Dissertation
Dissertation
Jens Ruckert
1 Introduction 1
3 Newtons method 19
3.1 The modified Newton algorithm . . . . . . . . . . . . . . . . . . . . . 19
3.2 Second linearization of the energy functional . . . . . . . . . . . . . . 20
i
Contents
7 Numerical examples 55
7.1 Plate deflection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
7.1.1 Approximation with FEM using BFS-elements . . . . . . . . . 56
7.1.2 Approximation with FEM using reduced HCT-elements . . . . 63
7.2 Bending-dominated deformation . . . . . . . . . . . . . . . . . . . . . 66
7.2.1 Approximation with FEM using BFS-elements . . . . . . . . . 66
7.2.1.1 1st example: Cylinder . . . . . . . . . . . . . . . . . 66
7.2.1.2 2nd example: Cylinder with further rotated edge
normals . . . . . . . . . . . . . . . . . . . . . . . . . 71
7.2.1.3 3rd example: Mobiusstrip . . . . . . . . . . . . . . . 74
7.2.1.4 4th example: Plate with twisted edge . . . . . . . . . 77
7.2.2 Approximation with FEM using reduced HCT-elements . . . . 80
7.2.2.1 1st example: Partly divided annular octagonal plate 80
7.2.2.2 2nd example: Divided annulus with rotated edge
normals . . . . . . . . . . . . . . . . . . . . . . . . . 84
Bibliography 91
Notation 97
Theses 99
List of Figures 101
ii
Acknowledgement
Financial support
The research I present in this thesis was financially supported by the German Federal
Ministry for Education and Research (BMBF) within the joint research project
GeoMec (grant 05M10OCC). Furthermore I am grateful for the employment at the
Department of Mathematics at Chemnitz University of Technology. Thus I had the
ability to held tutorials and to support lectures.
Personal thanks
First and foremost I like to express my deep gratitude to my supervisor Prof. Dr.
Arnd Meyer. He aroused my interest first for the FEM and later for plate deform-
ations and he was always willing to give me advice and help. He offered me the
opportunity to further my personal development and expertise. In addition, due
to him, I was able to participate in different scientific conferences in Germany and
abroad.
During my years at the university he was the most important person. Without
him I would not have been able to develop this dissertation. Thank you, very much.
Furthermore I am much obliged to Prof. Dr. Bernd Luderer and to Dipl.-Math.
Martin Stocker for their efforts and their time spent in the revision of this thesis
and for their advice.
Although they are not named separately, I thank all my colleagues at the Depart-
ment of Mathematics for the inspiring discussions and explanations, by what I got
an better overview of the mechanical background.
Moreover I owe a debt of gratitude to the members of staff of Fraunhofer ITWM
in Kaiserslautern, who were partners in the mentioned BMBF project GeoMec.
Particularly, I am indebted to Dr. Joachim Linn, whose suggestions and advice
broadened my prospectives concerning my research.
Last but not least, I would express my sincerest gratitude to my whole family
and my friends, who all gave me the support which I needed to author this work.
Especially, I thank my parents Ines and Willy R uckert, which are the best parents I
could imagine. Without them, I would not be the man, who I am, today. In addition
I want to mention my parents-in-law, who gave me the opportunity to shorten my
duty stroke enormously in the last months. Finally, my special thanks I extend to
my wife Nadja, for her patience, for her continuing support and not at least for her
great, infinite and unconditional love. At the end, my warmest words
Acknowledgement
are dedicated thankfully to my little daughter Arja, who brings unlimited pleasure
and happiness into my life.
iv
1 Introduction
In this thesis we present our research as a result of the joint project GeoMec (grant
05M10OCC) supported by the German Federal Ministry of Education and Research
(BMBF). In this project, large deformations of plates and shells are investigated from
different points of views. In that way, one project partner was examining a mathe-
matical theory, where large deformations can be calculated very fast, possibly with
low accuracy. Such methods are used, for example, in the animation of movements,
which becomes an integral part in modern animation movies or in the computer
game industry. Here, the main focus is on the simulation of movements in a way
that it looks good but not necessarily expressing the full mechanical truth.
By contrast to that we pursue the goal of mechanical accuracy as best as pos-
sible. For this approach much more aspects have to be taken into account. So, the
deformation is based on the minimization of the deformation energy. This espe-
cially concerns to some industrial project partners in automotive engineering. Here,
for instance, the bending of car headliners during the manufacturing process is of
interest.
For elasticity and deformations of plates or shells a very large number of literature
exists. The roots of linear plate theory go back far into the past. First approaches
have been made in the 17th century, but with some shortcomings (see [3]). Then
in 1850 Kirchhoff has published his linear plate theory and is considered as the
founder of modern plate theory. Quite nearly a century later, some extensions have
been made by Reissner (1944/45) and Mindlin (1951) to include transverse shear.
In the late seventies and eighties of the last century first non-linear plate theories
were established for example by Ciarlet (see [9]). For that only linear material laws
and so only small deformations were considered. A first approach for the calcula-
tion of large deformation of plates was the von Karman plate. Due to the rapid
development in the computer industry for the last 20 years, it is possible to com-
pute more complex problems considering non-linear theories and non-linear material
laws. Several approaches are found in [7]. In [14] the differential geometry is intro-
duced in a similar way as here, following by a restriction to linearized strain. Here,
the well known split of the total energy into a change of metric part and a change
of curvature part is used. Furthermore, the resulting theory is only adopted to a
cylindrical shell. Another approach one can find in [2]. Here, instead of the normal
of the shell mid surface an arbitrary direction vector is considered first and a non-
linear system of equations representing the equilibrium of forces is established. By
the identification of the direction vector as the normal of the deformed mid surface
the Kirchhoff hypothesis is assumed again. In the end the theory is constrained to
linearized strain.
1
1 Introduction
For our approach the Kirchhoff hypothesis is considered, too. By contrast with
the theories in the cited articles, we do not linearize the strain tensor and maintain
these non-linearities for its derivatives. Hence, the usual Kirchhoff assumption is
the one and only restriction. No other simplifications are done.
Therefore, at the beginning we consider the well-established 3D-theory for large
deformations. We discuss the inherent differential geometry and define the basic
principles like the deformation gradient, the right Cauchy-Green deformation tensor
or the Lagrangian strain tensor.
The equilibrium of forces corresponds to a minimum of an appropriate energy
functional, which is considered here. Its first derivative leads to the weak formulation
of the non-linear boundary value problem. For its solution we have to apply some
linearizations such as Newtons method.
In the third chapter Newtons method is explained and the incremental Newton
algorithm is given for the approximation of a resulting deformation. In this algorithm
the second derivative of the energy functional plays an important role. For the full
3D case it can be calculated comparatively easy.
We discuss the differential geometry for a shell and its simplification to a plate
in the fourth chapter. Generally, in the shell theory the displacement of the mid
surface is considered and then amplified to the whole plate. In our research the
Kirchhoff assumption is used for that. As a result a strain tensor of lower rank
(rank 2) occurs (mapping the tangential space into itself).
Furthermore, the emerging displacement vector depends on the displacement vec-
tor of the mid surface. Therefore, in Chapter 5 we establish the resulting energy
functional, which only depends on this mid surface displacement vector. The bound-
ary conditions are considered for hard and soft clamped shells.
For examining the weak formulation of the plate deformation problem the first
derivative of the energy functional is determined, which is now more complicated
than in the full 3D-theory.
For the resulting energy functional the thickness dependence is not eliminable as
usually in using linear strain, but the unknown vector function U only depends on
( 1 , 2 ), thus not on the thickness parameter . Therefore, we can use the 2D-FEM
for its approximation on the mid surface and we have to integrate numerically over
thickness direction . So, in Chapter 6 we deduce the second derivative of the energy
functional and take a closer examination of the difficulties in its calculation. With
this second derivative the Newton algorithm is established, which is integrated into
the F.E.-solution process. The method and both kinds of considered C1 -elements,
the BFS- and the reduced HCT-element, are explained. A very short excursion in
the field of efficient solvers follows, before we go into details of the implementation.
A short discussion about the use of Newtons method in the approximation process
brings Chapter 6 to an end.
After that we present two deflection examples, one using the BFS-elements and
one the reduced HCT-elements. Subsequently we take a closer look to bending
problems, which was the main focus of this work. For that, in the seventh chapter
four examples of bending-dominated large plate deformations are presented, using
2
1 Introduction
3
2 The deformation problem in the
three-dimensional space
In the following we consider vectors or vector fields written as bold capital or lower
case letters, respectively. Furthermore a vector is also called a first-order tensor.
Then we understand second-order tensors as linear maps of first-order tensors onto
first-order tensors, again. For the beginning a pair of first-order tensors such as
A1 A2 is understood as a second-order tensor. In general a second-order tensor is
any linear combination of such pairs.
Below, some linear operations for the tensor calculus are defined:
(A1 A2 ) U = A1 (A2 U )
U (A1 A2 ) = A2 (A1 U ).
tr(A1 A2 ) = A1 A2 .
In general, the definitions in (1) to (5) are true for every second-order tensor, because
all five definitions above are linear operations.
5
2 The deformation problem in the three-dimensional space
In the same way we use fourth-order tensors mapping second-order tensors onto
second-order tensors via the double contraction (5).
Note that in general a fourth-order tensor is a linear combination of such four-
tuple of vectors or of pairs of second-order tensors.
Throughout this paper capital letters are used for second-order tensors, displayed
in a flowing font. Vectors are written in bold and matrices in underlined letters.
Due to the fact that we consider large deformations, we have to distinguish between
the undeformed and the deformed configuration of the considered body using capital
and lower case letters, respectively. This is mainly used for the basis vectors.
0 = {X() : P R3 }, (2.1)
where X denotes the position of a material point in the undeformed body under
consideration. Herein, = ( 1 , 2 , 3 ) is a given, possibly curvilinear coordinate
system. 1 , 2 and 3 are called the coordinates of the point X.
Note that
X = ( 1 , 2 , 3 ) (2.2)
Gi = i X, i = 1, . . . , 3, (2.3)
form the co-variant tensor basis.
The elements of the matrix
Gi Gj = ij , i, j = 1, . . . , 3.
6
2.1 General differential geometry of deformation in the three-dimensional space
From now on we use the Einstein summation convention. For this, the indices
in this chapter run from 1 to 3. Later on, when the shell theory is considered, the
indices will run from 1 to 2, only.
