 4 - 5
4.2.2 Optimization techniques .................................................................................. 4 - 5
4.2.3 Mechanical approach ....................................................................................... 4 - 6

Finite Element Method............................................................................................ 5 - 1


5.1
Linear elastic plane stress: Principle of virtual work in matrix notation .............. 5 - 2
5.2
Short introduction to the Finite Element Method ................................................. 5 - 4
5.2.1 Simple 3- and 4-node isoparametric displacement elements ........................... 5 - 4
5.2.2 Convergence behavior..................................................................................... 5 - 6
5.3
Numerical integration ........................................................................................... 5 - 9
5.3.1 One Dimensional Rules ................................................................................... 5 - 9
5.3.2 Two Dimensional Rules................................................................................. 5 - 11
5.4
The assembly process ......................................................................................... 5 - 14
5.5
Virtual work of a surface stress field, non-linear formulation............................ 5 - 19

MM2: Numerical Theory

Index

5.6
Material Law....................................................................................................... 5 - 20
5.7
Internal and external virtual work....................................................................... 5 - 21
5.8
Soap film analogy to find minimal surfaces ....................................................... 5 - 21
5.9
Discretization of the governing equations .......................................................... 5 - 22
5.10 Linearization ....................................................................................................... 5 - 24
5.10.1 Membrane element......................................................................................... 5 - 24
5.10.2 Cable element................................................................................................. 5 - 26
5.11 Application: 3-node membrane element in 2D................................................... 5 - 28
5.12 Application: 3-node membrane element in 3D................................................... 5 - 32
5.13 Solution of non-linear finite element equations.................................................. 5 - 34
5.13.1 Newton-Raphson method............................................................................... 5 - 34
5.13.2 Time stepping, path following, numerical continuation ............................... 5 - 34

MM2: Numerical Theory

Literature

Literature
[1] Otto F. and Rasch, B., 1995. Finding Form. Deutscher Werkbund Bayern, Edition A.
Menges.
[2] Berger, H., 1996. Light structures - structures of light, Basel: Birkhuser.
[3] Holzapfel G, 2000. Nonlinear Solid Mechanics. Chichester: Wiley
[4] Onate E. and Krplin B. (Eds.), 2005. Textile Composites and Inflatable Structures.
Wien, Springer.
[5] Forster B. and Mollaert M. (Eds.), 2004. European Design Guide for Tensile Surface
Structures. Tensinet.
[6] Koch, K.-M. (ed.), 2004. Membrane Structures. Mnchen, Prestel.
[7] Carmo, M., 1976. Differential geometry of curves and surfaces. Englewood Cliffs,
Prentice-Hall.

1 Introduction to Form Finding

MM2: Numerical Theory

Introduction to Form Finding

1.1
Introduction
The main characteristic of membrane or cable net structures is their slenderness: One
dimension of a membrane structure is considerably smaller than the other ones. These
structures mainly extend to two dimensions and are therefore classified as surface structures
(in contrast to continua, where all the dimensions are of the same magnitude). For a cable
even two dimensions are smaller than the third. They are therefore consequently called line
structures.
This slenderness of these lightweight structures is also the reason for their characteristic
mechanical behavior: Membranes and cables are only capable of withstanding tensile forces,
while if they are subjected to compression forces, they lose their stability: If compressed, the
membrane tries to withdraw itself from carrying the load by wrinkling, a cable will buckle.
This can easily be illustrated with a simple experiment: If a piece of thread is pulled, one has
to pull the stronger the more the thread is stretched (for linearly elastic material the force and
the stretch are proportional to each other and linked over the Youngs modulus). In contrast to
that, almost no force at all has to be applied in order to compress it (Fig. 1.1).
F>0
(tension)

F<0
(compression)

load

cable

displacement

Fig. 1.1: Mechanical behavior of a cable under tension/compression


At the very least, a possible loss of stability through compression forces is only local and just
affects the visual appearance in a negative way. But it can also have worse effects: Because of
bending and frequent changes of curvature the outer layers of a wrinkled membrane are
subjected to higher strains than usual. Their mechanical properties may suffer, if they are
exposed to this situation for a longer period of time. The worst case is a global system failure,
if the whole structure loses its stability: This is especially dangerous for statically determinate
structures, where the failure of one element is enough to make the whole system collapse.
It is therefore obvious that this obliges the designer to take special care in determining the
right form of the structure: A form has to be found, which is not only capable of
withstanding the significant load cases under certain restrictions such as maximum allowed
deflection and stress, but also shows an even usage of the material: This leads to a
minimization of necessary building material, which has the nice side effect that it leads to a
reduction of building costs as well.
For a membrane structure made of an isotropic material (like foils) the optimum shape is
reached, when the stress distribution over the whole structure is uniform. For an orthotropic
material it may be advantageous to match the prestress with the different material properties
for each fiber direction.

1-2

MM2: Numerical Theory

Introduction to Form Finding

This form finding can be done in various ways: Before the large availability of
computational calculation power, mechanical models - such as soap films or stretched tissues
- were used to find a form, which is the equilibrium shape of a certain stress distribution in the
structure with external loads. Nowadays the numerical simulation has overcome these
approaches, since the increasing complexity of the structures, although mechanical
simulations are still sometimes used for verification of the numerical calculations.
In the process of form finding the resulting stresses under a given load case (mainly
mechanical prestress, but also external loads like pressure) are prescribed. Now the still
unknown final shape, on which these stresses act, has to be determined. This is done
considering large deformations, so that the static and mechanical formulations become fully
nonlinear. In this process the numerical calculation needs to be stabilized as can be seen later
on. Until now no material parameters were needed: The presence of the prescribed stresses is
assumed, no matter how they were generated. The material properties come into play, when
the reference geometry is needed. For membrane structures this reference geometry are the
cutting patterns. They have to be cut out in such a way, so that when they are stretched in the
boundary conditions, the resulting stress state is as close as possible to the desired one.
It can easily be seen that the form finding procedure is of reversed order compared to a
normal static analysis: Here, the reference geometry (e.g. the dimensions of a concrete bar)
and the external loads (self-weight, wind, snow, etc.), for which the static analysis shall be
carried out, are given. With the help of the governing equations of the material, the kinematics
and the equilibrium it is possible to determine the deformed geometry. In a last step, the
stresses inside the structure can be calculated with the usage of the material law.
Because of this reversed order form finding is often referred to as an inverse process.
Static analysis
external loads,
reference geometry

material, kinematics
equilibrium

deformed
geometry

Form finding
stresses,
external loads

material

stresses

Patterning
kinematics
equilibrium

deformed
geometry

material

reference
geometry

Fig. 1.2: Comparison: Static analysis Form finding


1.2
Static problem of form finding
The final shape of the form finding process is the equilibrium shape for a given load case,
which is in general only mechanical prestress. The governing equations for the calculation
and their characteristic features are presented exemplarily for a simple plane cable structure,
but they can be expanded for more complicated, three-dimensional cable net and membrane
structures in the same manner.

1-3

MM2: Numerical Theory

Introduction to Form Finding

In our example, a plane weightless cable with only mechanical prestress is given. Two rigid
supports, which are separated by the distance L, keep the cable fixed at the edges. The
discretization of the cable is approximately done with two linear elements and one
intermediate nodal point K. The displacements of K in the horizontal (x) and vertical (z)
direction are the only degrees of freedom in this system. It is quite obvious that the cable must
be straight to achieve equilibrium.
l2 l2 l2 l2

L
x
z

l1

l2

x1 = x

x2 = L - x

Fig. 1.3: Plane prestressed cable in reference configuration


Formally the equilibrium condition states that the sum of horizontal as well as vertical forces
acting on node K must vanish. The results are two nonlinear equations for the two unknown
variables x and z:
S1

S2

x1

x2

H = 0:

x1
x
S1 2 S 2 =
l1
l2

V = 0:

z
l1

S1 +

z
l2

S2 =

x
x2 + z2

z2
x2 + z2

S0

S0 +

Lx

(L x )2 + z 2
z2

(L x )

+z

S0 = 0

(1.1)

S0 = 0

(1.2)

S1 and S2 stand for the respective cable forces, which are assumed to be constant for every
configuration and equal to the desired prestress force S0. The cable forces can be calculated
from the cross section area A and the prestress 0.
S1 = S 2 = S0 = A 0

(1.3)

It is easy to see that the second equation is only fulfilled, if the vertical displacement z = 0,
which is in accordance with our first assumption. Now we have only the first equation left to
determine the horizontal displacement x. Using our result z = 0 we obtain:

x
x2

S0

Lx

(L x )

S0 = (1 1)S0 = 0

1-4

(1.4)

MM2: Numerical Theory

Introduction to Form Finding

This equation is always fulfilled, that means it is independent of x. Therefore we cant use it
to determine the horizontal position of the node K. Mechanically this can be interpreted that
every position of the node K is allowed, as long as the two cable elements form a straight line.
This means that there is not a unique solution for the form finding problem.
K

Fig. 1.4: Suitable solutions


If we expand our problem type now to three-dimensional surface structures (like membranes
or cable nets), we observe the same phenomenon: We cant determine a unique position of the
nodes (nodes of the cables net or nodes of a FE discretization) on the equilibrium surface with
only the help of the equilibrium conditions: The nodes can float on the surface and still they
describe the same geometry, only out-of-plane movements are not allowed. Out of the
originally three equations only one equation remains to determine the three coordinates of
each point. Therefore we need to modify the equation system, such that a unique solution can
be found.
3D view

top view

Fig. 1.5: Suitable solutions for a hypar like shape (four point tent)
But before that, a little rewriting of the governing equations needs to be done: The
equilibrium equations are identified as the derivatives of the virtual work of the whole system
w.r.t. to the various degrees of freedom.

H = 0:

A l1

V = 0 :

A l1

W
=0
x

(1.5)

W
z
z
+ A l 2 2 0 =
=0
2 0
z
l1
l2

(1.6)

x1
l 12

0 A l 2

1-5

x2
l 22

0 =

MM2: Numerical Theory

Introduction to Form Finding

Both equations can now be combined to one equation for the virtual work W:
W =

x
x
z
z
W
W
x +
z = A l 1 21 0 A l 2 22 0 x + A l 1 2 0 + A l 2 2 0 z (1.7)
x
z
l1
l2
l1
l2

The equation above states that in a state of equilibrium no work is done, if the system
undergoes virtual (infinitesimal small) displacements.
Furthermore all the terms of the virtual work are sorted in such a way, that the corresponding
terms of each cable element are combined:
x

z
z
W = W1 + W2 = A l 1 21 x + 2 z 0 + A l 2 22 x + 2 z 0 =
l1
l2
l2
l1
= A l 1 ea 1 0 + A l 2 ea 2 0 = A ea 1 0 dl + A ea 2 0 dl = 0
l1

(1.8)

l2

In this representation all quantities refer only to the current configuration (and not to the
reference or any other configuration): li are the actual cable lengths, ea i are the so called
virtual Euler-Almansi strains and 0 is the Cauchy stress. The Euler-Almansi strains and
the Cauchy stresses form an energetically conjugated stress-strain couple and are both
referring to the actual configuration.
Since we have started our calculation from a fixed reference configuration, where the lengths
of the cables were Li, we can again rewrite the equilibrium equations, so that these initial
lengths appear:

H = 0:

x
x L

A L1 21 1 0 A L 2 22

L2
L1 l 1

L 2
W

l 0 = x = 0

(1.9)

V = 0:

z
z L

A L1 2 1 0 + A L 2 2

L2
L1 l 1

L 2
W

l 0 = z = 0

(1.10)

Now the virtual work equation is rewritten in the same manner:


x
L
x
L

z
z
W = A L1 21 x + 2 z 1 0 + A L 2 22 x + 2 z 2 0 =
L2 l 2
L1 l 1

L1
L2
= A L1 gl1 pk 1 + A L 2 gl 2 pk 2 = A gl1 pk 1 dL + A gl 2 pk 2 dL = 0
L1

(1.11)

L2

Here all quantities (as well as the integrals) refer to the initial reference configuration: The
energetically conjugated stress-strain couple in the reference configuration consists of the
virtual Green-Lagrange strains gl i and the 2nd Piola-Kirchhoff stresses pk i .

1-6

MM2: Numerical Theory

Introduction to Form Finding

Both the stress and strain measures in the reference as well as in the actual configuration are
well known in nonlinear continuum mechanics and serve as a foundation for the large
deformation analysis.
While the Cauchy stresses are real physical stresses, the 2nd Piola-Kirchhoff stresses are
quantities, which are only used for calculation, but dont have any direct physical meaning.
They can be transformed into Cauchy stresses over the ratio of reference to actual length:

pk i =

Li
0
li

If there is a difference between the cable lengths of the reference and actual configuration,
there is also a difference between the Cauchy and 2nd Piola-Kirchhoff stresses. For small
deformations (1st order theory) the ratio Li/li is negligible and therefore the difference between
Cauchy and 2nd Piola-Kirchhoff stresses is often omitted. But as we are dealing with large
deformations in the form finding procedure, this difference has to be taken into account.