So, the gradient operator is defined by
Grad = Gi i . (2.5)
If we apply the gradient operator (2.5) to a given function (), we obtain a first-
order tensor Grad = Gi i . In the same way, we have a second-order tensor
i
Grad U () = G i U = Gi U ,i , if we apply (2.5) to the vector function U () 0 .
Here and in the following we use the abbreviation U ,i for the partial derivatives
U , i = 1, . . . , 3.
i
Let a deformation be described by the isomorphism
: 0 7 , x = (X). (2.6)
x = X + U : 0 7 . (2.7)
x = ( 1 , 2 , 3 ) (2.8)
to be a unique mapping between the material point of the deformed body and its
coordinates. Therefore, with the parametrization (2.7) of , we can denote a co-
variant and a contra-variant tensor basis again:
Let
g i = i x, i = 1, . . . , 3,
be the basis vectors belonging to the co-variant tensor basis in . Then, there exists
a contra-variant tensor basis g j , j = 1, . . . , 3, such that g i g j = ij , i, j = 1, . . . , 3.
The co-variant tensor basis of the deformed domain is connected to the co-
variant tensor basis of the initial domain 0 via the displacement vector field U :
g i = i x = i (X + U ) = Gi + i U = Gi + U ,i , (2.9)
7
2 The deformation problem in the three-dimensional space
and the gradient operator in the deformed domain is defined analogously to (2.5):
grad = g i i .
Obviously a Taylor expansion of the isomorphism (2.6) is supposed to be such
that
(X + V ) = (X) + F V + O(kV k2 )
Hence,
F Gk = Gk + U ,i Gi Gk
= Gk + U ,k = g k (2.10)
>
= Gk F .
[g 1 , g 2 , g 3 ] = [F G1 , F G2 , F G3 ] (2.13)
= det F[G1 , G2 , G3 ].
8
2.1 General differential geometry of deformation in the three-dimensional space
[G1 , G2 , G3 ] 6= 0
and
[g 1 , g 2 , g 3 ] 6= 0.
d0 = [G1 , G2 , G3 ] d 1 d 2 d 3
and
d = [g 1 , g 2 , g 3 ] d 1 d 2 d 3 .
Obviously, with (2.13) the volume element of the deformed domain, d, can be re-
presented by the product of the determinant of F with the volume element of the
undeformed domain, d0 .
(1) Hence
3 3
V () L2 () V () L2 (0 ) , (2.15)
because
Z Z
V V d = V V [g 1 , g 2 , g 3 ] d 1 d 2 d 3
Z
= V V det F [G1 , G2 , G3 ] d 1 d 2 d 3 .
0
kV k2L2 () kV k2L2 (0 )
and both function spaces have the same set of vector functions V ().
9
2 The deformation problem in the three-dimensional space
Due to equation (2.14), there exist two constants 0 < < such that in sense
of positive definiteness
I F > F = C I
and therefore
1 1
I F 1 F > = C 1 I
are true.
Hence, together with (2.15) we obtain
kV k2H1 () kV k2H1 (0 )
and both function spaces have the same set of vector functions V () again.
10
2.2 Equilibrium of forces
11
2 The deformation problem in the three-dimensional space
2
Because of the symmetry of T and from equation (2.18) we have
Z 2
T : (I + Grad U ) Grad V > d0
0
Z 2
= T : [(I + Grad U ) Grad V > ]> d0 (2.19)
0
Z 2
= T : E 0 (U ; V ) d0 V () H.
0
Therefore,
with
Due to the mass conservation in 0 and , the equation (det F) = 0 holds, where
0 is the density in 0 .
For the second integral of the right-hand side we have to consider the boundary
conditions.
12
2.2 Equilibrium of forces
Let
0 = 0,N D
and
= N D
with U (X) = 0 X D . Hence x and X coincide on D (fixed boundary). In
that way, we define
3
1
H = V H (0 ) : V = 0 on D .
Then
Z Z
n V dS = n V dS V () H.
N
By fixing one of the coordinates, for instance, 3 = c, we obtain (at least a part of)
the boundary N or 0,N , respectively, with
n o
N = x( 1 , 2 , c) : ( 1 , 2 ) P N .
Here, the vectors g 1 and g 2 of the appropriate co-variant tensor basis act as tangen-
tial vectors of this surface and with
dS = kg 1 g 2 k d 1 d 2
we obtain the surface element. In that way, the normal vector of the deformed mid
surface is given by
g1 g2
n= .
kg 1 g 2 k
Therefore
Z Z
[g 1 , g 2 , V ]
n V dS = kg 1 g 2 k d 1 d 2
kg 1 g 2 k
N PN
Z
= [g 1 , g 2 , V ] d 1 d 2
PN
Z
1 1
= [F G1 , F G2 , F T V ] d 1 d 2 (2.22)
det F
PN
Z 1
= [G1 , G2 , T V ] d 1 d 2
PN
Z 1
= N T V dS 0
0,N
13
2 The deformation problem in the three-dimensional space
1
where ~g = N T are the given traction on 0,N .
We are still not able to solve this non-linear equation system in (2.23), due to the
2
fact that, until yet, the Piola-Kirchhoff stress tensor T is unknown. In literature
these equations are called not to be closed. By consideration of the material laws
this closure is done in the next section.
Herein (C(U )) denotes the energy density depending on the invariants of C. These
invariants we get from the characteristic polynomial
det(C I) = 3 + IC 2 IIC + IIIC .
Here IC , IIC , IIIC are called the principal invariants of C and if 1 , 2 , 3 are the
three eigenvalues of the tensor C, these invariants can be worked out as
IC = tr(C) = 1 + 2 + 3 ,
1h i
IIC = (trC)2 tr(C 2 ) = 1 2 + 2 3 + 3 1 ,
2
IIIC = det C = 1 2 3 .
For convenience, we introduce the functions
1
ak = tr(C k ) (k = 1, . . . , 3) (2.25)
k
14
2.3 Material laws
and use these ak instead of the principal invariants of C to express material laws,
i.e.
IC = a1 ,
1
IIC = (a1 )2 a2 , (2.26)
2
IIIC = (a31 6a1 a2 + 6a3 )/6.
0 (U ; V ) = 0 V H. (2.27)
(U + V ) = (U ) + 0 (U ; V ) + O(kV k2 ) V H. (2.28)
Furthermore,
(C(U + V )) = C(U ) + C 0 (U ; V ) + O(kCk2 ) V H. (2.29)
C = 2E + I, (2.30)
C 0 (U ; V ) = 2E 0 (U ; V ) V H.
15
2 The deformation problem in the three-dimensional space
and, therefore, it can be calculated for X 0 (which is needed for the calculation
at the integration points of a FE-discretization, later on).
The materials with linear elastic behaviour represent a special case. For these
materials we assume that there exists a linear map between the Right-Cauchy-Green
deformation tensor and the second Piola-Kirchhoff stress tensor. So the density
function can be calculated separately by (2.33). From Hooks law we have
2
T = 2E + tr(E)I
with the Lame coefficients and . Using (2.30) we get
3
2 1 3 ! i1
T = C 1 + C 0 ( + a1 ) = 2
X
C .
2 2 i=1 ai
16
2.4 The weak formulation
and
Z Z
f (V ) = 0 p V d0 + ~g V dS 0 V () H (2.35)
0 0,N
a(U ; V ) = f (V ) V H. (2.36)
0 (U ; V ) = a(U ; V ) f (V ) V H (2.37)
holds.
17
3 Newtons method
The purpose of this chapter is to solve the non-linear equation system (2.36) using
Newtons method. Therefore, the second derivative of the energy functional and
the first derivative of the second Piola Kirchhoff stress tensor are needed. Due to
the fact that Newtons method converges in a small ball around the unknown exact
solution only, it is essential to choose a sufficiently well starting approximation. For
small deformations U = 0 as the initial approximation was adequate, but in the case
of large deformations the displacement vector differs very much from zero. Hence,
for Newtons method the well known incrementation of the forces f (U ) is used. In
that way, we get a series of Newton iterations with an incremental increase of the
outer forces f (U ). The proof of convergence for an incremental solving of equation
systems and some more information about are given in, for example, [11].
D(U ) = 0
D0 (U ; U ) = D(U )
19
3 Newtons method
is fulfilled.
Applying Newtons method to the problem (2.36) means solving (2.37). Then, in
every Newton step we have to calculate U from
00 (U ; U , V ) = 0 (U ; V ) V H.
0 (U + U , V ) = 0 (U ; V ) + 00 (U ; U , V ) + O(kU k2 ).
(1) Start: U = 0, t := t
(3) update U := U + U
With the algorithm above, a series of Newton methods is applied, due to the
incrementation of the force vector via t.
2
For computing the derivative T 0 we have to linearize the second Piola-Kirchhoff
stress tensor itself:
2 2
T (C + C) = T (C) + C : C + O(kCk2 ) (3.1)
20
3.2 Second linearization of the energy functional
Furthermore, we consider
2 2
T C(U + U ) = T C(U ) + 2E 0 (U ; U ) + O(kU k2 ) .
Obviously, we achieve
2
T 0 = C(U ) : E 0 (U ; U ).
Thereby, the fourth-order tensor C arises from the equations (2.33, 3.1) as
3 X
3
C(U ) = 4
X 2 hC i1 C j1 i + 4 J + 4 C
b (3.2)
ai aj a2 a3
i=1 j=1
with the fourth-order unity tensor J, mapping any second order tensor to itself, and
the fourth-order tensor C
b with the property
b :X =CX +X C
C
for all symmetric second-order tensors X . C b emerges from the expansion of the
square of the right Cauchy Green deformation tensor C 2 (U ).
The second derivative of the Lagrangian strain tensor is easily obtained by the
linearization of its first derivative (2.20):
To sum up, due to the symmetry of the fourth-order tensor C and of the second-
order tensor E 0 (U ; U ), the second derivative of the energy functional can be written
as
Z Z 2
0 0
00
(U ; U , V ) = E (U ; U ) : C : E (U ; V ) d0 + T : E 00 (U ; U , V ) d0 .(3.4)
0 0
In the literature of mechanics the tensor C(U ) is called the material tangent. With
(3.4) now the Algorithm 3.1 is complete.
21
4 Differential geometry of shells
In the geometrical characterization of shells we go along with the majority of mech-
anical literature. Therein, an initial shell 0 is characterized as a static, isotropic
three-dimensional object. One regards a mid surface m 3
0 R of the undeformed
shell and a thickness h diam(m 0 ). Now the deformation of the shell is approx-
imated by a deformation of the mid surface.
This chapter begins with the differential geometry of shells in general. Later on,
we consider the initial configuration as a plane shell, called plate. At the end of this
chapter the Kirchhoff assumption is introduced and than combined with the theory
for the plane shell.
In what follows all indices now run between 1 and 2. The content of this chapter
refers to the explanations in [21] and [24].
Ai = i Y , i = 1, 2,
and the surface normal vector is described by
A1 A2
A3 = A3 = . (4.2)
|A1 A2 |
The two tangential vectors together with the surface normal vector establish a co-
variant tensor basis in R3 .
The first fundamental forms A and the second fundamental forms B are written
as (2 2)-matrices:
A = (Aij )2i,j=1 , Aij = Ai Aj ,
23
4 Differential geometry of shells
= A(I h A1 B)2
= A I 2h A1 B + h2 2 (A1 B)(A1 B) (4.4)
= A 2h B + h2 2 BA1 B
= (A h B)A1 (A h B).
24
4.3 The plate as an exception of a shell
Y ( 1 , 2 ) = e1 1 + e2 2 , (4.5)
A3 = e 3
for the normal vector of the undeformed mid surface independent of ( 1 , 2 ). To-
gether with the definition of the co-variant tensor basis (4.3) of the initial shell we
get the co-variant tensor basis with the normal vector for the initial plate as
1
Gi = Gi = ei , i = 1, 2, G3 = he3 , G3 = e3 . (4.6)
h
Because of the independence of the normal vector from the coordinates ( 1 , 2 ),
the second fundamental forms at the mid surface of the initial plate vanish:
B = O.
25
4 Differential geometry of shells
26
4.4 Kirchhoff assumption and the deformed shell
U KH ( 1 , 2 , ) = U + h (a3 (U ) e3 ), (4.14)
Fb = g i Gi + a3 A3
= (a1 + h a3,1 )e1 + (a2 + h a3,2 )e2 + a3 e3 .
Then, the coefficients of the right Cauchy-Green deformation tensor (2.11) are given
by
Cb = Fb Fb = cij ei ej + e3 e3 (4.15)
with
Cb = gij ei ej + e3 e3 (4.17)
h i
= (ai + ha3,i ) (aj + ha3,j ) ei ej + e3 e3
h i
= (ei + U ,i + ha3,i ) (ej + U ,j + ha3,j ) ei ej + e3 e3
h
= ij ei ej + e3 e3 + (U ,i + ha3,i ) ej + ei (U ,j + ha3,j )
i
+ (U ,i + ha3,i ) (U ,j + ha3,j ) ei ej
h
= I + (U ,i + ha3,i ) ej + ei (U ,j + ha3,j )
i
+ (U ,i + ha3,i ) (U ,j + ha3,j ) ei ej .
Thus, with (2.12) we can calculate the Lagrangian strain tensor of the plate as
Eb = ij ei ej (4.18)
with
27
4 Differential geometry of shells
In that,
GradS := Gi i , i = 1, 2
defines the surface gradient operator. Then, with the orthoprojector into the tan-
gential space A = ei ei the Lagrangian strain tensor is written as
2E(U
b ) = GradS (U + ha3 ) A + A GradS (U + ha3 )>
(4.19)
+ GradS (U + ha3 ) GradS (U + ha3 )> .
For using matrix syntax we define the matrices containing the coefficients of the
associated tensors in the chosen tensor basis.
In case of the right Cauchy-Green deformation tensor Cb given in (4.15) and (4.16),
the appropriate matrix is
c = (cij )2i,j=1
= [(ai + h a3,i ) (aj + h a3,j )]2i,j=1
= a + h [ai a3,j + a3,i aj ]2i,j=1 + (h )2 [a3,i a3,j ]2i,j=1 . (4.20)
y ,i a3 = ai a3 0, i = 1, 2.
Consequently,
(ai a3 ),j 0 i = 1, 2
is also true.
Hence, by using the product rule and the equation (4.11) we can deduce
28
4.4 Kirchhoff assumption and the deformed shell
a3,i aj = bij i = 1, 2.
Therefore
a3,i = bij aj
= bij aj (ak ak )
= bij (aj ak ) ak
= bij ajk ak .
Consequently, we achieve
Obviously
and with (4.20), (4.21) and (4.22) the matrix of the right-Cauchy-Green deformation
tensor can be calculated:
c = a 2h b + (h )2 ba1 b
= (a h b)a1 (a h b).
Analogously, the matrix with the coefficients of the strain tensor (4.18) is
e = (ij )2i,j=1
with
2e = (a h b)a1 (a h b) I. (4.23)
29
5 Shell energy and boundary
conditions
In this chapter we combine the three-dimensional theory for large deformations,
described in the sections 2.2 and 2.3, with the new displacement vector U KH from
(4.14) as a result out of the Kirchhoff hypothesis. Consequently, we achieve a new
a-form used in the variational formulation.
where H10 is the Sobolev function space with functions vanishing at the Dirichlet
boundary D .
As seen in (2.35), in the right-hand side of this new energy functional (5.1), for
the boundary forces f (U KH ) we have to integrate over 0 . Within, the integration
over the thickness parameter leads to zero for the second part in U KH (4.14).
Consequently, for the right-hand side in (5.1) we obtain
f (U KH ) = f (U ).
(U KH ) = (U + h (a3 (U ) e3 ))
Z
= (U + h (a3 (U ) e3 )) dV f (U )
0
Z (5.2)
= (U
b ) dV f (U )
0
=: (U
b ).
31
5 Shell energy and boundary conditions
Remark 5.2 With the displacement vector (4.14) from the Kirchhoff hypothesis and
remark 5.1 we have to consider the function space for the displacement vector U of
the mid surface and the arbitrary test functions V , respectively.
Therefore we define U H b where
3
b := { H2 (m )
H with appropriate boundary conditions} (5.3)
0
!
b0 (U ; V ) = 0.
(U
b + V ) = (U
b ) + b0 (U ; V ) + h.o.t.
2
Here, the second Piola-Kirchhoff stress tensor T (U KH ) is double the derivative of
(U KH ) w.r.t. C. Hence
(U KH )
b0 (U , V ) = V
U ( 1 , 2 )
" 3 #
KH
X ak : CKH U
= ak C U U V
k=1
2
= T (C) : E 0 U KH ; L(V ) V H
b
with
KH
L(V ) := UU V .
The symbol denotes the application of the second part of the chain rule to an
arbitrary vector function V in the way that L(V ) is a linear operator with respect
to V arising from the variation of U KH :
U KH (U + V ) = U KH (U ) + L(V ) + h.o.t. (5.4)
32
5.1 The resulting Kirchhoff deformation energy
Obviously,
L(V ) = V + h a03 (U ; V ) V H.
b
of vector functions being defined in the same way as the displacement vector (4.14),
resulting from the Kirchhoff hypothesis. Hence, with (4.18) the strain tensor E(U )
forall U K is of lower rank (rank = 2) without components of ei e3 or e3 e3 .
Now,
(5.4) iswritten formally in a correct way but it seems to result in a tensor
0 KH
E U , L(V ) , which is of rank greater than two. Apparently, the derivative L(V )
of the vector function U KH (U ) in direction of a vector V is no longer in K (5.5) and
lest in the considered non-linear set (4.14), over which the energy functional (5.1)
should be minimized.
Contrary, if we consider the first derivative of the energy functional (2.32) directly
and insist on the vector functions to be from K (5.5), we first need the linearization
of the strain tensor E(U
b ) (4.19):
2E(U
b + V ) = GradS U + V + ha3 (U + V ) A
>
+ A GradS U + V + ha3 (U + V )
>
+ GradS U + V + ha3 (U + V ) GradS U + V + ha3 (U + V )
h i
= GradS U + ha3 (U ) + GradS V + ha03 (U ; V ) + h.o.t. A
h i>
+ A GradS U + ha3 (U ) + GradS V + ha03 (U ; V ) + h.o.t.
h i
+ GradS U + ha3 (U ) + GradS V + ha03 (U ; V ) + h.o.t.
h i>
GradS U + ha3 (U ) + GradS V + ha03 (U ; V ) + h.o.t.
>
= 2E(U
b ) + GradS V + ha03 (U ; V ) A + A GradS V + ha03 (U ; V )
>
+ GradS U + ha3 (U ) GradS V + ha03 (U ; V )
>
+ GradS V + ha03 (U ; V ) GradS U + ha3 (U ) + h.o.t.
Hence
0
2Eb (U ; V ) = GradS V + ha03 (U ; V ) A
>
+ A GradS V + ha03 (U ; V )
> (5.6)
+ GradS U + ha3 (U ) GradS V + ha03 (U ; V )
>
+ GradS V + ha03 (U ; V ) GradS U + ha3 (U ) .
33
5 Shell energy and boundary conditions
This linearization of the strain tensor (5.6) is a second-order tensor of rank two with
a03 (U ; V ) being a differential operator, which can be retrieved from the variation of
a3 :
a3 (U + V ) = a3 (U ) + a03 (U ; V ) + h.o.t.
a1 (U + V ) a2 (U + V )
=
ka1 (U + V ) a2 (U + V )k
[a1 (U ) + V ,1 ] [a2 (U ) + V ,2 ]
=
k[a1 (U ) + V ,1 ] [a2 (U ) + V ,2 ]k
a1 a2 + V ,1 a2 + a1 V ,2 + h.o.t.
=
ka1 a2 + V ,1 a2 + a1 V ,2 + h.o.t.k
The abbreviations
b := a1 a2
(5.7)
:= V ,1 a2 + a1 V ,2
v
yield
b+v + h.o.t.
a3 (U + V ) = .
kb + v
+ h.o.t.k
Consequently
!
1 bb 1
a03 (U ; V )= I v
= (I a3 a3 ) v
(5.8)
kbk kbk2 ka1 a2 k
34
5.2 Boundary conditions
with (I a3 a3 ) being the orthoprojector onto the tangential space of the deformed
mid surface. In mechanical literature this projector is also called the metric tensor
of the deformed mid surface.
Concluding, we can write the first derivative of the new energy functional as
Z 2 Z Z
0
b0 (U ; V ) = T (C) : Eb (U ; V )d0 0 p V d0 + ~g V dS 0 . (5.9)
0 0 0,N
Until now, we have not considered any boundary conditions of the function space
b Therefore we take a closer look onto this space in the following paragraph.