1.3
Methods for numerical form finding
Since it is not possible to determine a unique solution for the position of each node purely out
of the equilibrium conditions, several different strategies have been developed to overcome
this rank deficiency.
1.3.1 Reduction of equations
As seen in the previous example the three equilibrium conditions provide us only one useful
equation for the calculation of the nodal positions regarding the out-of-plane movements, but
not for the tangential movement. Some strategies make use of this and imply the restriction
that a node is allowed to move only perpendicular to the surface. By that the equation system
is diminished and can be solved, although it is still non-linear and needs therefore to be solved
iteratively. After convergence, the equilibrium regarding the other two independent spatial
directions should be checked.

initial mesh

overlapping
edge cable - surface

Fig. 1.6: Tangential adjustment of the surface mesh

1-7

tangential adjustment
of the surface mesh

MM2: Numerical Theory

Introduction to Form Finding

This simple method works fine with fixed boundaries, while it is not suitable for structures
with edge cables: In general cables, which form the boundary of the membrane, have to move
also tangentially to the surface in the form finding process in order to find an equilibrium
position. If on the other hand the membrane is only allowed to move perpendicular to the
surface, an overlapping between the edge cables and the surface mesh will occur, which
makes a tangential adjustment of the surface mesh necessary.
1.3.2 Force density method and updated reference strategy
This method is one of the earliest methods for numerical form finding. It has been originally
developed for the calculation of the roof of the Olympic stadium in Munich 1972. Although it
was only applicable for cable nets, it is still widely used for form finding of membrane
structures: The particular membrane is then approximated by a cable net.

The basic idea behind the force density method is the assumption that the ratio of the force in
a cable Si to its length li the so called force density qi = Si/li - is constant throughout the
form finding process. By that the system of equations is modified and all equations become
linear and can be solved in one step.
If we recall the first example, the governing equations are now:

q1 x 1 q 2 x 2 = 0

(1.12)

q1 z + q 2 z = 0

(1.13)

The force density is typically evaluated in the reference configuration for a given cable force
and length. Now for every other configuration the resulting stresses in the cable can be
determined by the following equation:

i =

q i l i Si l i
=
Ai
A i Li

(1.14)

If the actual length of the cable li differs from the reference length Li, the actual stress state
differs from the desired stress state as well. Therefore a new iteration step has to made in
which the resulting configuration of the previous step is the reference configuration of the
actual step.
Formally the assumption of constant force densities coincides with the assumption of constant
2nd Piola-Kirchhoff stresses during the form finding process.
A pk 1
L1

x1

A pk 1
L1

A pk 2

z +

L2

x2 = 0

A pk 2
L2

1-8

z =0

(1.15)

(1.16)

MM2: Numerical Theory

Introduction to Form Finding

With that knowledge, the force density method can also be applied to membranes. Since it is
necessary to update the reference (for evaluating new force densities or 2nd Piola-Kirchhoff
stresses) in each iteration step, the generalized method is called updated reference strategy.
1.3.3 Dynamic relaxation
Another approach to overcome the indeterminacy of the tangential positions of the nodes is
applied in the dynamic relaxation method, which is physically based on the second Newtonian
law of motion: The structure is modeled by nodes with concentrated masses on which not
only static forces, but also inertia and damping forces act. The nodes are linked by discrete
finite elements representing membranes and cables. The equilibrium shape is seen as the rest
position of a fictive damped oscillation of the system.

For our example the static equations are expanded with inertia and damping terms (resulting
from dAlemberts principle). In the time step tn they can be written as follows:

m a nx + C v nx +

S1 n S 2 n
x1
x2 = 0
l1
l2

(1.17)

m a nz + C v nz +

S1 n S 2 n
z +
z =0
l1
l2

(1.18)

For the mass m and the damping coefficient C suitable assumptions are made. This is
possible, since we dont want to analyze the real dynamical behavior of the system. The
accelerations ai and the velocities vi are approximated with a finite difference scheme for a
fixed time increment t:

1 n +1
x xn
t
1 n +1
=
z zn
t

v nx +1 2 =
v nz +1 2

)
)

1 n +1 2
vi
v in 1 2
t
1
v in = v in +1 2 + v in 1 2
2
a in =

(1.19)

)
(1.20)

After some manipulation of the equations, we get the following recursive formulas for the
velocities vx and vz:
v nx +1 2 =

S1 S 2 n S 2
2m C t n 1 2
2t
L

vx
+ x
l2
2m + C t
2m + C t l 1 l 2
1444
424444
3
R nx

1-9

(1.21)

MM2: Numerical Theory

v nz +1 2 =

Introduction to Form Finding


S1 S 2 n
2m C t n 1 2
2t

vz
+ z
2m + C t
2m + C t l 1 l 2
1442443

(1.22)

R nz

Finally the position of the node in the next time step can be calculated by extrapolation:
x n +1 = x n + t v nx +1 2

mit v1x 2 =

t R 0x
2 m

(1.23)

This scheme can be applied to bigger systems in a similar way: The simple recursive formulas
still persist and only relatively small equation systems have to be solved. The disadvantage of
this approach is the dependency of the numerical stability to the choice of mass, damping
coefficient and incremental time step: For example, the incremental time step has to be
smaller than its critical value, which is the inverse of the eigenfrequency of the undamped
system: t < t krit = 2 m K max .
An alternative to the viscous damping was proposed by Barnes: The nodal velocities are set to
zero as the kinetic energy reaches a certain maximum.

1 - 10

2 Differential Geometry of Surfaces


and Continuum Mechanical Basics

MM2: Numerical Theory

2.1

Differential Geometry and Continuum Mechanical Basics

Differential geometry

2.1.1 Description of a point in space


In order to uniquely describe the position of a point in space, we need to introduce a fixed
reference frame: This is usually a global Cartesian coordinate system, which consists of three
base vectors ei (i = 1,2,3) of unit length which are orthogonal to each other: The scalar
product of two different base vectors is 0, while it is 1 between the same base vectors.
0
ei e j =
1

,i j
,i = j

(2.1)

This type of basis is called orthonormal basis.


Another requirement is the orientation: The base vectors have to be orientated in such a way
that they form a right handed coordinate system: If the thumb of your right hand points in the
direction of the first base vector and your index finger in the direction of the second, the
direction of the third base vector is described by your middle finger. In a mathematical way
this can be written as:
e 3 = e1 e 2

(2.2)

Now the position vector x of a point P (pointing from the origin of the coordinate system to
the point P) can be uniquely described through a linear combination of multiples of these base
vectors:
3

x = x 1e1 + x 2 e 2 + x 3e 3 = x i e i = x i e i

(2.3)

i =1

x e3
e3
1

x e1

e1

e2

x e2

Fig. 2.1: Position vector x of a point P


The three scalars xi are the global Cartesian coordinates of the respective point (it is important
to know that coordinates are only meaningful in a combination with the corresponding
basis!).

2-2

MM2: Numerical Theory

Differential Geometry and Continuum Mechanical Basics

In the equation above Einsteins summation convention was used: If an index is repeated
(only once) in the same term, it is called a dummy index and a summation over the range of
this index is implied (unless otherwise indicated). Note, that Latin dummy indices usually run
from 1 to 3, while Greek run from 1 to 2.
2.1.2 Description of a spatially curved surface in space
One possibility to describe a spatially curved surface, which is still a two-dimensional entity
although it lies in the three-dimensional space, is the so called parametrical description: To
every point P on the surface two independent surface coordinates or surface parameters 1
and 2 are assigned. The position vector x, which points to the point P, is therefore a function
of these two surface parameters:

x = x 1 , 2

(2.4)

g1

g2

P(1,2)

x( , )
e3
e1

e2

Fig. 2.2: Parametric description of a spatially curved surface


If one parameter is kept constant and only the other one is varied, all points P, which fulfill
this restriction, form a coordinate line of the varied coordinate. These lines are generally
spatially curved.
In order to describe the local behavior of the surface at each point, a local reference frame is
needed: The so called covariant base vectors g1 and g2 are defined by differentiation of x with
respect to 1 and 2:

g1 =

x
;
1

g2 =

x
2

(2.5)

The covariant base vectors are tangential to the corresponding coordinate lines, e.g. g1 is
tangential to the coordinate line, which is generated by varying 1 and keeping 2 constant.
The tangential plane of the surface at point P is spanned up by g1 and g2. The surface normal
is determined by g3 which is defined by the normalized cross product of g1 and g2:
g3 =

g1 g 2
;
g1 g 2
2-3

g3 = 1

(2.6)

MM2: Numerical Theory

Differential Geometry and Continuum Mechanical Basics


g3

g2

g1

Fig. 2.3: Surface normal direction g3


The covariant base vectors, which span up the tangential plane, are in general not orthogonal
and not of unit length. The 1 and 2 are consequently called curvilinear coordinates.
The scalar products gij of the covariant base vectors the components of the covariant metric
tensor reflect the metric of the surface, i.e. the length of the covariant base vectors and the
angle between them:
g11 = g1 g1 = g1

2
2

g 22 = g 2 g 2 = g 2

(2.7)

g12 = g 21 = g1 g 2 = g1 g 2 cos (g1 , g 2 )


The three covariant base vectors gi form the local covariant basis. A local contravariant basis
consisting of three contravariant base vectors gi is defined by the rule
1 , i = j
g i g j = ij =
0 , i j

(2.8)

where ij is the Kronecker delta: If the indices of the Kronecker delta are identical, its value is
1, if they are different, its value is 0.
g

g2

g1
g

Fig. 2.4: Covariant and contravariant base vectors


This means that e.g. g1 is orthogonal to g2 (vice versa g2 is orthogonal to g1). The scalar
products g1 g1 and g 2 g 2 are 1. The contravariant base vectors g1 and g2 span up the same
plane (the tangential plane) as the covariant base vectors g1 and g2.

2-4

MM2: Numerical Theory

Differential Geometry and Continuum Mechanical Basics

Alternatively, the contravariant base vectors are defined as the derivative of the surface
coordinates with respect to the geometry of the surface:
1
g =
;
x

2
g =
x

(2.9)

Since g1 and g2 are orthogonal to g3 it follows that g 3 = g 3 and g 3 = 1 .


It can be proven that the contravariant metric tensor is the inverse of the covariant metric
tensor and vice versa:
g 11
g = 21
g
ij

g12
= g ij
g 22

( )

g
= 11
g 21

g12
g 22

(2.10)

Co- and contravariant base vectors can be transformed into each by use of the metric tensors:
g i = g ijg j ;

g i = g ijg j

(2.11)

2.1.2.1 Differential line element


The differential line element ds is linking the point P on the surface with the position vector
x(1,2) with another point in the vicinity of P with the position vector x(1+d1,2+d2). It
can be calculated as follows:

ds =

x 1 x
d + 2 d 2 = g1d1 + g 2 d 2
1

(2.12)

The differential length ds can be obtained by evaluating the length of the vector ds:

( )

ds = ds ds = g11 d1

( )

+ 2g12 d1d 2 + g 22 d 2

(2.13)

2.1.2.2 Differential area element


The differential piece of area da is defined as the area of the vector parallelogram which is
given by the vectors g1d1 and g2d2:

da = g1d1 g 2 d 2 sin g1d1 , g 2 d 2 = g1 g 2 d1d 2

( )

= (g1 g 2 ) g 3 d1d 2 = det g ij d1d 2

(2.14)

The total area of a given surface can be expressed in terms of the surface coordinates as
a=

j d d
1

1 2

g1 g 2
1 2

where j is the determinant of the Jacobian matrix.

2-5

d1d 2

(2.15)

MM2: Numerical Theory

Differential Geometry and Continuum Mechanical Basics

ds

g2d

g1d
1

P( , )

x( +d , +d )
1

x( , )

Fig. 2.5: Differential line and area


In the following we will describe any vector or tensor as the weighted sum of the co- or
contravariant base vectors. E.g. the vector a or the second order tensor T may be written in
terms of either gi or gi.
a = a igi = a igi ;

T = T ijg i g j = Tijg i g j

(2.16)

The normal letters represent the coefficient of a tensor, the bold faced letters the
corresponding basis. is the tensor product.
If a tensor is written in matrix notation, it implies that the corresponding basis is a Cartesian
basis. The entries of the matrix are then the tensor coefficients.
E.g., the second order unit tensor I is defined in terms of the co- or contravariant base vectors
as
I = g ijg i g j = g ijg i g j

(2.17)

2.2
Continuum mechanical basics
In continuum mechanics we want to observe how a body deforms during a certain process:
Starting from a known reference configuration, we trace each of the material points of the
body throughout the deformation process until the current or actual configuration. By that we
are able to calculate deformations, stretches, etc.
This material description is often referred to as Lagrangian description. The so called
Eulerian description on the other hand doesnt pay attention to a material point, but to a fixed
point in space: We look at this specific point and study what happens there during the
deformation process (e.g. we observe the velocities of the particles passing that point). This
approach is widely used in fluid mechanics.