H.
35
5 Shell energy and boundary conditions
Hence,
(3) (3)
U,1 = U,2 = 0
and therefore
(3)
U,12 = 0
are the resulting boundary conditions for the displacement vector.
Consistently, the displacement vector has to be chosen out of the appropriate
function space
3
2 (3) (3) (3)
H
b = H (0 ) : U,1 = U,2 = U,12 = 0, X D , j = 1, 2 .
In case of considering a soft clamped boundary part, one out of the two partial
(3)
derivatives U,i , i = 1, 2, is free.
V H.
b
In this way
Z 2
0
ab(U ; V ) = T (C) : Eb (U ; V ) dV, V H
b (5.10)
0
and
Z Z
f (V ) = 0 p V d0 + ~g V dS 0 V H.
b (5.11)
0 0,N
36
6 Newtons method and
implementation
6.1 Newton algorithm
In the Chapter 3 we have introduced Newtons method and we have defined the incre-
mental algorithm (3.1). Therein the displacement vector U of the three-dimensional
theory is used. We have replaced this displacement vector by the appropriate U KH in
(4.14), resulting from the Kirchhoff hypothesis (4.13). As a consequence, we get the
new energy functional (5.2) after recalculating the right Cauchy-Green deformation
tensor Cb from (4.15) and, therefore, the Lagrangian strain tensor Eb in (4.19). By
reason of this recalculation it takes much more effort to obtain the first linearization
0
Eb (U ; V ) ((5.6) in conjunction with (5.8)) with the vector V so, that U KH (U + V )
is from K, defined in (5.5).
In that way, we obtain the incremental Newton algorithm for this problem as
(1) Start: U = 0, t := t
(3) update U := U + U
Here
2 0 0 0
T 0 (U ; U ) : Eb (U ; V ) = Eb (U ; U ) : C(U ) : Eb (U ; V ) (6.2)
37
6 Newtons method and implementation
In that way
0
2Eb (U ; U , V ) = GradS ha003 (U ; U , V ) A
>
+ A GradS ha003 (U ; U , V )
>
+ GradS U + ha3 (U ) GradS ha003 (U ; U , V )
38
6.1 Newton algorithm
>
+ GradS ha003 (U ; U , V ) GradS U + ha3 (U )
>
+ GradS U + ha03 (U ; U ) GradS V + ha03 (U ; V )
>
+ GradS V + ha03 (U ; V ) GradS U + ha03 (U ; U ) .
Obviously, for the completion of the calculation of the second derivative of the
energy functional b00 (U ; U , V ) the second linearization of the normal vector of the
deformed mid surface a003 (U ; U , V ) has to be calculated.
In addition to the abbreviations (5.7) we define
g := U a + a U .
U ,1 2 1 ,2
Furthermore, the first derivative of the normal of the deformed mid surface is already
given in (5.8). Then, the variation of the orthoprojector in (5.8) yields
I a3 (U + U )a3 (U + U )
= (I a3 a3 ) a3 (U )a03 (U ; U ) a03 (U ; U )a3 (U ) + h.o.t.
Hence, its linearization is
a3 (U )a03 (U ; U ) a03 (U ; U )a3 (U )
1 g 1 (I a a ) U
= a3 (I a3 a3 ) U 3 3
ga
3
kbk kbk
1 h g g i (6.3)
= a3 U + U a3 (a3 Ug )a a (U
3 3
g a )a a
3 3 3
kbk
1 h i
= 2(a3 U
g )a a a U
3 3 3
g U
ga .
3
kbk
Therefore, the variation of the first derivative of the normal of the deformed mid
surface a03 (U ; V ) is obtained as
I a3 (U + U )a3 (U + U )
a03 (U +U ; V ) =
ka1 (U + U ) a2 (U + U )k
(V ,1 a2 (U + U ) + a1 (U + U ) V ,2 )
1
=
(I a3 a3 ) a3 (U )a03 (U ; U )
b + U
g + h.o.t.
a03 (U ; U )a3 (U ) + h.o.t. (ve + V ,1 U ,2 + U ,1 V ,2 )
!
1 g b
U
= 1 (I a3 a3 ) a3 (U )a03 (U ; U )
kbk kbk2
a03 (U ; U )a3 (U ) + h.o.t. (ve + V ,1 U ,2 + U ,1 V ,2 )
!
1 1
= (I a3 a3 ) ve + (I a3 a3 )
kbk kbk
1
(V ,1 U ,2 + U ,1 V ,2 ) a3 (U )a03 (U ; U )
kbk
39
6 Newtons method and implementation
g b
U
+ a03 (U ; U )a3 (U ) ve (I a3 a3 ) ve + h.o.t.
kbk3
= a03 (U ; V ) + a003 (U ; U , V ) + h.o.t.
In that way, the second linearization a003 (U ; U , V ) can be identified as
1
a003 (U ; U , V ) = (I a3 a3 ) (V ,1 U ,2 + U ,1 V ,2 )
kbk
1
a3 (U )a03 (U ; U ) + a03 (U ; U )a3 (U ) ve
kbk
g b
U
(I a3 a3 ) ve
kbk3
1
= (V ,1 U ,2 + U ,1 V ,2 )
kbk
1
+ a3 a3 (U ,2 V ,1 + V ,2 U ,1 )
kbk
1 h i
(6.3) + 2 2(a3 U )a3 a3 a3 U U a3 v
g g g e
kbk
1 g
2 (U a3 ) (I a3 a3 ) v
e
kbk
1
= (V ,1 U ,2 + U ,1 V ,2 )
kbk
1
+ [U ,2 , V ,1 , a3 ] + [V ,2 , U ,1 , a3 ] a3
kbk
2 1 g
+ 2 (a3 U )(a3 v)a3 (U v)a
g e 3
kbk2
e
kbk
1 1 g
2 (v a3 )U
g (U a3 )ve
kbk2
e
kbk
1
+ (a3 U
g )(a v)a
3 e 3
kbk2
1
= (V ,1 U ,2 + U ,1 V ,2 )
kbk
1 1 g
2 (v a3 )U
g (U a3 )ve
kbk2
e
kbk
(6.4)
1
+ [U ,2 , V ,1 , a3 ] + [V ,2 , U ,1 , a3 ]
kbk
3 1 g
+ (a3 U
g )(a v)
3 e ( U v)
e a3 .
kbk kbk
With the linear functional
1
1 (U ) = U a3
kbk2
40
6.2 Finite Element Method (FEM)
We have seen, that the calculation of the second derivative of the Lagrangian strain
00
tensor Eb (U ; U , V ) is much more labour-intensive as for the appropriate derivative
in the 3D theory (3.3).
m
[
0 = T, Ti Tj = Ti , Tj Th , i6=j.
T Th
In general, finite elements are triangles or quadrilaterals for R2 and for example
tetrahedrons or hexahedrons for R3 , respectively.
A finite element mesh is called a conform mesh, if every side of a finite element
Ti Th is either part of the boundary or a side of another finite element Tj Th .
We speak of isotropic finite elements T , if there exists a constant c such that
hT cT
41
6 Newtons method and implementation
with the element diameter hT and T the diameter of the inner element circle or of
the inner element sphere, respectively. A finite element mesh is called shape regular,
if it only consists of isotropic finite elements.
Now, with h = maxT Th hT we consider the biggest element diameter of the mesh
Th . Then, we call a mesh quasi uniform, if the constants c1 and c2 exist such that
c1 h hT c2 h.
In that way, for example locally refined meshes are not quasi uniform.
3N 3N 3N
Uh =
P
pi ui , V h =
P
pi vi , U h =
P
pi dui . (6.6)
i=1 i=1 i=1
Algorithm 6.2 (Discretized Newton solution process for one fixed increment t)
3N
(1) b = tf (pi ) a(U h , pi ) ,
i=1
" #3N
0 0 2 00
Eb (U Eb (U +T : Eb (U
R
(2) K(U h ) = 0 h ; pj ) :C: h ; pi ) h ; pj , pi ) d0
i,j=1
pj |T =: qk,T .
42
6.2 Finite Element Method (FEM)
With this form functions qk,T we can define the element stiffness matrix KT (U h )
and the element right hand side bT as
1
Z2 Z
0 0
K T (U h ) = Eb (U h ; qj,T ek ) : C : Eb (U h ; qi,T el )
12 T
NT
(6.7)
3
2 00
+T : Eb (U h ; qj,T ek , qi,T el ) dS 0 hd ,
i,j=1 k,l=1
NT 3
bT = f (qi,T el ) a(U h , qi,T el ) .
i=1 l=1
3NT is the number of the local degrees of freedom in the element T . After calculating
the element stiffness matrix and the element right hand side in every element T Th
the stiffness matrix K(U h ) and the right hand side b are assembled in the common
way.
After calculating the solution on the actual mesh, a mesh refinement may possibly
improve the approximation. Then, the interpolated solution of the coarser mesh is
the starting solution for the calculation on the refined triangulation.
Due to the problem (5.12) with functions in H2 , the conformal FE-approach requires
C 1 -continuity of the ansatz functions pj as proved in [5].
For axis-parallel rectangular elements, the Bogner-Fox-Schmidt elements can be
used and are considered. For a more general shape of the plate some generalizations
using triangular meshes are known, such as the (reduced) Hsiegh-Clough-Tocher
element or similar.
As written in the section above, we use ansatz functions with a local support.
ab
For that purpose, in the BFS-element T 16 bi-cubic form functions qk,T , k =
1, . . . 4, a, b = 0, 1 have to be defined. To avoid the definition of these form func-
tions on every finite element we consider a reference element Tb [1, 1]2 R2 , also
43
6 Newtons method and implementation
1
!
1 h 0
1
x= = x
b + x0 , (6.8)
2
2
0 h2
!
0
:= x4 x1 = x3 x2 ,
h2
and
1
x0 = (x1 + x2 + x3 + x4 )
4
is the element mid point.
As proved in [5], for a C 1 -continuous FE-approach we need bi-cubic form functions
in the master element Tb . For this purpose, we consider the four cubic functions paL
and paR (a = 0, 1), displayed in Figure 6.1.
16
0.8
p0L (s) p0R (s)
0.6
0.4
p1L (s)
0.2
0 -
1 p1R (s) 1 s
0.2
44
6.2 Finite Element Method (FEM)
45
6 Newtons method and implementation
Hence,
c d qkab (xl ) = kl ca db
uab a b
k = (1 ) (2 ) U h (xk ), k = 1, . . . , n; a, b = 0, 1, (6.13)
with uab ab ab ab
1 , u2 , u3 and u4 the nodal values of T after resorting their indices. Hence,
for the BFS-element we have 12 degrees of freedom per node and globally
3N = 3 4 n = 12n.
Let us now consider the approximation of U . Then, with the same rearrangement
(6.13) of the appropriate expansion coefficients du R3N on each element we have
4 X
1
duab ab
X
U h (x)|T = k qk (x).
k=1 a,b=0
The form functions are not polynomials (as usual) in the whole element T , but
polynomial within three sub-triangles. Hence, we subdivide T into three sub-
triangles Tk . Therefore, an arbitrary internal point x0 T has to be connected
46
6.3 Efficient solution of the linear systems of equation
qia , a = 0, 1, 2,
associated with the vertex xi , which are cubic inside of each sub-triangles fulfilling
the following properties:
q a is linear along each edge of T , with n the outer normal vector of the
n i
considered edge.
As result we get nine nodal base functions qia . For further information we refer to
[22].
47
6 Newtons method and implementation
for a fixed level l of the mesh refinement, are linear combinations of the next finer
functions
10 01 11 f iner
(p00
i , pi , pi , pi )
for nodes xi around the coarse-mesh node xk of the next level (l + 1).
Based on this refinement formula a BPX-preconditioner [6] has been investigated
in [17],[27] and [28] for the case of a linear Koiter-shell equation. All the details of
the refinement and the resulting base transformation are found in [27].
Here, we have used a slightly simpler approach, a hierarchical basis preconditioner
due to [29], which is a simple matrix multiply C 1 = Q J Q> , where Q is three times
the (N N )-matrix of basis transformation of the hierarchical basis of the finest
FE-space into the given nodal basis. (This matrix multiplication is extremely cheep
from using the refinement formula above.)
The diagonal matrix J contains important scaling factors, which are different for
the 4 functions pab
k (a, b = 0, 1) on each node xk . These factors can be obtained dur-
ing the mesh refinement from the diagonals of the actual stiffness matrices together
with the nodal hierarchy used in the Q-multiply.
So, these ingredients would remain valid when our uniform mesh refinement is
replaced by an adaptive regime.
6.4 Implementation
In this section we want to give an overview of the ideas for the realization of the the-
oretical aspects from the last chapters. The basic principles of the implementation
for computing the element stiffness matrix will follow.
For simplicity we restrict ourselves to BFS-elements. For the reduced HCT-
element the implementation is nearly the same, but the number of form functions
and the dimension of the element stiffness matrix differs. The mathematical pro-
cedures are equal.
For the implementation of the element stiffness matrix K T from (6.7) the most
0
interesting second-order tensors are Eb (U ; V ) in the first part and E(U
b ; U , V ) in
the second part of the integral. Note, that these tensors are of rank 2. There-
fore, their Cartesian coefficients are displayed as the (2 2)-matrices e0 (U , V ) and
e00 (U ; U , V ), respectively. These matrices are obtained by applying the chain rule
to (4.23):
2
e0 (U ; V ) = ij
i,j=1
= a0 (U ; V ) b0 (U ; V ) a1 (U ) a(U ) h b(U ) (I)
(6.14)
a(U ) h b(U ) a1 (U )a0 (U ; V )a1 (U ) a(U ) h b(U ) (II)
+ a(U ) h b(U ) a1 (U ) a0 (U ; V ) b0 (U ; V ) . (III)
48
6.4 Implementation
Obviously, the matrix of the first derivative of the strain tensor is a result of multi-
plications of the matrix of the first fundamental forms a(U ) from (4.10), the matrix
of the second fundamental forms b(U ) from (4.11) as well as the matrix of their lin-
earization a0 (U ; V ) and b0 (U ; V ), respectively. These linearizations are obtained by
utilizing the chain rule to (4.10) and (4.11), respectively. For the first fundamental
forms we achieve
a0ij (U ; V ) = ai (U ) V ,j + aj (U ) V ,i , i, j = 1, 2. (6.15)
and obtain
0
b0ij (U ; V ) = D0 (U ; V )ebij (U ) + D(U )ebij (U ; V ) (i, j = 1, 2) (6.16)
with
0
b ij (U ; V ) = [a1 , V ,2 , U ,ij ] + [V ,1 , a2 , U ,ij ] + [a1 , a2 , V ,ij ], i, j = 1, 2
e
For a simple implementation and due to the fact that we have symmetric (2 2)-
matrices only, we introduce Voigts notation. With that a (2 2)-matrix is written
as a vector with three components. In that way, the vector notation of the matrix
of the first derivative of the strain tensor (6.14) is
>
E0 (U ; V ) = 11 (U ; V ), 22 (U ; V ), 212 (U ; V ) .
With these linear operators, applied to from (6.12), the first derivative of the first
fundamental forms can be computed in Voigts notation by
A0 (U ; V ) = DA ()V + DA (U , )V .
49
6 Newtons method and implementation
which are (3 3)-matrices, too. Then, using the abbreviations from (6.12), the
0
vector notation of eb (U ; V ) is computable as
0
e (U ; V ) = D ()V + D (U , )V D (U , )V .
B B B1 2 B2 1
Hence, the vector notation of the first derivative of the strain tensor E0 (U ; V ) is a
linear combination of the components of the first derivative of the first fundamental
forms in vector notation A0 (U ; V ) and the one of the second fundamental forms
B0 (U ; V ), following the matrix multiplication rules, using A(U ) as the vector nota-
tion of the first fundamental forms and B(U ) the one of the second fundamental
forms, respectively.
In that way, the first part of the integral in (6.7)
Z
0
Eb (U ; V 1 ) : C(U ) : E(U
b ; V 2 ) dV
0
is the same as
Z
E0 (U ; V 1 )> C E0 (U ; V 2 ) dV (6.19)
0
D(U , )V = E0 (U , V )
the derivation operator D(U , ) defines all the linear combinations of the derivation
operators (6.17) and (6.18) for calculating E 0 (U , ), following the matrix calculus.
Then, there exists the (3 48)-matrix
M = D(U , )(Q I)
50
6.4 Implementation
with denoting the Kronecker product. With this matrix M above, the first part
of the element stiffness matrix (6.19), written in Voigts notation, yields
Z
M > C(U ) M dV.
0
As written before, for the second part of equation (6.7) for the element stiffness
matrix K T the matrix of the second linearization of the strain tensor is needed. The
reapplication of the chain rule to (6.14) yields
e00 (U ; U , V ) = a00 (U ; U , V ) b00 (U ; U , V ) a1 (U ) a(U ) h b(U )
a0 (U ; V ) b0 (U ; V ) a1 (U )a0 (U ; U )a1 (U ) a(U ) h b(U ) (I)
+ a0 (U ; V ) b0 (U ; V ) a1 (U ) a0 (U ; U ) b0 (U ; U )
a0 (U ; U ) b0 (U ; U ) a1 (U )a0 (U ; V )a1 (U ) a(U ) h b(U )
+ a(U ) h b(U ) a1 (U )a0 (U ; U )a1 (U )a0 (U ; V )a1 (U ) a(U ) h b(U )
a(U ) h b(U ) a1 (U )a00 (U ; U , V )a1 (U ) a(U ) h b(U ) (II)
+ a(U ) h b(U ) a1 (U )a0 (U ; V )a1 (U )a0 (U ; U )a1 (U ) a(U ) h b(U )
a(U ) h b(U ) a1 (U )a0 (U ; V )a1 (U ) a0 (U ; U ) b0 (U ; U )
+ a0 (U ; U ) b0 (U ; U ) a1 (U ) a0 (U ; V ) b0 (U ; V )
a(U ) h b(U ) a1 (U )a0 (U ; U )a1 (U ) a0 (U ; V ) b0 (U ; V ) (III)
+ a(U ) h b(U ) a1 (U ) a00 (U ; U , V ) b00 (U ; U , V ) .
Obviously, the matrix of the second linearization of the strain tensor is a result of
multiple matrix multiplications of the matrices of the first and second fundamental
forms from (4.10) and (4.11), respectively, with their first linearizations from (6.15)
and (6.16) as well as their second derivatives. These second derivatives we obtain
by applying the chain rule to their first derivatives, respectively. Hence,
a00ij (U ; U , V ) = U ,i V ,j + U ,j V ,i (i, j = 1, 2)
and
0
b00 (U ; U , V ) = D00 (U ; U , V )eb(U ) + D0 (U ; V )eb (U ; U )
0 00
+ D0 (U ; U )eb (U ; V ) + D(U )eb (U ; U , V )
with
00
b (U ; U , V
e ) = [V ,1 , U ,2 , U ,ij ] [V ,2 , U ,1 , U ,ij ] + [V ,2 , U ,ij , a1 ]
51
6 Newtons method and implementation
2
[V ,ij , U ,2 , a1 ] + [V ,ij , U ,1 , a2 ] [V ,1 , U ,ij , a2 ] .
i,j=1
i,j=1
2 2 2 2
with T = T (U ) = T ij the coefficients of the second Piola-Kirchhoff stress
i,j=1
2
tensor as well as e00 (U ; U , V ) = e00ij (U ; U , V ) .
i,j=1
For calculating the element stiffness matrix the vectors V and U now pass
through all 48 vector valued form functions of the BFS element. Therefore, every
component Hij = e00ij (U ; Q I, Q I) is understood as a (48 48)-matrix, itself.
Hence, the element stiffness matrix (6.7) is
Z Z 2 2
>
X
K T (U ) = M C(U ) M dV + T ij H ji dV. (6.20)
0 0 i,j=1
The integration over the undeformed plate 0 in (6.20) is realized in two steps.
First, the integration over the undeformed mid surface m 0 is accomplished by a
Gaussian integration with 4 4 points and weights. For the integration over the
thickness numerical experiments have shown that we achieve good results in a ac-
ceptable time for the Gaussian integration with 7 points and weights.
For each Gaussian point the matrix multiplications above, have to be done. Hence,
in the realization of the theory we reduce the calculations to the Kronecker products,
which are needed. After that, we multiply them with all scalars, which result from
the matrix multiplications. In that way, the performing time for the assembly of
the element stiffness matrices has been reduced to less than a fourth in contrast to
apply the full matrix multiplications in (6.20).