2-6

MM2: Numerical Theory

Differential Geometry and Continuum Mechanical Basics

2.2.1 Deformation gradient F


We now look at a surface (membrane) which is parametrized with the two convective surface
coordinates 1 and 2. The coordinate lines are quasi carved in the surface - the
parameterization doesnt change during the deformation process. As a simplification we
assume that the thickness of the membrane stays constant throughout the deformation.
We adopt the convention: All quantities indicated with a capital letter refer to the initial or
reference configuration, quantities with a lower case letter to the current configuration:
reference configuration
deformation

current configuration

G2

g2

G1

u( , )
1

g1
x( , )
1

X( , )
1

e3
e1

e2

Fig. 2.6: Deformation of a surface


A point P on the actual surface with position x(1,2) had originally the position X(1,2) on
the reference surface. The point is identically identified by its surface coordinates 1, 2 which
stay invariant during deformation. The displacement u of point P is defined as:

) (

) (

u 1 , 2 = x 1 , 2 X 1 , 2

(2.18)

The definition of the base vectors Gi and Gi is as before and stems from the differentiation of
X with respect to the surface coordinates.
The reference configuration is transformed into the actual configuration by the deformation
gradient F. It can be calculated by deriving the actual geometry with respect to the reference
geometry.
F=

x
= gi Gi ; FT = Gi gi
X

(2.19)

The inverse of the deformation gradient F transforms the actual into the reference geometry:

2-7

MM2: Numerical Theory

Differential Geometry and Continuum Mechanical Basics


F 1 =

X
= G i g i ; F T = g i G i
x

(2.20)

E.g., base vectors are transformed from one configuration to the other one as follows:
g i = F G i = g j G j G i = ijg j = g i
G i = F 1 g i = G j g j g i = ijG j = G i
g i = F T G i = g j G j G i = ijg j = g i

(2.21)

G i = F T g i = G j g j g i = ijG j = G i

A differential piece of area transforms by


da = jd1d 2 = g1 g 2 d1d 2
= det FdA = det FJd1d 2 = det F G 1 G 2 d1d 2

g1 g 2
da
j
= det F = =
dA
J
G1 G 2

(2.22)

(2.23)

The surface area of the actual configuration can now be written with respect to the reference
configuration:
a = da =
a

jd d
1

1 2

j
= dA = det FdA
J
A
A

(2.24)

2.2.2 Nonlinear strain measures


The stiffness of membrane structures is relatively low compared with conventional structures
(e.g. structures made of concrete or steel). Thus large deformations can be observed when the
structure is loaded. In order to simulate this mechanical behavior of a membrane in a correct
way, nonlinear strain measures have to be applied: Their main feature is that rigid body
movements mustnt lead to any strains.
One strain measure fulfilling all these conditions is the Green-Lagrange strain tensor E:
E=

) (

1 T
1
F F I = g ij G ij G i G j
2
2

(2.25)

It can be calculated with the deformation gradient or with the difference of the metric
coefficients. In order to get physical strains the usually contravariant basis has to be
transformed into a Cartesian basis: The corresponding coefficients represent the strains.
Example
We look at a straight cable linking the points P1 and P2. The cable is parametrized with one
coordinate 1 ranging from 0 (P1) to 1 (P2). The initial cable length is L, the deformed cable
length is l.

2-8

MM2: Numerical Theory

Differential Geometry and Continuum Mechanical Basics

reference configuration
P2

deformation

G1

current configuration

P1

P1

g1

X2=X( =1)

P2

x1=x( =0)
1

X1=X( =0)

x2=x( =1)
e3
e1

e2

Fig. 2.7: Example: Strain calculation at a cable


The position vectors of the cable points in the reference and actual configuration are:

( ) ( )
x( ) = (1 )x

X 1 = 1 1 X1 + 1 X 2
1

+ 1x 2

(2.26)

Since a cable is only a one-dimensional entity, we have only one meaningful covariant base
vector (the other two are perpendicular to the first, but their direction is not uniquely defined):
X
= X 2 X1
1
x
g 1 = 1 = x 2 x1

G1 =

(2.27)

The covariant metric coefficients are:


G 11 = X 2 X1
g11 = x 2 x1

= L2

= l2

It can be seen that the length of the covariant base vector is the length of the cable of the
corresponding configuration.
The contravariant metric coefficients can be calculated from the inverse of the covariant
metric tensor (which reduces to a scalar for this example):

2-9

MM2: Numerical Theory

Differential Geometry and Continuum Mechanical Basics


G 11 =
g

11

1
L2
1

(2.28)

l2

A local Cartesian basis ~ei is introduced in the reference configuration, whose first base vector
is parallel to the cable. The contravariant base vectors can therefore be written as:

G 1 = G 11G 1 =

1
(X 2 X1 ) = 12 L~e1 = 1 ~e1
2
L
L
L

(2.29)

Now we can evaluate the Green-Lagrange strain tensor:

1
1 2
1 l 2 L2 ~ ~
1
1
2
1
1
E = (g11 G 11 )G G = l L G G =
e1 e1
2
2
2 L2

(2.30)

The coefficients of the Green-Lagrange strain tensor, whose basis is Cartesian, represent the
physical strains: In this example we have only one strain component (parallel to the cable
1 l 2 L2
axis), which is constant over the cable length and has the value
.
2 L2
2.2.3 Stress measures
2.2.3.1 Cauchy stress tensor
In general when we talk about physical stresses, we mean the Cauchy stress tensor without
actually knowing it: The Cauchy stress tensor lives in actual configuration: Its natural basis
is the covariant basis in the actual configuration (and therefore components are contravariant):
= ij g i g j

(2.31)

Using the Cauchy stress tensor we can calculate the force dt acting on a differential piece of
area da on the boundary of a body with the normal vector n as follows:

dt = n da

2 - 10

(2.32)

MM2: Numerical Theory

Differential Geometry and Continuum Mechanical Basics

reference configuration

current configuration
n
dT=PN dA=dt
dt=n da

F
dA

-1

da

-1

dT=SN dA=F dt

Fig. 2.8: Various stress measures


2.2.3.2 1st Piola-Kirchhoff stress tensor P
It is often helpful to have a stress measure which is based not in the actual, but in the
reference geometry. A first step is the 1st Piola-Kirchhoff stress tensor P. Using P you can
calculate the force dT acting on a differential piece of area dA with the normal vector N:

dT = P N dA

(2.33)

This force dT has the same magnitude and direction as the force dt.
The 1st Piola-Kirchhoff stress tensor can be obtained by a transformation of the Cauchy stress
tensor:
P = det F F T = P ijg i G j

(2.34)

It can be seen that P lives with its first leg in the actual configuration, while it lives with its
second leg in the reference configuration. It is therefore called a two point tensor. The 1st
Piola-Kirchhoff stress tensor is in general not symmetric, which can be a disadvantage.
Because of that we introduce a symmetrical stress tensor in the reference configuration:
2.2.3.3 2nd Piola-Kirchhoff stress tensor S
The 2nd Piola-Kirchhoff stress tensor S is generated by a pull-back operation of the Cauchy
stress tensor:
S = det FF 1 F T = SijG i G j

(2.35)

The 2nd Piola-Kirchhoff stress tensor is in contrast to the 1st Piola-Kirchhoff stress tensor a
symmetrical tensor.
Using S a force dT acting on a differential piece of area dA with the normal vector N can be
calculated:

dT' = S N dA = F 1 dt

2 - 11

(2.36)

MM2: Numerical Theory

Differential Geometry and Continuum Mechanical Basics

Note, that the force dT in general hasnt the same magnitude and direction as dt, although
they are linked by the deformation gradient.
2.2.3.4 Summary of the transformation rules
The transformation rules of the various stress measures are:
1
1
P FT =
F S FT
det F
det F
ij
T
P = P g i G j = det F F = F S

= ij g i g j =

(2.37)

S = SijG i G j = F 1 P = det FF 1 F T

The coefficients transform as follows:


1
1 ij
P ij =
S
det F
det F
P ij = det F ij = Sij

ij =

Sij = P ij = det F ij

2 - 12

(2.38)

3 Plane Stress State and


Principal Stresses

MM2: Numerical Theory

Plane Stress State and Principal Stresses

3.1
Plane stress state
The mechanical model of a membrane is typically a two dimensional surface, nevertheless the
real membrane is of course a three dimensional body: The main reason for this reduction of
one dimension is to reduce the number of degrees of freedom and thus simplify the model.
Whenever a structural model with less dimensions than the real structural element is used, one
has to know the mechanical behavior in the dimension which is not included in the model!
This information is then used to construct a two dimensional constitutive model of an
originally three dimensional body.
The usual assumptions for a 2D membrane element are:

Thickness is much smaller than the other two dimensions of the membrane

Stresses are constant over the thickness (only normal forces as stress resultants, no
bending moments)

No stresses perpendicular to the midplane of the membrane (simulation of force


transmission is not possible)

The membrane can freely deform itself in thickness direction

All these assumptions can be summarized as plane stress state.

midplane

reference geometry

deformed geometry

Fig. 3.1: Plane stress state of a differential membrane element


The opposite of the plane stress state is the plane strain state: In this case the strains
perpendicular to the midplane are zero instead of the stresses. This model is e.g. applicable to
structures, which are much longer in one direction than in the other two and which are
continuously loaded and supported perpendicular to this direction (e.g. a dam). It is usually
not necessary to analyze the whole structure, but sufficient to cut out a two dimensional
segment, as the strains along the longitudinal axis have to be zero.

3-2

MM2: Numerical Theory

Plane Stress State and Principal Stresses

plane stress: membrane

plane strain: dam

3
1

2 1
3

Fig. 3.2: Comparison: plane stress - plane strain


3.1.1 Reduced stress tensor
We now look at a point P in the midplane of a membrane: At this particular point we
introduce a local Cartesian coordinate system in such a way that the third base vector e3 is
perpendicular to the midplane (e1 and e2 lie in the midplane).

22

21

12

11

e3
e1

e2
P
12

11

22

21

Fig. 3.3: Definition of local Cartesian coordinate system and positive stress components
The stress state at this point P is defined by the stress tensor .
11 12

= ijei e j = 21 22
31 32

13

23
33

(3.1)

The result of a post multiplication of the stress tensor with a unit vector n is a traction
vector t acting on a cutting plane with the normal vector n:
t = n

(3.2)

The traction vector t can be additively decomposed into a component which is perpendicular
to the cutting plane (thus parallel to the normal vector n) and a parallel component:
t = t + t ||

3-3

(3.3)

MM2: Numerical Theory

Plane Stress State and Principal Stresses

The length of t is the normal stress , the length of t || the shear stress (note that stresses
are scalar quantities).
Another possibility to evaluate the normal stress is to project the traction vector t on the
normal vector n.
= t = t n = ( n ) n
= t || = t t n

(3.4)

The tensor coefficients ij directly represent the physical stresses, if a Cartesian basis was
used for representation. The first index i indicates the direction of the stress, the second index
j the normal vector of the cutting plane on which the stress acts (note that this convention may
vary along different authors!). Two coinciding indices indicate normal stress, two different
indices shear stress.
As we assume to have a plane stress state, all stress components in the third direction vanish:

13 = 31 = 23 = 32 = 33 = 0

(3.5)

We additionally know that the stress tensor has to be symmetric to guarantee equilibrium of
momentum:

12 = 21

(3.6)

Thus only three significant stress components are left which can be written in a vector instead
of a matrix, which is widely known as Voigt notation.
11

= 22
12

(3.7)

3.1.2 Reduced strain tensor


A similar procedure is done for the strain tensor:
The full strain tensor at the point P is:

11 12
= ije i e j = 21 22
31 32

13
23
33

(3.8)

No shear deformations perpendicular to the midplane occur:

13 = 31 = 0
Again, the strain tensor has to be symmetric:

3-4

(3.9)

MM2: Numerical Theory

Plane Stress State and Principal Stresses

12 = 21

(3.10)

We now have four strain components left: The three components, which correspond with the
components of the reduced stress tensor, are again written in a vector:

11
= 22
212

(3.11)

Note the factor 2 in front of component 12 describing the shear deformation. This is due to
the convention of the engineering community to work with the shear angle instead of the
tensor components 12!

= 212

(3.12)

The strain component 33 describing the thickness change is a function of the components of
the reduced strain tensor and the material. It is therefore not an independent parameter.
3.1.3 Hookean law for the plane stress state
The stresses in a body are linked to the strains over a constitutive model. For the simplest case
of an isotropic, linear elastic material this relation can by written as:

1
0

22
E
1
0
=
2

1
12

0 0

11

11

22
212

(3.13)

C is the so called constitutive or elasticity matrix (in general it is a fourth-order elasticity


tensor). E represents the Youngs modulus, the Poissons ratio.
The inverse relationship is as follows:

C 1
0
1
1
0
1

E
0
0 2(1 + )

=
11
=
22
212

11
22

12

(3.14)

The strain in the thickness direction can be determined as:


33 =

11
(11 + 22 )
+ 22 =
E
1

3-5

(3.15)

MM2: Numerical Theory

Plane Stress State and Principal Stresses

3.2
Principal stresses
In the previous chapters we have seen that the magnitude of the normal and shear stresses
depends on the considered cutting plane: If the cutting plane is changed, the normal and shear
stresses change, although we still have the same stress tensor .

There exists a cutting plane, on which no shear stresses act and the normal stresses become a
maximum: The cutting plane is rotated until the traction vector t acting on this plane is
parallel to the normal vector n, which is necessary to fulfill the condition of vanishing shear
stresses.
cutting plane with
arbitrary stress state

cutting plane with


principal stress state
t = 0n

n
t

t||

Fig. 3.4: Different stress states due to rotation of cutting plane


The traction vector t can thus be written as a multiple of the normal vector n (which has unit
length):
t = on

(3.16)

0 represents the so called principal stress, the direction of the normal vector the principal
stress direction.
3.2.1 Principal stress determination
We now want to determine the principal stresses and their direction of a stress tensor which is
given in matrix notation referring to an arbitrary local Cartesian basis:
11
= 21

12

22

(3.17)

We have seen that the traction vector t has to be parallel to the normal vector n to fulfill the
condition of vanishing shear stresses. This can be written as:
t = n = 0 n = ( 0 I ) n

(3.18)

I is the identity tensor, which can be written in matrix notation as the identity matrix:
1 0
I=

0 1

3-6

(3.19)

MM2: Numerical Theory

Plane Stress State and Principal Stresses

The governing equation can be rearranged and leads to an eigenvalue problem for the
principal stress 0:

( 0 I ) n = 0

(3.20)

The principal stresses 0 are the eigenvalues, the principal stress directions the eigenvectors
of the stress tensor.
Since the vector n is defined to have unit length, the equation can only be fulfilled if the
tensor ( 0 I ) is singular: A tensor is then singular if its determinant vanishes:
11 0
det ( 0 I ) =
21

=
0

12
22

( )

2
= ( 0 )2 11 + 22 0 + 11 22 12 =

(3.21)

= ( 0 )2 tr 0 + det = 0
The resulting equation is the characteristic polynomial, which is quadratic for a plane stress
state. The generally two independent solutions are the principal stresses of the stress tensor .
The principal stress direction n for a particular principal stress can be obtained by solving the
governing equations of the eigenvalue problem (0 has now a fixed value and isnt a variable
anymore). Note that only the line of action of the eigenvectors is defined, but not their length
and orientation.
It can be proven that the principal stress directions are orthogonal to each other (as all
eigenvectors). So if the cutting plane is parallel to e.g. principal stress direction 1, the stress
acting on that plane is principal stress 2 (and vice versa).
3.2.2 Mohrs circle of stress
If the normal and shear stresses on two cutting planes which are perpendicular to each other
are known, it is possible to determine the principal stresses (or every possible stress state)
graphically via Mohrs circle of stress.
As an example the following situation is given:
2 = 5.0
2 = 1.5
e2
1
1

e1

2
2

Fig. 3.5: Example - Given stress state

3-7

1 = 1.5
1 = 2.0

MM2: Numerical Theory

Plane Stress State and Principal Stresses

We now denote the stress states of the two different cutting planes in a diagram: The
horizontal axis of the diagram represents the normal stresses , the vertical axis the shear
stresses . If the shear stress is turning anti clockwise it is defined as positive shear stress.
In the next step Mohrs circle of stress is constructed: The center of the circle M is situated at
the intersection of the normal stress axis and the line connecting the two points which
describe the two stress states. The radius is the distance from the center to one of the two
points.
Every point on the circle is describing a possible stress state: The extremal normal stresses are
obtained at the intersection of the circle with the normal stress axis (these are the principal
stresses). The maximum shear stress is equal to the radius of the circle.

(1=2.0,1 =1.5)

r
M

min=1.38

max=5.62

(2=5.0,2 =1.5)

Fig. 3.6: Example - Mohrs circle of stress


The results can be proved analytically. The characteristic polynomial of the eigenvalue
problem is:
2.0 1.5
1 0
det ( 0 I ) = det
0

=
1
.
5
5
.
0
0
1

= 0 2 7.0 0 + 7.75 = 0
The two principal stresses are the solutions of this characteristic polynomial, which
correspond with the solutions obtained by Mohrs circle of stress:
7.0 7.0 2 4 7.75
0 1,2 =
2
min = 1.38,
max = 5.62
The principal stress directions can also be determined over Mohrs circle of stress, but the
procedure is not as straight forward as the procedure for determining the principal stresses.
Therefore only the analytical method is presented:
We now want to determine the direction of the first principal stress min. After plugging it in
the governing equation system we get:
3-8

MM2: Numerical Theory

Plane Stress State and Principal Stresses

2.0 1.5
1 0 n 1 0.62 1.5 n 1 0

1.5 5.0 1.380 1 n = 1.5 3.62 n = 0


2
2

The two equations are linearly dependent (multiplying the first equation with 2.41 results the
second equation). Therefore we can choose any of these two equations. Since initially the
length of the eigenvectors is not defined we just set n1 = 1 and use the first equation:

[0.62

1
1.5] = 0.62 + 1.5n 2 = 0
n 2

By that we obtain the second component of the direction vector:


n2 =

0.62
= 0.41
1 .5

The vector is now normalized to have unit length:


1

n min =

12 + ( 0.41)2

1 0.93
0.41 = 0.38

As we know that the eigenvectors are orthogonal to each other we can easily determine the
principal stress direction for the maximum normal stress max as:
0.38
n max =

0.93
2 = 5.0

max = 5.62

2 = 1.5
e2
1
1

e1

min

1 = 1.5

nmax

1 = 2.0
nmin

2
max

Fig. 3.7: Example Principal stresses and principal stress directions

3-9

min = 1.38

4 Numerical cutting pattern


generation

MM2: Numerical Theory

Numerical cutting pattern generation

The result of the form finding process is generally a structure with a doubly curved surface
with an inherent prestress state. The building material e.g. fabrics, foils, etc. on the other
hand comes in plane unstretched panels. This discrepancy has to be overcome in the
patterning process: Patches are cut out of the material according to certain cutting patterns
such that the resulting form and prestress state is as close as possible to the desired one when
they are attached together and moved into their boundary conditions.
The general procedure for cutting pattern generation is as follows:
4.1
Partitioning of the structure into strips
The three dimensional structure is cut into strips along certain cutting lines. The governing
factors for the placement of these cutting lines are:

material properties (especially the shear stiffness)

available panel width and length

curvature of the surface

main load carrying paths

aesthetical reasons

etc.

4.1.1 Definition of cutting lines


The simplest way to define cutting lines is to intersect the surface with a plane. The
disadvantage of this approach is the banana-like shape of the resulting cutting patterns, which
increases the cut-off of membrane material.
Therefore it is proposed by several authors that cutting lines should follow geodesic paths on
the surface: A geodesic path or line on a doubly curved surface is the equivalent to a straight
line on a plane. They are commonly defined as the line with the minimum distance between
two points on the surface (although this is not the exact mathematical definition): It has to be
paid attention to the fact that this includes also local minima. There may exist more than one
geodesic line with different lengths connecting two points P1 and P2 (e.g. on a cylinder):
While the shortest distance represents the global minimum, the other distances represent the
local minima.

4- 2

MM2: Numerical Theory

Numerical cutting pattern generation

P2

geodesic line 2
(local distance minimum)
geodesic line 1
(global distance minimum)

P1

Fig. 4.1: Several geodesic lines on a cylinder


4.1.2 Geodesic line calculation during form finding
It is possible to include the task of geodesic line calculation in the form finding process: Cable
elements with high prestress are included in the FE mesh along an initial guess for the desired
geodesic line. For each iteration step the resulting residual forces of these cable elements are
calculated for each node of the FE mesh and all components normal to the surface are set to
zero. If the node is either the starting or end point of the geodesic line the whole residual force
vector is set to zero. As a result of this the cable elements representing the geodesic line affect
only the in plane positions of the nodes and not the general form finding process.
The advantage of this integrated procedure is that all finite elements are automatically pulled
in the right places. That means the geodesic line doesnt intersect any elements and
therefore no remeshing is necessary.
The disadvantage is that an initial guess for the geodesic line has to be made on a shape which
may be far from the final one. This makes it often necessary to make two form finding runs:
First without and the second time with geodesic line calculation
4.1.3 Geodesic line calculation on a found form
If a geodesic line is to be calculated on an already existing shape, one can make use of a
physical analogy: An elastic cable with high prestress is fixed on the starting point of the
geodesic line and pulled to the end point. By allowing it to slide over the surface it proceeds
to a configuration with a minimum of potential energy. The final position of the cable
represents the geodesic line. The problem here is that the cable nodes must have the
possibility to move freely on the surface described with the FE mesh. Therefore a contact
formulation between the cable and the surface is necessary which makes it numerically more
complicated.
Another possibility, which was originally developed in the computer graphics sector, is the
geodesic line calculation over triangulated surfaces: As indicated in the name it is only
applicable if the structure is meshed with linear triangles. The advantage of linear triangles is
their well defined geometry which is essentially a plane between the nodal points.
At first a discrete geodesic line consisting only of element edges of the FE mesh is created.
This is usually done with a fast forward marching algorithm: For each node the minimum

4- 3

MM2: Numerical Theory

Numerical cutting pattern generation

geodesic distance to the starting node is calculated. Then one starts from the end node and
looks at all neighboring nodes: The one with the minimum distance to the starting node is
now included in the geodesic line. This new node is now set to be the end node of the
geodesic and the procedure of finding the neighboring node with minimum distance and
including it to the geodesic line is continued until the starting point is reached.
This discrete geodesic line is then optimized using an angular approach: The positions of the
nodes of this geodesic are now not restricted anymore to coincide with the positions of the FE
nodes. They are also allowed to be part of the edges linking two FE nodes. The nodes of the
geodesic line are moved along the edges of the FE mesh in such a way that the two angles
between the edge of the FE mesh containing the node of the geodesic line and the two
adjoining elements of the geodesic line are identical. Geometrically this can be explained as
follows: When the two finite elements containing the edge with the node are developed into a
plane the geodesic line should form a straight line.
discrete geodesic on
triangulated surface

surface
Pend

optimized geodesic on
triangulated surface
Pend

Pend

Pstart

n
Pstart

Pstart

development
into a plane

Fig. 4.2: Geodesic line calculation on triangulated surfaces


Both of the presented methods for geodesic line calculation on a found form have in common
that the resulting geodesic line is generally intersecting the FE mesh which requires at least a
remeshing of the intersected elements. It has to be paid attention to the fact that the quality of
the FE mesh may suffer if only the intersected elements are remeshed and not the whole
mesh.
4.2
Flattening and compensation of the strips
The strips which are now obtained by cutting the three dimensional structure along the
generated cutting lines are still doubly curved and prestressed. The following cutting pattern
generation generally consists of two parts: The doubly curved strip is first flattened into a
plane and then compensated (scaled down) to generate the prestress in the three dimensional
structure. This flattening and compensation can be combined into a single process.
4- 4

MM2: Numerical Theory

Numerical cutting pattern generation

The main difficulty here is the general non-developability of the membrane strips. A spatial
surface is only then developable into a plane without any distortions when the Gaussian
curvature K, which is the product of the two principal curvatures at each point, is zero at
every point of the structure. Since this is generally not the case several methods have been
developed to minimize the resulting distortions which can lead to deviations from the desired
prestress state.
4.2.1 Simple triangulization technique
The idea behind this technique is to make a generally non-developable surface developable:
This is done by remeshing the membrane strips with linear triangles: All nodes except the
nodes along the cutting lines of the longer edge are deleted. Then the patch is remeshed with
triangles according to the initial segmentation of the cutting lines (only one element along the
shorter edge of the strip). Now it is possible to develop the geometry described by this mesh
into a plane: The mesh is simply unfolded.
The disadvantage of this method is that it doesnt take into account any material parameters
and the inside geometry. Therefore it is only suitable for strips where the curvature of the
membrane parallel to the shorter edge is relatively low.
The compensation is usually done with empirically determined scale factors.
spatial surface

FE mesh

simple triangulization

planar development

Fig. 4.3: Simple triangulization technique


4.2.2 Optimization techniques
More sophisticated approaches use optimization techniques for the cutting pattern generation.
Since a huge variety of different methods exists, only a few are presented in note form:
4.2.2.1 Minimizing the cable length difference
In this approach the FE mesh is seen as a cable net. The edges of the elements represent the
cables. If the cable net is then flattened, it is impossible to keep all cable lengths constant in
the case of an originally doubly curved surface. A least-squares minimization technique is
applied to minimize the sum of the squares of the length difference between the 3D and 2D
configuration.
The objective function, which is to be minimized, is:

4- 5

MM2: Numerical Theory

Numerical cutting pattern generation

P=

n edges

(l 3D,i l 2D,i )

minimize

(4.1)

i =1

l3D,i represents the length of the ith cable element in the three dimensional configuration, l2D,i
the length in the flattened two dimensional configuration.
Again it is not possible to take into account any material parameters.
4.2.2.2 Minimizing the stress difference
The difference between the resulting and desired stress state is minimized in these
approaches. The constraints are the equilibrium conditions. Here it is possible to include the
effects of anisotropic material. Since these methods need more mechanical and mathematical
background, this is beyond the scope of the course.
4.2.3 Mechanical approach
A mechanical analogy is used here for cutting pattern generation: The doubly curved
membrane strip is assumed to consist of an elastic material.
In a first step, the membrane is pressed into an arbitrary plane. It has just to be taken care that
the flattened membrane mustnt overlap itself. Due to this deformation elastic stresses arise in
the membrane which are not in equilibrium anymore. In a second step, the membrane is
released in the plane while keeping the boundary conditions statically determinate. Thus the
membrane can relax itself until it reaches a residual eigenstress state. The corresponding
equilibrium shape is the cutting pattern.
It is possible to do simultaneous compensation: Then the prestress is included, which causes a
contraction of the membrane.
The mechanical modeling has to be done under the hypothesis of large deformations.
Therefore the resulting equations become non-linear and must be solved iteratively.
The advantage of this method is the possibility to take into account anisotropic material
behavior and prestress states.