52
6.5 Newtons method and mesh refinement
(2) After the forces are fully applied to the considered device on a relatively coarse
mesh, we may start mesh refinements to obtain an appropriate approximation
of displacements, stresses or strains. Then, in every refinement stage we apply
Newtons Algorithm 6.2 on the actual finite element mesh. Hence, on a mesh
of level l we achieve an approximated vector function U h . After the next
refinement is performed, on the finer mesh with the level (l + 1) we need an
initial vector function U h/2 (x) for the continuation of Newtons method. For
that, the vector function U h (x) is interpolated with the ansatz functions of
the finer mesh and we claim pointwise
at all nodes xi of the finer mesh. Now, on this finer mesh Newtons Algorithm
6.2 starts with the interpolated
vector function U h/2 (x). Note, that the stiff-
ness matrix K U h/2 (x) now is of a higher dimension than it was on the
coarser mesh of level l, due to the increased number of elements.
For BFS-elements we achieve nested function spaces for the ansatz functions,
V1 V2 . . . Vl V(l+1) .
Therefore, the ansatz functions in the mesh of level l can be displayed as linear
combinations of the ansatz functions of the next finer mesh. For that purpose
the demand (6.21) is not only fulfilled pointwise at xi , but the vector functions
are obviously identical.
Unfortunately, the mesh refinement does not generate nested function spaces
in the case of HCT-elements. That means
Vl 6 V(l+1) . (6.22)
Here, the vector function U h/2 (x) defined by (6.21) in the finer mesh differs
from the vector function U h (x) in the mesh of level l. Therefore, the interpo-
lated vector function U h/2 (x) as a starting vector function in Algorithm 6.2 is
not as good as in the case of BFS-elements. Consistently, Newtons method
needs some more steps in the HCT-case.
Additionally from using shape regular triangles, the HCT-elements are more
flexible and may be used for approximating curved boundaries. In this case,
after the refinement the new boundary nodes are corrected to the true boun-
dary position (another reason for (6.22) and for more complications with New-
tons method).
There are some approaches in [13] or [26] for generalizing HCT-elements in a
way that nested function spaces for the ansatz functions arise. But for these
elements more degrees of freedom have to be regarded.
53
7 Numerical examples
In this chapter we present some numerical examples. For that, we consider plates
in different forms and sizes. The goal of the theory, which was presented in the
last chapters, is the calculation of large deformations for thin plates. Hence, for all
devices a thickness is chosen, which is one eightieth of the half of the sum of its
width and its length.
Furthermore, we use the Neo-Hooke material law for all plates. The density
function of this material was already defined in (2.34).
The examples are grouped by their kind of plate deformation as membrane-
dominated and bending-dominated examples. For the deflection of a plate we apply
a volume force, perpendicular to the mid surface. The second group consists of
examples with bending-dominated plate deformations, where one or more edges will
follow a prescribed movement.
The procedure of the simulation in nearly all examples is the same. First of all the
forces are applied to the plate incrementally. For each increment a Newton iteration
has to be done. The incremental stage is followed by a number of mesh refinements
to improve the approximation.
For all examples we present some pictures in different approximation stages of
the plates mid surface. With the exception of the pictures of the initial states, the
surfaces are coloured by the Frobenius norm of the mid surface strain tensor or by
one component of the displacement vector. The colour bar is adapted particularly to
each example. For this postprocessing the open source software ParaView, version
3.98.0 nightly, was used.
A word of caution on the graphical output: Note, that for the display
of the deformed mid surface we have chosen the following simplification.
Even though the obtained solutions are piecewise (bi)cubic functions,
the figures only display their linear interpolation with the nodal values
(additionally calculated nodal derivatives are ignored). Especially for
coarser meshes the reader should keep in mind this simplification for the
graphics only the calculated functions are correctly C1 -continuous!
55
7 Numerical examples
1-axis
56
7.1 Plate deflection
Newton iteration is determined to = 108 . For this the Newton iteration ends,
if the calculated correction term U is smaller than 108 of the actual computed
displacement vector. Figure 7.2 shows the deformed mid surface after one Newton
iteration in the first increment.
strain
1.3045
1.2
0.8
0.4
2 -axis
1-axis
Figure 7.2: Deformed mid surface after applying the first increment, 16 elements.
Expectedly, the values in Figure 7.2 of each element are still small and do not
differ very much over the different elements.
strain
1.3045
1.2
0.8
0.4
2-axis
1 -axis
Figure 7.3: Deformed mid surface after applying the full force vector, 16 elements.
57
7 Numerical examples
Note that for the graphical outputs a constant colour bar for the values of the
norm of the strain tensor is defined to make the colouring for all pictures of one
example comparable.
Following the algorithm, the increment would be increased until the full force
vector is applied. The deformed mid surface, approximated in the coarse mesh, is
illustrated in Figure 7.3. As can be seen, the interpolated value of the strain in the
elements in the center of the surface increased a little bit but the resolution of the
mid surface is far to coarse for a good approximation.
The proceeding before, was necessary to execute the full force to the plate. For the
improvement of the approximation we adopt five mesh refinements. The following
Table 7.1 shows the number of Newton steps per iteration in the single stages.
..................
Table 7.1: Example 7.1.1; number of Newton steps at the single periods.
The higher number of Newton steps in the first iteration is easily explained by
the used starting vector U 0 = 0. Obviously, it is not as close to the solution in the
first increment as later the approximated values from the Newton iteration of the
incremental step before, which are used as initial solution for the Newton iteration
of the next increment. The number of Newton steps in the single refinement stages
decreases, due to the fact, that the error for the approximation of the displacement
vector in the finer mesh is smaller than in the coarser triangulation.
58
7.1 Plate deflection
Note, that the number of Newton steps in each iteration is small, by reason of the
fact that we get nested function spaces for BFS-elements.
strain
1.3045
1.2
0.8
0.4
0
3-axis
1 -axis
Figure 7.4: Deformed mid surface, 16.384 elements, full force applied.
strain
1.3045
1.2
0.8
0.4
2 -axis
1-axis
Figure 7.5: Deformed mid surface, 16.384 elements, full force applied, birds eye
view.
At this point, let us consider the deformed mid surface after five refinements,
approximated by 16.384 elements. As seen in Figure 7.4 and Figure 7.5 the highest
59
7 Numerical examples
strain occurs in the center of the surface boundaries. Due to the symmetry of the
example, the side view in Figure 7.4 looks identically for all four sides.
Apart from the performance test of Newtons method, a second thought for con-
sidering this example above is the verification of our code. For that, we consider the
third component of the displacement vector in Figure 7.6.
U^3
1.1756001
1
0.75
0.5
0.25
2 -axis
1 -axis
Figure 7.6: Deformed mid surface, 16.384 elements, third component of displace-
ment vector U (3) .
Here, a very small boundary layer is observable as it is typical for large deflection
examples.
In contrast to that let us consider a small deformation example. For that purpose
we apply a very small force vector to the plate. Due to that, we do not need to
apply the forces incrementally. Figure 7.7 shows the resulting deformed mid surface
after five refinements, which looks as similar as in the small deformation case.
As expected, in Figure 7.7 the values of the strain tensor as well as for the dis-
placement vector are very small compared to the appliance of large deformations
(Figure 7.4).
Another difference towards the results of large deformation are the boundary
layers. So, in Figure 7.4 they have a steeper slope than in Figure 7.7. This fact is
also seen directly in Figure 7.6 in comparison with Figure 7.8, which both display the
third component of the displacement vector in the case of large and of small strain,
respectively. In Figure 7.6 the coloured ranges get finer towards the boundary. On
the contrary, in Figure 7.8 the coloured ranges are thicker at the boundary, in order
to get finer in direction to the center of the plate, before they grow wider near this
center again.
60
7.1 Plate deflection
strain
6.107e-5
6e-5
4e-5
2e-5
0
3 -axis
1-axis
U^3
0.0028221
0.002
0.001
2 -axis
1 -axis
Figure 7.8: Deformed mid surface, 16.384 elements, small strain, third component
of the displacement vector U (3) .
Last, we consider the first and the second component of the displacement vector
for small deformation, seen in Figure 7.9 and 7.10, respectively. In addition to the
symmetry of the displacements in the corresponding direction 1 or 2 , the values
are very close to zero.
In contrast to this results U 1 and U 2 are obviously zero in the small strain theory.
61
7 Numerical examples
With the use of our code, the first and the second component of the displacement
vector differ from zero, because of the non-linear parts, which are not discarded as
in the small deformation theory.
U^1
1.0382e-5
8e-6
4e-6
0
-4e-6
-8e-6
2 -axis
-1.038e-5
1 -axis
Figure 7.9: Deformed mid surface, 16.384 elements, small strain, first component of
the displacement vector U (1) .
U^2
1.0382e-5
8e-6
4e-6
0
-4e-6
-8e-6
2 -axis
-1.038e-5
1 -axis
Figure 7.10: Deformed mid surface, 16.384 elements, small strain, second compo-
nent of the displacement vector U (2) .
62
7.1 Plate deflection
strain
0.43587
0.1
0.01
2-axis
0.005
1 -axis
Figure 7.11: Deformed mid surface after applying first increment, 24 elements,
birds eye view.
The deformed mid surface after the first Newton iteration is shown in Figure 7.11,
displayed with 24 elements. The values for the strain norm are very small. As before
in the corresponding mesh of the quadrilateral plate, the mesh is much to coarse for
a well suited approximation.
For comparison, Figure 7.12 depicts the view heading to the ( 1 , 3 ) - plane of
the deformed mid surface in the same mesh as seen in Figure 7.11. In that way, the
third component of the displacement vector U (3) is shown at the same time.
Following the algorithm, at this point the increments are increased and for every
increment a Newton iteration is computed. After 10 increments the full force vector
is now applied and Figure 7.13 displays the deformed mid surface in this stage.
For approximating the displacement vector good enough, the mesh is refined four
times, uniformly.
63
7 Numerical examples
strain
0.43587
0.1
0.01
0.005
3 -axis
1-axis
Figure 7.12: Deformed mid surface after applying the first increment, 24 elements.
strain
0.43587
0.1
0.01
0.005
3 -axis
1-axis
Figure 7.13: Deformed mid surface after applying the full force vector, 24 elements.