Fig. 4.4: Cutting pattern generation using mechanical approach

4- 6

5 Finite Element Method

MM2: Numerical Theory

Finite Element Method

5.1
Linear elastic plane stress: Principle of virtual work in matrix notation
We start from the principle of virtual work for a 3-dimensional body:

w = (w int + w ext ) = ij ij d b i u i d g i u i d

(5.1)

where ij and ij are the components of stress and strain tensors, bi and gi the components
of body and surface loads, respectively.

z, w

g
y, v

z, w

y, v

x, u

x, u
3D continuum
plane stress/strain continuum
Fig. 5.1: Reduction of general 3D continuum to plane stress/strain
The special case of a thin 2-dimensional plane stress/strain structure is considered by
splitting the integrals of 5.1 into an integral over the thickness and into one over the midsurface A or along the edges E:
w = ij ij dA dt b i u i dA dt g i u i dE dt = 0
t A

t A

t E

As the stresses are constant through the thickness the integration over t can be done in
advance (pre-integration) which transfers stresses into stress resultants and external loads to
area q or line loads p, respectively. Using a matrix notation the principle of virtual work now
displays as:

w = n T dA q T u dA p T u dE = 0
A

where

n x t x
x


p x t g x
q x t b x
n = n y = t y = y dt ; q = =
; p = =
; t = const.
p y t g y
q y t b y

t
n xy t xy
xy
x
x

u
u
= y ; u =
= y ; u = ;
v
v

xy
xy
5-2

MM2: Numerical Theory

Finite Element Method


ny

nyx
px

qy

nxy
nx
qx
x
py

Fig. 5.2: In-plane loading and internal forces of 2D membrane


The constitutive equations are also written in matrix notation, introducing the constitutive
matrix C:
n = tC

plane stress

plane strain
1

1
0
C=

(1 + )(1 2 )
1 2
0
0
2

1
0

1
0
C=
2
1
1

0 0
2

At any point the strain is related to the displacement u by

x
= Lu = 0

0
u
y v

x

(5.2)

The matrix L is a differential operator matrix.


Putting all together:
n = t C = t CLu
= L u

and, finally:

w = t u T LTC L u dA q T u dA p T u dE = 0
A

5-3

(5.3)

MM2: Numerical Theory

Finite Element Method

This equation represents equilibrium of plane stress / plane strain states in terms of the
unknown displacement field uT = (u(x,y), v(x,y)). It is the basis for finite element procedures
to determine an approximate solution of u. The virtual work equation is also called the weak
form of equilibrium.
5.2

Short introduction to the Finite Element Method

5.2.1 Simple 3- and 4-node isoparametric displacement elements


The principal idea of the finite element method is to reduce a continuous problem to a
problem of a finite number of discrete parameters. The solution of the discrete problem gives
an approximation of the continuous one. Here, we assume that the displacement field can be
described by number of discrete displacement values which are defined at the finite element
nodes. Several nodes together form a finite element (e.g. 3 node triangle, 4 node rectangle,
Fig. 5.3). Inside the element, i.e. between the nodes, the displacement field is approximated
by a linear combination of shape functions, each of them related to one node of the element.
3-node triangle
y

4-node rectangle
y

1
m

x
Fig. 5.3: Two simple plane stress/strain finite elements

b
4
x

The shape functions can be defined with respect to the x,y-coordinate system as:
3-node triangle:

1
[x 2 y 3 x 3 y 2 + (y 2 y 3 )x + (x 3 x 2 )y]
2
1
[x 3 y1 x 1 y 3 + (y 3 y1 )x + (x 1 x 3 )y]
N 2 ( x , y) =
2
1
[x 1 y 2 x 2 y1 + (y1 y 2 )x + (x 2 x 1 )y]
N 3 ( x , y) =
2
N 1 ( x , y) =

where is the element area: = 1 / 2[x 1 (y 2 y 3 ) + x 2 (y 3 y1 ) + x 3 (y1 y 2 )]


4-node rectangle:
N 1 ( x, y) = 1 / 4(1 + )(1 + )

N 2 ( x , y) = 1 / 4(1 )(1 + )

N 3 ( x, y) = 1 / 4(1 )(1 )

N 4 ( x , y) = 1 / 4(1 + )(1 )

5-4

MM2: Numerical Theory

Finite Element Method


2
(x x m )
=
a
1
(x 1 + x 2 ) y m =
xm =
2
=

with

a = x1 x 2

2
(y y m )
b
1
(y1 + y 4 )
2

b = y1 y 4

For the triangle the displacement field is defined as:

u (x, y ) N1 (x, y ) u 1 + N 2 (x, y ) u 2 + N 3 (x, y ) u 3


u (x , y ) =
=

v(x, y ) N1 (x, y ) v1 + N 2 (x, y ) v 2 + N 3 (x , y ) v 3


or by separation of shape functions Ni and nodal displacements ui, vi using a matrix formulation:

N1
u (x , y ) =
0

N2

N3

N1

N2

u1

v1

0 u 2
= N (x , y ) v
N 3 v 2

u 3

v 3

(5.4)

Inserting 5.30 into the strain-displacement relation (5.2) yields


(x , y ) = L u(x, y ) = L N(x, y ) v
x

(x , y ) = 0

N (x , y ) v =
y


x
0
y 3 y1
y2 y3

= 21 0
x3 x2
0

x 3 x 2 y 2 y 3 x 1 x 3

y1 y 2

x1 x 3

y 3 y1

x 2 x1

x 2 x1 v

y1 y 2
0

(x , y ) = B v

The differential operator matrix B is the discrete equivalent of L, now relating discrete
nodal displacement with an approximation of the strain field . The procedure is equivalent
for the 4-node rectangle or any other displacement finite element.

5-5

MM2: Numerical Theory

Finite Element Method

All expressions are inserted into the virtual work equation (5.3):
w = (w int + w ext ) = t u T LT C L u dA q T u dA p T u dE = 0
A

w int = t v T B T C B v dA = v T t B T C B dA v = v T k v
A

w ext = q T (x , y ) N(x , y ) dA + p T (x , y ) N(x , y ) dE v = f T v


E

And, finally, the element stiffness matrix k and the equivalent nodal force vector f are defined as:

k = t B T C B dA
A

f = N T q dA + N T p dE
A

which contribute to the system stiffness matrix and force vector.

5.2.2 Convergence behavior


The quality of a finite element analysis is shown for the example of a cantilever beam discretized by 4-node elements:
P=1

t=1
h=1

l = 10

w exact =

P l3
= 10 3
EI

w
max =

Pl
= 60
W

Fig. 5.4: Cantilever beam with concentrated load

5-6

E = 4106; = 0

MM2: Numerical Theory

Finite Element Method

Element discretization:
20 elements, 33 nodes, 60 degrees of freedom (dof)

40 elements, 63 nodes, 120 degrees of freedom (dof)

160 elements, 205 nodes, 400 degrees of freedom (dof)

3
Fig. 5.5: Several different discretizations
Convergence behavior:

w/w exact

1
0,8
0,6
0,4
0,2
0
0

200

400

dof

/ exact

Fig. 5.6: Displacement of cantilever trip


60
50
40
30
20
10
0
0

200

400

dof

Fig. 5.7: Convergence of surface stresses

5-7

MM2: Numerical Theory


10

Finite Element Method


20

30

40

50

60

surface

2
1

exact solution

10
x
Fig. 5.8: Distribution of surface stress

5-8

MM2: Numerical Theory

Finite Element Method

5.3
Numerical integration
Integration is essential to determine stiffness matrices and load vectors, e.g. (5.48) and (5.49).
It is only possible, however, to evaluate the integration analytically for some few geometric
types of finite elements, as e.g. linear triangular, rectangular and rhombic, plane geometries.
The reason is that the nominators of the contravariant base vectors are functions of the natural
coordinates and . As a consequence, the integrand is a non-uniform rational function
which cannot be integrated analytically in a closed form.

For the most of elements, in particular isoparametric elements, the standard practice has been
to use numerical Gaussian integration rules because such rules use a minimal number of sample points to achieve a desired level of accuracy. This property is important for efficient element calculations because we shall see that at each sample point we must evaluate a matrix.
The fact that the location of the sample points in the Gauss rules is usually given by nonrational numbers is of no concern in digital computation.

5.3.1 One Dimensional Rules


Thee standard Gauss integration rules in one dimension are:
1

i =1

f () d w i f (i )

(5.5)

Here p 1 is the number of Gauss Integration points, wi are the integration weights, and i are
sample-point abscissas in the interval [-1, 1]. The use of the interval [-1, 1] is no restriction,
because an integral over another range, say [a, b] can be transformed to the standard interval
via a simple linear transformation of the independent variable, as shown below. The first four
one-dimensional Gauss rules, also depicted in Figure 5.9, are:
1

one point :

f () d 2 f (0)

f () d f ( 1

3 +f 1

) (

f () d 9 f (

5
8
3 5 + f (0) + f
9
9

two points :

three points :

35

(5.6)

four points :

f () d w f ( ) + w f ( ) + w f ( ) + w f ( )
1

3 = 2 =

(3 2

6 5 7 , 4 = 1 =

w 1 = w 4 = 12 16 5 6 ,

(3 + 2

65 7

w 2 = w 3 = 12 + 16 5 6

The four rules integrate exactly polynomials in of orders up to 1. 3, 5 and 7, respectively. A


one-dimensional rule with p points integrates exactly polynomials of order up to 2p - 1.

5-9

MM2: Numerical Theory

Finite Element Method

A more general integral, such as f over [a, b] in which l = b a > 0, is transformed to the canonical interval [-1, 1] as follows:
b

f (x ) dx = f () det J d

(5.7)

in which and detJ are defined by the mappings:


x = 12 a (1 ) + 12 b (1 + ) =
i.e.

1
2

(a + b ) + 12 (b a )

2
(x 12 (a + b )) and det J = dx =
d
ba

1
2

(b a ) = 12 l

(5.8)

Fig. 5.9: The first four one-dimensional Gauss rules p = 1, 2, 3,4 illustrated over the line segment = -1 to = +1. Gauss point locations arc marked with black bullets. The radii of these
bullets are proportional to the integration weights. (Figure from C. Felippa)
Example:
consider the function:

f = 4 + 12 x 7 x 2 + 1.25 x 3

5 - 10

MM2: Numerical Theory

Finite Element Method


b

F = f (x ) dx ;

determine the integral

a = 1; b = 4

N1 =

shape functions:

1
2

(1 );

N1 =

x = N 1a + N 2 b =
b

1
2

1
2

(1 + )

(a + b ) 12 (a b ) = 5 + 3
2

F = f (x ) dx = f ( ) det J d

mapping:

d etJ =

dx 3
=
d 2

Fp = w i f ( i ) det J

numerical integration:

i =1

results: f is cubic function, p must be p 2 to get exact results


p
exact
1
2
3

x2

x1
2.5
1.6340
1.3381

3.3660
2.5

x3

3.6619

f(x1)

f(x2)

f(x3)

1.7813
2.3717
2.5185

4.7533
1.7813

7.4565

Fp
10.6875
5.3438
10.6875
10.6875

5.3.2 Two Dimensional Rules


Thee simplest two-dimensional Gauss rules arc called product rules and are obtained by applying one-dimensional rules to each independent variable in turn. To be able to apply these
rules we must first reduce the integrand to the canonical form:
1 1

f (, ) d d

(5.9)

1 1

Then, we can process numerically each integral in turn:


1 1

p1

1 1

i =1 j=1

p2

f (, ) d d w i w j f (i , j )

(5.10)

where p1 and p2 are the number of Gauss points in the and directions, respectively. Usually the same number p = p1 = p2 is chosen if the shape functions are taken to be the same in
the x and y directions. This is the case for all quadrilateral elements. The first four twodimensional Gauss product rules are illustrated in Fig. 5.10.

5 - 11

MM2: Numerical Theory

Finite Element Method

Fig. 5.10: The first four two-dimensional Gauss product rules p = 1, 2, 3, 4 depicted over a
straight-sided quadrilateral region. Gauss points are marked with black bullets. The areas of
these bullets are proportional to the integration weights. (Figure from C. Felippa)
A general surface integral is transformed to the canonical form as follows:
1 1

f (x, y ) dx dy = f (, ) det J d d

(5.11)

1 1

x y

where detJ is the determinant of the Jacobian matrix, e.g. in 2D:


x x
J=
;
y y

det J = x y x y

(5.12)

The columns of J are the components of the covariant base vectors. More general in 3D:

x x x

J = y y y = [g1 g 2

z z z

g 3 ]; det J = (g1 g 2 ) g 3

(5.13)

In the special case of surfaces in 3D space, where g3 is orthogonal to g1 and g2 and of unit
length, det J can be determined as det J = g1 g 2 .
The determinant detJ transforms the differential piece of area from natural (, ) to spatial
coordinates (x, y):
dA = dx dy = det J d d

5 - 12

(5.14)

MM2: Numerical Theory

Finite Element Method

Examples: determine Jacobian matrix and base vectors for given mappings
surface in 2D

surface in 3D

5
1
x = 3+ +
2
2
1
3
y = 2+ +
2
2

5
1
x = 3+ +
2
2
1
3
y = 2+ +
2
2
3
z = 1+ +
2

1 1
1 5
g1 = ; g 2 =
2 3
2 1

1
5
1

g g
1
1
1
g1 = 1; g 2 = 3; g 3 = 1 2 =
1
g1 g 2
2
2
6
2
2
3

J=

1 5 1

;
2 1 3

det J =

x
d

5 1 2 6

1
7
J = 1 3 2 6 ; det J = g1 g 2 =
6

2
4
3 2 4 6

7
2

g2

y
d

dA= detJ d d

g1

y
d

x
d

Fig. 5.11: Geometrical interpretation of detJ

5 - 13

MM2: Numerical Theory

Finite Element Method

5.4
The assembly process
The assembly process is considerably simplified if the finite element implementation has the
following properties:

1. All elements are of the same type. For example: all elements are plane stress elements.
2. The configuration of degrees of freedom (dof) at each node is in principle the same: First,
the horizontal than the vertical dof.
3. There are no "gaps" in the dof numbering sequence.
4. The system stiffness matrix is stored as a full symmetric matrix.
If the first four conditions are met the implementation is quite simple because the element
freedom table described below can be constructed "on the fly" from the element node numbers. The last condition simplifies the merge process.
7

12
5

7
5

9
6

10
3

2
1

11

8
1

(a) Element and node numbers

7
3

(b) Degrees of freedom

Fig. 5.12: Sample structure to illustrate assembly procedure


The plane membrane structure, a short cantilever, shown in Figure 5.12 will be used to illustrate the details of the assembly process. The structure has 8 elements, 9 nodes and 12 degrees
of freedom. Note, that the support conditions (built in left edge of the cantilever) are directly
considered. There are no degrees of freedom specified to nodes 1, 4 and 7.
Begin by clearing all entries of the 12 x 12 system stiffness matrix K to zero, so that we effectively start with the null matrix:

5 - 14

MM2: Numerical Theory

Finite Element Method

0
0

0
0

0
K =
0

0
0

0
0

0
0
0
0
0
0
0
0
0
0
0
0

0
0
0
0
0
0
0
0
0
0
0
0

0
0
0
0
0
0
0
0
0
0
0
0

0
0
0
0
0
0
0
0
0
0
0
0

0
0
0
0
0
0
0
0
0
0
0
0

0
0
0
0
0
0
0
0
0
0
0
0

0
0
0
0
0
0
0
0
0
0
0
0

0
0
0
0
0
0
0
0
0
0
0
0

0
0
0
0
0
0
0
0
0
0
0
0

0
0
0
0
0
0
0
0
0
0
0
0

0
0
0

0
0

0
0

0
0

0
0

1
2
3
4
5
6
7
8
9
10
11
12

(5.15)

The numbers written after each row of K are the global freedom numbers, Fig. 5.12 (b).
Element (1) goes from node 2 over 5 to 1. The local numbering is w.r.t. the origin of the natural coordinate system (, ). In the sequence the node at the origin is the first one, than that in
direction of and at last that in direction of .
Element
1
2
3
4
5
6
7
8

1st node
2
4
4
8
2
6
6
8

2nd node
5
1
5
7
3
5
9
5

3rd node
1
5
7
5
5
3
5
9

The degrees of freedom of the nodes have the global numbers:


Node 1st dof 2nd dof
1
2
1
2
3
7
8
4
5
3
4
6
9
10
7
8
5
6
9
11
12

5 - 15

MM2: Numerical Theory

Finite Element Method

These numbers are collected into an array called the element freedom table, or EFT for short.
Element freedom tables, EFT
Element 1 dof 2nd dof 3rd dof 4th dof 5th dof 6th dof
1
1
2
3
4
2
3
4
3
3
4
4
5
6
3
4
5
1
2
7
8
3
4
6
9
10
3
4
1
2
7
9
10
11
12
3
4
8
5
6
3
4
11
12
st

The element stiffness matrix of element (1) is listed below with the EFT entries listed after the
matrix rows:

K (1)

73 31 21 10 52 21 1

31 73
21 52 10 21 2

21 21
21
0
0
21 3

=
10 52 0
52 10
0 4

52 10
0
0
10 52

21 21 21 0
0
21

Merging the element into (5.15) gives


73 31 21 10

31 73
21 52

21 21
21
0

10 52 0
52

0
0
0
0

0
0
0
0
K=
0
0
0
0

0
0
0
0

0
0
0
0

0
0
0
0

0
0
0
0

0
0
0
0

0 0 0 0 0 0 0 0

0 0 0 0 0 0 0 0

0 0 0 0 0 0 0 0

0 0 0 0 0 0 0 0

0 0 0 0 0 0 0 0

0 0 0 0 0 0 0 0

0 0 0 0 0 0 0 0

0 0 0 0 0 0 0 0

0 0 0 0 0 0 0 0

0 0 0 0 0 0 0 0

0 0 0 0 0 0 0 0

0 0 0 0 0 0 0 0

5 - 16

1
2
3
4
5
6
7
8
9
10
11
12

MM2: Numerical Theory

Finite Element Method

Element stiffness matrix of element 2

K ( 2)

73 31 21 10 52 21

31 73
21 52 10 21

21 21
21
0
0
21

=
10 52 0
52 10
0

52 10
0
0 3
10 52

21 21 21 0
0
21 4

merged into (5.15):


73 31 21 10

31 73
21 52

21 21
73
0

10 52 0
73

0
0
0
0

0
0
0
0
K=
0
0
0
0

0
0
0
0

0
0
0
0

0
0
0
0

0
0
0
0

0
0
0
0

0 0 0 0 0 0 0 0

0 0 0 0 0 0 0 0

0 0 0 0 0 0 0 0

0 0 0 0 0 0 0 0

0 0 0 0 0 0 0 0

0 0 0 0 0 0 0 0

0 0 0 0 0 0 0 0

0 0 0 0 0 0 0 0

0 0 0 0 0 0 0 0

0 0 0 0 0 0 0 0

0 0 0 0 0 0 0 0

0 0 0 0 0 0 0 0

Element stiffness matrix of element 8

K (8)

73 31 21 10 52 21 5

31 73
21 52 10 21 6

21 21
21
0
0
21 3

=
10 52 0
52 10
0 4

52 10
0
0 11
10 52

21 21 21 0
0
21 12

5 - 17

1
2
3
4
5
6
7
8
9
10
11
12

MM2: Numerical Theory

Finite Element Method

merged into (5.15):


73 31 21

31 73
21

21 21
94

10 52 0

0
0
21

0
0
21
K=
0
0
0

0
0
0

0
0
0

0
0
0

0
0
0

0
21
0

0
0
0 0 0 0
0
0
52

0
0
21 21 0 0 0 0
21

125 10 52 0 0 0 0 10
0

10
73 31 0 0 0 0 52 21

52 31 73 0 0 0 0 10 21

0
0
0
0 0 0 0
0
0

0
0
0
0 0 0 0
0
0

0
0
0
0 0 0 0
0
0

0
0
0
0 0 0 0
0
0

0
10 52 10 0 0 0 0 52

0
21 21 0 0 0 0
0
21
10

0 0 0 0

1
2
3
4
5
6
7
8
9
10
11
12

By this procedure all element stiffness matrices contribute to the over all system stiffness matrix (which is not given here). The same procedure is applied to build the system load vector
R. In the context of non-linear analysis R is called unbalanced force vector.
Finally, we can state the system stiffness equations:
Ku = R

(5.16)

from which we can solve for the system vector of node displacements u which are the problem discretization parameters. Later, we will call the vector of discretization parameters also
b. Again using the element freedom table the node displacements are related to the individual
elements, forming the element vector v of node displacements. Through the B-matrices the
element strains are determined, e.g. the linearized strain (theory of first order):
=Bv

(5.17)

Applying the material law the stresses are determined by

= C = CB v

(5.18)

1212
C

(5.19)

Note, the special arrangement of , and C in vectors or matrices:


11


= 22 ;
12

11

= 22 ;

12

C1111

C = C 2211

5 - 18

C1122
C 2222
0

MM2: Numerical Theory

Finite Element Method

5.5
Virtual work of a surface stress field, non-linear formulation
Consider a stress field which acts tangential to a surface and which is assumed to be in selfequilibrium (Fig. 5.13). E.g., one can imagine the stress field as the resulting stresses caused
by initial pre-stress and stresses due to deformation of a textile membrane within a rigid
boundary. However, at this point it is not necessary to specify any special material since we
consider the stress field as given, no matter how it was generated. The problem is to find the
geometry of the related surface which allows the stress field to be in equilibrium. The governing equation is the principle of virtual work

w = t :
a

(u )
da = t : u, x da = 0
x
a

(5.20)

which states that the virtual work of a stress field in equilibrium vanishes. is the prescribed
Cauchy stress tensor which acts on the surface in equilibrium, u,x is the derivative of the
virtual displacement with respect to the geometry of the actual surface. The thickness of the
membrane is denoted by t. It is comparatively thin and assumed to be constant during deformation, i.e. the Poisson effect in thickness direction is neglected. This is in accordance with
the behavior of available membrane material.

y
x

Fig. 5.13: Tangential surface stress

u,x can be expressed in terms of the deformation gradient. With u = x we get:


u, x =

(u ) (x ) X
x 1
1
=

=
F = F F
x
X x
X

(5.21)

(5.21) inserted into (5.20) and integration over the reference area A:
w = t : u, x da = t : u, x det F dA =t : (F F 1 ) detF dA = 0
a

(5.22)

Rearrangement of tensors gives the alternative equivalent formulation:

w = t : (F F 1 ) detF dA = t detF ( F T ) : F dA = 0
A

5 - 19

(5.23)

MM2: Numerical Theory

Finite Element Method

Further rearrangement of (5.23) leads to


w = t detF(F F 1 F T ) : F dA
A

= t (F (detFF 1 F T )) : F dA

(5.24)

= t (F S ) : F dA = 0
A

where S is the 2nd Piola-Kirchhoff stress tensor. S is related to by:


= g g
S = detF F 1 F T = detF G G = S G G

(5.25)

The components S of the 2nd Piola-Kirchhoff stress tensor are identical to the components
of the Cauchy stress tensor multiplied by detF. However, they are related to the base vectors of the reference configuration whereas the Cauchy stress components are related to the
actual configuration. S and are identical, if actual and reference configurations are identical.
Note, that all stress components normal to the surface are zero, i.e.:
i 3 = 3 i = Si 3 = S3 i = 0

(5.26)

Introducing the Green-Lagrange strain tensor E


E=

1
2

(F

F I)

(5.27)

and making use of the symmetry of S the virtual work can be written in terms of S and E:

w = t S : E dA = 0

(5.28)

5.6
Material Law
In general, a material law describes the relations between strains and stresses. For a theory to
be consistent, specific strain and stress measurements must be energetically conjugated. For
example, the following pairs are conjugated:

strain
displacement field derivative

stress
u
x

Cauchy stress tensor

Green Lagrange strain tensor E

2nd Piola-Kirchhoff stress tensor S

Deformation gradient F

1st Piola-Kirchhoff stress tensor SPK1

Linearized strain

Linearized stress lin

5 - 20

MM2: Numerical Theory

Finite Element Method

The well known Hookes law states a linear relation between the linearized values and lin:

lin = C : lin

(5.29)

where C is the material or constitutive tensor which is of order 4. It is valid for the linearized
theory of first order where infinitesimally small displacements are assumed. The coefficients
of C are defined by two independent material parameters, e.g. by Youngs modulus E and
Poissons ratio . The constitutive tensor takes special structures in the case of plane stress or
plane strain.
Elastic behaviour in the case of simultaneously large displacements and small strains can be
described the St. Venant-Kirchhoff material law which assumes linear relations between
Green-Lagrange strain and 2nd Piola-Kirchhoff stresses. Formally, it is identically with
Hookes law:

S = C:E

(5.30)

It is valid for non-linear structural behaviour and strains not larger than, lets say, 5%. It fits
very well for the geometrically non-linear analysis of elastic membrane structures.
Making use of (5.30) and inserted into the equation of virtual work (5.28) we get:

w = t S : E dA = t E : C : E dA = 0
A

(5.31)

Usually, material parameters are defined as physical coordinates. Therefore, it is straight forward to define the constitutive tensor C w.r.t. to a Cartesian basis of base vectors ei:

C = Cijklei e j e k e l ;

ei e j = ij

(5.32)

For anisotropic material (textiles) the directions of ei define the material directions (weft and
warp). In the case of plane stress e3 must be normal to the surface.
5.7

Internal and external virtual work

Considering the most general case of an elastic membrane subjected to pre-stress S and surface load q the equation of virtual work (5.28) writes in terms of the reference configuration
as:

w = t S : E dA q u det F dA = t S + E : C : E dA q u det F dA = 0 (5.33)


A

5.8
Soap film analogy to find minimal surfaces
Minimal surfaces are defined by an isotropic, homogeneous surface stress field of constant
magnitude. More precisely we look at the deformed, actual configuration and define the prestress in that state. In other words, we are talking about Cauchy type pre-stress . As in this
context elastic deformation and external loading are not considered the governing equation is
given by:

5 - 21

MM2: Numerical Theory

Finite Element Method

w = t detF ( F T ) : F dA = 0

(5.34)

As an isotropic homogeneous stress field is prescribed, i.e. = I is a scalar multiple of the


unit tensor I, (5.23) reduces to

w = t detF F T : F dA = 0

(5.35)

(5.23) is the theoretical basis of the form finding procedure of membrane structures. The manipulation of allows for a broad range of structural shapes other than minimal surfaces. It is
the governing degree of freedom in the design procedure.
5.9
Discretization of the governing equations
The equations (5.23) or (5.33) are solved numerically by the finite element method. Geometry
and displacements are discretized by standard displacement elements. I.e. the surface geometry and the displacement field are piecewise approximated by the interpolation of nodal coordinates or displacements, respectively (Fig. 5.14):
X = N k (1 , 2 ) X k ; u = N k (1 , 2 )uk ; x = N k (1 , 2 ) (X k + uk )
nel

nel

nel

k =1

k =1

k =1

u1

1
2

3
X3

(5.36)

u( , )
2

X( , )

z, Z

u4

X4

y, Y
x, X
Fig. 5.14: Finite element surface

The upper bar denotes nodal values, e.g. X k is the reference position vector of the k-th element node. Nk are the standard C0-continous shape functions, and nel is the number of nodes
of one finite element. For the sake of simplicity no difference in notation between real and
approximated fields is made.