The deformed mid surface after all refinement steps is shown in Figure 7.14,
above. It consists of 6.144 elements. Just like in the example for the BFS-elements
the boundary layers are small, as it is typical for this kind of large deformations.
Figure 7.15 displays the birds eye view of the deformed mid surface in the finest
triangulation. In both figures we can see that the maximum of the norm of the
strain tensor is approximated near the boundary.
64
7.1 Plate deflection
strain
0.43587
0.1
0.01
0.005
3 -axis
1-axis
Figure 7.14: Deformed mid surface, 6.144 elements, full force applied.
strain
0.43587
0.1
0.01
2-axis
0.005
1 -axis
Figure 7.15: Deformed mid surface, 6.144 elements, full force applied, birds eye
view.
In Table 7.2 the number of Newton steps per iteration is displayed in the respective
situation.
65
7 Numerical examples
.................
Table 7.2: Example 7.1.2; number of Newton steps at the single periods
The number of Newton steps per iteration in the single increment stages is simi-
larly small as in the example before. However, we observe a slightly increase of the
numbers in the individual refinement stages. This is due to the refinement technique
of the HCT-elements. Furthermore, we have a curved boundary. Nevertheless, the
algorithm converges, due to the fact that the boundary is hard clamped and only
the lateral adjustment has to be put to the mid point of each outer edge.
66
7.2 Bending-dominated deformation
2-axis
1-axis
As mentioned before, for this group of examples we do not apply any force vector
to the plate but simulate an incremental transport of a part of the boundary.
Here, both boundaries in 2 -direction at 1 = 0 and 1 = 1.5, respectively, are hard
clamped with inhomogeneous Dirichlet type boundary conditions. The remaining
edges are left free (do nothing condition).
For the left edge (0, 2 , 0)T we apply a prescribed displacement
h i
respectively, with t 0, 2 . The outer normals of the edges will be rotated through-
out their transport and are defined by
respectively.
The edge displacements are applied in 10 equidistant steps of t. The result of the
first increment is displayed by Figure 7.17.
67
7 Numerical examples
strain
0.051177
0.04
0.02
3 -axis
2 xi
s
1-axis -a
Figure 7.17: Deformed mid surface after applying the first increment, 8 elements.
strain
0.051177
0.04
0.02
3 -axis
2 xi
s
1-axis -a
At that, as usual for the algorithm, we start the mesh refinement. After every
refinement a Newton iteration is carried out.
68
7.2 Bending-dominated deformation
For the full simulation progress the number of Newton steps per Newton iteration
in every incremental stage and in the refinement states are displayed in Table 7.3.
..................
Table 7.3: Example 7.2.1.1; number of Newton steps in the single periods.
Obviously, in contrast to the example in Section 7.1.1 the iteration numbers in-
creases a little bit for each individual incremental stage. This may depend on the
fact that for bending deformations the change of the displacement vector and of its
partial derivatives from one to the next incremental state is much higher than in
the deflection examples.
In the refinement stage the number of Newton steps per iteration are expectedly
small, due to the nested function spaces of the ansatz functions for the refined
triangulations.
As result of the six refinements we achieve a cylindrical shell. Its mid surface is
shown in Figure 7.19, which consists of 32.768 elements. Furthermore, the strain is
69
7 Numerical examples
strain
0.051177
0.04
0.02
3 -axis
2 xi
s
1-axis -a
For a graphical depiction of the approximation process the graphical output for
each incrementation and the one for the finest mesh is displayed in the following
figure (Figure 7.20).
strain
0.051177
0.04
0.02
3 -axis
2 xi
s
1-axis -a
70
7.2 Bending-dominated deformation
In the second example we consider the same plate as in example 7.2.1.1, before, with
equal boundary conditions. The transport of the two outer edges in 2 -direction
follows the same prescribed path, as given in (7.1) and (7.2).
In contrast to example 7.2.1.1 we apply a further rotation of the outer normals of
both edges, which now is defined by
strain
0.058776
0.04
0.02
3 -axis
2 x
is
1-axis -a
After applying further nine increments, the deformed mid surface looks like in
Figure 7.22, displayed still with a mesh consisting of eight elements.
71
7 Numerical examples
strain
0.058776
0.04
0.02
3 -axis
2 x
is
1-axis -a
At this point of the simulation, the complete edge transport and the full edge
rotation is computed. For a useful approximation of the deformation the initial
mesh for the mid surface has to be refined. As in the example before, six refinement
steps are carried out. The graphical result can be seen beneath in Figure 7.23.
strain
0.058776
0.04
0.02
3 -axis
2 x
is
1-axis -a
Here, we have plotted the mid surface with 32.768 elements. The mid-surface
strain is slightly higher than in the example before, due to the further rotation
72
7.2 Bending-dominated deformation
of the edge normals. But in contrast to the corresponding value in the deflection
example 7.1.1 the coloured value here is much smaller.
In the following Table 7.4 the number of Newton steps per iteration are displayed.
..................
Table 7.4: Example 7.2.1.2; number of Newton steps in the single periods.
The number of Newton steps in each single incremental situation does not differ
very much from the corresponding number of steps in example 7.2.1.1. Obviously,
the further edge rotation has only a small influence on the numbers of Newton
steps for these stages. By comparison to the example before, the numbers in the
refinement stages, respectively, increase a little bit, in particular for the first two
refinement states. This shows that the approximation of the displacement vector and
its partial derivatives in the coarser meshes is still far away from the approximations
73
7 Numerical examples
For the next example we consider a new, longer plate with a mid surface
m0 = [0, 64][2, 2] and a thickness of 0.425. The boundary conditions are compar-
able to the one in Example 7.2.1.1 above. In contrast to that example, additionally
to the boundary edge transport and the boundary edge normal rotation, a second
rotation with an angle () around the 1 -axis is defined to both boundary edges
of the mid surface in 2 -direction.
Therefore, a further inhomogeneous Dirichlet boundary condition arises.
The transport and the rotation of the two edges as well as the rotation of the outer
normals are simulated at once. Due to that, we have twice as much increments as
in all examples before, applying the full prescribed displacement vectors to the two
boundaries. Using 20 increments, for the mid surface of the plate after the first
Newton iteration we get the slightly deformed mid surface.
To guarantee the convergence of all Newton iterations during the computing pro-
cess, we apply two mesh refinements. So, in Figure 7.24 the slightly deformed mid
surface approximated with 256 elements is displayed.
strain
0.043872
0.04
0.03
0.02
0.01
2-axis
1-axis
After applying all increments we achieve the fully deformed mid surface, approx-
imated with 256 elements in Figure 7.25.
74
7.2 Bending-dominated deformation
strain
0.043872
0.04
0.03
0.02
0.01
2-axis
1-axis
Figure 7.25: Deformed mid surface, 256 elements, full force applied.
Subsequently, three further refinements are carried out. In Figure 7.26 we can see
that the computed mid-surface strain is very small over the whole plate.
strain
0.043872
0.04
0.03
0.02
0.01
2-axis
1-axis
Although we apply the transport as well as both rotation angles for the edges
in 2 -direction at the same time, it seems that the simulation does not need more
Newton steps per Newton iteration prior reaching the stopping criterion as in the
examples before. Certainly, the smaller increments have a positive influence on the
75
7 Numerical examples
number of Newton steps per Newton iteration. Because we have chosen a mesh,
which was twice refined before the incrementation takes place, we have a very fast
convergence in the refinement stage as can also be seen in the following Table 7.5.
..................
Table 7.5: Example 7.2.1.3; number of Newton steps in the single periods.
At the end of this example we show the deformation progress in Figure 7.27, in
that the approximated mid surface in all increment stages as well as the most refined
mid surface is displayed.
76
7.2 Bending-dominated deformation
strain
0.043872
0.04
0.03
0.02
0.01
2-axis
1-axis
Figure 7.27: M
obiusstrip, deformation progress.
This example considers the same initial plate as it was considered in Example 7.2.1.3,
before. But we change the boundary conditions. For the left edge of the mid surface
the values for the boundary displacement vector and its derivatives are fixed to zero,
except for its first component U (1) , which is left free.
strain
0.11832
3-axis
0.1
0.075
0.05
0.025
1-a
xis
s
xi
-a
2
77
7 Numerical examples
For the right boundary edge (64, 2 , 0) of the mid surface we apply inhomogeneous
Dirichlet boundary conditions so that a rotation around the 1 -axis until 4 is carried
out.
This rotation is applied with 100 increments, which is much more than in all
examples, above. The deformed mid surface approximated with again 256 elements
in the first increment is displayed in Figure 7.28.
strain
0.11832
3-axis
0.1
0.075
0.05
0.025
1-a
xis
s
xi
-a
2
Figure 7.29: Deformed mid surface, 256 elements, full rotation of 4 applied.
strain
0.11832
3-axis
0.1
0.075
0.05
0.025
1-a
xis
s
xi
-a
2
78
7.2 Bending-dominated deformation
After the application of all increments the fully deformed mid surface, consisting
of 256 elements, can be seen in Figure 7.29.
After further three refinements the deformed mid surface now is approximated
with 16.384 elements. It is to be seen in Figure 7.30.
As before, we observe the norm of the strain tensor being small over the whole
plate. However, the maximum of the strain tensor is nearly twice the corresponding
value in the other bending-dominated examples. The largest strain is generated at
the free boundary edges in 1 -direction. Expectedly, the left edge in 2 -direction is
shifted around 1.5 in direction of the 1 -axis.
..................
..................
..................
Table 7.6: Example 7.2.1.4; number of Newton steps in the single periods.
79
7 Numerical examples
1 -axis
The mid surface of the plate is cut along the edge from (0, 0.5, 0)> to (0, 1, 0)> ,
yielding two edges with inhomogeneous Dirichlet type boundary conditions. Only
the outer normals of these both boundaries are rotated throughout the simulation
by
>
U ,1 = (1 cos t), 0, sin t (7.3)
h i
with the incremental parameter t 0, 2 . The full rotation angle is applied in 10
incremental steps. The remaining edges are left free.
After applying the first increment, we get a slightly deformed mid surface, in that
the outer normals of the emerged boundary edges have rotated a little bit. Due to
that, the outer boundary is bending upwards. We have applied one mesh refinement
in order to secure the convergence of Newtons method during the incremental stage.
The mid-surface strain is very small in this stage, which can be seen in Figure 7.32.