5 - 22

MM2: Numerical Theory

Finite Element Method

The shape functions of a three node element are:


X1
X1k N 1 = 1 1 2 N1 / 1 = 1 N1 / 2 = 1
3

2
2
1
k
=
X
N
; N 2 / 1 = 1 ; N 2 / 2 = 0
X k ; N 2 =
3 k =1 3
2
N 3 / 1 = 0
N 3 / 2 = 1
X
X k N 3 =

(5.37)

and of a four node element:


1
1
(1 1 ) (1 2 ); N 2 = (1 + 1 ) (1 2 )
4
4
1
1
N 3 = (1 + 1 ) (1 + 2 ); N 4 = (1 1 ) (1 + 2 )
4
4
N1 =

(5.38)

The covariant base vectors are determined straight forward to be:


nel
X
X
,
N, k X k
=
=

k =1
nel
x
g = = x, = N,k (X k + uk )

k =1

G =

(5.39)

which gives for the three node triangle

G1 = X 2 X1 ; g1 = (X 2 + u2 ) (X1 + u1 )

G 2 = X3 X1 ; g 2 = (X3 + u3 ) (X1 + u1 )

(5.40)

The nodal displacement components are defined to be the free discretization parameters of
form finding or structural analysis. They are arranged in a column matrix b of dimension ndof
(number degrees of freedom). Variation of any entity, e.g. the deformation gradient F, means
now variation with respect to the free parameters. By use of the chain rule of differentiation
we get:

F =

g
F
b r = g G = b r G ; r = 1, K , ndof
b r
b r

(5.41)

and
nel
g
u
g = x, =
b r = N, k k b r ; r = 1, K, ndof
b r
b r
k =1

(5.42)

where br is the r-th component of b, i.e. the r-th degree of freedom of the discretized problem.
Discretization of (5.33) for non-linear structural analysis or (5.35) for form finding yields:

E
u
dA q
det F dA b r = 0
w = t S + E : C :
b r
b r
A

A
5 - 23

(5.43)

MM2: Numerical Theory

Finite Element Method


w = b r t detF F T :
A

F
dA = 0
b r

(5.44)

which have to be fulfilled for any choice of b. Finally, we arrive at the non-linear systems of
ndof equations:

Structural analysis of membranes:

w
E
u
= t S + E : C :
dA q
det F dA = 0 (5.45)
b r

b
b
r
r
A
A

Form finding of minimal surfaces:

w
F
= t detF F T :
dA = 0
b r
b r
A

5.10

(5.46)

Linearization

5.10.1 Membrane element


The governing discrete systems of equations (5.45) and (5.46) are non-linear in terms of the
discretization parameters br. They are solved iteratively by subsequent linearization using the
Newton-Raphson method. Linearization yields:
Structural analysis (5.45):

w
E
u

E
= t S + E : C :
LIN
dA q
det F dA + b s t
S + E:C :
dA = 0 (5.47)
b r
b r
b s
b r
b r A
A
A
Here, to reduce complexity, we assumed that the direction of loading q does not follow the
deformation (as it is the case for pneumatic structures). Furthermore, E is linear dependent
from discretization parameters br, i.e. the second derivatives w.r.t. br vanish.


E
dA
S + E:C :
b r
b s
A

Stiffness matrix:

K rs = t

Unbalanced force vector:

Rr = q
A

(5.48)

u
E
det F dA t S + E : C :
dA
b r
b r
A

(5.49)

Further evaluation of the stiffness matrix making use of the covariant base vectors g of the
actual configuration, yields:

2w
= t Ti T (g , r g )Tj Cijkl Tk T (g g ,s )Tl + S + S (g ,r g ,s ) dA
b r b s
A

K rs =

),r = and
b r

),s =
b s

(5.50)

Note, that Cijkl are the components of the constitutive tensor w.r.t. to the Cartesian base vectors ei which define the material directions. The transformation matrix Ti moderates between
the material directions ei and the contravariant base vectors G of reference configuration:

5 - 24

MM2: Numerical Theory

Finite Element Method


Ti = G ei

(5.51)

The fully non-linear stiffness matrix (5.50) can be decomposed into linear and non-linear contributions:
(i) Linear partition, theory of first order, elastic stiffness matrix Ke, defined w.r.t. to base vectors G of the undeformed reference configuration:

K e rs = t Ti T (g ,r G )Tj Cijkl Tk T (G g ,s )Tl dA

(5.52)

(ii) Non-linear, initial displacement matrix Ku:


K u rs = t Ti T (g ,r g )Tj Cijkl Tk T (g g ,s )Tl dA
A

+ t Ti T (g ,r g )Tj Cijkl Tk T (G g ,s )Tl dA

(5.53)

+ t Ti T (g ,r G )Tj Cijkl Tk T (g g ,s )Tl dA


A

with g = g G
(iii) Non-linear, geometrical or initial stress matrix Kg:

K g rs = t S + S (g ,r g ,s ) dA;

S = C E

(5.54)

Form finding of minimal surfaces (5.46):

w
F

F
= t det F F T :
detF F T :
dA = 0
LIN
dA + b s t
b r
b s
b r
b r
A
A
r, s = 1,K, ndof

Stiffness matrix:


F
det F F T :
dA
b s
b r
A

K rs = t

= t det F (g , r g

)(g

,s

g ) (g ,r g )(g ,s g

)] dA

(5.55)

(5.56)

Unbalanced force vector:

R r = t det F F T :
A

5 - 25

F
dA = t det F (g , r g ) dA
b r
A

(5.57)

MM2: Numerical Theory

Finite Element Method

Form finding of membranes with generalized Cauchy type pre-stress:

w
LIN
b r

= t det F F T :
dA + b s t
b r
A
A b s

F
det F F T :
b r

dA (5.58)

F
det F F T :
dA
b r

= t det F (g ,r g ,s ) + det F , s (g ,r g ) dA

A b s

K rs = t

Stiffness matrix:

(5.59)

Unbalanced force vector: R r = t det F F T :


A

F
dA = t det F g ,r g dA (5.60)
b r
A

5.10.2 Cable element


The flexible edges of membranes are reinforced by cables. During the form finding procedure
the cable forces are assumed to be given. The same principle ideas as above apply to determine the unbalanced force vector and the stiffness matrix of a cable element. First, we take
(5.23) and reduce it to one dimension. The remaining relevant covariant base vector g1 is tangential to the cable. The other two base vectors g2 and g3 are orthogonal to g1 and of unit
length. As a consequence, det F reduces to be det F = ||g1|| / ||G1||. The remaining stress component acts in direction of g1 along the cable axis. Again, the stress tensor can be set to be a
multiple of the unit tensor: = I .
The virtual work of a cable element displays as:
w = A c det F ( F T ) : F dS = A c
S

= A c
S

g1
G1

g1
G1

( g

g1 g1 g1 G1 ) : g1 G1 dS

11

(5.61)

g11 (g1 g1 ) dS

where integration is performed along the arc length S of the reference configuration and the
cable cross section area Ac is assumed to be constant during deformation, i.e. the Poisson effect is neglected.
We further discretize the cable by simple two-node elements (Fig. 5.15). By this assumption
we can set ||g1|| = l, ||G1|| = L, g11 = 1/l2, and perform the integration explicitly. We get:

w = A c

S A c
A c
l 1
(
)
(
)
(g1,r g1 ) b r
b
L
b
g

=
g

=
1, r
1
r
1, r
1
r
L
L l2
l

(5.62)

where the last expression of (5.62) reflects the alternative formulation using the 2nd PiolaKirchhoff pre-stress S :

5 - 26

MM2: Numerical Theory


from

Finite Element Method

L
S = S I = det F F 1 I F T = I
l

we get

L
S =
l

(5.63)

The unbalanced force vector and the stiffness matrix display as


A c
(g1,r g1 )
l
A c
(g1,r g1,s ) 12 (g1,r g1 )(g1,s g1 )
K rs =

l
l

Rr =

(5.64)

Note the ratio / l which is identified to be the force density.

k
G1

i
Z, z

Xi

ui
xi

g1

Y, y
X, x
Fig. 5.15: 2-node cable finite element

5 - 27

MM2: Numerical Theory

Finite Element Method

5.11 Application: 3-node membrane element in 2D


The previous results will be applied for the special case of 2D plane stress.

We consider a 3-node element of the given geometry:


y,

c
3

b
2

d
1

x,

Fig. 5.16: 3-node element


Shape functions

N1 = 1 ; N 2 = ; N 3 =

Position vectors, ref. config

X1T = {0 0 0}; X 2T = {a

Nodal displacement vectors

u1T = {u1

v1

0}; u 2T = {u 2

Discretization parameters

b T = {u 1

v1

u2

Position vectors, act. config

x1T = X1T + u1T ; x 2T = X 2T + u2T ; x3T = X3T + u3T

x1T = {u1

v1 0};

x3T = {c + u 3

Coordinate fields

Xh =

v2

b + v3

0}

0}; u3T = {u 3

v2

x 2T = {a + u 2

d + v2

k =1

y h = v1 + (d v1 + v 2 ) + (b v1 + v 3 )

G 1T = {a d 0}; G T2 = {c b 0}
g1T = {a u1 + u 2

d v1 + v 2

g T2 = {c u1 + u 3

Element area

A = G1 G 2 =

Contravariant base vectors

G1 T =

1
2A

{b

b v1 + v 3
1
2

0}

0}

(ab cd ) = det J

c 0}; G 2 T =

5 - 28

0}

Yh = N k Yk = d + b

x h = u1 + (a u1 + u 2 ) + (c u1 + u 3 )

Covariant base vectors

v3

v3 }

u3

N k X k = a + c ;
k =1

d 0}; X 3T = {c b 0}

1
2A

{ d

a 0}

0}

MM2: Numerical Theory


Metric tensors

Finite Element Method

a 2 + d 2 ac + bd
; G ij =
G ij =
2
2
ac + bd b + c

1
4A2

b2 + c2
(ac + bd )

(ac + bd )
a 2 + d 2

Material directions
e3 normal to surface

e1T = {1 0 0}; e T2 = {0 1 0}; e 3T = {0 0 1}

Transformation matrix

G1 e1 G1 e 2
=
T=
G 2 e1 G 2 e 2

1
2A

b c

d a

Basis derivatives

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

g 2
g1
= 0 1 0 0 0 1
= 0 1 0 1 0 0 ;
b

b
0 0 0 0 0 0
0 0 0 0 0 0

Strain derivatives
linear part,

g1 ~11
= { a
=
b
b
g
~
G1 2 = 12 = { a
b
b
g
~
G 2 1 = 21 = { c
b
b
g
~
G 2 2 = 22 = { c
b
b

Transformation to material
directions

G1

d a d 0 0}
d 0 0 a d}
b c b 0 0}
b 0 0 c b}

,b = ~ ,b G G = ij ,b ei e j ;
, b1 = , u1 =

1
2A

ij ,b = Ti T ~ Tj

d b c a

; , b 2 = , v1 =
0
0

1
2A

0
0

;
d b c a

etc.
B-matrix, rearranged strain
derivatives

Constitutive matrix

11 ,b

B = 22 ,b =

12 ,b + 21 ,b

0
b
0 d 0
d b

0
c
a
0
c
0
a
2A

c a d b c b
a d

C1111 C1122

C = C 2211 C 2222

0
0

E
0 =
1 2
1212
C

5 - 29

1
0

0
1
1

0 0
2

MM2: Numerical Theory

Finite Element Method

Elastic stiffness matrix:


1 1

Ke = t

B C B det J d d
T

= 0 = 0

special case c = d = 0:
2b 2 + a 2 (1 )

2
2(1 )
ab

2(1 )

2
b
Et
(1 2 )
Ke =
ab
2ab

2(1 + )

a2

2(1 + )

ab

(1 2 )

2a 2 +b 2 (1 )
2(1 2 )
ab

(1 2 )
b2

2(1 + )
ab

2(1 + )
a2

(1 2 )

a2
(1 2 )

sym.
b2
(1 2 )
0
0
ab
(1 2 )

b2
2(1 + )
ab
2(1 + )

a2
2(1 + )

(g1 G1 ) g1 = 11 = {u1 u 2

Strain derivatives
non-linear part,
initial displ.

v1 v 2 (u1 u 2 ) (v1 v 2 ) 0 0}
b
b
~
(g1 G1 ) g 2 = 12 = {u1 u 2 v1 v 2 0 0 (u1 u 2 ) (v1 v 2 )}
b
b
~
(g 2 G 2 ) g1 = 21 = {u1 u 3 v1 v3 (u1 u 3 ) (v1 v3 ) 0 0}
b
b
~
(g 2 G 2 ) g 2 = 22 = {u1 u 3 v1 v3 0 0 (u1 u 3 ) (v1 v3 )}
b
b

Transformation
and arrangement
in Bu-Matrix,
c=d=0

b 2 u12
b 2 v12
b 2 u12
1
a 2 u13
a 2 v13
0
Bu =
2
4A
ab (u12 + u13 ) ab (v12 + v13 ) ab u13

where u ij = u i u j ; v ij = v i v j

Initial displacement matrix:

(B C B
1 1

Ku = t

b 2 v12

a 2 u13

ab v13

ab u12

+ B u C B + B u C B u det J d d

= 0 = 0

5 - 30

a 2 v13

ab v12
0

MM2: Numerical Theory

Finite Element Method


1 0 1 0

0 1 0 1

1 0 1 0
B g11 = g1T ,b g1 ,b =
0 1 0 1

0 0 0 0

0 0 0 0

Strain derivatives
non-linear part,
initial displ.

1 0

0 1

1 0
= g1T ,b g 2 ,b =
0 1

0 0

0 0

B g12

0 0

0 0

0 0

0 0

0 0

0 0

0 0 1
0 0

0 0

0 0

0 0

0 0

1 0 1 0

0 1 0 1

0 0 0 0
B g 21 = g T2 ,b g1 ,b =
0 0 0 0

1 0 1 0

0 1 0 1

1 0

0 1

0 0
= g T2 ,b g 2 ,b =
0 0

1 0

0 1

B g 22

Geometrical
stiffness matrix

0 0

0 0

0 0

0 0

0 0

0 0

0 0 1
0 0

0 0

0 0

0 0

0 0

1 1

Kg = t

S:B

det J d d

= 0 = 0

1 1

=t

(S

11

B g11 + S12 B g12 + S21 B g 21 + S22 B g 22 )det J d d

= 0 = 0

S are contravariant components of the stress tensor w.r.t. to covariant reference base vectors G. For the form finding of minimal
surface it is: S = S I = S G G G

5 - 31

MM2: Numerical Theory

Finite Element Method

5.12 Application: 3-node membrane element in 3D


As a further example, the elastic stiffness matrix of a membrane element in 3D space is developed analogously. The approach is general and can be applied in the identical manner to
any kind of element. For distorted geometry numerical integration must be applied.

1
2

Fig. 5.17: 3-node element in 3D space


Shape functions

N1 = 1 ; N 2 = ; N 3 =

Position vectors, ref. config

X1T = {0 0 0}; X 2T = {4 1 1}; X 3T = {1 4 2}

Nodal displacement vectors

u1T = {u 1

v1

w 1 }; u2T = {u 2

Discretization parameters

b T = {u1

v1

w1

Coordinate fields

X h = 4 + ;

Covariant base vectors

G1T = {4 1 1}; G T2 = {1 4 2}

Element area

A = G 1 G 2 = 16.67 = det J

Contravariant base vectors

G1 T = {0.2662 - 0.0683 0.0036}

u2

v2

Yh = + 4 ;

w 2 }; u3T = {u 3

v2
w2

u3

Zh = + 2

G 2 T = { 0.0791 0.2230 0.0935}


Material directions:
e3 normal to surface
e1 aligned with G1 (arbitrary
choice for isotropic material)

e1T = {0.943 0.236 0.236}

Transformation matrix

0.2357 0.1414
T=

0
0.2544

e T2 = { 0.120 0.420 0.899}


e 3T = { 0.311 0.876 0.367}

5 - 32

v3

w 3}

v3

w 3}

MM2: Numerical Theory

Finite Element Method

Basis derivatives
g1
b

g 2
b

1 0 0

= 0 1 0

0 0 1
1 0 0

= 0 1 0

0 0 1

1 0 0 0 0 0

0 1 0 0 0 0

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

0 0 0 0 1 0

0 0 0 0 0 1

B-matrix, rearranged strain derivatives


.0
.0
.0
-.222 -.055 -.055 .222 .055 .055

B = .035 -.099 -.041 .043 -.123 -.051 -.079 .223 .093

-.033 -.233 -.113 -.206 .173 .053 .239 .059 .059


Constitutive matrix

0
104.16 20.83

C = 20.83 104.16
0 ;

0
0
41.67

E = 10 000
= 0.2
t = 0.01

Elastic stiffness matrix:


41.63 , 13.88 , 12.03 , -39.49 , -11.39 , -10.58 , -2.14 , -2.49 , -1.45

13.88 , 32.02 , 16.79 , -2.02 , -5.81 , -2.98 , -11.86 , -26.21 , -13.81

12.03 , 16.79 , 9.44 , -6.21 , -4.23 , -2.80 , -5.82 , -12.56 , -6.64

-39.49 , -2.02 , -6.21 , 62.78 , -10.80 , 3.33 , -23.29 , 12.82 , 2.88

Ke := -11.39 , -5.81 , -4.23 , -10.80 , 24.05 , 9.78 , 22.19 , -18.23 , -5.55

-10.58 , -2.98 , -2.80 , 3.33 , 9.78 , 5.01 , 7.25 , -6.80 , -2.21

-2.14 , -11.86 , -5.82 , -23.29 , 22.19 , 7.25 , 25.43 , -10.33 , -1.43

-2.49 , -26.21 , -12.56 , 12.82 , -18.23 , -6.80 , -10.33 , 44.44 , 19.36

-1.45 , -13.81 , -6.64 , 2.88 , -5.55 , -2.21 , -1.43 , 19.36 , 8.85

5 - 33

MM2: Numerical Theory

5.13

Finite Element Method

Solution of non-linear finite element equations

5.13.1 Newton-Raphson method


Our goal is to solve the non-linear equations (5.45) and (5.46) to determine the state of displacements with regard to the applied state of loading and/or pre-stress. As we had realized
the equations are non-linear in terms of the discretization parameters b which, typically, are
nodal displacements of the chosen finite element discretization. The formal appearance of the
governing equations can be generalized as:

r (b, ) = 0

(5.65)

where is a continuation, time or load factor which controls the progress of simulation. For
load control it is assumed that the actual loading state is a linear multiple of a reference state
of loading q 0 , 0 , S , respectively:
q = q0
load factor
= 0 (Cauchy)
load factor
S = S0 (2 nd Piola Kirchhoff )
~
homotopy mapping ( URS) f (b ) = f (b ) + (1 )f (b )
load factor

(5.66)

The problem is to find the root of a non-linear function. The basic idea of the NewtonRaphson method is to replace the original non-linear problem by a sequence of linear problems which easily can be solved. Their solutions converge to the solution of the original problem. That is illustrated in Fig. (5.18) for a 1D problem.
r
r

(0)

b*

(k)

(1)

(0)

(1)

Fig. 5.18: 1D Illustration of Newton-Raphson method

5 - 34

MM2: Numerical Theory

Finite Element Method

Given is a non-linear function r(b). We want to determine the root b* where r(b*) = 0. We
start at a first trial solution b(0) and replace the original function r(b) by a linear approximation
r(0):

r ( 0) = r (b ( 0) ) +

dr
(b b(0) ) = r 0 + r0 (b b(0) )
d b b( 0)

(5.67)

It is determined by a truncated Taylor series expansion where the function value and its first
derivative at the trial solution b(0) are used.
The next trial solution b(1) is defined as the root of the actual approximation r(0) = 0:
0 = r 0 + r0 (b (1) b ( 0 ) ) = r 0 + r0 b (1)
b
or

(1)

=b

( 0)

b (1) =

r0
0
r

r0
;
r 0

b (1) = b ( 0 ) + b (1)

(5.68)

generally, for iteration step (k ) :


b ( k +1) =

rk
;
rk

b ( k +1) = b ( k ) + b ( k +1)

The procedure is repeated until the solution converges to b*, the root of the original problem.
The extension to more dimensional problems is straight forward:

( )

Approximation r(k) in step (k):

r ( k ) = r k + r k, b b b ( k ) ;

Root of approximation r(k):

0 = r k + r k,b (b ( k +1) b ( k ) ) = r k + r k,b b ( k +1)

Next trial solution:

b ( k +1) = r k,b

[ ]

rk ;

r k = r b ( k ) ; r k, b =

r
b b ( k )

b ( k +1) = b ( k ) + b ( k +1)

In the case of non-linear structural analysis or form finding we can identify


rk
k

r ,b

to be the negative unbalanced force vector R


to be the stiffness matrix K

They are evaluated for the current state b(k) of discretization parameters. The unbalanced force
vector R is the difference of external and internal forces at load state , e.g. for structural
analysis:
R = q 0
A

u
E
det F dA t S 0 + E : C :
dA
b
b

5 - 35

(5.69)

MM2: Numerical Theory

Finite Element Method

Principal 1D example:
Determine the root of:

k
0
1
2
3
4
5
6

1
1
r = 1 b + b 2 + b 3 ;
3
3

b
3,000000
1,909091
1,302094
1,050652
1,001817
1,000002
1,000000

1
r = + 2b + b 2
3

Iteration history
r
r'
16,000000 14,666667
4,327573
7,129477
0,997294
3,966302
0,140246
2,871840
0,004852
2,673938
0,000007
2,666677
0,000000
2,666667

b
-1,090909
-0,606997
-0,251442
-0,048835
-0,001814
-0,000002
0,000000

Fig. 5.19: Visualization of Newton-Raphson method


Note the convergence rate. It is called quadratic if the number of zero digits in b increases
by factor 2 as can be seen from iteration step 3 on. For more dimensional problems the norm
of the difference vector is used to check the convergence rate. The Newton-Raphson method
should converge quadratically near the solution. If not, something is wrong with the implementation of the algorithm, or, the problem exhibits special problems (e.g. singularity) near
the solution. The latter may be the case in form finding if the non-linear equations are linearized without care. That was the reason that frequently Newton-Raphson methods were
thought to be not suited for form finding. However, methods like Updated Reference Strategy
(URS) use a time stepping scheme and converge quadratically in each time step because the
initially singular problem is properly stabilized.

5 - 36

MM2: Numerical Theory

Finite Element Method

5.13.2 Time stepping, path following, numerical continuation


Consider a load controlled experiment. A structure is loaded by subsequent load steps of increasing magnitude. The load increment is the difference between two load steps (k+1) =
(k+1) - (k). If the structure behaves non-linear (e.g. large deformations, material softening) the
relation between load factor and a representative displacement value is also non-linear,
Fig. 5.20. For a given load state (k) the related displacements u(k) have to be found by iteratively solving for equilibrium (Newton-Raphson method). As the pairs ((k), u(k)) define points
of equilibrium on the load-displacement-curve the procedure is also called path following
method. Several extensions are available where instead of a displacement component (displacement control) or combinations of and displacement (arc length control) are controlled.
Path following methods like these define the state of the art of non-linear structural analysis
and form finding.

Newton iteration
in load step 4

zoom

(4)

(3)

(2)

(1)

u
(1)

(4)

Fig. 5.20: Load-displacement diagram, load steps


A more general interpretation of path following is to introduce a curve or arc length parameter t. As the procedure of loading can be understood as a sequence of load steps the curve parameter t is also called pseudo time (no dynamic analysis) as it defines the time which is
needed to follow the load-displacement-curve.

(3)

t
(2)

t
(1)

t
t

Fig. 5.21: Pseudo time, time steps

5 - 37

MM2: Numerical Theory

Finite Element Method

Finally, the procedure of path following is also called numerical continuation. That is the
most general interpretation used in mathematics. If the idea of continuation is applied to trace
~
the path between two distinct states or interpretations of a certain problem, e.g. f and f ,

~
f (b ) = f (b ) + (1 )f (b )

(5.70)

the method is known as homotopy method. Then the continuation or homotopy factor can be
interpreted besides pseudo time also as a blending factor. Most recently, numerical continuation became very important in optimization as interior point methods. The idea here is to
sneak into a converged solution by following a sequence of approximated sub-problems and
their solutions (central path). That technique appears to be superior to directly solve the original, typically heavily non-linear problem because the sub-problems are much simpler or free
from the deficits of the original problem. Analogously, that idea is applied to form finding by
the Updated Reference Strategy (URS) as already introduced earlier.

5 - 38

You might also like