80
7.2 Bending-dominated deformation
strain
0.026407
s
axi
0.01
1- 0.0026407
is
3 -ax
2 -axis
Now we have to increase the increment until the full rotation angle is applied. In
that way, we get a heavily deformed mid surface, which is displayed in Figure 7.33.
strain
0.026407
s
axi
0.01
1-
0.0026407
is
3 -ax
2 -axis
After three consecutive refinements we obtain a subdivided mid surface with 3.072
elements. As seen in Figure 7.34 the norm of the strain tensor is bounded by about
0.02.
81
7 Numerical examples
strain
0.026407
s
axi
0.01
1-
0.0026407
is
3 -ax
2 -axis
Figure 7.34: Deformed mid surface, 3.072 elements, full force applied.
Another refinement step and a further solved Newton iteration yields the deformed
mid surface, displayed with 12.288 elements in Figure 7.35.
strain
0.026407
s
axi
0.01
1-
0.0026407
is
3 -ax
2 -axis
Figure 7.35: Deformed mid surface, 12.288 elements, full force applied.
Here, the maximum of the norm of the strain tensor is about 0.0264. This is much
less than in the deflection example in Section 7.1.2.
For the approximation progress the following numbers of Newton steps were
needed for the Newton iteration in every incremental and every refinement stage.
82
7.2 Bending-dominated deformation
Table 7.7: Example 7.2.2.1; number of Newton steps in the single periods.
For this example, the number of Newton steps is similar to the numbers in the
second example for bending problems using BFS-elements, Example 7.2.1.2, al-
though the rotation angle here is only one third the rotation angle in the named
example, respectively.
83
7 Numerical examples
For the last example we consider an annulus. This is similar to the plate in Example
7.2.2.1, above, but with a curved inner and outer boundary. The diameter is 2 again
and the annulus is cut at the same edge, as the octagonal plate, before. The two
new edges have inhomogeneous Dirichlet boundary conditions. All other edges are
left free.
The undeformed mid surface looks alike the respective one in the example above,
displayed in Figure 7.31.
The outer normals of the two arose straight edges are rotated in the same way as
in the example before, applying equation (7.3).
For this rotation a serious problem occurs. The incrementation was successfully
done for 80 or more increments. Unfortunately, after the full rotation angle was
applied, the Newton iteration fails for the first refinement, due to both, the special
refinement technique and the curved boundary.
A way out was changing the procedural method. So, first we apply four refine-
ments in the first incremental stage. After that, we get a slightly deformed mid
surface approximated with 3.072 elements, displayed in Figure 7.36.
strain
0.014533
s
axi
0.01
3-
0.0014533
1
-a
xi
s
2-axis
From the same reason as discussed in section 6.5, here the incrementation starts
at a more refined mesh than before. The result can be seen in Figure 7.37.
84
7.2 Bending-dominated deformation
strain
0.014533
s
axi
0.01
3-
0.0014533
1
-a
xi
s
2-axis
Figure 7.37: Deformed mid surface, 3.072 elements, full force applied.
The maximum of the norm of the strain tensor is about 0.0145. This is a little
bit shorter than in Example 7.2.2.1 using 12.288 elements (Figure 7.35). It appears
in the same region as in the example, above. Unfortunately, the computing time is
much higher than in the example before, due to the large number of increments and
the change of the sequence of incrementation and refinement.
strain
0.014533
s
axi
0.01
3-
0.0014533
1
-a
xi
s
2-axis
In Figure 7.38 the mid surface is displayed in the undeformed state using 12
85
7 Numerical examples
elements and in the first incremental stage as well as in the refined period with
3.072 elements, respectively, for the imagination of the deformation progress.
The number of Newton steps for each stage of the simulation is shown in the
following Table 7.8.
.................
.................
.................
Table 7.8: Example 7.2.2.2; number of Newton steps in the single periods.
Due to the problems discussed in section 6.5, here it was necessary to perform more
mesh refinements in the beginning, before the incrementation starts. Unfortunately,
this enlarges the total operating expense drastically. Nevertheless, the number of
Newton steps is not much higher for one Newton iteration as in the examples before.
86
7.2 Bending-dominated deformation
Contrary, the HCT-elements are more flexible and can also be used for curved
boundaries. But due to the discussion in section 6.5 there occur other problems to
overcome. Because of the many different demands to a finite element, one would
never find an element, which is the best for all the examples, at the same time.
87
8 Outlook and open questions
The main focus of this work is on the numerical simulation of large deformations
of thin plates. Therefore, we consider a non-linear plate theory with a non-linear
material law. Additional simplifications have been set aside. The displacement
vector function is derived from a non-linear system of equations, which emerges from
the equilibrium of forces. The reduction of dimension to the mid surface requires
a plate assumption. Here, we use the Kirchhoff hypothesis, where no change in
thickness direction is allowed during the deformation process. The basic principle for
the simulation is the minimization of the energy functional over the deformed plate.
This implies that the first derivative of this energy functional vanishes. For solving
this non-linear system of equations Newtons method is applied, which necessitates
the calculation of the second derivative of the energy functional. Its calculation and
the efficient use in the F.E.-simulation was one of the most difficult parts within
this work. As a result of each Newton iteration we obtain a correction term for the
displacement vector function. Some examples for plates of different sizes and for
different kinds of deformation are presented.
One starting point for additional research can be the weakening of the Kirchhoff
hypothesis. Therefore let us consider the deformed plate as
x() = Y ( 1 , 2 ) + U + h d
Another research issue is the development of an adaptive approach for the FEM.
For that an error estimator is needed. In [20] a basis for an error functional in the
linear and in the non-linear case is provided for the full 3D theory. At present, the
development of an error estimator for shells and plates and linear problems is in
progress. In future, potential results have to be transferred to the non-linear case.
89
Bibliography
[1] S. S. Antman. Nonlinear Problems of Elasticity. Springer, Berlin, Heidelberg,
New York, 2nd edition, 2005.
[2] Y. Basar and W. Kratzig. A Consistant Shell Theory for Finite Deformations.
Acta Mechanica, 76:7387, 1989.
[3] M. Bischoff, W. A. Wall, K.-U. Bletzinger, and E. Ramm. Models and Finite
Elements for Thin-walled Structures. Encyclopedia of Computational Mechan-
ics, 2:59137, 2004.
[5] D. Braess. Finite Elements: Theory, Fast Solvers and Applications in Solid
Mechanics. Cambridge University Press, Cambridge, 3rd edition, 2007.
[7] D. Chapelle and K.-J. Bathe. The Finite Element Analysis of Shells - Funda-
mentals. Springer, Berlin, Heidelberg, 2003.
[10] P.G. Ciarlet. The Finite Element Method for Elliptic Problems. North-Holland,
Amsterdam, 1978.
[12] P.G. Ciarlet. Mathematical Elasticity Volume II: Theory of Plates. North-
Holland/Elsevier, Amsterdam, 1997.
91
Bibliography
[15] Ch. Gromann and H.-G. Roos. Numerical Treatment of Partial Differential
Equations. Springer, Berlin, Heidelberg, New York, 2007.
[20] A. Meyer. Error Estimators and the Adaptive Finite Element Method on Large
Strain Deformation Problems. Mathematical Methods in the Applied Sciences,
32:21482159, 2009.
[21] A. Meyer. The Koiter Shell Equation in a Coordinate Free Description. Chem-
nitz Scientific Computing Preprints, 02:14, 2012.
[23] A. Meyer and J. Ruckert. C1 -continuous FEM for Kirchhoff Plates and Large
Deformation. Journal of Applied Mathematics and Mechanics, (ZAMM), sub-
mitted, 92, 2012.
[24] A. Meyer and J. Ruckert. Kirchhoff Plates and Large Deformation. Chemnitz
Scientific Computing Preprints, 01, 2012.
[26] P. Oswald. Hierarchical Conforming Finite Element Methods for the Bihar-
monic Equation. Siam Journal Numerical Analysis, 29(6):16101625, 1992.
92
Bibliography
[27] M. Thess. Parallel Multilevel Preconditioners for Thin Smooth Shell Finite Ele-
ment Analysis. Numerical Linear Algebra with Applications, 5:401440, 1998.
[28] M. Thess. Parallel Multilevel Preconditioners for Thin Smooth Shell Problems.
PhD thesis, TU Chemnitz, Chemnitz, 1998.
93
Notation
= ( 1 , 2 , 3 ) coordinate system
i i-th coordinate
= 3 thickness coordinate
m m
0 ( ) domain of the mid surface of a undeformed (deformed)
shell/plate
U , V , U KH , U vector functions
F deformation gradient
95
Notation
material density
p outer acceleration
96
Notation
T finite element
97
Theses
Theses
(1) We introduce the 3D-theory for large deformations and describe the common
non-linear shell theory. We restrict us to the plate, a shell, which is flat in the
initial state. Both theories combined with the Kirchhoff hypothesis lead to a
strain tensor E of lower rank 2 and to the non-linear weak formulation of the
boundary value problem.
(2) The numerical solution of the non-linear boundary value problem obtained
in the weak formulation requires both a Newton-linearization and a Finite
Element-approximation. For a correct Newton iteration the second derivative
of the energy functional has to be established. Its implementation in generating
the element matrices after the F.E.-discretization is the main difficulty for
obtaining an efficient simulation.
(3) For the F.E.-discretization the BFS-element as well as the reduced HCT-
element have been considered, which both satisfy the special requirements
for a C1 -continuous treatment.
(5) This work can be used as a starting point for additional research. For example,
the use of other C1 -continuous elements or a non-conformal approach could be
investigated. Additionally, the weakening of the Kirchhoff hypothesis would
require C0 -continuous elements only, but implies other numerical difficulties.
Furthermore, an error estimator would allow adaptive mesh refinement, which
increases the total efficiency.
99
List of Figures
101
List of Figures
102
List of Tables
7.1 Example 7.1.1; number of Newton steps at the single periods. . . . . 58
7.2 Example 7.1.2; number of Newton steps at the single periods . . . . . 66
7.3 Example 7.2.1.1; number of Newton steps in the single periods. . . . . 69
7.4 Example 7.2.1.2; number of Newton steps in the single periods. . . . . 73
7.5 Example 7.2.1.3; number of Newton steps in the single periods. . . . . 76
7.6 Example 7.2.1.4; number of Newton steps in the single periods. . . . . 79
7.7 Example 7.2.2.1; number of Newton steps in the single periods. . . . . 83
7.8 Example 7.2.2.2; number of Newton steps in the single periods. . . . . 86
103